Asymptotic Methods in Fluid Mechanics by Steinruck PDF
Asymptotic Methods in Fluid Mechanics by Steinruck PDF
Asymptotic Methods in Fluid Mechanics by Steinruck PDF
&,60&2856(6$1'/(&785(6
6HULHV(GLWRUV
7KH5HFWRUV
*LXOLR0DLHU0LODQ
)UDQ]*5DPPHUVWRUIHU:LHQ
-HDQ6DOHQoRQ3DODLVHDX
7KH6HFUHWDU\*HQHUDO
%HUQKDUG6FKUHÁHU3DGXD
([HFXWLYH(GLWRU
3DROR6HUDÀQL8GLQH
7KHVHULHVSUHVHQWVOHFWXUHQRWHVPRQRJUDSKVHGLWHGZRUNVDQG
SURFHHGLQJVLQWKHÀHOGRI0HFKDQLFV(QJLQHHULQJ&RPSXWHU6FLHQFH
DQG$SSOLHG0DWKHPDWLFV
3XUSRVHRIWKHVHULHVLVWRPDNHNQRZQLQWKHLQWHUQDWLRQDOVFLHQWLÀF
DQGWHFKQLFDOFRPPXQLW\UHVXOWVREWDLQHGLQVRPHRIWKHDFWLYLWLHV
RUJDQL]HGE\&,60WKH,QWHUQDWLRQDO&HQWUHIRU0HFKDQLFDO6FLHQFHV
,17(51$7,21$/&(175()250(&+$1,&$/6&,(1&(6
&2856(6$1'/(&785(61R
$6<03727,&0(7+2'6
,1)/8,'0(&+$1,&6
6859(<$1'5(&(17$'9$1&(6
(',7(' %<
+(5%(5767(,15h&.
9,(11$81,9(56,7<2)7(&+12/2*<$8675,$
SpringerWienNewYork
7KLVYROXPHFRQWDLQVLOOXVWUDWLRQV
7KLVZRUNLVVXEMHFWWRFRS\ULJKW
$OOULJKWVDUHUHVHUYHG
ZKHWKHUWKHZKROHRUSDUWRIWKHPDWHULDOLVFRQFHUQHG
VSHFLÀFDOO\WKRVHRIWUDQVODWLRQUHSULQWLQJUHXVHRILOOXVWUDWLRQV
EURDGFDVWLQJUHSURGXFWLRQE\SKRWRFRS\LQJPDFKLQH
RUVLPLODUPHDQVDQGVWRUDJHLQGDWDEDQNV
E\&,608GLQH
3ULQWHGLQ,WDO\
63,1
$OOFRQWULEXWLRQVKDYHEHHQW\SHVHWE\WKHDXWKRUV
,6%16SULQJHU:LHQ1HZ<RUN
PREFACE
Herbert Steinrück
CONTENTS
Herbert Steinrück*
*
Vienna University of Technology, Institute of Fluid Mechanics and Heat
Transfer, Vienna, Austria
Abstract The method of matched asymptotic expansions will be
presented by applying it to three examples showing the wide appli-
cability of the method.
1 Introduction
The governing equations describing a flow field are in general a set of non-
linear partial differential equations. Only in few situation exact solutions
mostly in the form of similarity solutions exist. Thus asymptotic expan-
sions with respect to an appropriate dimensionless parameter (e.g. Reynolds
number, Mach number, thickness ratio, ...) which tends to a limiting value
(zero or infinity) are sought. Let φ(x, ε) with x ∈ D ⊂ R3 be a function of
a variable x depending on a small, positive parameter ε with 0 < ε 1.
We call
[φ](n) = δ1 (ε)φ1 (x) + δ2 (ε)φ2 (x) + · · · + δn (ε)φn (x) (1)
a n-term asymptotic series of φ with respect to ε 1 if the gauge functions
δk (ε) form an asymptotic series, i.e. δk+1 (ε) = o(δk (ε)) for k = 1, · · · , n − 1
and φ(x, ε) − [φ(n) ] = o(δn (ε)).
Note a function f (ε) is called a small ‘o’ of the function g(ε), f (ε) =
o(g(ε)) if limε→0 f (ε)/g(ε) = 0 holds, see Van Dyke (1975). The expansion
(1) is called uniformly valid if there exist constants c1 ,...,cn independent of
x with
φ(x, ε) − [φ](m)
< cm , m = 1, ..., n. (2)
δm+1 (ε)
However, the solutions of many perturbation problems in fluid mechan-
ics do not permit an approximation by an asymptotic series of type (1).
Such problems are called singularly perturbed. Most of these problems are
characterized by two different (length) scales. For example consider the at-
tached high Reynolds number flow. A regular (outer) expansion fails near a
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
2 H. Steinrück
solid surface and a local variable has to be introduced to describe the local
behavior near the wall and a local (inner) expansion of the flow field can
be found. To get a uniformly valid approximation of the entire flow region
both expansions have to match. Thus one speaks of matched asymptotic
expansion.
Boundary layers have been first introduced by Prandtl (1904) by explain-
ing the role of viscosity in large Reynolds number flows. As a mathemati-
cal tool the method of matched asymptotic expansions has been developed
systematically in between the 50s and 70s of the last century, see Kaplun
(1967), Lagerström and Van Dyke (1975), Fraenkel (1969)
Matched asymptotic expansions are used if a regular asymptotic expan-
sion fails near located singularities. Then the problem has to be rescaled
appropriately by using local variables before expanding its solution asymp-
totically. Both expansion have to agree in some overlap region, i.e. a region
where both expansions hold. We will demonstrate the method by consider-
ing three typical examples showing the wide applicability of the method.
-ε x
-1 1
φy (x, ±εT (x)) = ±εT (x)φx (x, ±εT (x)), −1 < x < 1 (4)
φ → x, for x2 + y 2 → ∞. (5)
φi (x, y) → 0, x, y → ∞. (9)
(q)
Note the flow potential of a source at the origin of strength q is φ =
q
2π ln x2 + y 2 . The perturbation potentials φk can be obtained by placing
distributed sources along the centerline of the airfoil on the interval (−1, 1)
and one can verify that the corresponding velocity fields can be represented
by:
1 1 x−ξ
uk (x, y) = φk,x (x, y) = φk,y (ξ, 0) dξ, (10a)
π −1 (x − ξ)2 + y 2
1 1 y
vk (x, y) = φk,y (x, y) = φk,y (ξ, 0) dξ, (10b)
π −1 (x − ξ)2 + y 2
cf. Van Dyke (1975). Using the perturbation potentials φi the surface
velocity us has the expansion
us (x) = φ2x (x, ±εT (x)) + φ2y (x, ±εT (x)) ∼
(11)
2 1 2
+εφ1 (x, 0) + ε φ2x (x, 0) + T (x)T (x) + T (x) + · · · .
2
Introduction to Matched Asymptotic Expansions 5
WHUPVRXWHUH[SDQVLRQ
ï ï ï ï ï ï
x
Figure 2. Asymptotic expansion of surface velocity
To be more
√ specific we consider an elliptical airfoil with the shape func-
tion T (x) = 1 − x2 . Using the notation of complex variables z = x + iy
the perturbation potentials φ1 , φ2 are
φ1 = φ2 =
z − z 2 − 1 , (12)
where
z denotes the real part of a complex number z. In order to make
the square root unique the complex plane is sliced along the interval (-1,1).
We have φ1,x (x, ±0) = φ2,x (x, ±0) = ±1. Thus the expansion of the surface
velocity
ε2 x2
us (x) ∼ 1 + ε − + ··· (13)
2 1 − x2
turns out to be not uniformly valid. If x is close to the leading or trailing
edge, say |x + 1| ε2 the second order correction term will become larger
than the first, (see figure 2).
It is interesting to note that the flow potential of a source or sink flow
at the leading or trailing edge can be added to the perturbation potential
φ1 . In particular the flow potential
2
C z+1
φ1 =
z − z − 1 − ln (14)
2π z − 1
satisfies all required conditions for an arbitrary constant C.
scale near the leading edge is the radius of curvature of the profile, which
in case of the elliptical airfoil is ε2 . Thus we define the local coordinates
1+x y
X= , Y = , (15)
ε2 ε2
and the local flow potential by Φ(X, Y ) by
The local potential satisfies again the Laplace equation. The contour of
the airfoil written in local coordinates is given by
√ ε2
2 2
Y = ± 2X + ε X ∼ 2X 1 + X + ... . (17)
4
and obtain for the first two terms the kinematic boundary condition
√ 1 √
Φi,Y X, 2X − √ Φi,X X, 2X = 0, i = 0, 1. (20)
2X
This can be interpreted as the kinematic boundary condition for the inviscid
flow around a parabola. Due to symmetry the stagnation point is in the
apex of the parabola. The flow potential can be determined by conformal
mapping, cf. Betz (1964). It is given by
√
Φi = Uloc,i
Z − 1 + 1 − 2Z , i = 0, 1, (21)
where the velocities of the free stream Uloc,i with i = 1, 2 are unknown. They
have to be determined by matching with the outer (global) expansion. The
expansion of the local surface velocity is given by
2X
Us ∼ (Uloc,0 + εUloc,1 ) . (22)
1 + 2X
Introduction to Matched Asymptotic Expansions 7
y
Y =
ε2
√
Y = ± 2X + ε2 X 2
Uloc
1+x
X=
ε2
The local expansion of the velocity field in the overlap region is given by
1
Φ ∼ (Uloc,0 + εUloc,1 ) 1 + √
1 − 2Z
(26)
1
∼ Uloc,0 1 + √ + εUloc,1 .
−2ξ
8 H. Steinrück
holds.
Van Dyke (1975) has formalized the matching procedure in the match-
ing principle. Fraenkel (1969) discussed criteria on the inner and outer
expansion for the validity of the matching principle. For example when the
gauge functions in the outer and inner expansion are powers of the expan-
sion parameter, which is defined as the ratio of the scales of the inner and
outer variable, the matching principle holds. Problems may arise when the
gauge functions are a combination of powers and logarithmic terms of the
perturbation parameter.
On the other hand two terms of the inner expansion rewritten in the outer
variables gives
(2) (2)
(2) 2X ε2
us = (1 + ε) =1+ε− . (31)
in out 1 + 2X 4(1 + x)
out
2X 2 x2 1
= (1 + ε) − 2
+ .
1 + 2X 2 1−x 4X
In figure 2 the outer, the inner expansion and the composite approxima-
tion of the surface velocity are shown for ε = 0.3 .
z̃
Ω̃ − ω̃
ũ
r̃
2L̃
Ekman-layers Ω̃ + ω̃
The distance between the two discs is 2L̃. Here and in the following
we denote dimensional quantities with a tilde. The upper disc rotates with
speed Ω̃ − ω̃ and the lower with Ω̃ + ω̃. We choose a cylindrical coordinate
system with the axis of rotation as the z-axis and its origin in the mid-
10 H. Steinrück
dle between the two plates. The fluid between the discs is assumed to be
incompressible and its kinematic viscosity ν̃ to be constant.
Thus two independent dimensionless group can be formed: The Ekman
number Ek = ν̃/Ω̃L̃2 and the Rossby number Ro = ω̃/Ω̃. If the Rossby
number is zero both discs rotate at the same angular speed. In that case the
fluid between the two discs would do the same. Thus the Rossby number is
a measure of the deviation of the solid body rotation of the fluid. Here we
will assume that the Rossby number is small.
The Ekman number can be interpreted as the reciprocal value of a
Reynolds number based on the reference velocity Ω̃L̃. Here we are interested
in the limit of small Ekman numbers.
The governing Navier-Stokes equation written in cylindrical coordinates
can be found in Schlichting (2000). We use the dimensionless vertical coor-
dinate Z = L̃z̃ and introduce the similarity ansatz
Ro (U 2 − V 2 + W UZ ) = 2V − A + Ek UZZ , (34a)
These are a set of ordinary differential equations for the velocity profiles
U , V , W , the pressure profile B and the constant A. At first glance one
might think that the six no-slip boundary conditions are not enough. But
we have to consider equation (34a), (34b) as second order equations for U
and V , respectively. The continuity equation (34d) can be considered as
first order equation for the vertical velocity profile W and equation (34c)
can be considered as an algebraic equation for the vertical pressure gradient
BZ . Thus in total six boundary conditions are needed to determine U , V ,
W , BZ and A .
Introduction to Matched Asymptotic Expansions 11
We remark that for the local behavior near the upper disc we obtain a sim-
ilar differential equation. Only the sign in front of Ŵ has to be changed.
Inspecting the local differential equation and setting Ek = 0 we see imme-
diately that a nontrivial differential equation is obtained if β = 1/2. Thus
we expand
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
U Û0 (ζ) √ Û1 (ζ)
⎝ V ⎠ ∼ ⎝ V̂0 (ζ) ⎠ + Ek ⎝ V̂1 (ζ) ⎠ + · · · . (41)
W Ŵ0 (ζ) Ŵ1 (ζ)
From the boundary condition at the lower disc (Z = −1) we obtain the
boundary condition for the local expansion
Û0 (0) = Û1 (0), V̂0 (0) = 1, V̂1 (0) = 0, Ŵ0 (0) = Ŵ1 (0). (42)
Inserting the expansion (41) into (40) and after some elementary manipu-
lations we obtain a fourth-order differential equation for V̇0
(iv)
V̂0 + 4V̂0 = 2A0 , V̂0 (0) = 1, V̂0 (0) = 0, (43)
with the solution
A0 A0
V̂0 (ξ) = + 1− e−ξ cos ξ + c1 sinh ξ cos ξ + c2 cosh ξ sin ξ. (44)
2 2
For the radial velocity component we obtain from Û0 = − 12 V̂0
A0
Û0 (ξ) = 1 − e−ξ sin ξ − 2c1 cosh ξ sin ξ + 2c2 sinh ξ cos ξ (45)
2
and for the vertical component Ŵ0 = 0 and
A0
Ŵ1 (ξ) = 1 − [1 − e−ξ (sin ξ + cos ξ)]− (46)
2
c1 − c2 c 1 + c2
sinh ξ sin ξ − (cosh ξ cos ξ − 1)
2 2
follows.
3.4 Matching
Applying the matching principle yields that all quantities in the overlap
region between the core layer and the boundary layer have to be constant.
Thus we conclude that c1 = c2 = 0. Furthermore we obtain
A0
Ū0 = 0, W̄0 = −1 + . (47)
2
Introduction to Matched Asymptotic Expansions 13
Ro = 0
Ek = 0.1 Ro = 1
Ek = 0.001
U
E
k
ï =
0.0
ï 1
ï
ï
ï ï ï ï ï
Z
a) radial velocity component
Ek = 0.1 Ro = 0
Ro = 1
Ek = 0.01
V
ï
ï
ï Ek = 0.001
ï
ï
ï ï ï ï ï
Z
b) azimuthal velocity component
Ro = 0
ï Ro = 1
W Ek = 0.1
ï
ï
ï
ï
ï
Ek = 0.001 Ek = 0.01
ï ï ï ï ï
Z
c) vertical velocity component
Figure 5. Velocity profiles between two rotating discs (solution of differen-
tial eq. (33)) for Ek = 0.1, 0.01, 0.001 and Ro = 0, 1.
14 H. Steinrück
The analysis of the boundary layer at Z = 1 follows the same lines as
indicated above. Only the sign in front of the first derivatives and in the
boundary condition for the azimuthal velocity V has to be changed.
After matching with the upper layer with the core region we obtain
A0
Ū0 = 0, W̄0 = −1 − . (48)
2
Thus we conclude that A0 = 0 and we have determined all leading order
terms of the asymptotic expansion of the velocity profile.
Combining the expansion of the core region with that of the Ekman-
layers we obtain the uniformly valid approximation.
dp̃ dτ̃
0=− + . (50)
dx̃ dỹ
The averaged shear stress τ̃ is the sum of the Reynolds shear stress τ̃t =
−ρ̃(ũ ṽ ) and the viscous stress
dũ
τ̃ = τ̃t + μ̃ . (51)
dỹ
Introduction to Matched Asymptotic Expansions 15
In case of a channel flow an expression for the mixing length l̃ = ˜l(ỹ) = dl(y)
˜
as a function of the dimensionless distance y = ỹ/l̃ from the wall can be
found in Schlichting (2000)
κ κ
l(y) = c0 − 2c0 − (1 − y)2 − − c0 (1 − y)4 . (53)
2 2
Note that the mixing length l̃ vanishes at the wall ỹ = 0 and that
l(y) ∼ κy + l2 y 2 /2 + O(y 3 ) for y 1 holds.
We introduce dimensionless variables by referring the velocity to the
center line velocity ũc , the shear stress to the double stagnation pressure
ρ̃ũ2c , the unknown pressure gradient to ρ̃ũ2c /d.
˜
We define
where we have made use of the force balance −dp̃/dx̃ = τ̃w /d˜ for a fully
developed
flow and the definition of the wall shear stress velocity ũτ =
τ̃w /ρ̃ with τ̃w denoting the wall shear stress. The dimensionless equations
reduce to the stress balance
dτ
0 = γ2 + , (55)
dy
and stress relation
2
du du
τ =ε + l(y)2 , (56)
dy dy
16 H. Steinrück
Integration yields
(D) 1 (D)
u1 (y) = ln y + F (D) (y) + u1 (1), (61)
κ
where
√
y
1 − y 1
F (D) (y) = − dy (62)
1 l(y ) κy
is a smooth bounded function of y on the interval (0, 1).
(D)
Now it is obvious to see that u1 is smooth at the centerline y = 1 and
(D) (D)
thus u0 = 1 and u1 (1) = 0. The velocity profile deviates only by a
small velocity defect of order γ from its maximum value at the center line.
Therefore this layer is called defect layer. Near the wall y = 0 the velocity
(D)
component u1 (y) is singular. Its asymptotic behavior is given by
(D) 1 κ + l2
u1 (y) ∼ ln y +CD − y, with CD = F (D) (0) as y → 0. (63)
κ 2κ2
Introduction to Matched Asymptotic Expansions 17
(v) 1 1
u1 (η) = ln η + ln 2κ + 4κ2 + 1/η 2 . (69)
κ κ
In order to match the viscous layer to the defect layer we consider the
(v) (v)
asymptotic behavior of u1 (η) and u2 (η) for η → ∞.
(v)
du1 1 1 1
∼ − + 3 3 + ··· , (70a)
dη κη 2κ2 η2 8κ η
(v) 1 1 1
u1 (η) = ln η + CV + 2 − + ··· ,
κ 2κ η 16κ2 η 2
(70b)
1
with CV = (ln 4κ − 1)
κ
18 H. Steinrück
du(v)
du2
(v) −η − κl η3 ( 1 )2 1 l2 l2 3l2
dη
=− (v)
∼− − 2+ 3 − 4 2 (70c)
dη du 2κ 2κ 2κ η 8κ η
1 + 2κ2 η 2 1
dη
(v) 1 l2 l2 3l2
u2 (η) ∼ − + 2 η + CV,2 + 3 ln η + , (70d)
2κ 2κ 2κ 16κ4 η
where CV,2 is an appropriate constant.
4.3 Matching
Finally it remains to match the velocity profile of the viscous layer
with that of the defect layer. Applying the matching principle we have
to care when counting the number of terms in the asymptotic expansions.
In Fraenkel (1969) it had been shown that the Van Dykes matching princi-
ple is still valid when the asymptotic expansion contain besides powers of
the perturbation parameter products of powers of ε and and powers of ln ε.
Than all logarithmic term multiplied by the same power of ε have to be
considered as one term.
In the present example it will turn out that γ(ε) ∼ O(1/ ln(1/ε)). Thus
(D) (D)
the two term expansion of the defect layer is 1 + γU1 + εu2 . Expanding
it in the viscous layer variable into two terms and using γσ = ε we obtain
(2) (2)
(2) (D) (D)
[u]D = 1 + γu1 + εu2 =
V V
1 1 1 1
=1+γ ln σ + ln η + CD + 2 + (71)
κ κ 2κ η
κ + l2 l2 l2
+ε CD,2 − η + ln η + ln σ .
2κ2 2κ3 2κ3
(v) (v)
Taking two terms of the inner (viscous)-layer expansion γu1 + γσu2 ,
rewriting it in the outer variables and expanding it into two term yields
(2) (2)
(2) (v) (v)
[u]V = γu1 + γσu2 =
D D
1 1 κ + l2
=γ ln y + CV − ln σ − y + (72)
κ κ 2κ2
l2 σ 1
+γσ CV,2 + 3 ln y − 3 ln σ + 2 .
2κ 2κ 2κ y
Introduction to Matched Asymptotic Expansions 19
Both expressions agree if the matching condition
1 1 l2
= − ln σ + CV − CD − σ ln σ + CD,2 − CV,2 (73)
γ κ 2κ3
is satisfied. Taking only the first order terms of both expansions the well
known friction law
1 1 γ
= ln + CV − CD (74)
γ κ ε
is obtained. It can be interpreted as a relation between the dimensionless
wall shear stress γ 2 and the Reynolds number ε−1 .
u
YLVFRXVVXEOD\HUSURILOH
ï GHIHFWOD\HUSURILOH
ï YHORFLW\SURILOH
ï
y
Figure 6. solution of model problem and viscous and defect layer approx-
imation for ε = 10−4
In figure 6 velocity profile as the solution of the force balance with the
mixing length model (53) for ε = 10−4, the approximation in the viscous
sub-layer and the defect layer is shown in a logarithmic plot. It can be clearly
seen that in the overlap region (here from 0.02 to 0.2) viscous and defect
expansion agree. In the overlap region both expansions can be represented
by a logarithmic velocity profile.
ỹ dũ ỹ ũτ ỹ
= Φ(y, y+ ), with y+ = , y= . (75)
ũτ dỹ ν̃ d˜
1
lim Φ(y, y+ ) = Φ(∞, 0) = . (76)
y+ →∞,y→0 κ
Integration of (76) yields the logarithmic velocity profile in the overlap re-
gion. We emphasize that the existence of the limit (76) from a theoretical
point of view is a nontrivial assumption (similarity of the first kind, see
Barenblatt (1996). However, the logarithmic law, if interpreted correctly,
is in excellent agreement with measured velocity profiles. Thus it can be
considered as an empirical fact. On the other hand there are authors , e. g.
Barenblatt (1996), who question the logarithmic law. Barenblatt (1996)
considers that the limit (76) does not exist, but that the function Φ is a
sophisticated power function of the Reynolds number (similarity of the sec-
ond kind). Instead of the logarithmic velocity profile these authors obtain
a power-law with an Reynolds-number dependent exponent. Although ac-
cording to Barenblatt (1996) the power law seems to reproduce some data
even better than the log-law it is a dead end from the asymptotic point of
view since it does not comply with the requirements of a rational asymptotic
expansion.
In modern papers concerning turbulence asymptotics the order of argu-
ments is reversed, see Walker (1998), Kluwick and Scheichl (2009). Usually
the dimensionless wall shear stress velocity γ = uτ /Uref is considered as
a small perturbation parameter and the existence of a viscous sub-layer
together with the log-law in the overlap region is postulated.
5 Conclusions
We have given an introduction to the method of matched asymptotic expan-
sion by analyzing three different problems of fluid mechanics. Characteristic
to all examples is the appearance of different length scales and that a uni-
formly valid asymptotic approximation can be constructed employing the
matching principle.
Introduction to Matched Asymptotic Expansions 21
Bibliography
G. I. Barenblatt. Scaling, self-similarity and intermediate asymptotics.
Cambridge Univ. Press, 1996.
A. Betz. Konforme Abbildungen. Springer, 2nd edition, 1964.
L. E. Fraenkel. On the method of matched asymptotic expansions, part i:
A matching principle. Proc. Camb. Phil. Soc., 65:209–231, 1969.
S. Kaplun. In P.A. Lagerstrom, L. N. Howard, and C. S. Liu, editors, Fluid
Mechanics and Singular Perturbation. Academic Press, 1967.
A. Kluwick and B. Scheichl. High-Reynolds-Number Asymptotics of Tur-
bulent Boundary Layers: From Fully Attached to Marginally Separated
Flows. In A. Hegarty, N. Kopteva, E. O’Riordan, and M. Stynes, editors,
BAIL 2008 - Boundary and Interior Layers, volume 69 of Lecture Notes
in Computational Science and Engineering, pages 3–22, 2009.
L. Prandtl. Über Flüssigkeitsbewegung bei sehr kleiner Reibung. In Ver-
handl. d. III. Intern. Mathem. Kongresses, Heidelberg, pages 484–491,
1904.
K. Schlichting, H. Gersten. Boundary-layer theory. Springer, 8th edition,
2000.
M. Ungarish. Hydromechanics of Suspensions. Springer, 1993.
M. Van Dyke. Perturbation Methods in Fluid Mechanics. The Parabolic
Press, Stanford, annoted edition edition, 1975.
Th. von Karman. Mechanische Ähnlichkeit und Turbulenz. Ges. Wiss.
Göttingen, Math.-Phys. Kl., pages 58–76, 1930.
J. D. A. Walker. Turbulent boundary layers ii: Further developments. In
A. Kluwick, editor, Recent Advances in Boundary Layer Therory, volume
390 of CISM Courses and Lectures, pages 145–230. Springer, 1998.
Asymptotic Methods For PDE Problems In
Fluid Mechanics and Related Systems With
Strong Localized Perturbations In
Two-Dimensional Domains
1 Introduction
The method of matched asymptotic expansions is a powerful systematic an-
alytical method for asymptotically calculating solutions to singularly per-
turbed PDE problems. It has been successfully used in a wide variety
of applications (cf. Kevorkian and Cole (1993), Lagerstrom (1988), Dyke
(1975)). However, there are certain special classes of problems where this
method has some apparent limitations.
In particular, for singular perturbation PDE problems leading to infinite
logarithmic series in powers of ν = −1/ log ε, where ε is a small positive pa-
rameter, it is well-known that this method may be of only limited practical
use in approximating the exact solution accurately. This difficulty stems
from the fact that ν → 0 very slowly as ε decreases. Therefore, unless
many coefficients in the infinite logarithmic series can be obtained analyti-
cally, the resulting low order truncation of this series will typically not be
very accurate unless ε is very small. Singular perturbation problems in-
volving infinite logarithmic expansions arise in many areas of application
in two-dimensional spatial domains including, low Reynolds number fluid
flow past bodies of cylindrical cross-section, low Peclet number convection-
diffusion problems with localized obstacles, and the calculation of the mean
first passage time for Brownian motion in the presence of small traps, etc.
In this article we survey consider various singularly perturbed PDE prob-
lems in two-dimensional spatial domains where hybrid asymptotic-numerical
methods have been formulated and implemented to effectively ‘sum’ infinite
logarithmic expansions. Some of the problems considered herein directly re-
late to fluid mechanics, whereas other problems arise in different scientific
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
24 M.J. Ward and M.-C. Kropinski
w = −β , x ∈ Ω\Ωε , (1a)
w = 0, x ∈ ∂Ω , (1b)
w = 0, x ∈ ∂Ωε . (1c)
analysis below will show how to calculate the sum of all the logarithmic
terms for w in in the limit ε → 0 of small core radius.
In the outer region we expand the solution to (1) as
w(x; ε) = W0 (x; ν) + σ(ε)W1 (x; ν) + · · · . (3)
The unique solution to (7) has the following far-field asymptotic behavior:
p·y
vc (y) ∼ log |y| − log d + + ··· , as |y| → ∞ . (7c)
|y|2
The constant d > 0, called the logarithmic capacitance of Ω 1 , depends on
the shape of Ω1 but not on its orientation. The vector p is called the
dipole vector. Numerical values for d for different shapes of Ω1 are given in
Ransford (1995), and some of these are reproduced in Table 1. A boundary
integral method to compute d for arbitrarily-shaped domains Ω1 is described
and implemented in Dijkstra and Hochstenbach (2008).
The leading-order matching condition between the inner and outer so-
lutions will determine the constant γ in (6b). Upon writing (7c) in outer
variables and substituting into (6b), we get the far-field behavior
Choosing
ν(ε) = −1/ log(εd) , (9)
and matching (8) to the outer expansion (3) for W , we obtain the singularity
condition for W0 ,
The singularity behavior in (10) specifies both the regular and singular
part of a Coulomb singularity. As such, it provides one constraint for the
determination of γ. More specifically, the solution to (4) together with
(10) must determine γ, since for a singularity condition of the form W0 ∼
Asymptotic Methods for PDE Problems 27
Here W0H (x) is the smooth function satisfying the unperturbed problem
Gd = −δ(x − x0 ) , x ∈ Ω ; Gd = 0 , x ∈ ∂Ω , (13a)
1
Gd (x; x0 ) = − log |x − x0 | + Rd (x0 ; x0 ) + o(1) , as x → x0 . (13b)
2π
Here Rd00 ≡ Rd (x0 ; x0 ) is the regular part of the Dirichlet Green’s function
Gd (x; x0 ) at x = x0 . This regular part is also known as either the self-
interaction term or the Robin constant (cf. Bandle and Flucher (1996)).
Upon substituting (13b) into (11) and letting x → x0 , we compare the
resulting expression with (10) to obtain that γ is given by
W0H (x0 )
γ= . (14)
1 + 2πνRd00
Therefore, for this problem, γ is determined as the sum of a geometric
series in ν. The range of validity of (14) is limited to values of ε for which
2πν|Rd00 | < 1. This yields,
1
0 < ε < εc , εc ≡ exp [2πRd00 ] . (15)
d
We summarize our result as follows:
Principal Result 1: For ε 1, the outer expansion for (1) is
2πνW0H (x0 )
w ∼ W0 (x; ν) = W0H (x) − Gd (x; x0 ) , for |x − x0 | = O(1) ,
1 + 2πνRd00
(16a)
and the inner expansion with y = ε−1 (x − x0 ) is
νW0H (x0 )
w ∼ V0 (y; ν) = vc (y) , for |x − x0 | = O(ε) . (16b)
1 + 2πνRd00
Here ν = −1/ log(εd), d is defined in (7c), vc (y) satisfies (7), and W0H sat-
isfies the unperturbed problem (12). Also Gd (x; x0 ) and Rd00 ≡ Rd (x0 ; x0 )
are the Dirichlet Green’s function and its regular part satisfying (13).
28 M.J. Ward and M.-C. Kropinski
∞
W ∼ W0H (x) + ν j W0j (x) + μ0 (ε)W1 + · · · , (17)
j=1
with μ0 (ε) ν k for any k > 0. By formulating a similar series for the inner
solution, we will derive a recursive set of problems for the W0j for j ≥ 0
from the asymptotic matching of the inner and outer solutions. We will
then sum this series to re-derive the result in Principal Result 1.
In the outer region we expand the solution to (1) as in (17). In (17),
ν = O(1/ log ε) is a gauge function to be chosen, while the smooth function
W0H satisfies the unperturbed problem (12) in the unperturbed domain.
By substituting (17) into (1a) and (1b), and letting Ωε → x0 as ε → 0, we
Asymptotic Methods for PDE Problems 29
The matching of the outer and inner expansions will determine a singularity
behavior for W0j as x → x0 for each j ≥ 1.
In the inner region near Ωε we introduce the inner variables
The higher-order matching condition, from (22), shows that the solution
W0j to (18) must have the singularity behavior
Finally, to obtain the outer solution we substitute (25) and (28) into
(17) to obtain
∞
∞
w − W0H (x) ∼ ν j (−2πγj−1 ) Gd (x; x0 ) = −2πνGd (x; x0 ) ν j γj
j=1 j=0
∞
∼ −2πνW0H (x0 )Gd (x; x0 ) [−2πνRd00 ]j
j=0
2πνW0H (x0 )
∼− Gd (x0 ; x0 ) . (29a)
1 + 2πνRd00
Equation (29a) agrees with equation (16a) of Principal Result 1. Similarly,
upon substituting (28) into the infinite-order inner expansion (20), we obtain
∞
j νW0H (x0 )
v(y; ε) = νW0H (x0 )vc (y) [−2πRd00 ν] = vc (y) , (30)
1 + 2πνRd00
j=0
core, an elliptical core, and an equilateral triangular core. Using the no-
tation in the table, we set the major and minor semi-axes of the ellipse as
a = 2 and b = 1, and both the side of the square and the equilateral triangle
as h = 1. To compute the hybrid method solution, we readily calculate that
the Green’s function is Gd = −(2π)−1 log(r/r0 ) and that the unperturbed
solution is W0H = β(r02 − r2 )/4. The outer hybrid method solution, as
obtained from (16a) of Principal Result 1, is simply
β 2 2 2 log(r0 /r)
w(x; ε) = r − r − r0 , r = |x| . (31)
4 0 log(r0 /[εd])
From (31), we can compute the asymptotic mean flow rate using (2).
0.42 0.45
Hybrid Hybrid
Full Numerical Exact
0.4 0.44
Square
W
W
0.36 0.42
0.34 0.41
Ellipse
0.32 0.4
0.3 0.39
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
ε ε
To validate the asymptotic results for w̄, we compare them with corre-
sponding direct numerical results computed from the full problem (1) using
the PDE Toolbox of MATLAB (1996). For a circular pipe of radius r0 = 2
containing a concentric core and with β = 1, Fig. 1(a) contains curves of
mean flow velocity, w̄, versus ε, a measure of the size of the core, for three
32 M.J. Ward and M.-C. Kropinski
x2
x3
x1
Capillary
Cross-section
2-D cut
∂n p = 0, x ∈ ∂Ω . (34b)
ε∂n p + κj (p − pcj ) = 0 , x ∈ ∂Ωεj , j = 1, . . . , N . (34c)
The condition (34c) models the capillary wall as a finitely permeable mem-
brane, where κi > 0 is the permeability coefficient of the ith capillary and
pci is the oxygen partial pressure within the ith capillary (assumed con-
stant). In (34c) and (34b), ∂n is the outward normal derivative to the
tissue domain. In deriving (34) we have assumed that the oxygen diffusiv-
ity is spatially constant, and so the oxygen consumption rate M has been
normalized by this constant value. To incorporate skeletal muscle tissue het-
erogeneities, such as localized oxygen-consuming mitochondria, we assume
that M is spatially-dependent and has the form
m
|x − xi |2
M(x) = M0 + Mi exp − , (35)
i=1
σi2
P0 = M , x ∈ Ω\{x1 , . . . , xN } , (38a)
∂n P0 = 0 , x ∈ ∂Ω , (38b)
P0 is singular as x → xj . (38c)
The matching of the outer and inner expansions will determine singularity
behaviors for P0 as x → xj for j = 1, . . . , N .
In the inner region near the j th capillary Ωεj we introduce the inner
variables
y = ε−1 (x − xj ) , p(y; ε) = qj (xj + εy; ε) , (39)
together with the local expansion
Here we assume that μ νjk for any k > 0. We then write q0j in the form
y qcj = 0 , y∈
/ Ωj ; ∂n qcj + κj qc = 0 , y ∈ ∂Ωj , (42a)
qcj ∼ log |y| , as |y| → ∞ , (42b)
where Ωj ≡ ε−1 Ωεj . The unique solution to (42) has the following far-field
asymptotic behavior:
1
qcj (y) ∼ log |y| − log dj + O , |y| 1 . (42c)
|y|
In comparing (42) with (7) for the pipe problem of §2, we observe that here
dj = dj (κj ). For a particular cross-sectional shape of the capillary and
for a given value of κj , one must compute dj = dj (κj ) numerically from a
boundary integral method applied to (42). For a circular capillary of radius
ε, for which qcj can be found analytically, we readily calculate that
Moreover, by comparing (6b) with (41) we observe that here we have intro-
duced a slight change in the definition of the inner solution. In the analysis
below, we will show that Aj = O(1) as ε → 0 in (41), which is a direct
consequence of the Neumann boundary condition in (34b).
By using (40) and (42c), we re-write the far-field form for |y| 1 of the
inner solution in terms of the outer variables as
Aj
qj ∼ pcj + Aj log |x − xj | + . (44a)
νj
The matching condition is that the far-field form (44a) of the inner solution
must agree with the near-field behavior of the outer solution for p. Thus,
P0 satisfies (38) subject to the following singularity behavior
Aj
P0 ∼ pcj + Aj log |x − xj | + , as x → xj , j = 1, . . . , N . (45)
νj
N
1
Aj = − M(x) dx . (46)
2π Ω
j=1
N
P0 = PR (x) − 2π Ai GN (x; xi ) + χ . (47)
i=1
where qcj satisfies (42). In the outer region, defined at O(1) distances from
the centers of the capillaries, we have
N
p ∼ PR (x) − 2π Ai GN (x; xi ) + χ . (51b)
i=1
4.95
j=3
4.9
j=2
min
ε j=1 4.85
p
4.8 j=1 (+)
j=2 (x)
4.75 j=3 (*)
4.7
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
ε
The analysis of the solution in the inner region is the same as that in §2
since the effect of the nonlinear term in the inner region is O(ε 2 ), which is
transcendentally small compared to the logarithmic terms. Hence, W0 must
have the following singular behavior (see equation (10)):
W0 ∼ S log |x − x0 | , as x → x0 . (56)
Asymptotic Methods for PDE Problems 41
S
γ = R(S; x0 ) − α , ν= . (58)
R(S; x0 ) − α
1
W0 + W0 + eW0 = 0 , 0 < r ≤ 1 ; W0 = 0 , on r = 1 , (59a)
r
W0 ∼ S log r , as r → 0 , (59b)
where r = |x|. This problem (59) can be solved analytically by first intro-
ducing the new variables v and η defined by
When S > −2, we then obtain that v = v(η) is smooth and satisfies
−2
1 S
v + v + 1 + ev = 0 , 0 ≤ η ≤ 1; v = 0, on η = 1 . (61)
η 2
42 M.J. Ward and M.-C. Kropinski
The well-known solution v = v(η) to (61) (see Ward et al. (1993)) can be
written in parametric form as
−2
1+ρ S 8ρ
v(η) = 2 log , 1+ = . (62)
1 + ρη 2 2 (1 + ρ)2
The maximum of the right-hand side of the implicit expression for ρ(S) in
(62) is 2 and occurs when ρ = 1. Therefore, for the existence√ of a solution
2
to (59)
√ we require that (1 + S/2) > 1/2, which yields S > 2 − 2. When
S > 2 − 2, then ρ(S) from (62) has two roots for ρ, and hence (59) has
two solutions. Let us consider the smaller root, labeled by ρ− (S), given by
1/2
ρ− (S) = (S + 1)(S + 3) − (S + 2) (S + 2)2 − 2 . (63)
This expression (69c) yields a singularity structure for the outer Oseen so-
lution as r → 0. This behavior suggests that we introduce the new variable
Ψ by Ψ(r, θ) = εψ(ε−1 r, θ), and that we expand Ψ as
∞
Ψ(r, θ) = r sin θ + νΨ1 (r, θ) + ν j Ψj (r, θ) + · · · , (70)
j=2
The limiting conditions (71b) and (71d) are the required singularity be-
haviors of Ψ1 and Ψj for j ≥ 2, respectively, as r → 0. For (71b) the
strength of the singular part r log r sin θ is set to unity. In terms of the solu-
tion to (71a) with Ψ1 ∼ r log r sin θ as r → 0, we then calculate the constant
a2 of the regular part of the singularity structure from the limiting process
Ψ1 − r log r sin θ ∼ a2 r sin θ as r → 0. Then, with a2 determined in this
way, we solve for Ψ2 from (71c) with singular behavior Ψ2 ∼ a2 r log r sin θ
as r → 0, The constant a3 in the regular part of (71d) is then determined
from the limiting process Ψ2 − a2 r log r sin θ ∼ a3 r sin θ as r → 0.
Hence, the coefficients aj for j = 2, 3, .., which are independent of ε and
of the shape of the body, are determined recursively from (71), in a similar
way as in §2. The first two coefficients are (cf. Kaplun (1957), Proudman
and Pearson (1957))
∞
cn (r/2)
Ψ1 (r, θ) = − r sin(nθ) , cn (s) ≡ 2 [K1 (s)In (s) + K0 (s)In (s)] .
n=1
n
(73)
As r → 0, then cn (r/2) = O(rn−1 ) for n > 1, and c1 (r/2) ∼ 1−log(ρ/4)−γe ,
where γe is Euler’s constant. Therefore, we conclude that Ψ1 −r log r sin θ →
r (γe − log 4 − 1) sin θ as r → 0. Hence, from the regular part in (71b), we
obtain that a2 = γe − log 4 − 1. In contrast, the derivation in Kaplun (1957)
of the explicit formula for a3 given in (72b) is considerably more involved.
Explicit analytical formulae for aj when j ≥ 4 are not available.
A formula for the drag coefficient CD is given in Imai (1951) in terms of
an arbitrary closed contour around the body. From this formula, and from
Asymptotic Methods for PDE Problems 47
4π
−1
CD ∼ ν̂(εdf ) 1 − 0.8669 ν̂ 2(εdf ) , ν̂(z) ≡ [log (3.7027/z)] .
ε
(76)
For a circular cylinder, the explicit form (76) provides a rather poor
determination of the experimental drag coefficient unless ε is rather small
(cf. Dyke (1975)). One way to overcome this difficulty would be to com-
pute numerically further coefficients aj , for j ≥ 4, from the the infinite
sequence of PDE’s (71c) with singularity structures (71d). This would still
require truncating the series (75) at some finite j. As an alternative to series
truncation, we now follow Kropinski et al. (1995) and formulate a hybrid
asymptotic-numerical method that has the effect of summing all the terms
on the right-hand side of (75), but which avoids computing the coefficients
aj for j ≥ 1 individually.
To do so, we let A (z) denote a function that is asymptotic to the sum
of the terms written explicitly in the brackets on the right-hand side of (75):
∞
A (z) ∼
ν j−1 (z) aj , z ≡ εdf . (77)
j=1
48 M.J. Ward and M.-C. Kropinski
0.5
R(S) 0.0
−0.5
−1.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7
S
4π −1
CD = [ν(z)A (z) + · · · ] , ν(z) =
, z = εdf . (84)
ε log z e1/2
Kaplun’s equivalence principle follows from the fact that the curve A (z)
versus z can be used for a cylinder of arbitrary cross-section. To determine
A (εdf ) for a body of a specific shape, we need only compute the single
constant df from the numerical solution to the canonical Stokes problem
(68). This feature provides a significant advantage over a direct numerical
approach on the full problem (65).
For a few simple cross-sectional shapes, the constant df can be deter-
mined analytically from (68). In particular, for a circular cross-section,
50 M.J. Ward and M.-C. Kropinski
δ θT k c df b
.050 0◦ 2.000 0.961 0.328 0.040
.080 5◦ 1.972 0.952 0.344 0.066
.100 13◦ 1.928 0.960 0.354 0.082
.120 16◦ 1.910 0.954 0.364 0.098
.120 20◦ 1.889 0.968 0.363 0.096
.200 25◦ 1.861 0.915 0.410 0.170
20.0
c
cc
15.0 cc
cc
c
c
cc
c c
c c
CD 10.0 c
cc c
c cc
c c
c c
c
c
c c c c
5.0 c
0.0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
We will calculate the exact solution to this problem, and then show how a
hybrid method similar to that used for the low Reynolds number flow prob-
lem can be readily formulated and implemented to calculate an approximate
solution to (88) that contains all logarithmic correction terms.
54 M.J. Ward and M.-C. Kropinski
1 B
+A+ = κε2 , (91)
2 2
where κ is an O(1) constant to be found. Substituting (91) into (90), and
neglecting the higher order Aε3 and 3Aε2 terms in (90), we obtain the
approximate system
1 B 1 B
B log ε + −2A + − ≈ −κ , B + B log ε + −2A + − ≈ κ.
2 2 2 2
(92)
By adding the two equations above to eliminate κ, we obtain that
From (93), together with A ∼ −(1 + B)/2 from (91), we obtain that
3ν 3 −1
B∼ , A=1− , where ν ≡
. (94)
2−ν 2−ν log εe1/2
Finally, substituting (94) into (89), we obtain that the outer solution has
the asymptotics
u ∼ (1 − Ã)r3 + ν Ãr log r + Ãr sin θ , r O(ε) . (95a)
where à is defined by
3 −1
à ≡ , ν≡
. (95b)
2−ν log εe1/2
Asymptotic Methods for PDE Problems 55
κT = U · ∇T , X ∈ R2 \ ∪j=1 Ωj , (102a)
T = Tj , X ∈ ∂Ωj , j = 1, 2 , (102b)
T = T∞ , |X| → ∞ . (102c)
We also define the dimensionless centers of the two circular disks by xj for
j = 1, 2, and their constant boundary temperatures αj for j = 1, 2, by
xj = Xj /γ , αj = wj /T∞ , j = 1, 2 . (104)
|y2 − y1 | = 2l . (109)
v = v0 + νAvc , (110)
y v0 = 0 , y ∈ R2 \ ∪2j=1 Dj , (111a)
v0 = αj , y ∈ ∂Dj , j = 1, 2 , (111b)
v0 bounded as |y| → ∞ . (111c)
y vc = 0 , y ∈ R2 \ ∪2j=1 Dj , (112a)
vc = 0 , y ∈ ∂Dj , j = 1, 2 , (112b)
vc ∼ log |y| , as |y| → ∞ . (112c)
where d is given by
∞
ξc e−mξc
log d = log (2β) − + . (115)
2 m=1
m cosh(mξc )
G = u · ∇G − δ(x − ξ) , x ∈ R2 , (119a)
1
G(x; ξ) ∼ − log |x − ξ| + R(ξ; ξ) + o(1) , x → ξ , (119b)
2π
with G(x; ξ) → 0 as |x| → ∞. Here R(ξ; ξ) is the regular part of this
Green’s function at x = ξ.
The solution to (118) with singular behavior V0 ∼ νA log |x| as x → 0 is
V0 = 1 − 2πνAG(x; 0) . (120)
the average heat flux across the bodies, by using the divergence theorem
together with the form (117) of the far-field behavior in the inner region.
We leave this simple calculation to the reader.
Here Gij = G(xj ; xi ) and Rjj = R(xj ; xj ) are the Green’s function and its
regular part as defined by (119).
Finally, we remark that for the case of a uniform flow where u = (1, 0),
then the explicit solution to (119) is
1 x1 − ξ1
G(x; ξ) = exp K0 (|x − ξ|) , (126a)
2π 2
where x = (x1 , x2 ) and ξ = (ξ1 , ξ2 ). By letting x → ξ, and using K0 (r) ∼
− log r + log 2 − γe , as r → 0+ , where γe is Euler’s constant, we readily
calculate that
1
R(ξ, ξ) = (log 2 − γe ) . (126b)
2π
These results for G and its regular part can be used in the results of either
(121) or (125) for Case I or Case II, respectively.
60 M.J. Ward and M.-C. Kropinski
∂n u = 0 , x ∈ ∂Ω ; u = 0, x ∈ ∂Ωp ≡ ∪K
i=1 ∂Ωεi . (127b)
wandering particle
O(ε )
x2
N small absorbing holes
x1
n
reflecting
walls
Here, ν(ε) = −1/ log(εd) where d is the logarithmic capacitance of the hole.
For the unperturbed problem with ε = 0, we have λ00 = 0. In the O(ν)
term, λ01 is independent of the location of the hole at x = x0 (cf. Ward
et al. (1993)). Therefore, we need the higher-order coefficient λ02 in order
to determine the location of the hole that maximizes λ0 .
For the case of one hole, an infinite logarithmic expansion for λ0 (ε) has
the form
ε 1
λ0 (ε) = λ∗ (ν) + O , ν≡− .
log ε log(εd)
To calculate λ∗ (ν) we use the hybrid method of Kolokolnikov et al. (2005).
Near the hole, we identify an inner (local) region in terms of a local spatial
variable y = ε−1 (x−x0 ), and where the hole is rescaled so that Ω0 ≡ ε−1 Ωε .
Denoting the inner (local) solution by v(y, ε) = u(x0 + εy, ε), we then
expand v(y, ε) as
v(y, ε) = A ν vc (y) + · · · . (130)
62 M.J. Ward and M.-C. Kropinski
Δy vc = 0 , y ∈ Ω0 ;
vc = 0 , y ∈ ∂Ω0 , (131a)
p·y
vc ∼ log |y| − log d + , as |y| → ∞ . (131b)
|y|2
where μ O(ν k ) for any k > 0. Substituting (132) into (127a) and the
boundary condition (127b) on ∂Ω, we obtain the full problem in a domain
punctured by the point x0 ,
Δu∗ + λ∗ u∗ = 0 , x ∈ Ω\{x0 } ; (u∗ )2 dx = 1 ; ∂n u∗ = 0 , x ∈ ∂Ω .
Ω
(133)
The singularity condition for (133) as x → x0 given below arises from
matching u∗ to the inner solution. Substituting (131b) into (130), and
expressing the result in global variables, we obtain
p · (x − x0 )
v(y, ε) ∼ A ν log |x−x0 |+A+εA ν +· · · , as y → ∞ . (134)
|x − x0 |2
Comparing the terms in (134) and (132) at the next order, we see that
μ = O(εν).
Next, we must determine u∗ (x, ν) and λ∗ (ν) satisfying (133) and (135).
To do so, we introduce the Helmholtz Green’s function, Gh (x; x0 , λ∗ ), and
its regular part, Rh (x0 ; x0 , λ∗ ), satisfying
Substituting this expression into (138), we get the following two-term asymp-
totic result:
Principal Result 3:(One Hole) For ε → 0, the first eigenvalue λ0 of (127)
has the two-term asymptotic behavior
2πν 2πν 4π2 ν 2
λ0 (ε) = +O(ν 3 ) = − RN (x0 ; x0 )+O(ν 3 ) .
|Ω| (1 + 2πνRN (x0 ; x0 )) |Ω| |Ω|
(143)
Here ν = −1/ log(εd), and d is the logarithmic capacitance determined from
the inner problem (131). An infinite-order logarithmic expansion for λ0 is
given by λ0 ∼ λ∗ , where λ∗ is the first positive root of (138).
Next, we extend the asymptotic framework to the case of K holes. Much
of the analysis above remains the same, except that now the single hole x 0
is replaced by xi , for i = 1, . . . , K. The hybrid formulation for K holes is
K
u∗ = −2π Aj νj Gh (x; xj , λ∗ ) ∼ Ai νi (log |x − xi | − 2πνi Rh (xi ; xi , λ∗ ))
j=1
K
− 2π Aj νj Gh (xi ; xj , λ∗ ). (145)
j=1
j=i
The matching condition is that the expressions in (144b) and (145) agree.
The logarithmic terms agree, and from the remaining terms, we obtain a
K × K homogeneous linear system to solve for the Ai , given by
K
Ai (1 + 2πνi Rh (xi ; xi , λ∗ )) + 2π Aj νj Gh (xi ; xj , λ∗ ) = 0 , i = 1, . . . , K .
j=1
j=i
(146)
Asymptotic Methods for PDE Problems 65
A solution to (146) exists only when the determinant associated with the
linear system (146) vanishes. This condition provides an expression for
λ∗ (ν1 , . . . , νK ) that sums all the logarithmic terms in the asymptotic ex-
pansion of λ0 (ε).
As with the case for one hole in the domain, we can derive an asymptotic
formula for λ∗ that has an error of O(ν 3 ). This formula is again determined
in terms of the Neumann Green’s function GN and its regular part RN ,
defined in (49). By using (141) and (142) in (146), we obtain a homogeneous
linear system for the Ai for i = 1, . . . , K, given by
K
2πνi 1
Ai 1 + 2πνi RN (xj ; xj ) − +2π A ν
j j − + G N (x j ; xi ) = 0.
|Ω|λ∗ j=1
|Ω|λ∗
j=i
(147)
It is convenient to write (147) in matrix form as
2π
Ca = BVa , C ≡ I + 2πGN V , (148a)
|Ω|λ∗
where
⎛ ⎞ ⎛ ⎞
ν1 0 ··· 0 1 1 ··· 1 ⎛ ⎞
⎜ .. ⎟ ⎜ .. ⎟ A1
⎜ ··· 0 ⎟ ⎜ ··· 1 ⎟
V ≡⎜
0 . ⎟, B≡⎜ 1 . ⎟, a≡⎜ . ⎟
⎝ .. ⎠ .
⎜ .. .. .. .. ⎟ ⎜ .. .. .. . ⎟
⎝ . . . . ⎠ ⎝ . . . .. ⎠
AK
0 0 · · · νK 1 1 ··· 1
(148b)
In (148a), the Neumann Green’s function matrix GN is the K ×K symmetric
matrix with entries
(GN )ij ≡ GN (xi ; xj ) , i = j ; (GN )jj = RN (xj ; xj ) . (148c)
Let νm = max νj . Then, for νm sufficiently small, we can invert C ap-
j=1,...,K
proximately, to obtain that λ∗ is an eigenvalue of the matrix eigenvalue
problem
2π −1
Aa = λ∗ a , A= C BV . (149)
|Ω|
By using this representation of λ∗ we obtain the following result:
Principal Result 4:(K Holes) For ε → 0, the first eigenvalue λ0 of (127) has
the explicit two-term asymptotic behavior
⎛ ⎞
2π K K K
λ0 (ε) ∼ λ∗ , λ∗ = ⎝ νj − 2π νj νi (G)N ij ⎠ + O(νm
3
).
|Ω| j=1 j=1 i=1
(150)
66 M.J. Ward and M.-C. Kropinski
Here (G)N ij are the entries of the Neumann Green’s function matrix GN
defined in (148c).
Proof. We first notice that the matrix BV has rank one, since V is diagonal
and B = e0 et0 , where et0 = (1, 1, . . . , 1). This implies that A has rank one,
and so λ∗ is the unique nonzero eigenvalue of A. Hence, λ∗ = TraceA. By
using the structure of A in (149), we readily calculate that
2π $ %
K K
∗
λ = νj cij , cij ≡ C −1 ij . (151)
|Ω| j=1 i=1
2πKν 4π 2 ν 2
λ0 (ε) ∼ λ∗ , λ∗ = − p(x1 , . . . , xK ) + O(ν 3 ) , (152)
|Ω| |Ω|
K
K
K
K
K
F(x1 , . . . , xK ) = − log |xj −xk |− log |1−xj x̄k |+K |xj |2 ,
j=1 k=1 j=1 k=1 j=1
k=j
(154)
for |xij| = 1 and xj = xk when j = k. Here x̄k denotes the complex
conjugate of xk .
An interesting open problem is to determine the optimal arrangement of
K 1 traps in the dilute fraction limit Kε2 1. In particular, does the
optimal arrangement approach a hexagonal lattice structure with a bound-
ary layer near the rim of the unit disk?
6 Conclusion
In this article we have surveyed the development and application of a hybrid
asymptotic-numerical method for solving linear and nonlinear PDE mod-
els in two-dimensional domains that have small inclusions or obstructions.
Related theoretical approaches have also been developed to treat similar
strongly localized perturbation problems including, an eigenvalue perturba-
tion problem in a three-dimensional domain (cf. Ward and Keller (1993)),
cell-signaling problems in mathematical biology (cf. Bressloff et al. (2008),
Straube and Ward (2009)), the narrow escape problem from a sphere or a
disk that has small absorbing windows on its boundary (cf. Cheviakov et al.
(to appear, 2010), Pillay et al. (to appear, 2010)), and the mean first pas-
sage time in a three-dimensional domain with interior traps (cf. Cheviakov
and Ward (to appear, 2010)).
Acknowledgments
M. J. W. gratefully acknowledges the fundamental contributions of my other
co-authors D. Coombs (UBC), J. B. Keller (Stanford), T. Kolokolnikov
(Dalhousie), A. Peirce (UBC), S. Pillay (J.P. Morgan), R. Straube (Max
Planck Institute, Magdeburg), and M. Titcombe (Champlain College) to
some of the work surveyed herein. M. J. W. and M. C. K acknowledge the
grant support of NSERC (Canada).
Bibliography
C. Bandle and M. Flucher. Harmonic radius and concentration of energy;
hyperbolic radius and liouville’s equation. SIAM Review, 38(2):191–238,
68 M.J. Ward and M.-C. Kropinski
1996.
P. C. Bressloff, D. Earnshaw, and M. J. Ward. Diffusion of protein recep-
tors on a cylindrical dendritic membrane with partially absorbing traps.
SIAM J. Appl. Math., 68(5):1223–1246, 2008.
A. Cheviakov and M. J. Ward. Optimizing the fundamental eigenvalue of
the laplacian in a sphere with interior traps. Mathematical and Computer
Modeling, to appear, 2010.
A. Cheviakov, M. J. Ward, and R. Straube. An asymptotic analysis of the
mean first passage time for narrow escape problems: Part ii: The sphere.
SIAM J. Multiscale Modeling, to appear, 2010.
J. Choi, D. Margetis, T. M. Squires, and M. Z. Bazant. Steady advection-
diffusion around finite absorbers in two-dimensional potential flows. J.
Fluid Mech., 536:155–184, 2005.
D. Coombs, R. Straube, and M. J. Ward. Diffusion on a sphere with local-
ized traps: Mean first passage time, eigenvalue asymptotics, and fekete
points. SIAM J. Appl. Math., 70(1):302–332, 2009.
A. M. Davis and S. G. Llewellyn-Smith. Perturbation of eigenvalues due
to gaps in two-dimensional boundaries. Proc. Roy. Soc. A, 463:759–786,
2007.
W. W. Dijkstra and M. E. Hochstenbach. Numerical approximation of
the logarithmic capacity. CASA Report 08-09, Technical U. Eindhoven,
2008.
M. Van Dyke. Perturbations Methods in Fluid Mechanics. Parabolic Press,
1975.
J. E. Fletcher. Mathematical modeling of the microcirculation. Math Bio-
sciences, 38:159–202, 1978.
I. Imai. On the asymptotic behavior of viscous fluid flow at a great distance
from a cylindrical body, with special reference to filon’s paradox. Proc.
Roy. Soc. A, 208:487–516, 1951.
S. Kaplun. Low reynolds number flow past a circular cylinder. J. Math.
Mech., 6(5):52–60, 1957.
J. B. Keller and M. J. Ward. Asymptotics beyond all orders for a low
reynolds number flow. J. Engrg. Math., 30(1-2):253–265, 1996.
J. Kevorkian and J. Cole. Multiple Scale and Singular Perturbation Methods.
Applied Mathematical Sciences, Vol. 114, Springer-Verlag, 1993.
T. Kolokolnikov, M. Titcombe, and M. J. Ward. Optimizing the fundamen-
tal neumann eigenvalue for the laplacian in a domain with small traps.
Europ. J. Appl. Math., 16(2):161–200, 2005.
A. Krogh. The number and distribution of capillaries in muscles with cal-
culations of the oxygen pressure head necessary for supplying the tissue.
J. Physiology (London), 52:409–415, 1919.
Asymptotic Methods for PDE Problems 69
*
Roger Grimshaw
*
Loughborough University, UK
1 Introduction
An inverse power series is asymptotic to a function f (x) if, for fixed N and
sufficiently large x > 0,
N
an 1
f (x) − n
= O( N +1 ) . (1)
n=0
x x
∞
an
f (x) ∼ . (2)
n=0
xn
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
72 R. Grimshaw
This function is well-defined for all x > 0. Now integrate by parts N times,
to get that
N
n!
I(x) = (−1)n n+1 + N (x) , (4)
n=0
x
∞
(N + 1)! exp (−xt)
N (x) = (−1)N +1 dt . (5)
xN +1 0 (1 + t)N +1
Since the integral term in (5) is bounded by 1/x, it is easily shown that
N (x) = O(x−N −2 ) for each fixed N , and so
∞
n!
I(x) ∼ (−1)n . (6)
n=0
xn+1
Note that the series diverges for all x > 0, but nonetheless is a useful
approximation to I(x) as x → ∞.
For each fixed x, as x → ∞, we can minimize |N (x)| with respect to
N . Using Stirling’s formula for large N , N ! ≈ (2π)1/2 exp (−N )N N +1/2 , we
find that the minimum occurs when N ≈ x, and then the error
where F̂ ∗ (−k) = F̂ (k) for all real k, in order that F (x) is real-valued (the
superscript ∗ denotes the complex conjugate). Suppose that the Fourier
transform F̂ (k) and all its derivatives vanish as k → ±∞. Then we can
estimate F (x) as x → ±∞ by integrating (7) by parts. The outcome is
for all integers N . The asymptotic series (2) has zero coefficients, and we
infer that F (x) is exponentially small. For x > 0, the contour of integration
is moved into the upper half of the complex k-plane, that is Im k > 0. Then
if the nearest singularity of F̂ (k) to the real k-axis is at k = a + ib, b > 0 we
can estimate that F (x) = O(exp (−bx)) as x → ∞. An analogous procedure
74 R. Grimshaw
is used if x < 0, when the contour is moved into the lower half of the complex
k-plane.
For example, suppose that F̂ (k) = sech(γk) so that
∞
F (x) = sech(γk) exp (ikx) dk . (9)
−∞
2π π|x|
F (x) ∼ exp (− ). (10)
γ 2γ
Indeed, in this case the integral (9) can be explicitly evaluated to give
π πx
F (x) = sech( ) ,
γ 2γ
d2 w
= zw , (11)
dz 2
expressed here for a complex-valued function w(z) of the complex variable
z = x+iy. Note that on the real positive-axis, z = x > 0, we expect the two
linearly independent solutions to be either exponentiall growing or exponen-
tially decaying, but on the real negative-axis, z = x < 0, both solutions will
be oscillatory. The task is to connect these contrasting behaviours, and we
will show that this can be accomplished using ideas of exponential asymp-
totics in the complex z-plane. On the real axis z = x, the Airy function of
the first kind can be defined by
1 ∞ t3
Ai(x) = cos (xt + ) dt . (12)
π 0 3
This is real-valued on the real axis, and is extended to all complex z by
1 t3
Ai(z) = exp ( − zt) dt , (13)
2πi C 3
Note that, like Ai(z), Bi(z) (15) is real-valued on the real axis. Our interest
here is in the asymptotic expansions of Ai(z), Bi(z) in the complex z-plane.
These can be obtained from the integral (13) using the method of steepest
descent. Here we shall combine that with WKB asymptotic expansions
directly from Airy’s differential equation (11) in order to exhibit the Stokes
phenomenon. First, from the integral (13) we find that (using the saddle
point at t = z 1/2 )
1 2z 3/2
Ai(z) ∼ exp (− ), as |z| → ∞ , |arg(z)| < π . (16)
2π 1/2 z 1/4 3
which is a Riccati equation for φz . The WKB procedure for the limit |z| →
∞ yields the solutions
1
φz = ∓z 1/2 − + O(z −5/2 ) . (19)
4z
The outcome are the WKB asymptotic solutions,
1 2z 3/2
w1,2 = exp (∓ )(1 + O(z −3/2 )) . (20)
z 1/4 3
76 R. Grimshaw
1 2z 3/2
Ai(z) = a1 w1 (z) + a2 w2 (z) , where w1,2 ∼ exp (∓ ). (21)
z 1/4 3
The issue is to find the constants a1,2 . The key requirement is to find
the regions where w1,2 are either exponentially decaying or growing. These
regions are defined by anti-Stokes lines where the exponential is purely
oscillatory, while the Stokes lines are where the exponent is real-valued.
Here the anti-Stokes lines are arg = ±π/3, ±π (shown in red), and the
Stokes lines are arg = 0, ±2π/3 (shown in blue).
Figure 2. Plot of the ant-Stokes lines (red) and the Stokes lines (blue).
(16) and so a1 = 1/2π 1/2 . The final task then is to find the value of a2
in terms of a1 . The fact that a2 = 0 when we evaluate the asymptotic
expression (21) for Ai(z) as z → x < 0 is crucial. The evaluation of a2 can
be achieved by recourse to the integral expression (13), or more simply by
the requirement that Ai(z) should be real-valued when z = x < 0. This
latter yields a2 = ±ia1 as z = x < 0, arg(z) = ±iπ. Substitution into (21)
yields the asymptotic expression
1 2(−x)3/2 π
Ai(x) ∼ sin ( + ), as x → −∞ . (22)
π 1/2 (−x)1/4 3 4
The corresponding expressions for Bi(z) can now be found from the formula
(15). Thus, the counterparts of (16, 22) for Ai(z) are
1 2z 3/2 π
Bi(z) ∼ exp ( ), as |z| → ∞ , |arg(z)| < , (23)
π 1/2 z 1/4 3 3
1 2(−x)3/2 π
Bi(x) ∼ cos ( + ), as x → −∞ . (24)
π 1/2 (−x)1/4 3 4
where c is the constant wave speed, and F (x, t) is a specified forcing term.
Suppose F (x, t) = f (x) exp (−iωt) and seek a time-harmonic solution U (x, t) =
u(x) exp (−iωt). Then we get a forced harmonic oscillator equation for u(x),
where k = ω/c. We shall suppose that f (x) is localized, so that f (x) and
all its derivatives vanish as x → ±∞. In the absence of any forcing, the
general solution is
The radiation coefficients α± (k) are not zero, but are exponentially small
∗
as k → ∞. Note that α− = −α+ (the superscript ∗ denotes the complex
conjugate). Hence it is sufficient to evaluate α+ . The contour of integration
is moved into the lower half of the complex x̂-plane, that is Im x̂ < 0. Then
if the nearest singularity of f (x̂) to the real x̂-axis is at x̂ = K − iL, L > 0
we can estimate that α+ = O(exp(−kL)) as k → ∞. For example, fˆ(x̂) =
sech(γ x̂) has poles at x̂ = ±iπ/2γ, ±3iπ/2γ, · · · . Evaluating the residue at
the nearest pole to the real axis yields
π πk
α+ ∼ exp (− ) . (32)
ikγ 2γ
Exponential Asymptotics and Generalized Solitary Waves 79
Indeed, in this case the integral (31) can be explicitly evaluated to give
π πk
α+ = sech( ) ,
2ikγ 2γ
cannot easily be moved from the real axis. Instead, in this case, we can
evaluate the integrals explicitly to get that
π 1/2 k2
α± = exp (− 2 ) .
2ikγ 4γ
xt = −y , yt = x + buv . (34)
The variables (u, v, w) represent the large-scale flow, and the variables (x, y)
represent the internal gravity waves. The parameter b represents a Froude
number, and we shall consider the case when b << 1. When b = 0, the
system uncouples into a set of equations describing the slow manifold, a
resonant triad of large-scale planetary (Rossby) waves, and an equation for
fast oscillations, representing gravity waves. Hence, seek an asymptotic
expansion
∞
∞
(u, v, w) ∼ (un , vn , wn )bn , (x, y) ∼ (xn , yn )bn . (35)
n=0 n=1
80 R. Grimshaw
Thus there is no pure slow manifold, as gravity waves are inevitably gen-
erated as t → ∞. Note that the oscillations are exponentially small as the
parameter F → 0.
where the speed c = c(x) varies spatially. Now seek a time-harmonic solu-
tion U (x, t) = u(x) exp (−iωt) and then (41) becomes
(c−1/2 cX )X cXX c2
Q(X) == 1/2
= − X2 . (45)
2c 2c 4c
In the limit ω → ∞, the right-hand side of (2.3) can be ignored at
the leading order, assuming that Q(X) is a smooth function. Then WKB
solutions are found by putting
Q̃(X) Q
v± (X) = exp (±i[ωX − ]){1 + 2 + O(ω −3 )} . (49)
2ω 4ω
The general solution of (42) is
Now suppose that Q(X) (45) is a localized function such that Q(X) → 0
as X ± ∞. That is, the variable speed c(x) is such that c → c± as x → ±∞,
where c± are constants. Then consider a wave incident on the variable
medium from the left. That is, we impose the boundary conditions that
v ∗ vX − vvX
∗
= constant , (53)
|R|2 + |T |2 = 1 . (54)
Next, again from the the exact equation for v(X) (44) we can show that
Equation (55) is an integral equation for v(X), while the expressions (56)
allow the determination of the reflection coefficients. Since v(X) ∼ v+ (X) to
all orders O(ω −N ) (52), it follows that the reflection coefficient R can now be
obtained from (56). Keeping just the leading term for v+ (X) ∼ exp (iωX),
∞
1
R∼− exp (2iω X̂)Q(X̂)dX̂ . (57)
2iω −∞
Since we are assuming that Q(X̂) is a smooth function of X we can infer that
R is O(ω −N ), and is exponentially small, determined by the singularities of
Q(X̂) in the complex X̂-plane. This if the nearest singularity to the real
axis is at X̂ = K + iL, L > 0, then R = O(exp (−2ωL) as ω → ∞.
Now recall that
x ˆ
(c−1/2 cX )X cXX c2X d(x)
Q(X) = 1/2
= − 2
, X = ,
2c 2c 4c x0 c(x̂)
A balance is now required between the two small parameters, and this
depends on the form of f (x). Since we expect that the structure of the
84 R. Grimshaw
μ2 − −i
vqq + v − v 2 = 2 2 for case (a) , for case (b) , (63)
λ2 λ q λq
respectively. In this inner equation all terms must be in balance, and so
1 1 7 1 2
us ∼ (− + 4 + · · · ) + ( − 2 + · · · ) + O() . (68)
q2 q 3 3q
1
vqq + v − v 2 = − 2
1/2
∼ − 2 + + ··· . (69)
sinh ( q) q 3
Im v(q) = 0 , on Re q = 0 . (70)
1 1
so that v0qq + v0 − v02 = − .v1qq + v1 − 2v0 v1 = − . (72)
q2 3
Matching with the outer expansion (68) for us yields
1 7 1 2
v0 ∼ − + 4 + ··· , v1 ∼ − + ··· , (73)
q2 q 3 3q 2
1
v0qq + v0 − v02 = −
q2
Now seek an asymptotic solution
∞
bn
v0 ∼ , as q → ∞ , Re q > 0, Im q < 0 (74)
n=1
q 2n
Figure 3. The complex s-plane for Req > 0, Im q < 0. The line Re (sq) = 0
is blue, and the contour Γ is red.
The aim now is to sum the series (74) using Borel summation We seek a
solution of equation (72) in the form of a Laplace transform
v0 = exp (−sq) V (s) ds , (76)
Γ
where the contour Γ runs from zero to infinity in the complex s-plane such
that Re(sq) > 0.
Substitution of the Laplace transform into equation (72) for v0 yields a
Fredholm integral equation for V (s),
s
(s2 + 1)V (s) − V (ŝ)V (s − ŝ) dŝ = −s . (77)
0
Next, substitution of (78) into the Laplace transform (76) recovers the
asymptotic series (74) with
Thus solving the recurrence relation (79) for an effectively sums the asymp-
totic series (74) and yields the solution of equation (72) for v0 as a Laplace
transform. Examination of the recurrence relation (79)
n−1
(2j + 1)!(2n − 2j − 1)!
an + an−1 − aj an−j = 0, n = 1, 2, · · · .
j=0
(2n + 1)!
converges for |s| < 1. Analytic continuation into the complex s-plane, using
the integral equation (77), yields a complete solution for V (s) and hence
v0 (q) as the Laplace transform (76). )
Since an ∼ (−1)n K as n → ∞ we see that V (s) = ∞ n=0 an s
2n+1
has
singularities at s = ±i given by
Ks
V (s) ≈ , for|s| ≈ 1 . (81)
s2 + 1
There are similar poles at s = 2i, 3i, · · · , but these will be seen to generate
higher harmonics in the tail oscillations, and so are not our immediate
concern. We must know make a specific choice of the contour Γ in the
Laplace transform (76). Since we are seeking a symmeric solution, which
satisfies the condition (70), it is sufficient to suppose that at first Re q > 0
and Im q < 0. Then choose the contour Γ to lie initially in Re s > 0, Im s ≥
0, so that Re sq > 0 and the Laplace transform is well defined for the allowed
values of q. In particular, since V (s) can be represented by the power series
(78), which generates the asymptotic series (74), which in turn is equivalent
to the asymptotic series (66), we conclude that v0 ∼ vs = us .
88 R. Grimshaw
The next step is to deform the contour Γ onto the imaginary s-axis. in
this process we will need to deform around the poles at s = i, 2i, 3i, · · · and
collect the (half) residues. Hence we find that (76) becomes, on putting
s = iy on the deformed contour Γ,
∞
iπK
v0 = exp (−iyq) V (iy) i dy + exp (−iq) + · · · , (82)
0 2
where the dots denote terms proportional to exp (−2iq), exp (−3iq), · · · .
The integral is interpreted as a principal value integral at the singularities
at y = 1, 2, 3, · · · . This holds in Re q > 0, Im q < 0. To apply the symmetry
condition (70) that v0 should be real-valued on the imaginary q-axis, we
must now let Re q → 0, and put)q = −iQ, Q > 0 in the expression (82).
∞
From the series (78), V (iy) = i n=0 an (−1)n y 2n is pure imaginary since
the coefficients an are all real-valued. Hence the integral term is real-valued
as required. But the contributions from the poles are pure imaginary, and
hence the expression (82) cannot satisfy the symmetry condition. The rem-
edy is to note that the term exp (−iq) is exponentially small in the sector
Re q > 0, Im q < 0 and hence is subdominant, so that we are allowed to add
such terms asymptotically to (82).
Thus we replace (82) with
∞
iπK ib
v0 = exp (−iyq) V (iy) i dy + exp (−iq) + exp (−iq + iδ) + · · · ,
0 2 2
(83)
where b, δ are real constants. Now application of the symmetry condition
shows that
b cos δ = −πK . (84)
Thus the final solution for v0 in Re q > 0, Im q < 0 is
ib
v0 ∼ exp (−sq) V (s) ds + + exp (−iq + iδ) + · · · , (85)
Γ 2
where us ∼ sech2 (x) + O() is given by (66). The tail oscillations form a
one-parameter family characterized by the phase shift δ, 0 ≤ δ < π/2, where
the minimum amplitude occurs for δ = 0.
Suppose that instead of the symmetry condition (70) we look for one-
sided solutions such that u(x) → 0 as x → ∞. The same procedure can
be followed, and again the solution for Re q > 0 is given by the Laplace
transform (76), with the same contour Γ, which ensures that there are no
oscillations as x → ∞. But now, to find the behaviour as x → −∞, the
contour Γ must be moved across the imaginary q-axis to Re q < 0, and in
doing so the solution collects the residue at the poles s = i, 2i, 3i, · · · . The
residue at s = i is iπK exp (−iq), exactly twice the contribution from the
half residue shown in (82). Then, bringing the solution back to the real
x-axis, in x < 0, we find that
2πK π x
u ∼ us + exp (− 1/2 ) sin ( 1/2 ) . (87)
2
Again, for real x this asymptotic solution is symmetric, and no tail oscil-
lations emerge to all orders in n . The singularties of us in the complex
x-plane, are again located at x = ±iπ/2, ±3iπ/2, · · · , and we now write in
place of (67), see (62),
iπ
x= + q . (90)
2
Substitution into (89) and evalaution in the limit q → 0 yields
1 i 1 iq 1
us ∼ (− − 2 + · · · ) + ( − + · · · ) + O(3 ) . (91)
q q 6 3q
Here we replace x with q (90), and put v(q) = u(x) (62), so that
i i iq
vqq + v − v 2 = − ∼− + + ··· . (92)
sinh (q) q 6
i 1
so that v0qq + v0 − v02 = − .v1qq + v1 − 2v0 v1 = − . (94)
q 3
Matching with the outer expansion (91) for us yields
i 1 1 2
v0 ∼ − − 2 + · · · , v1 ∼ − + ··· , (95)
q q 3 3q 2
The aim now is to sum the series (96) using Borel summation We again
seek a solution of equation (94) in the form of a Laplace transform (76).
As before, we substitute the Laplace transform into equation (94) to get
now the Fredholm integral equation
s
(s2 + 1)V (s) − V (ŝ)V (s − ŝ) dŝ = −i . (98)
0
an ∼ Cn + O(n−1 ) , as n → ∞,
The essential difference here from case (a) is the double pole at s = i. This
leads to a half-residue of
v0 ∼ vs + vw , (104)
Then, as |q| → ∞,
for some real constants B, δ. The secular term arises, because in contrast
to case (a), vs ∼ q −1 rather than q −2 .
Then we choose the same contour Γ as in case (a), move the contour to
the imaginary q-axis, collect the half-residues and add the allowed subdom-
inant terms. The outcome is that (83) is replaced by
It remains to bring this solution back to the real x-axis, for x > 0, using
(90), x = iπ/2 + q. Here we must also collect a similar contribution from
the singularities in the lower half of the complex x-plane. The outcome will
be
u ∼ us + uw , (111)
where us ∼ sech(x) + O() is given by (89). Substitution into (88) and
linearization yields
2 uwxx + uw − 2uw us ≈ 0 . (112)
Just as for the corresponding equation (105) for vw , a WKB analysis yields
where both A(x), φ(x) are real-valued for real x. We find that
1
φx ∼ − us + O() , A ∼ A0 + O() , (114)
Using the leading expression for us ,
x
φ∼ − 2 tan−1 (exp (x)) + φ0 + O() . (115)
In order to find the constants A0 , φ0 , the WKB expression for uw (113)
should be matched with the corresponding expression for vw = uw (106)
when x = iπ/2 + q, for Re q > 0, Im q < 0. Thus, we find that
iπ q
φ∼ + q − π + i ln [tanh ( )] + φ0 + O() ,
2 2
so that, as |q| → ∞,
A0 q π
uw ∼ − exp ( − iq − iφ0 ) , (116)
2 2
noting that in Im q > 0, exp (−iq) is dominant over exp (iq). The matching
condition now shows that
2B π
A0 = − exp (− ) , φ0 = −δ . (117)
2 2
Finally we get that, as x → ∞, the tail oscillations are given by
4B π x
uw ∼ 2
exp (− ) cos ( − δ) , (118)
2
where B cos δ = −πC (109). Again, they form a one-parameter family.
94 R. Grimshaw
The linearized equations will then yield the linear dispersion relation
c = c(k) , (124)
written here for the phase speed c rather than the frequency ω = ck.
Whereas usually this dispersion relation (which may have several branches)
is considered as an equation for c given a real wavenumber k, for solitary
wave tails it needs to be considered as an equation for a complex-valued k
given a real speed c. Indeed, it is immediately clear that if there exist real-
valued solutions of (124) for the given value of c, then it is unlikely that the
solitary wave can decay to zero in its tail region. Instead, it will probably be
accompanied by a non-decaying co-propagating oscillatory wave field. This
consideration leads to the notion that solitary waves generally can only exist
in the gaps in the linear spectrum.
For instance, for water waves, the dispersion relation in the presence of
surface tension is
c2 (1 + Bq 2 )
= tanh q , q = kh , (125)
gh q
Exponential Asymptotics and Generalized Solitary Waves 97
∂ω ∂(ck) ∂c
cg = = =c+k
∂k ∂k ∂k
is the group velocity. Generically, there are four possibilities:
Case (1) corresponds to a KdV-type solitary wave, and case (3) corresponds
to an envelope (NLS) solitary wave. Case (2) corresponds to a so-called
generalized solitary wave, which does not decay at infinity, but instead is
accompanied there be small-amplitude co-propagating oscillations. Case
(4) has only rarely been studied and corresponds to a transition from a
KdV-type solitary wave to an envelope solitary wave.
The full system is now projected onto the appropriate four-dimensional
subspace, and the resulting bifurcation analyzed within the framework of
this subspace. Of course, rigorous results require a delicate and sophisti-
cated justification of this process. Here we shall instead briefly describe the
structure of the subspace, which we suppose is represented by the 4-vector
W(ξ). It satisfies an equation of the form
Wξ = L( W;
) + N ( W) . (126)
Here L( W;
) is a linear operator and N ( W) contains all the nonlinear
terms. The bifurcation parameter is
, and is such that the spectrum of L
at
= 0 reproduces one of the cases (1) to (4) describe above. That is, the
eigenvalues λ = ik of the linear operator L( W; 0) are respectively :
0 1
. (127)
0
Aξ = B + a1 A2 + a2 AB + a3 B 2 + O(α3 ) , (129)
Bξ =
A + b1 A2 + b2 AB + b3 B 2 + O(α3 ) . (130)
à = A + α1 A2 + β1 AB + γ1 B 2 , B̃ = B + α2 A2 + β2 AB + γ2 B 2 , (131)
Aξ = B + · · · ,
Bξ =
A + μA2 + · · · , (132)
2(c − c(0))
=− . (133)
ckk (0)
100 R. Grimshaw
It follows that for solitary waves, c > (<)c(0) according as ckk (0) < (>)0,
as expected. When the error terms in (132) are omitted, it becomes the
steady-state KdV equation (122) with the “sech2 ” solution (121) It is then
a delicate and intricate task to establish that this solitary wave solution
persists when the error terms are restored.
Aξ = B + · · · ,
Bξ =
A + μA2 + ν|C|2 · · · ,
Cξ = iγ(1 + δA)C + · · · . (135)
> 0 (when case (1) is a KdV-type solitary wave), the solution is a one-
parameter family of homoclinic-to-periodic solutions, with |C| = C0 con-
stant and (A, B) → (A0 , ) as ξ → ±∞ where A0 is a real constant, given
by
A0 + μA20 + νC02 = 0. The solution is a generalized solitary wave which
typically has a “sech2 ” core, and decays at infinity to non-zero oscillations
of constant amplitude C0 and wavenumber γ, see Figure 4. A delicate anal-
ysis of the full system (126) with the the small error terms shows that at
least two of these solutions persist; the minimal amplitude C0 being expo-
nentially small, that is O(exp (−K/|
|1/2 )) where K is a positive real con-
stant. Although such waves are permissible as solutions of the steady-state
equations, they have infinite energy and their associated group velocity is
Exponential Asymptotics and Generalized Solitary Waves 101
inevitably inward at one end and outward at the other end. Hence, they
cannot be realised in a physical system from any localized initial condition.
Instead localized initial conditions will typically generate a one-sided gener-
alized solitary wave, whose central core is accompanied by small-amplitude
outgoing waves on one side only. Such waves cannot be steady, and instead
will slowly decay with time.
Aξ = iβA + B + iAP (
, |A|2 , K) + · · · ,
Bξ = iβB + iBP (
, |A|2 , K) + AQ(
, |A|2 , K) + · · · . (137)
where K = i(AB ∗ − A∗ B) , (138)
P (
, |A|2 , K) =
+ ν1 |A|2 + ν2 K ,
Q(
, |A|2 , K) = 2
β + μ1 |A|2 + μ2 K (139)
eigenvalues, λ ≈ iβ + (2
β)1/2 have split off the imaginary axis, and so pro-
vide the conditions needed for exponential decay at infinity; the condition
μ1 < 0 depends on the particular physical system being considered.
The solution of the truncated system is
4
β
A = a exp (i[β +
]ξ)sech(γξ) , where γ = (2
β)1/2 , |a|2 = − . (142)
μ1
ut + 6uux + uxxx +
2 uxxxxx = 0 . (144)
When
= 0 this is the standard KdV equation. But when
= 0 it has a
linear dispersion relation
c = −k2 +
2 k 4 , (145)
and so there is a resonance at c = 0 between k = 0 and k = ±
−1 .
We seek solutions of the form
Im v(q) = 0 , on Re q = 0 . (154)
v0qq + v0 + 3v02 = 0
Now seek an asymptotic solution
∞
bn
v0 ∼ 2n
, as q → ∞ , Re q > 0 Im q < 0 . (158)
n=1
q
The aim now is to sum the series (158) using Borel summation We seek a
solution of equation (156) in the form of a Laplace transform
v0 = exp (−sq) V (s) ds , (160)
Γ
where the contour Γ runs from zero to infinity in the complex s-plane such
that Re(sq) > 0, see Figure 3.
Substitution of the Laplace transform into the equation (156) for v0
yields the Fredholm integral equation for V (s),
s
(s2 + 1)V (s) + 3 V (ŝ)V (s − ŝ) dŝ = 0 . (161)
0
(2n − 1)(2n + 6)
an + an−1 +
(2n + 2)(2n + 3)
n−1
(163)
(2j + 1)!(2n − 2j + 1)!
3 aj an−j = 0, n = 1, 2, · · · .
j=1
(2n + 3)!
106 R. Grimshaw
Next, substitution of (162) into the Laplace transform (160) recovers the
asymptotic series (158) with
Thus solving the recurrence relation (163) for an effectively sums the asymp-
totic series (158) and yields the solution of equation (156) for v0 as a
Laplace transform. Examination of the recurrence relation (163) shows
that as n → ∞, the nonlinear terms drop out, and so an ∼ (−1)n K where
K = −19.97 is a constant found numerically. Hence the series (162)
∞
V (s) = an s2n+1 .
n=0
converges for |s| < 1. Analytic continuation into the complex s-plane, using
the integral equation (161), yields a complete solution for V (s) and hence
v0 (q) as the Laplace transform (160). )
Since an ∼ (−1)n K as n → ∞ we see that V (s) = ∞ n=0 an s
2n+1
has
singularities at s = ±i given by
Ks
V (s) ≈ , for|s| ≈ 1 . (165)
s2 + 1
There are similar poles at s = 2i, 3i, · · · , but these generate higher harmon-
ics in the tail oscillations, and so are not our immediate concern. We must
know make a specific choice of the contour Γ in the Laplace transform (160)
v0 = exp (−sq) V (s) ds .
Γ
Since we are seeking a symmeric solution, which satisfies the condition (154),
it is sufficient to suppose that at first Re q > 0 and Im q < 0. Then choose
the contour Γ to lie initially in Re s > 0, Im s ≥ 0, so that Re sq > 0 and the
Laplace transform is well defined for the allowed values of q. In particular,
since V (s) can be represented by the power series (162), which generates
the asymptotic series (158), which in turn is equivalent to the asymptotic
series (148), we conclude that v0 ∼ vs =
2 us .
The next step is to deform the contour Γ onto the imaginary s-axis. In
this process we will need to deform around the poles at s = i, 2i, 3i, · · · and
collect the (half) residues. Hence we find that (160) becomes, on putting
s = iy on the deformed contour Γ,
∞
iπK
v0 = exp (−iyq) V (iy) i dy + exp (−iq) + · · · , (166)
0 2
Exponential Asymptotics and Generalized Solitary Waves 107
where the dots denote terms proportional to exp (−2iq), exp (−3iq), · · · .
The integral is interpreted as a principal value integral at the singulari-
ties at y = 1, 2, 3, · · · . This holds in Re q > 0, Im q < 0. To apply the
symmetry condition (154) that v0 should be real-valued on the imaginary
q-axis, we must now let Re q → 0, and put q = )−iQ, Q > 0 in the ex-
∞
pression (166). From the series (162), V (iy) = i n=0 an (−1)n y 2n is pure
imaginary since the coefficients an are all real-valued. Hence the integral
term is real-valued as required. But the contributions from the poles are
pure imaginary, and hence the expression (166) cannot satisfy the symmetry
condition. The remedy is to note that the term exp (−iq) is exponentially
small in the sector Re q > 0, Im q < 0 and hence is subdominant, so that we
are allowed to add such terms asymptotically to (166).
Thus we replace (166) with
∞
iπK
v0 = exp (−iyq) V (iy) i dy + exp (−iq)
0 2
(167)
ib
+ exp (−iq + iδ) + · · · ,
2
where b, δ are real constants. Now application of the symmetry condition
shows that
b cos δ = −πK . (168)
Thus the final solution for v0 in Re q > 0, Im q < 0 is
ib
v0 ∼ exp (−sq) V (s) ds + + exp (−iq + iδ) + · · · , (169)
Γ 2
b π x
u ∼ us + 2
exp (− ) sin ( − δ) , b cos δ = −πK , (170)
2
γ
et al (1988) (for the case δ = 0), and with the numerical solutions of Boyd
(1991). Amick and Toland (1992) have established theoretically that the
nonlocal solutions of the fourth order ordinary differential equation (147)
form a one-parameter family homoclinic to periodic solutions for (147).
∞
v(q) ∼ vn (q)
2n ,
n=0
and examine the effect of the next term v1 , where we recall from (152) that
the matching condition is that v1 → 2γ 3 /3 as q → ∞ in Re q > 0, Imq < 0.
First, we re-examine the tail oscillations and seek a solution of (147) of the
form
u ∼ us + uw , (172)
Exponential Asymptotics and Generalized Solitary Waves 109
where us is the outer expansion (148) and uw are the tail oscillations. Sub-
stitution into (147) and linearization about us yields
kx
uw ∼ α sin ( − δ) , k 4 − k 2 =
2 c . (174)
kx
uw ∼ α(1 +
2 R(x)) sin ( − δ +
φ(x)) . (175)
2γ 2
v1 = + w1 , (183)
3
It is then readily shown that w1 satisfies the same equation (179 as that for
w0 , and hence can be absorbed into the same solution (180) by allowing the
Exponential Asymptotics and Generalized Solitary Waves 111
Hence the main effect of the higher-order terms is to replace k = 1 with the
full expression (174) for k.
Figure 8. Plot of a typical set dispersion curves for internal waves for the
phase speed c = c(k) in terms of wavenumber k: mode 1 (blue), mode 2
(red), mode 3 (green) and the surface mode (violet).
(1988) and Kruskal and Segur (1991). The full system was analyzed us-
ing this approach by Akylas and Grimshaw (1992). Here, for simplicity,
we consider instead two coupled Korteweg-de Vries (KdV) equations, which
can be shown to describe the interaction between two weakly nonlinear long
internal waves whose linear long wave speeds are nearly equal,
1
ut + 6uux + uxxx + (pvxx + quv + rv 2 )x = 0 , (186)
2
1
vt + Δvx + 6vvx + vxxx + λ(puxx + ruv + qu2 )x = 0 . (187)
2
speed c = c(k),
Δ Δ2 1/2
c= − k2 ± {λp2 k 4 + } . (188)
2 4
If we let the coupling parameter λ → 0 these linear modes uncouple into a
u-mode with spectrum c = −k2 and a v-mode with spectrum c = Δ − k 2 .
This situation persists for λ > 0, and there is a resonance between the long
wave (u-mode, in red) and a short wave (v-mode, in blue), with a resonant
wavenumber k0 = (Δ/1 − λp2 )1/2 provided that λp2 < 1, see Figure 9.
form an outer boundary condition. The outcome is just the same system
(190, 191) with x replaced by z. Note that c = O(
2 ), and can be omitted
at the leading order.
We seek a solution of this inner problem as a Laplace transform
[u, v] = exp (−zs)[U (s), V (s)] ds , (200)
Γ
where the contour Γ runs from 0 to ∞ in the half-plane Re(sz) > 0, see
Figure 3 and note that here z replaces q. We seek a power series solution
∞
[U (s), V (s)] = [an , bn ]s2n−1 , (201)
n=1
This agrees with the asymptotic series (198, 199), and in effect the Laplace
transform is a Borel summation of the asymptotic series.
Substitution of the Laplace transform (200) and the series (201) into
the differential equation system (190, 191) yields a recurrence relation for
[an , bn ]. Putting Δ[An , Bn )] = (−k02 )n [an , bn ], we get
(n + 1)(2n + 5) q
An−1 + {p − }Bn−1 = Fn , (203)
(n − 1)(2n − 1) (n − 1)(2n − 1)
rBn−1 + qAn−1
(1 − λp2 )Bn − Bn−1 − λpAn−1 + λ = Gn , (204)
(n − 1)(2n − 1)
where Fn , Gn are quadratic convolution expressions in A2 , · · · An−2 , B2 · · · ,
Bn−2 . As n → ∞, these nonlinear terms can be neglected, and we find that
where [U (s), V (s)] has a pole singularity at s = ik0 , also at the complex
conjugate point s = −ik0 and at all their harmonics s = ±ink0 , n = 2, 3
etc. Hence the contour Γ should be chosen to avoid the imaginary s-axis,
and to be explicit we choose it to lie in Re s > 0. But we seek a symmetric
solution, which in the z-variable requires that Im [u.v] = 0 when Re z = 0.
But the presence of the pole prevents (200) from satisfying this condition,
and so, as in section 5, we must correct it by adding a subdominant term
ib
[u, v] = exp (−zs)[U (s), V (s)] ds + [p, −1] exp (−ik0 z + iδ). (207)
Γ 2
Here b, δ are real constants, and note that | exp (−ik0 z)| is smaller than of
z −n as z → ∞ in Re z > 0 , Im z < 0, recalling that x = (iπ/2
γ) + z. The
symmetry condition is now applied by bringing the contour Γ onto Re s = 0
and deforming around the pole at s = ik0 .
The outcome is
b cos δ = πK . (208)
which is substituted into
ib
[u, v] = exp (−zs)[U (s), V (s)] ds + [p, −1] exp (−ik0 z + iδ).
Γ 2
The final step is to bring this solution (207) back to the real axis, using x =
(iπ/2
γ) + z. Taking account of the corresponding singularity at s = −ik0 ,
we finally get that
[u, v] ∼ [us , vs ] + bΔ[−p, −1] exp(−πk0 /2
γ) sin(k0 |x| − δ) . (209)
Here we recall (193) that us ∼ 2
2 γ 2 sech2 (γ
x), vs ∼ O(
4 ). This is a two-
parameter family, the parameters being
γ, δ, 0 < δ < π/2. The minimum
tail amplitude occurs at δ = 0. Note that the constant in the exponential
term is determined by the location of the singularity, but the amplitude
needs the exponential asymptotics.
The constant K is determined by the recurrence relations (203, 204).
It is a function of the system parameters λ, p, q, r and in general is found
numerically. But K = 0 for q = 6p (see (193, 194)), and in general it
was found by Grimshaw and Cook (1996) that there are many parameter
combinations where K = 0, see their Figures 1-4. In particular
λ(6p − q)
K≈ as λ → 0. (210)
3Δ
Exponential Asymptotics and Generalized Solitary Waves 117
These special values imply that the solitary wave decays to zero at infin-
ity, even although its speed lies inside the linear spectrum, at least in this
asymptotic limit. These are called embedded solitons. They are usually
not stable, but are instead metastable, or are said to exhibit semi-stability.
Nevertheless they are found useful in several applications, such as nonlin-
ear optics and solid state physics. For water waves with surface tension,
generalized solitary waves exist for Bond numbers 0 < B < 1/3, but from
numerical simulations it seems there are no embedded solitons.
These symmetric solitary waves cannot be realized in practice, since they
require an energy source and sink at infinity. Instead, they are replaced
by solitary waves with radiating tails on one side only, determined by the
group velocity. That is, in x > 0 for cg > c, or in x < 0 for cg < c,
where cg is the group velocity at the resonant wavenumber. For the present
case, the linear dispersion relation is (188) and so for the relevant u-mode,
cg = Δ − 3k2 < c = Δ − k 2 . Hence there are no oscillations in x > 0, but
they will appear in x < 0. Thus, in this case for one-sided oscillations, in
x > 0, or more generally in Re z > 0, the solution is completely defined by
the Laplace transform integral (200), with the contour Γ lying in Re s > 0.
Then for x < 0, or Re z < 0, the contour Γ must be moved to Re z < 0
across the axis Re s = 0. In this process the solution collects a contribution
from the pole at s = ik0 , which generates the tail oscillation. The final
outcome is that (209) is replaced by
where H(·) is the Heaviside function. That is, in effect the phase shift δ = 0,
there are no oscillations in x > 0 and the amplitude in x < 0 is exactly twice
the amplitude of the symmetric solution.
q
(Δ − c0 )v1 + v1xx + pu0xx + u20 = 0 . (215)
2
u20
(Δ − c0 )v1 + v1xx = f (x) = −pc0 u0 + (6p − q) . (216)
2
Note that in this limit λ → 0, the resonant wavenumber k0 ≈ (Δ−c0 )1/2 and
takes account of the finite speed of the wave. We must now take c0 < Δ to
get tail oscillations, and for c0 > Δ the expansion yields a genuine solitary
wave. The general solution of (216) is
∞
1
v1 = A sin k0 x + B cos k0 x + f (x ) sin (k0 |x − x |)dx . (217)
2k0 −∞
(Δ − c0 )
u1 ∼ −p b1 sin (k0 |x| − δ) , as |x| → ∞ , (220)
Δ
(Δ − c0 )
[u1 , v1 ] ∼ [−p , 1] b1 sin (k0 |x| − δ) , as |x| → ∞ ,
c0
v1 ∼ b1 sin (k0 |x| − δ) as|x| → ∞ ,
∞
1
b1 cos δ = L = f (x) cos (k0 x)dx ,
2k0 −∞
u20
f (x) = −pc0 u0 + (6p − q) .
2
We find that
∞
6k0 2
L=− {k (q − 6p) + 4β 2 q} sech2 (βx) cos (k0 x)dx . (221)
β2 0 −∞
Then as β =
γ → 0, this reduces to
πk02
L∼ (6p − q) exp (−πk0 /2
γ) , (222)
3
Exponential Asymptotics and Generalized Solitary Waves 119
which agrees with the previous result (210) from the exponential asymp-
totics, since L = πK. The one-sided solutions are obtained by setting
δ = 0, and replacing b1 in (218, 220) by 0, 2b1 for x > 0, x < 0.
Bibliography
T. R. Akylas and R. H. J. Grimshaw Solitary internal waves with oscillatory
tails J.Fluid Mech., vol 242, 279-298, 1992.
T. R. Akylas and T-S. Yang On short-scale oscillatory tails of long-wave
disturbances Stud. Appl. Math., vol 94, 1-20, 1995.
C. Amick and J. Toland Solitary waves with surface tension I: trajecto-
ries homoclininic to periodic orbits in four dimensions Arch. Rat. Mech.
Anal., vol 118, 37-69, 1992.
C.M. Bender and S. A. Orzsag, Advanced mathematical methods for scien-
tists and engineers, McGraw-Hill, (ISBN 0-07-066173-1) 1978,
M.V. Berry and C. J. Howls, Hyperasymptotics, Proc. Roy. Soc., vol A 430,
653-668,1990.
M. V. Berry, Asymptotics, superasymptotics, hyperasymptotics, “Asymp-
totics beyond all orders” , ed. H Segur and S Tanveer, Plenum, 1-14,
1992.
J. Boyd Weakly nonlocal solitons for capillary-gravity waves: fifth-degree
Korteweg-de Vries equation Physica D, vol 48, 37-69, 1991.
J. P. Boyd Weakly nonlocal solitary waves and beyond-all-order asymptotics,
Kluwer, Amsterdamr, (ISBN 0-7923-5072-3), 1998.
J.P. Boyd The devils invention: asymptotic, superasymptotic and hyper-
asymptotic series Acta Applicandae Mathematicae, vol 56, 1-98, 1999.
F. Dias and G. Iooss Water waves as a spatial dynamical system Handbook of
Mathematical Fluid Dynamics, ed, S.Friedlander and D.Serre Elsevier,
North Holland, Chapter 10, 443 -499, 2003.
R. Grimshaw The reflection of internal gravity waves from a shear layer
Quart. J. Appl. Math., vol 29, 511-525, 1976.
R. Grimshaw Solitary Waves in Fluids, Advances in Fluid Mechanics, ed.
R. Grimshaw, vol 47, WIT press, UK, Ch. 1 1-17, Ch. 7 159-179, 2007.
R. Grimshaw and P. Cook Solitary waves with oscillatory tails Proceed-
ings of Second International Conference on Hydrodynamics, Hong Kong,
1996, ”Hydrodynamics: Theory and Applications”, Vol. 1, ed. A.T.
Chwang, J.H.W. Lee and D.Y.C. Leung, A.A. Balkema, Rotterdam, 327-
336, 1996.
R. Grimshaw and G. Iooss Solitary waves of a coupled Korteweg-de Vries
system Mathematics and Computers in Simulation, vol 62, 31 - 40. 2003.
120 R. Grimshaw
Philippe H. Trinh
University of Oxford, UK
1 Introduction
The Stokes Phenomenon describes the puzzling event in which exponen-
tially small terms can suddenly appear or disappear when an asymptotic
expansion is analytically continued across key lines (Stokes lines) in the
Argand plane—“as it were into a mist,” Stokes once remarked in 1902.
Fortunately, much of the inherent vagueness of this phenomenon, as well
as its deep implications for the study of asymptotic approximations has been
examined since Stokes’ time (see Boyd (1999) for a comprehensive review).
In another paper of this volume by Grimshaw—henceforth referred to as
[Grimshaw]—it was shown how Borel summation can be used to reveal
the exponentially small waves found in the fifth-order Korteweg-de Vries
equation (5KdV).
In this review paper, we will show how the methodology outlined in
Olde Daalhuis et al. (1995) and Chapman et al. (1998) can be used as an
alternative treatment of the 5KdV equation. The procedure is as follows:
(1) Expand the solution as a typical asymptotic expansion, (2) find the
behaviour of the late-order terms (n → ∞), and (3) optimally truncate the
expansion and examine the remainder as the Stokes lines are crossed.
The location of the Stokes lines, as well as the details of the Stokes
Phenomenon and resultant exponentials are intrinsically linked to the late-
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
122 P. Trinh
u(x)
|S|
(x)
- 10 -5 5 10
η
x = πi/2γ
672.(6/,1(
(x)
x = −πi/2γ
into Equation (1). This yields the first two orders as,
while at O(2n ),
Q(z)Γ(2n + γ)
un ∼ , as n → ∞. (5)
[χ(z)]2n+γ
Here, γ is a constant, while Q(z) and χ(z) are functions to be determined.
Substituting this ansatz into Equation (4) yields at leading order,
4
2
dχ dχ
− + = 0, as n → ∞. (6)
dz dz
Now from the above discussion, we would expect that χ = 0 at the relevant
singularities, x = σi for some i; we then conclude that χ = ±1 and thus
without loss of generality, χ = x − σi . In general, un will be a sum of
terms of the form (5), one for each singularity. However along the real
axis, the behaviour of un will be dominated by those singularities closest to
124 P. Trinh
the axis and thus we need only concern ourselves with the singularities at
x = ±σ ≡ ±iπ/2γ. Finally, at next order as n → ∞, we find that Q(z) = Λ,
a constant.
The determination of γ, Λ, and in fact, the Stokes line smoothing in
the next section will require an analysis near each of the two singularities,
x = ±σ; for brevity, we will henceforth focus on the singularity at x = σ in
the upper-half plane.
First, since by Equation (2), u0 ∼ −2/(x − σ)2 as x → σ, we must
require that γ = 2. Second, in order to determine the final constant Λ, we
need to re-scale near the singularity x = σ, express the leading-order inner
solution as a power series (in inner coordinates) and match with the outer
solutions. In the end, however, Λ is determined by the numerical solution
of a canonical inner problem. As was shown in [Grimshaw], Λ ≈ −19.97.
Finally, let us discuss the significance of χ. Following Dingle (1973), we
expect there to be a Stokes line wherever un and un+1 have the same phase
as n → ∞, or in this case where,
−χ2 = 0 and −χ2 ≥ 0. (7)
Thus there exist Stokes lines from x = πi/2γ down the imaginary axis and
from x = −πi/2γ up the imaginary axis (as illustrated in Figure 1). In
the next section, we will optimally truncate the asymptotic expansion and
examine the switching-on of exponentially small terms as these two Stokes
lines are crossed.
N −1
u= 2n un + RN (x).
n=0
2 R N
+ RN + 6u0 RN − c0 RN + . . . ∼ 2N uN , (8)
d ie−iθ d
χ = x − σ = reiθ and =− , (11)
dx r dθ
where now, since N is fixed (and thus also the modulus, r), we are only
interested in the “fast” variation in θ across the Stokes line. Then using
Equations (9) and (11) in (8) gives
√ r/+2ρ+γ+2 r/+2ρ
dS Λ rπ iθ
∼√ e−r/ eire / e−iθ e−πi/2 eiθ
dθ 2γ+1/2
√ ,
Λ rπ r iθ πi
=√ × exp − 1 − ie + iθ +
2γ+1/2 2
- π .
+i −2ρ θ + − θ(γ + 1) (12)
2
From the terms within the curly braces, we see that the change in S is
exponentially small, √except near the Stokes line θ = −π/2. Here, we will
re-scale θ = −π/2 + η and integrate Equation (12) from left (η → −∞)
to right to show that
√ √rη
Λ π πi(γ+1)/2 2
S ∼ const + √ e e−s /2 ds. (13)
2γ
−∞
This integral (the error function) precisely illustrates the smoothing of the
Stokes line in Figure 1. Thus the jump in the Stokes multiplier and conse-
quently, the remainder is
126 P. Trinh
Λπ 3πi/2 Λπ
S ∼ e =⇒ R N ∼ 2 e3πi/2 e−i(x−σ)/ . (14)
Stokes γ Stokes
We must remember that the analysis must be repeated for analytic continua-
tion into the lower-half x-plane and thus near the singularity at x = −πi/2γ.
The result is another exponentially small contribution which is the complex
conjugate of Equation (14) and thus along the real axis, the sum of contri-
butions from crossing the pair of Stokes lines is,
2Λπ −π/2γ
uexp ∼ − e sin (x/) . (15)
2
Let us recap our analysis: (1) The singular nature of the 5KdV equation
produces singularities in the early terms, (2) As more and more terms are
taken, the effects of the singularities grow, eventually producing factorial
over power divergence in the late terms, (3) Stokes lines emerge from each
of the singularities, and (4) By optimally truncation and examining the
jump in the remainder as the Stokes lines are crossed, we see the Stokes
Phenomenon and thus the appearance of exponentially small terms.
So finally, we are ready to answer the original question: Do there exist
classical solitary wave solutions of the 5KdV equation? No. For suppose
that we did impose the condition that only the base (non-oscillatory) asymp-
totic solution applies at x = −∞. Then as we pass through x = 0, the term
in Equation (15) necessarily switches on and u ∼ u0 + uexp for x > 0.
We have thus passed through Stokes’ mist and subsequently, realized
that there do not exist classical solitary wave solutions of the 5KdV.
Bibliography
J.P. Boyd. The Devil’s invention: Asymptotics, superasymptotics and hy-
perasymptotics. Acta Applicandae, 56:1–98, 1999.
S.J. Chapman, J.R. King, and K.L. Adams. Exponential asymptotics and
Stokes lines in nonlinear ordinary differential equations. Proc. R. Soc.
Lond. A, 454:2733–2755, 1998.
R.B. Dingle. Asymptotic Expansions: Their Derivation and Interpretation.
Academic Press, London, 1973.
A.B. Olde Daalhuis, S.J. Chapman, J.R. King, J.R. Ockendon, and R.H.
Tew. Stokes Phenomenon and matched asymptotic expansions. SIAM
J. Appl. Math., 55(6):1469–1483, 1995.
G.G. Stokes. On the discontinuity of arbitrary constants that appear as
multiples of semi-convergent series. Acta. Math. Stockholm, 26:393–397,
1902.
Multiple scales methods in meteorology
1 Overview
In Chapter 2 we will derive the fluid mechanical conservation laws. We
explore the basic principles considering “pure” fluid mechanics, i.e., we ne-
glect the influences of gravity, Earth’s rotation (Coriolis force), molecular
transport, and of the so called “diabatic effects”. The latter subsume all
processes that involve external energy supply by radiation or conversion
of energy due to condensation, chemical reactions, etc.. Gravity and the
Coriolis force will be included in subsequet sections. The chapter concludes
with a summary of the governing equations, now extended to also include
a general set of species transport equations. These will be important, e.g.,
in describing (atmospheric) chemistry or moist processes.
Chapter 3 introduces the technique of multiple scales asymptotics using
the (almost trivial) example of a linear oscillator. After deriving analyti-
cal solutions, we will focus on a situation which, in many ways, resembles
situations arising frequently in geophysical problems: a slow background
motion caused by an external force is accompanied by rapid oscillations
around it, with the oscillation amplitudes generally not being small. To
give some meaning to the notions of “smallness” and “rapidity”, we will
first nondimensionalize the oscillator equations and identify small parame-
ters that lend themselves for comparison. By means of single and multiple
scales analyses we will then try to derive simplified approximate solutions
that become more and more accurate as the small parameters vanish.
One important aim of theoretical meteorology is the development of sim-
plified model equations that describe the large variety of scale-dependent
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
128 R. Klein et al.
∂Ω
Δt v
Ω
Δσ
Ω n
v Δt (v · n)
n
This mass will change during a time interval Δt if mass parcels cross in-
terface ∂Ω, being carried along by the flow velocity, v (see Fig. 1). The
change of mass, ΔM , associated with the passage of parcels across a con-
trol surface segment Δσ ⊂ ∂Ω, is then equal to − (Δt v) · n Δσ, where n is
the outward pointing normal on Δσ. The scalar multiplication v · n selects
the component of (Δt v) perpendicular to Δσ as the one relevant for mass
transport across the surface element. By summation (integration) along the
entire boundary of the control volume, and for time increments covering a
finite interval t ∈ [t1 , t2 ] we find
t2
M (t2 ; Ω) − M (t1 ; Ω) = − (v) · n dσ dt . (2)
t1 ∂Ω
This is the most general formulation of the law of mass conservation which
holds for arbitrary control volumes for which the integrals in (1), (2) are
meaningfully defined. Notice that the above definitions merely require suit-
able integrability for the mass and momentum densities, and v. These
quantities need not be differentiable in either space or time for the mass
balance in (2) to make sense!
If, however, we may assume differentiability of M (t; Ω) w.r.t. time, t, we
may let (t2 − t1 ) → 0 to find
dM
=− (v) · n dσ . (3)
dt ∂Ω
130 R. Klein et al.
This equation can hold, for continuously differentiable fields , v and for ar-
bitrary control volumes Ω, only if pointwise the following partial differential
equation is satisfied:
t + ∇ · (v) = 0 . (5)
Adopting this differential form restricts solutions to the class of continu-
ously differentiable fields. Yet, in practice one uses (5) as a short-hand
for (2), thereby implying that whereever t and ∇ · (v) are singular, their
spacio-temporal integrals remain well defined. For discussions of such weak
solutions of conservation laws see, e.g., LeVeque (1990) and Kröner (1997);
for a measure theoretical approach to conservation laws see Temam and
Miranville (2000).
If, in addition, u and f are sufficiently smooth, then they satisfy the
Partial Differential Equation in Conservation Form
ut + ∇ · f = 0 . (9)
Multiple Scales Methods in Meteorology 131
Remark: The integral form of the conservation law in (8) is the most general
basis for the formulation of numerical methods for problems in continuum
mechanics because, by construction, it allows for the correct representation
of non-smooth, e.g., discontinuous, solutions. See, e.g., LeVeque (1990),
Kröner (1997).
Remark: Partial differential equations which, in addition to divergence
terms, include other expressions that have no equivalent divergence form
cannot be cast in the more general integral form (8). Such equations do not
describe conservation in the present sense.
where e is the total energy per unit mass, or specific total energy, and is
again the mass density.
Energy is transported by the motion of fluid parcels in analogy with the
flux of mass considered in section 2.1. The associated contribution to the
energy flux density is e v. In addition, there are forces acting within a
fluid between adjacent fluid parcels. Those forces are represented by means
of a second order tensor field (p id + τ ), where p is the thermodynamic
pressure, and id, τ (t, x) ∈ IR3×3 are the unit tensor and the viscous stress
tensor, respectively. The interpretation of this tensor, (p id + τ ), and its
two contributions to the energy flux is as follows:
Consider the boundary, ∂Ω, of a control volume (or some similar surface
embedded in the flow domain). At any location x ∈ ∂Ω the vector (p id +
τ ) · n, with n the outward pointing normal on ∂Ω, denotes the force per
unit area, i.e., the stress, which the fluid within the control volume exerts
onto the fluid outside. If the fluid is in motion, with flow velocity v(t, x),
then v · (p id + τ ) · n is the work per unit time and unit area done by the
fluid inside the control volume on the fluid outside. This is the second
contribution to the energy flux density to be considered here. It consists
132 R. Klein et al.
of (i) the work done by the thermodynamic pressure forces, p v · n, and (ii)
the work done by the viscous stresses, v · τ · n.
We know from experience that two bodies of finite mass and different
temperature tend to exchange thermal energy so as to eventually approach
states of equal temperature. Let us denote the associated energy flux per
unit area by j.
Combining the three effects just discussed we obtain the energy conser-
vation law, written here in its differential form as
(e)t + ∇ · ([e + p] v + v · τ + j) = 0 . (11)
In much of the subsequent discussions we will neglect the terms, (v · τ + j),
which are associated with molecular transport processes, for simplicity of
exposition.
Remark: Because of the shape of the earth, g actually varies at sea level
around ±0.3 percent in north-south direction and around 0.3 percent with
a change of height of 10 km Gill (1982).
To account for the influence of gravity, we must endow the momentum
balance in (13) with a source term,
In the energy balance, (11), we must account for the potential energy as-
sociated with the position of the fluid mass in the Earths gravity field. This
is done by extending the constitutive law from (17), to include a potential
energy term, viz.,
v 2
e = (eth + ekin + epot ) = cv T + + Φ(x) . (23)
2
Here Φ(x) is the Earth’s geopotential, which in the present setting (tangent
plane approximation) we approximate by
Φ(x) = gz (24)
or
− g k = −∇ · Πhy . (27)
ΔX b
ΔΘ
X b (t + Δt)
|X b | sin((Ω, X b ))
X b (t)
(Ω, X b )
Δt → 0 we find
ΔX b dX b dθ Ω × X b
lim = = |X b | sin((Ω, X b )) , (31)
Δt→0 Δt dt dt |Ω × X b |
Ẋ b = Ω × X b . (32)
Both observers see the same vector X b but their perception of how it
changes is completely different.
Remark: The length of X b is constant, independent of the used coordinate
system. Because of X b ⊥(Ω × X b ) it follows that
d|X b |2 dX b
= 2 Xb · = 2 X b · (Ω × X b ) = 0 . (33)
dt dt
Thus we get
Here vectors and tensors are represented by their coordinate tupels and
matrices as indicated, and u is the column of cartesian coordinates of u.
With this notation, the mass balance in the intertial frame reads
t + ∇ · (v) = 0 . (39)
˜ )T we have
For the scalar quantity x̃T Ω(∇˜
T
˜ )T = x̃T Ω(∇˜
x̃T Ω(∇˜ ˜ )T ˜ ˜)ΩT x̃
= (∇ (42)
Thus the equation for mass conservation is invariant under the present co-
ordinate transformation into a rotating frame.
The momentum balance, in the present notation, reads
T
(v)t + ∇ · ( v v T ) + ∇p = −∇Φ − ∇ · τ . (44)
As only the first two terms do change under coordinate transformations
(the reader may want to verify this), we can neglect the others for the time
being. For the first two terms we use the product rule to obtain
(v)t = t v + v t and (45)
T
∇ · (v v T ) = (∇ · (v))v + (v T ∇T )v . (46)
Using mass conservation, we further have
T
(v)t + ∇ · (v v T ) = (t + ∇ · (v)) v + v t + (v T ∇T )v
= 0 + vt + (v T ∇T )v ,
(47)
and the transformation of these terms into the rotating coordinate system
yields
∂ ṽrel rel rel rel ˜ = −∇Φ− ˜ ˜ · τ̃ .
+ ((ṽ ) ∇ )ṽ + 2 (Ω ṽ ) + Ω(Ω x) +∇p
T T
∇
∂t
(48)
rel
Physically, the term 2 (Ω ṽ ) represents the Coriolis acceleration.
The term Ω(Ω x) expresses the centripetal acceleration due to the rota-
tion of the reference frame. As this term can be written as the density times
the gradient of a potential (namely which one?), it is often combined with
gravity term, thereby inducing a modified effective gravitational potential.
The order of magnitude of the centripetal inertia may be estimated by
m
(10−4 s−1 )2 · 6 · 106 m ≈ 10−2 2 . (49)
s
In contrast, the acceleration of gravity is of the order of g ≈ 10 m s−2 , and
the centripetal acceleration may be neglected for most practical purposes in
meteorology. Notice, however, that this may be a different issue in climate
models, because in long-time simulations, even small effects can accumulate
and eventually induce leading-order changes.
Like the equation for the mass conservation, the equation of energy con-
servation does not change when introducing a rotating coordinate system.
We leave the verification of this claim to the reader.
140 R. Klein et al.
is the time derivative which an observer moving with the fluid (at velocity
v) will measure.
The above is equivalent to
DΘ (p/p0 )1/γ
=0 where Θ = T0 (55)
Dt /0
t + ∇ · (v) = 0 (56a)
where
nspec nspec
cv (T ) = Yi cv,i (T ) , R= Yi R i , (58)
i=1 i=1
and T
hi = cp,i (T )dT + Qi . (59)
0
Here the constants Ri , Qi are the gas constants and formation enthalpies
of the species, cv,i , cp,i are their specific heat capacities at constant volume
and at constant pressure, respectively.
In addition, we have to adopt appropriate expressions for the stress ten-
sor, heat flux density, and species diffusion fluxes, τ , j, and d, respec-
tively. For a Navier-Stokes fluid, the former two are given by (19) and
(20). However, in practical meteorological modelling applications involving
turbulence, one often replaces these fluxes with effective turbulent closure
schemes so as to describe not the transport due to molecular motions but
rather the transport due to turbulent fluctuations. In that case, the func-
tional form of these “subscale” momentum and heat flux terms may take a
wide variety of forms which we will not address here in detail.
FF = −cx
mẍ
FD (t)
(Demon) FR = −k ẋ
There are two forces acting within the system: the restoring force of the
spring in the direction opposite to the displacement of the mass, and the
frictional force FR = −k ẋ of the viscous fluid which acts in the direction
opposite to the motion of the mass. Newton’s law, which says that the
temporal change of momentum equals the sum of all acting forces, yields
an equation of motion for the system,
mẍ = −cx − k ẋ + FD (t) . (60)
Here we have included a general external force FD = FD (t) which some
demon may exert on the mass.
The solution to this second-order ordinary differential equation (ODE)
is uniquely determined once initial conditions x(0) = x0 and ẋ(0) = ẋ0
are prescribed. Following the theory of linear ODEs [Walter (1996)] we
can construct all solutions x(t) as a superposition of the general solution
of the homogenous problem and one so-called particular solution of the
inhomogenous equation, i.e.,
x(t) = xh (t) + xp (t) . (61)
Free oscillations We will now derive the general solution of the spring-
mass system’s homogenous ODE ((60) with FD ≡ 0),
mẍ + k ẋ + cx = 0 . (62)
To solve this equation, we choose an exponential ansatz,
x(t) = exp(ωt) . (63)
Inserting, we have
m ω2 exp(ωt) + k (ω exp(ωt)) + c (exp(ωt)) = 0 , (64)
and after division by exp(ωt) = 0 we find
mω 2 + kω + c = 0 , (65)
which is the system’s characteristic equation. The solutions are
1k √ k2 c
ω1/2 = − ± D with D := − . (66)
2m 4m2 m
If ω1 = ω2 , i.e., if the discriminant D = 0, the solutions differ qualitatively
from those obtained when ω1 = ω2 . Generally, if ω is a k-fold multiple solu-
tion of the characteristic equation, it corresponds to k linearly independent
solutions of the form
eωt , teωt , . . . , tk−1 eωt (67)
Multiple Scales Methods in Meteorology 145
Because sine and cosine are linearly independent, the solution is real valued
if the coefficients (Ar − Br ) and (Ai + Bi ) satisfy
(Ar − Br ) = 0 ⇔ Ar = Br
(74)
(Ai + Bi ) = 0 ⇔ Ai = −Bi .
146 R. Klein et al.
x [m]
ï
ï
ï
time [s]
The constants a and b in (75) can be determined from the given initial
conditions and we obtain a particular solution for the considered system.
Example: Let us chose x0 = x(0) = 0, so that the mass is at its equilibrium
at time t = 0, and ẋ0 = dx
dt (0) = Cω0 , so that it has initial velocity Cω0 .
Inserting, we find from (75)
ẋ0
a=0 und b = =C, (76)
ω0
See Fig. 4.
k2 k2
a) D<0 4m2
< c
m
⇒ ω1,2 = − 2m
k
±i c
m
− 4m2
where
c k2
ω̃a = − . (79)
m 4m2
The coefficients ã and b̃ again have to be determined so as to satisfy the
required initial conditions. The mass now performs a damped oscillation
around its equilibrium position with a modified frequency compared to first
case. See Fig. 5.
k2 k2
b) D > 0 4m2
> c
m
⇒ ω1,2 = − 2m
k
± 4m2
− c
m
Again, we insert into (68), and A and B are constants that have to be
computed using given initial conditions. We find
k
x(t) = exp − t A exp(ω̃b t) + B exp(−ω̃b t) (80)
2m
where
k2 c
ω̃b = − . (81)
4m2 m
Comparison of the solutions in (78) and (80) shows that both cases describe
a damped motion, the term exp(ω̃b t) “losing” against exp(− 2mk
t) for t → ∞,
but there is a fundamental difference: in case a), the system with the ex-
k
ponential function exp(− 2m t) performs a damped harmonic oscillation and
passes the origin several times. In case b), however, the damping is so strong
that the mass is not oscillating at all and just moves back monotonically to
its equilibirium (creeping case).
k2
c) D=0 4m2 = c
m ⇒ ω1,2 = − 2m
k
The solution in this case is in line with the one in the case already mentioned
above with general solution (69). We find
k
x(t) = exp − t (A + Bt) , (82)
2m
148 R. Klein et al.
x [m]
ï
ï
time [s]
Figure 5. Exact solution for the oscillator with friction (k > 0): a) D < 0
(blue, line), b) D > 0 (green,dashed), c) D = 0 (red, point-dashed).
where again the constants A and B have to be determind from the initial
conditions.
Thus, case c) is exactly “between” cases a) and b). It is called aperiodic
limiting case. In case b) the damping is still stronger and faster than in
case c) but in both cases there is no oscillation in the system.
x [m]
ï
ï
time [s]
and the case k = 0 and (c − mΩ2 ) = 0 yields the resonant solution (see for
example Walter (1996)) with time-dependent coefficients
F0
Ap (t) = t and Bp = 0 . (90)
2mΩ
We have found particular solutions that satisfy the inhomogenous equa-
tion for each of the parameter regimes. The sum of the homogenous and
the particular solution is the general solution of the oscillator equation with
external periodic forcing. For example, in case 2a) this solution reads
k
x(t) = exp − t ã cos(ω̃a t) + b̃ sin(ω̃a t) + Ap sin(Ωt) + Bp cos(Ωt)
2m
(91)
The constants Ap and Bp have to be defined according to equation (88) and
only ã and b̃ are to be computed from the intial conditions. The solutions
for the other cases can be derived analogously.
and initial conditions x(0) = x0 and ẋ(0) = ẋ0 . This solution consists
of a homogeneous and an inhomogenous part. As we have seen in the
second case, the homogenous solution is decaying exponentially when there
is non-zero damping, so that in the longtime motion only the inhomogenous
(particular) part of the solution prevails.
In the present section, we will continue our analysis by studying situ-
ations in which the free oscillation, damping, and forcing act on very dif-
ferent characteristic time scales. (The notion of a characteristic scale will
hopefully be reasonably clear by the end of the section. The reader may
trust her or his intuition for the time being.) An important tool for deriv-
ing concise mathematical descriptions of such scale-separated processes is
Multiple-Scales Asymptotics, the main ideas and techniques of which will
be explained here using the linear oscillator as an example.
To determine the different time scales of the system we will nondimen-
sionalize the general equation of motion (92) in the first step.
initial data
x0 L
x˙0 L/T
For any given solution of the equations one can identify “characteristic
values” [φj,ref ]N
j=1 of the total of N physical quantities in the system which
roughly describe their orders of magnitude throughout the solution or at
least during a certain time interval (and within a selected region in space
for pdes). These dimensional characteristic quantities may be combined
into non-dimensional characteristic numbers
/
N
ajk
Πk = (φj,ref ) , (95)
j=1
with the exponents ajk chosen so as to guarantee that the Πk do not have a
physical dimension as will be explained shortly.
These numbers are extremely useful as they provide a comparison be-
tween various quantities that may have the same physical dimension but
very different physical origin. An example is the ratio of the oscillator’s
frequency of free, undamped oscillations, c/m, and that of its harmonic
excitation,
c/m
Π∗ = . (96)
Ω
For the non-dimensional Π’s to be actually non-dimensional, all the phys-
ical dimensions have to cancel exactly in the product. Using (93), we may
rephrase this statement as
3 ajk ⎡ ⎤ " #
3 3 P
N
/N / i / /
N
i j / bij ajk
Dim(Πk ) = b
(Xi ) j = ⎣ (Xi ) j k ⎦ =
b a
(Xi ) j=1 ≡ 1.
j=1 i=1 i=1 j=1 i=1
(97)
For this equation to hold, the respective powers of each of the fundamental
dimensions Xi must vanish independently, so that
N
bij ajk ≡ 0 (i = 1 . . . 3, k arbitrary) . (98)
j=1
(St. Petersburg 1911) for the first proof, and Görtler (1975) for a concise
formulation.
Remark: For further aspects of dimensional analysis, the reader may want
to consult Barenblatt (1996).
x T2 M
y = ; Dim (y) = L· · = 1
F0 /c ML T 2
1
τ = Ωt ; Dim (τ ) = ·T = 1
T
Characteristic numbers
(99)
Ω2 1 T2
μ = m ; Dim (μ) = M· 2 · = 1
c T M
Ω M 1 T2
κ = k ; Dim (κ) = · · = 1
c T T M
x0 T2 M
y0 = ; Dim (y0 ) = L· · = 1
F0 /c ML T 2
ẋ0 L T2 M
y0 = ; Dim (y0 ) = ·T · · = 1
ΩF0 /c T ML T 2
Here μ characterizes the system’s inertia, κ its damping, and y0 , y0 the
initial data that allow us to select a specific solution. In interpreting these
quantities, notice that F0 /c is the static displacement of a spring with stiff-
ness c under the effect of a (constant) force F0 .
Notice that y : IR+ → IR, is a function, not just a number, and τ varies
all over IR+ . The original unknowns x(t), t and the new ones, y(τ ), τ , must
154 R. Klein et al.
,,,
,,
satisfy
x(t)
= y(Ωt) and Ωt = τ . (100)
F0 /c
Using this identity, we have
F0 dy dτ F0 Ω
ẋ(t) = = y (τ ) (101)
c dτ dt c
2
F0 d2 y dτ F0 Ω2
ẍ(t) = = y (τ ) , (102)
c dτ 2 dt c
and (92) then allows us to specify a differential equation for y(τ ),
mΩ2 kΩ
y + y + y = cos τ (103)
c c
or
μy + κy + y = cos τ . (104)
The appropriate initial conditions read
x0 c ẋ0 c
y(0) = y0 = and y (0) = y0 = . (105)
F0 F0 Ω
for fixed initial data, but for smaller and smaller values of the inertia and
damping parameters μ, κ (i.e., μ, κ ∈ IR+ ; μ, κ 1).
where
∂y ∂y
, = grad(μ,κ) y (108)
∂μ ∂κ
is the gradient (or (Fréchet-) derivative) of the solution with respect to our
two small parameters. (For the definition of a Fréchet-derivative see, e.g.,
(Werner, 2000, S. 113)).
However, from our previous considerations regarding the path-dependence
of the solution behavior as μ → 0 and κ → 0 we conclude that even though
there is a limit solution, y|μ,κ=0 (τ ) = cos(τ ), it is not the limit of either
solution found along paths I or II in the parameter space of Fig. 7. Thus we
cannot decide whether the “proper behaviour” for very small μ, κ should be
purely oscillatory, purely damped, or something in between. We are led to
conclude that the gradient in (108) does not exist!
Fortunately, not all is lost. Analysis has it that even if a Fréchet-
derivative does not exist, linear approximations to the solution along straight
lines, i.e., directional derivatives, may still be well-defined. This would lead
156 R. Klein et al.
Then, dropping the solution’s explicit dependence on the initial data in the
notation for the moment, we Taylor-expand w.r.t. ε,
∂ ŷ
y(τ ; μ, κ) = ŷ(τ ; ε) = ŷ(τ ; μ̂(ε), κ̂(ε)) = ŷ(τ ; 0)+ε (τ ; 0)+O(ε) (ε → 0) .
∂ε
(110)
The mappings
(μ̂, κ̂) : IR → IR2
(111)
ε → (μ̂(ε), κ̂(ε))
comprise a distinguished limit.
Remark: Here ∂ ŷ/∂ε(τ ; 0) is a generalization of the directional or Gâteaux
derivative
∂
(y(τ ; αε, βε))ε=0 α, β = const. (112)
∂ε
In general we know from functional analysis that for some mapping f
⇒
f Fréchet-differentiable f Gâteaux-differentiable . (113)
allow us to compare the internal damping and oscillation time scales of the
oscillator with that of the external forcing. Notice that both are functions
of our dimensionless mass and damping parameters, (μ, κ).
Three examples shall underline the consequences of picking particular
distinguished limits in the oscillator problem:
For this case, Fig. 8 reveals that the system does no longer oscillate at all.
2
For small ε, the damping timescale KD√ = ε 3 /κ̂ is always much smaller
than the timescale of oscillation KO = ε. Therefore, the inertial motion
of the oscillator is already damped before the first overshoot due to the
incipient oscillation can take place. After a short initial transient, the mass
is essentially “slaved” by the external forcing.
It turns out that the regime μ ∼ κ2 is precisely the threshold that
separates the regions in μ-κ–space in which, as ε → 0, either an oscillatory
or a purely damped motion prevails. If μ vanishes slower than κ2 , the system
will oscillate, otherwise it is strongly damped as summarized in figure 9.
158 R. Klein et al.
\τ
ï
ï
τ
\τ
ï
ï
τ
\τ
ï
ï
τ
κ 1
κ ∼ μ3
√
κ = κ̂ μ
GDPSHG
RVFLOODWRU\
κ∼μ
Figure 9. Ratio between mass and damping in the problem and the result-
ing behaviour of the solution.
with φn+1 (ε) = O(φn (ε)) for ε → 0 is called an asymptotic N-term expansion
of the function u if
N
u(x; ε) − φn (ε) un (x) = O(φN (ε)) (121)
n=1
for ε → 0 (see for example Kevorkian and Cole (1996) and Schneider (1978)).
Notice that the result of the analysis will depend on the choice of the asymp-
totic sequence {φn (ε)}n∈IN .
The chosen ansatz is a Taylor expansion in ε = 0 of the desired solution
y = y(τ ; ε), i.e., we look for the coefficients of ε, ε2 etc. in
N
1 n ∂ny
y(τ ; ε) = ε n
(τ ; 0) + O εN
n=0
n! ∂ε (122)
ε2
= y(τ ; 0) + ε(∂ε y)(τ ; 0) + (∂ε2 y)(τ ; 0) + O ε2
2
Letting y (0) (τ ) := y(τ ; 0), y (1) (τ ) := (∂ε y)(τ ; 0) etc. naturally leads to (119).
We proceed to determine the asymptotic behaviour of the solution y for a
fixed τ and an arbitrary but small ε. Depending on the power of ε consid-
ered, one speaks of the behaviour of the solution at a certain order. Thus,
y (0) (τ ) describes the behaviour at leading or zeroeth order, y (1) (τ ) the be-
haviour at first order, etc.
Inserting the time derivatives from (119) (time means the dimensionless
time τ )
y (τ ; ε) = y (0) (τ ) + εy (1) (τ ) + ε2 y (2) (τ ) + O ε2
(123)
y (τ ; ε) = y (0) (τ ) + εy (1) (τ ) + ε2 y (2) (τ ) + O ε2
Multiple Scales Methods in Meteorology 161
or
y (0) − cos τ +ε y (0) + κ̂y (0) + y (1) +ε2 y (1) + κ̂y (1) + y (2) = O ε2 .
(125)
If this equation is to hold for arbitrary (but small) ε, each of the coeffi-
cients of εi for (i = 1, 2, . . . ) (the expressions in brackets) has to be zero
individually. Therefore,
y (0) = cos τ
y (1) = −y (0) − κ̂y (0) = cos τ + κ̂ sin τ (126)
y (2) = −y (1) − κ̂y (1) = (1 − κ̂2 ) cos τ + 2κ̂ sin τ .
Instead, the initial displacement and velocity from the leading-order asymp-
totics are y (0) (τ ) = 1 and y (0) (0) = 0. Any different values for y0 and y0 can
nowhere be accounted for. The reason is that with the present ansatz we can
only see the solution after any initial transient, which would be determined
by the initial data, has already decayed.
162 R. Klein et al.
\τ ε
ï
ï
ï
ï
τ
This can be verified in Fig. 10, even when the next two higher-order
terms are included. Adding the first- and second-order contributions, makes
merely a slight difference for the asymptotic approximation. It can do no
more than reproduce with better accuracy the background oscillation as
more and more expansion terms are included.
t τ τ
ϑ= = =√ (129)
m/c 2
mΩ /c ε
would remedy this problem. We will try out a new expansion scheme in
which the unknowns will depend on ϑ instead of on τ . Since each time
derivative d/dτ that√ appears in the governing equation in that case will
produce a √factor 1/ ε by the chain rule, we should expand the solution in
powers of ε instead of in powers of ε. (What happens if we don’t but use
ϑ as the independent variable while expanding in powers of ε?) Thus we
choose the new asymptotic expansion scheme
√
y(τ ; ε) =: y (0) (ϑ) + εy (1) (ϑ) + εy (2) (ϑ) + O(ε) , (130)
Multiple Scales Methods in Meteorology 163
which is equivalent to
(0) τ √ (1) τ (2) τ
y(τ ; ε) = y √ + εy √ + εy √ + O (ε) . (131)
ε ε ε
We will need to be aware of the latter form of writing the expansion when
we insert into the governing equation in (118), which is written in terms of
τ . In preparation, we compute the τ -derivatives of (131) for fixed ε,
∂y dy (0) dϑ √ dy (1) dϑ dy (2) dϑ dϑ
(ϑ; ε) = + ε +ε +O ε
∂τ ε dϑ dτ dϑ dτ dϑ dτ dτ
1 dy (0)
dy (1) √ dy (2) √
= √ + + ε +O ε (132)
ε dϑ dϑ dϑ
∂ 2 y 1 d2 y (0) 1 d2 y (1) d2 y (2)
(ϑ; ε) = + √ + + O(1) . (133)
∂τ 2 ε ε dϑ2 ε dϑ2 dϑ2
∂y
The notation ∂τ ε
shall underline that, for any solution of the oscillator
problem, ε is a fixed parameter and thus held constant when differentiating.
If, instead, we were to consider ε as varying in time, then the time scales, and
thus the mass, spring stiffness, and the damping coefficient, would change
in time, too.
We express√the forcing term in (118) in terms of ϑ, and then Taylor-
expand w.r.t. ε,
√ ε
cos τ = cos( εϑ) = 1 − ϑ2 + O ε2 . (134)
2
perturbation equations,
d2 y (0)
O(1) : + y (0) = 1
dϑ2
√ d2 y (1) dy (0)
O ε : + y (1) = −κ̂ (136)
dϑ2 dϑ
d2 y (2) dy (1) ϑ2
O(ε) : + y (2) = −κ̂ − .
dϑ2 dϑ 2
At leading order (i.e., at O(1)) we find the equation for an undamped os-
cillator with time-independent forcing. The solution is
(f¨ − 2ġ) sin ϑ + (g̈ + 2f˙) cos ϑ = −κ̂ (a0 cos ϑ − b0 sin ϑ) . (140)
The desired solutions are polynomials in ϑ of degree less than two, and we
let f (ϑ) = Af ϑ + Bf and g(ϑ) = Ag ϑ + Bg . Without loss of generality, we
may also assume Bf ≡ 0 and Bg ≡ 0, as these terms can be covered by the
homogenous part of the solution. Solving (141), yields
κ̂
yp(1) = −ϑ (a0 sin ϑ + b0 cos ϑ) (142)
2
Multiple Scales Methods in Meteorology 165
so that
(1) κ̂
y (1) = yh + yp(1) = a1 sin ϑ + b1 cos ϑ − ϑ (a0 sin ϑ + b0 cos ϑ) . (143)
2
In the same way one may compute higher-order solutions.
It turns out that this ansatz based on the short-time variable is still
not satisfactory. As seen in figure 11, although the solution improves with
increasing order, this is true only for the early stages of the evolution. With
increasing time the solution deteriorates dramatically, and this gets worse
the higher the approximation order!
\τ ε
ï
ï
ï
ï
τ
(145)
Inserting into (118) one has
√ √
(0) (0) (1) (0)
yϑϑ + y (0) − cos τ + ε 2yϑτ + yϑϑ + κ̂yϑ + y (1) +O ε = 0 . (146)
Multiple Scales Methods in Meteorology 167
ϑ
ϑ1
ϑ2
ϑ3
ϑ∗
τ∗ τ
Figure 12. Relation between the time coordinates τ and ϑ for chosen ε,
√
where ϑi = τ / εi (ε1 < ε2 < ε3 ). For every τ ∗ there is exactly one ϑ∗ .
For this equation to hold for arbitrary ε 1 it is sufficient that the dif-
ferent terms in brackets vanish.
√ Considering that, through the dependence
of the y (i) on τ and ϑ = τ / ε, these brackets implicitly do depend on ε, it
is not immediately clear that the vanishing of the brackets is also necessary.
But, if we can make all the coefficients disappear for arbitrary (ϑ, τ ), then
our equation is in any case satisfied.
√ Let’s give it a try!
For the different orders in ε this leads to
(0)
O(1) : y (0) + yϑϑ = cos τ
√ (147)
(1) (0) (0)
O ε : y (1) + yϑϑ = −2yϑτ − κ̂yϑ .
Assuming initial conditions of type y0 = O(1) and y0 = O(1) and an allow-
ing for small but otherwise arbitrary ε, we conclude that
y (0) (0, 0) = y0
(0) . (158)
yϑ (0, 0) = 0
(0) (1)
(yτ + yϑ )(0, 0) = y0
(i) (i+1)
(yτ + yϑ )(0, 0) = 0 (i = 1, 2, ...)
Inserting these results into the leading-order solution from (148), and using
(0)
yϑ = −A(τ ) sin ϑ + B(τ ) cos ϑ , (159)
Thus, with (155) and (148) we obtain the leading-order multiple-scales so-
lution,
(0) κ̂
y (ϑ, τ ) = (y0 − 1) exp − τ cos ϑ + cos τ . (161)
2
Remark: We have used the initial conditions exclusively at ϑ = τ = 0, but
not for ϑ = 0 and arbitrary τ . The latter would be incorrect, because in the
ϑ, τ –plane, for fixed system parameters μ, κ, i.e., for fixed ε, the solution
evolves along a straight line through the origin in the ϑ, τ –plane as shown
170 R. Klein et al.
\τ ε
ï
ï
ï
ï
ï
τ
without greenhouse gases, would render the mean near-surface air temper-
ature near 250K. The actual value in Table 1 is the freezing temperature of
water under standard conditions, i.e., Tref ∼ 273K, which is about midway
between the observed maximal and minimal near-surface air temperatures.
The equator-to-pole air temperature difference near the surface, ΔT |peq , is
a consequence of the latitudinal variation of the sun’s irradiation. The dry
air gas constant, R, and isentropic exponent, γ, as thermodynamic proper-
ties are also universally characteristic for atmospheric flows, because their
variations due to admixtures of water vapor, trace gases, and the like are
no larger than a few percent in general.
Based on these eight basic reference quantities, four independent di-
mensionless combinations can be composed. To define combinations with
intuitive interpretations, we introduce as auxiliary quantities the pressure
scale height, hsc , the characteristic speed cref of barotropic1 gravity waves,
and a reference density, ref , via
1
Atmospheric flow modes are called “barotropic” if their structure is homogeneous in
the vertical direction.
Multiple Scales Methods in Meteorology 173
Then we let
hsc
Π1 = ∼ 1.67 · 10−3 ,
a
ΔT |peq (163)
Π2 = ∼ 0.18 ,
Tref
cref
Π3 = ∼ 0.5 .
Ωa
These limits are compatible with the estimates in (163) for actual values
of ε ∈ [1/7 . . . 1/8]. We will adopt ε as the reference expansion parame-
ter for asymptotic analyses below, and any additional small or large non-
dimensional parameter that may be associated with singular perturbations
in the governing equations is subsequently tied to ε through suitable further
distinguished limits.
Remark: Before we proceed to do so, we notice that Keller and Ting
(1951) already proposed to use the acceleration ratio, ε ∼ (aΩ2 /g)1/3 =
(Π1 /Π23 )1/3 , as a basic expansion parameter for meteorological modelling.
When Π3 = O(1), this is equivalent to (164) above.
174 R. Klein et al.
ghsc ΔT |peq 3
uref = = cref Π2 Π3 ∼ ε 2 cref (165)
Ωa Tref
Questions
• What would be the scaling in terms of ε of the Froude number based
on a typical internal gravity wave speed∗ ?
• What would be the scaling of the Rossby number based on our refer-
ence length, hsc ∗ ?
• By what power of 1/ε is the internal Rossby radius larger than hsc ∗ ?
• By what power of 1/ε is the Obhoukhov or external Rossby radius
larger than hsc ?
• By what power of 1/ε is the Obhoukhov or external Rossby radius
larger than the internal Rossby radius? — Compare your result with
a related remark in Pedlosky (1987).
• Does the Oboukhov or external Rossby radius come out larger, com-
parable or smaller than the Earth radius, a, which is representative of
the planetary scale?
∗
It is safe to assume that the variation of potential temperature across
the troposphere is comparable to the equator-pole temperature difference,
ΔT |peq .
Θt + v · ∇ Θ + wΘz = QΘ ,
(168)
where
p1/γ
Θ= (169)
ρ
is the dimensionless potential temperature, k is the local vertical unit vector
indicating the direction of the acceleration of gravity. The terms Qv , Qp ,
176 R. Klein et al.
1 1
v ,t + v · ∇ v + wv ,z + ε 2(Ω × v) + ∇ p = 0,
ε3 ρ (172)
1 1 1
wt + v · ∇ w + wwz + ε 2(Ω × v)⊥ + 3 pz = − ,
ε ρ ε3
Θt + v · ∇ Θ + wΘz = 0,
and
p1/γ
Θ= . (173)
ρ
The quasi-geostrophic theory is designed to address flows on length scales
comparable to the internal Rossby radius, and on time scales corresponding
to horizontal advection across such distances. How can we access these
length and time scales within our multiple scales asymptotic scheme?
The internal Rossby radius is defined as the distance which an internal
gravity wave would travel during a characteristic Earth rotation time. This
is equivalent to requiring
N hsc
L Ro = , (174)
Ω
where
g ∂Θ
N= (175)
Θ ∂z
is the so-called Brunt-Väisälä or buoyancy frequency, and
hsc ∂Θ hsc ∂Θ
N hsc = ghsc = cref (176)
Θ ∂z Θ ∂z
is a typical travelling speed of internal gravity waves. The reader may want
to consult the established literature for corroboration.
178 R. Klein et al.
Here we have used the scalings of our fundamental parameters from Table 1
as discussed in (162)–(164), and the well established observation, [Schneider
(2006); Frierson (2008)], that the equator-to-pole temperature differences
are comparable to the vertical potential temperature variations across the
troposphere, so that
hsc ∂Θ ΔT |peq
∼ . (179)
Θ ∂z Tref
With the estimate in (178), if we want to describe horizontal variations
on scales comparable to LRo , we should use the dimensionless horizontal
coordinate
x hsc x
ξ= = = ε2 x . (180)
LRo LRo hsc
We will be interested here in phenomena associated with advection over
distances of LRo , so we will use the time variable
t hsc t
τ= = = ε2 t . (181)
LRo /uref LRo hsc /uref
Finally, in order to study phenomena which occupy the full depth of the
troposphere, we will use a vertical coordinate non-dimensionalized by the
pressure scale height, hsc , i.e., we use our original dimensionless coordinate
z
z= . (182)
hsc
These scalings are summarized in Fig. 14.
Multiple Scales Methods in Meteorology 179
1
LRo ∼ h
ε2 sc
hsc
t ∼ LRo
vref
∼ hsc
ε2 vref
(0)
O ε0 : 0 ∇ξ · v = 0,
(186)
(1) ∂
O(ε) : 0 ∇ξ · v + 0 w(3) = 0.
∂z
(0)
In writing down the terms of O(ε) we have already neglected (1) ∇ξ · v on
account of (186).
We will need the vertical component of Ω (see Fig. 15), which we expand
Multiple Scales Methods in Meteorology 181
Pol
Ω
k
Ω0
eΩ ϑ
eeq Äquator
Figure 15. Splitting of the coriolis term into a horizontal and a vertical
component.
as
Ω⊥ = k·Ω
= (eeq cos ϑ + eΩ sin ϑ) · eΩ |Ω|
= |Ω| sin ϑ
y
= |Ω| sin ϑ0 + (188)
a
= |Ω| sin(ϑ0 + ε ξ2 )
= |Ω| sin(ϑ0 ) +ε |Ω| cos(ϑ0 ) ξ2 + O(ε)
1 23 4 1 23 4
=:Ω0 =:β
= Ω0 + εβξ2 + O(ε) .
Here we have taken into account that ϑ is the arclength along a longitudinal
circle divided by the radius of the reference sphere, a, introduced deviations
from a reference latitude, so that ϑ = ϑ0 +y /a, and recalled that hsc /a = ε3
and ξ2 = ε2 y /hsc . The rest is Taylor expansion of the sine function about
the reference latitude.
Since w(0) ≡ w(1) ≡ 0 we also know that wΩ × k = O(ε) and, in
182 R. Klein et al.
summary, we find
(Ω × v) = (Ω0 + εβξ2 )k × v + O(ε) ,
(189)
(Ω × v)⊥ = Ω × v .
Using w(0) ≡ w(1) ≡ 0, and that p(0) ≡ p0 (z) and p(1) ≡ p1 (z), we im-
mediately move to the equation at O ε−1 where we find the geostrophic
balance,
(0) p(2)
Ω0 k × v + ∇ξ π (2) = 0 , where π (2) = , (191)
0
i.e., the balance of the horizontal Coriolis and pressure gradient forces.
Geostrophic balance implies that, at leading order, the horizontal flow di-
rection is perpendicular to the horizontal pressure gradient.
We verify for later purposes that
(0) (0) 1
∇ξ · v =0 and v = k × ∇ξ π (2) (192)
Ω0
by, respectively, applying (k · (∇ξ × [ · ])) and (k × [ · ]) to (191)1 .
Remark: With (191) we have found a time-independent constraint on the
leading order velocity and second order pressure fields. Such constraints
did not exist for the original system of the compressible flow equations!
The constraint implies that only if the initial data for a given flow problem
satisfy the constraint, at least at the given orders, can we hope that the
approximate asymptotic solution will remain close to the exact solution.
This kind of “change of type” of the asymptotic limit problem relative to
the original one is typical of singular perturbation problems. Also, we
recall that we encountered a similar issue with the slow-time expansion for
the linear oscillator in chapter 3.
1
O ε0 : 0 = p0 γ
1 p1
O(ε) : 1 + 0 Θ1 = p0 γ (194)
γp0
(2)
1 p (1 − γ)p1 2
O ε2 : (2) + 1 Θ1 + 0 Θ(2) = p0 γ +
γp0 2γ 2 p0 2
In a similar way one solves the first order equation explicitly for given Θ1 (z).
We leave this as an exercise.
dpi
= −i (i = 0, 1) . (198)
dz
The remaining primary unknowns for description of the flow field are
then
(0)
π (2) , v , w(3) , Θ(2) (τ, ξ, z) , (199)
where π (2) = p(2) /0 , and they satisfy the following balance and transport
equations:
184 R. Klein et al.
Hydrostatic Balance
∂π (2)
= Θ(2) (200)
∂z
Geostrophic Balance
(0)
Ω0 k × v + ∇ξ π (2) = 0 (201)
Anelastic Constraint
∂
(1)
0 ∇ξ · v + 0 w(3) = 0 (202)
∂z
where,
(0)
ζ (0) = k · (∇ξ × v ) , (205)
is the vorticity of the leading-order velocity field.
This completes the summary of the quasi-geostrophic model equations.
Remark: We have given the QG equations here in a somewhat unusual
form, sticking as closely as possible to the original equations. In this way, it
remains transparent that (200) and (201) are direct consequences of the
vertical and horizontal momentum balances, respectively, (202) emerges
from mass conservation, and (203) from the potential temperature transport
Multiple Scales Methods in Meteorology 185
equation. These equations can all directly be read off the original equations
at the appropriate orders in the asymptotic expansion.
(1)
closure for ∇ξ · v in (204) emerges as a solvability condition at
The
0
O ε in the horizontal momentum balance as will be shown in the next
section.
Remark: As can see clearly in the present summary of our asymptotic limit
equations, considering large spacial and long time scales only implies strong
constraints on the solutions. Instead of evolution equations for the primary
unknowns in the compressible flow equations (the densities of mass, momen-
tum, and energy) we find three time independent constraints or balances!
Only the potential temperature evolution equation in (203) and the vortic-
ity transport equation in (204) have maintained the original “prognostic”
(time evolution) character.
k · (∇ξ × ∇ξ φ) ≡ 0 (209)
for any scalar function φ(ξ) that is sufficiently smooth. By applying the
operator k · (∇ξ × [ · ]) to (208) we thus eliminate the terms involving π (2)
186 R. Klein et al.
= βv (0)
∂ (0)
= β + v · ∇ξ ξ2 .
∂τ
as announced in (204).
(0) 1
v = k × ∇ξ p(2) ,
Ω0
∂π (2)
Θ(2) = , (214)
∂z
(0) 1 2 (2)
ζ (0) = k · ∇ξ × v = ∇ π
Ω0 ξ
Appendix
Gauß’ Integral Theorem
Let Ω ⊂ IR be a compact subset with a smooth boundary, n : ∂Ω → IRn
the field of outer unit normal vectors and U ⊃ Ω an open subset of IRn . Then
for every continuous differentiable vector field F : U → IRn the following is
true
∇ · F (x) dV = F (x) · n dσ
Ω ∂Ω
Vector identities
For vectors A, B, C ∈ IRn and scalars ϕ, ψ ∈ IR the following general
identities are true :
∇ · (ϕ A) = ϕ∇ · A + A · ∇ϕ (217)
∇ × (ϕ A) = ϕ∇ × A + ∇ϕ × A (218)
∇(A · B) = A × (∇ × B) + B × (∇ × A)
+ (A · ∇)B + (B · ∇)A (221)
∇ × ∇ϕ = 0 (222)
∂p
= −g
∂z
Multiple Scales Methods in Meteorology 189
Starting from a reference pressure pref (e.g. a mean pressure on sea level)
we can define the pressure scale height by the height difference at which the
pressure in an atmosphere in hydrostatic balance and with constant density
changes of an order of magnitude of pref
pref
hsc := .
ref g
is
∞ ν
t
X b (t) = exp(Ωt) X b (0) , wobei exp(Ωt) := (Ω)ν .
ν=0
ν!
Because of
⎛ ⎞
−Ω22 Ω23 Ω21 Ω22 Ω21 Ω23
2
Ω = ⎝ Ω21 Ω22 −Ω21 Ω23 Ω22 Ω23 ⎠ = |Ω|2 (eΩ eTΩ − 1) ,
−Ω21 Ω23 Ω22 Ω23 −Ω21 Ω22
t4 4 t 5
4
− |Ω| (eΩ eΩ − 1) + |Ω| Ω + ... X b (0)
T
4! 5!
2 2
|Ω| t |Ω|4 t4
= (eΩ eΩ ) + 1 −
T
+ − ... (1 − eΩ eTΩ )
2! 4!
|Ω|3 t3 |Ω|5 t5 Ω
+ |Ω|t − + − ... X b (0)
3! 5! |Ω|
Ω
= (eΩ eTΩ ) + cos(|Ω|t)(1 − eΩ eTΩ ) + sin(|Ω|t) X b (0)
|Ω|
ΩT
RT (t) = (eΩ eTΩ )T + cos(|Ω|t)(1 − eΩ eTΩ )T + sin(|Ω|t)
|Ω|
−Ω
= (eΩ eTΩ ) + cos(−|Ω|t)(1 − eΩ eTΩ ) − sin(−|Ω|t) = R(−t)
|Ω|
Multiple Scales Methods in Meteorology 191
in the basis {ẽ1 , ẽ2 , ẽ3 } of the rotating coordinate system. The interrelation
between the coordinate systems is given in the following form :
ẽi = R(t)ei , i = 1, 2, 3 .
3
3
3
x= xi ei = x̃i ẽi = x̃i R(t)ei .
i=1 i=1 i=1
3
3
xk = x̃i (eTk R(t))ei = x̃i (R(t))ki .
i=1 i=1
Hence,
˜ 3
∂f ∂ t̃ ∂f ∂ x̃i ∂ f˜
= +
∂t x ∂t x ∂ t̃ ∂t x ∂ x̃i
x̃ i=1 t̃,xj (j=i)
3
∂ f˜ ∂ f˜
= + (−Ω x̃)i
∂ t̃ i=1
∂ x̃i
3
∂ f˜ ∂ f˜ (226)
= + (−Ωij x̃j )
∂ t̃ i,j=1
∂ x̃i
3
∂ f˜ ∂ f˜
= + (x̃j Ωji )
∂ t̃ i,j=1
∂ x̃i
∂ f˜ ˜ f˜)T
= + x̃T Ω(∇
∂ t̃
and
∂f ∂ t̃ ∂ f˜
= +
∂xi t̃,xk (k=i) ∂xi t,xk (k=i) ∂ t̃
x̃
3
∂ x̃j ∂ f˜
+
∂xi t,xk (k=i) ∂ x̃j
j=1 t̃,x̃k (k=j)
3
∂ f˜
= (R−1 (t))ji
j=1
∂ x̃j
3
∂ f˜
= (R(t))ij
j=1
∂ x̃j
or
˜ f˜)T
(∇f )T = R(t)(∇ .
3
3
xp (t) = xpi (t)ei = x̃pj (t)ẽj (t)
i=1 j=1
Multiple Scales Methods in Meteorology 193
3 3
dxp
= ẋpi (t)ei = (x̃˙ pj (t)ẽj (t) + x̃pj (t)ẽ˙ j (t))
dt i=1 j=1
3
= ẋrel
p (t) + x̃pj (t)Ω ẽj (t)
j=1
= ẋrel
p (t) + Ω xp (t) .
v = ẋ = v rel + Ω x
3
(227)
= ṽirel ẽi + Ω x
i=1
into a part arising from earth’s rotation and the relative wind speed. When
transforming the velocity divergence we find
∇ · v = ∇ · (v rel + Ω x) = ∇ · v rel + ∇ · (Ω x) ,
3
3
∂
∇ · (Ω x) = (Ωij xj ) = Ωij δij = 0 ,
∂xi
i,j=1 i,j=1
3
3 3
∂v rel ∂
∇ · v rel = i
= T
(ei ṽjrel ẽj )
i=1
∂xi i=1
∂x i j=1
3
∂
= ṽjrel(t̃(t, x), x̃(t, x)) eTi ẽj
i,j=1
∂xi
3 (228)
∂ rel
= Rik ṽ Rij
∂ x̃k j
i,j,k=1
3 3
∂ rel ∂ṽkrel ˜ · ṽ rel
= δik ṽj = =∇ .
∂ x̃k ∂ x̃k
j,k=1 k=1
194 R. Klein et al.
Because of
3
T
(v rel ∇T )v rel = ((ṽ rel )T ∇
˜ T )ṽirel ẽi ,
i=1
˜ T )ṽ rel = −(x̃T Ω∇
(Ω x)T ∇T v rel = (x̃T ΩT ∇ ˜ T )ṽ rel ,
3
3
3
T ∂ ∂
(v rel ∇T )Ω x = virel Ω xi ei = ṽjrel (Ω ei x̃i )
i=1
∂xi i=1 i,j=1
∂ x̃j
3
= Ωṽjrel δij ei = Ω ṽ rel
i,j=1
and
(Ω x)T ∇T (Ω x) = −(xT Ω∇T )Ω x
3
3
∂
= − xi Ωik Ωjl xl ej = xi Ωki Ωjk ej
∂xk
i,j,k,l=1 i,j,k=1
T
= xT ΩT ΩT = Ω(Ω x) = Ω × (Ω × x)
Multiple Scales Methods in Meteorology 195
we obtain
v t + (v T ∇T )v = v t + (v rel + Ω x)T ∇T v rel + Ω x
T
= v t + (v rel ∇T )v rel + (Ω x)T ∇T v rel
T
+(v rel ∇T )(Ω x) + (Ω x)T ∇T (Ω x)
3
3
∂ṽ rel
= i
ẽi + ((ṽ rel )T ∇
˜ T )ṽirel ẽi
i=1
∂t i=1
+2 Ω ṽ rel + Ω(Ω x)
∂ṽkrel
v t + (v T ∇)v k
= + ((ṽ rel )T ∇T )ṽkrel + 2 (Ω ṽ rel )k + Ω(Ω x) k
∂t
Bibliography
G. I. Barenblatt. Scaling, self-similarity, and intermediate asymptotics.
Cambridge Texts in Applied Mathematics. Cambridge University Press,
Cambridge, New York, Melbourne, 1996.
J.A. Biello and A.J. Majda. Transformations for temperature flux in mul-
tiscale models of the tropics. Theor. Comp. Fluid Dyn., under revision,
2006.
O. Forster. Analysis 3. Vieweg & Sohn Verlagsgesellschaft, Braunschweig,
Wiesbaden, 3. edition, 1984.
D. M. W. Frierson. Midlatitude static stability in simple and comprehensive
general circulation models. J. Atmos. Sci., 65:1049–1062, 2008.
A.E. Gill. Atmosphere-Ocean Dynamics, volume 30 of International geo-
physics series. Academic Press, New York, 1982.
H. Görtler. Dimensionsanalyse: Theorie der physikalischen Dimensio-
nen mit Anwendungen. Ingenieurwissenschaftliche Bibliothek. Springer,
Berlin, Heidelberg, New York, 1975.
J.B. Keller and L. Ting. Approximate equations for large scale atmospheric
motions. Internal Report, http://www.arxiv.org/abs/physics/0606114,
Inst. for Mathematics & Mechanics (renamed to Courant Institute of
Mathematical Sciences in 1962), NYU, 1951.
J. Kevorkian and J.D. Cole. Multiple Scale and Singular Perturbation Meth-
ods, volume 114 of Applied Mathematical Sciences. Springer-Verlag, New
York, 1996.
196 R. Klein et al.
Herbert Steinrück*
*
Vienna University of Technology, Institute of Fluid Mechanics and Heat
Transfer, Vienna, Austria
Abstract The undular hydraulic jump in turbulent open chan-
nel flow is considered in the double limit of very large Reynolds
numbers, and Froude numbers approaching the critical value, i.e.
F r = 1 + 32 ε with ε → 0+.
The undular jump is associated with a distinguished limit, which
is characterized by the similarity
√ parameters A and a. The square
root of the first parameter A is essentially the ratio of the dimen-
sionless friction velocity and the difference of the Froude number to
its critical value. The second parameter a is a scaled measure of
the difference of the incident turbulent flow to the fully developed
turbulent flow.
Since a wavy solution with a slowly varying amplitude is ex-
pected, a multiple scales expansion is performed. A new indepen-
dent variable is introduced such that the wave length becomes con-
stant and normalized to one. The perturbation equations of the
orders ε, ε3/2 , ε2 , and ε5/2 have to be considered in order to obtain
a complete first-order solution.
In case of fully developed incident flow analytical results for the
amplitude and wave length of the first wave are obtained. They
are compared with measured data and reasonable agreement is ob-
served.
1 Introduction
The undular hydraulic jump is a peculiar change of state that can be
observed in steady open channel flows if the upstream Froude number is
slightly above the critical value 1, cf. Böß (1927), Chow (1959), Hager and
Hutter (1984), Hager (1992), Henderson (1966), Chanson (1993), Chanson
and Montes (1995), Reinauer and Hager (1995), Ohtsu et al. (2001).
As far as the observations are concerned, the undular hydraulic jump is
characterized by a wavy shape of the free surface, with wave lengths that
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
198 H. Steinrück
are much larger than the depth of the liquid and amplitudes that decay
slowly in main flow direction.
From a theoretical point of view there is the difficulty that, though the
viscosity effects are very weak, an inviscid-flow solution does not exist, cf.
Benjamin and Lighthill (1954).
For laminar flow, Johnson (1972) was able to cope with these difficulties
and provide an asymptotic analysis in terms of Froude numbers near 1 and
large Reynolds numbers. His main result is a steady-state version of the
Korteweg-de Vries-Burgers equation that governs the perturbations of the
surface elevation.
For turbulent flow, attempts have been made to analyze the undular
hydraulic jump on the basis of various ad-hoc approximations, e.g. An-
dersen (1978), Hager and Hutter (1984), Kaufmann (1934), Lauffer (1935),
Chanson and Montes (1995). More detailed comments on those previous
investigations can be found in Grillhofer (2002).
In Grillhofer (2002), Grilllhofer and Schneider (2003) and Steinrück et al.
(2003) an asymptotic analysis of the undular jump in turbulent flow was
given. Plane flow over a bottom of constant slope was considered in the
double limit of very large Reynolds numbers, i.e. Reτ → ∞, and Froude
numbers approaching the critical value, i.e. F r = 1 + 32 ε with ε → 0.
Here we will follow the analysis in Steinrück et al. (2003) to study the
influence of small viscosity effects on near critical open channel flow. Based
on the shallow water approximation we will employ the method of multiple
scales to derive equations for the change of amplitude and wave length of
the undulated surface.
y g
ur l
hr
u(x r , y) h C
h(x)
x
α
The Reynolds stresses (u 2 ), (u v ) and (v 2 ) are scaled with the square
of the reference friction velocity, which is given by:
τwr
uτ2 r = . (1)
ρ
x y ū v̄ p̄ F r 2 h̄
X= , Y = , Ū = , V̄ = , P̄ = , H̄ = , (2a)
l hr ur δu r ρu2r hr
(u 2 ) (u v ) (v 2 ) uτ
(U 2 ) = , (U V ) = , (V 2 ) = , Uτ = , (2b)
uτ2 r uτ2 r uτ2 r uτ r
where the dimensionless parameters
ur hr uτ r
Fr = √ , δ= , α, γ= (3)
gh r l ūr
are the Froude number, the shallow water parameter, the angle of inclina-
tion and the ratio of the reference wall friction velocity and the volumetric
mean velocity, respectively. The later is a function of the Reynolds num-
ber Re = u r h r /ν only. If the fully developed flow in an open channel
with the inclination α is chosen as reference state the force balance yields
sin α = γ 2 F r 2 .
For high Reynolds numbers the flow field can be decomposed into a
defect layer and a viscous sub-layer. Since a universal solution exists for the
viscous sub-layer, cf. Kluwick (1998), Schlichting and Gersten (2000), it is
sufficient to consider the defect layer only.
200 H. Steinrück
Using the scalings (2a) and (2b), the continuity equation for the dimen-
sionless Reynolds averaged velocities becomes:
The momentum equations for the defect layer in dimensionless form are:
sin α γ2
F r2 Ū ŪX + V̄ ŪY = −P̄X + − F r2 δ(U 2 )X + (U V )Y , (5a)
δ δ
δ 2 F r 2 Ū V̄X + V̄ V̄Y = −P̄Y −cos α− F r2 γ 2 δ(U V )X + (V 2 )Y . (5b)
Using the above definition of the mean free surface, the common kinematic
boundary condition for the averaged velocities holds:
2 2 (U 2 ) (U V ) −δ H̄X δHX
γ Fr · = P̄ at Y = H̄.
(U V ) (V 2 ) 1 −1
(8)
1
ŪS (X) = γ Ūτ ln Reγ Ūτ H̄ + C + + C̄(X) , (10)
κ
Multiple Scales Analysis of the Turbulent Hydraulic Jump 201
where ŪS = Ū(X, H̄(X)) is the surface velocity, Ūτ2 = −(U V )Y =0 is the
dimensionless wall shear stress and the function C̄(X) is given by
H̄
1 dŪ 1
C̄(X) = − dY,
0 γUτ dY κY
where κ is von Karman’s constant. C + is another empirical constant, which,
as it will turn out, does not appear in the final results.
Let us denote ΔŪS = ŪS − ŪS r , ΔŪτ = Ūτ − 1, and ΔH̄ = H̄ − 1 the
difference of the dimensionless surface velocity, friction velocity and surface
height to its reference values at X = X r , respectively. Using the friction
law (10) the surface velocity ŪS r at the reference state, X = X r , is given
by
1
ŪS r = γ ln(Reγ) + C + + C̄(X r ) (11)
κ
and we can rewrite the friction law as
1
ΔŪS = ŪS r ΔŪτ + (1 + ΔŪτ )γ ln(1 + ΔH̄) + C̄(X) − C̄(X r ) . (12)
κ
Expanding (12) for small changes we obtain
δ = 3ε1/2 , (16)
where the coefficient 3 is introduced for the sake of simplifying the final
equations.
In order to define a distinguished limit we have to couple the angle of
inclination α and the wall shear stress γ to the perturbation parameter ε
in such a way that to the leading order the shallow water equation will be
recovered. Therefore we introduce the two similarity parameters a and A,
which are assumed to be of order one as
γ 2 F r 2 − sin α 3γ 2 F r2
a= , A= . (17)
δε5/2 δε3/2
As mentioned above the parameter a measures the deviation of the reference
state from the fully developed flow. With other words if the flow in the
reference state is fully developed the parameter a vanishes. This assumption
has been made in Grilllhofer and Schneider (2003) and Steinrück et al.
(2003). The other parameter A is a measure for the change of the wall
shear stress due to a change of the fluid height.
The reference state We assume that the reference state state is attained
at Xr and the flow is a locally fully developed turbulent open channel flow.
Thus in the defect layer the velocity profile can be written as
√ 1
Ū(Xr ) = 1 + γ ŪD = 1 + ε AŪD with ŪD (Y ) dY = 0, (18)
0
Multiple Scales Analysis of the Turbulent Hydraulic Jump 203
3 Asymptotic Analysis
3.1 Shallow water approximation
In a first attempt we try to use a regular expansion of the form
Ū = 1 + εŪ1 + ε2 Ū2 + · · · , (19a)
However the perturbation H̄1 of the fluid height remains undetermined from
the analysis of the O(ε)-terms.
The component of the gravity term in flow direction is of order O(ε3/2 ).
Due to the scaling assumptions the derivative of the Reynolds’ shear stress
is of the same order. Using the dynamic boundary condition we obtain
(U V )0 = 1 − Y. (22)
Considering the O(ε)2 -terms the Y -momentum equation yields
9
P̄2 = H̄2 + H̄1,XX (1 − Y 2 ) − A(V 2 )0 (23)
2
Using the X-momentum equation and integration of the continuity equation
yields the second order term of the vertical velocity component.
V̄2 (X, 1) = H̄2,X + 3H̄1,XXX +
√ √ 1 (24)
+ −3 + H̄1 + AŪD − 2 A ŪD dY H̄1,X .
0
Employing the kinematic boundary condition a second expression for V̄2 (X, 1)
can be obtained
√
V̄2 (X, 1) + V̄1,Y H̄1 = H̄2,x + (−H̄1 + AΔŪD )H̄1,x . (25)
204 H. Steinrück
Comparing both expressions for V̄2 (X, 1) we obtain the third order differ-
ential equation for H̄1
1
H̄1,XX + H̄12 − H̄1 = R. (27)
2
Multiplying (27) by H̄1,X and integration yields
2
−3H̄1,X = p(H̄1 , R, S) := H̄13 − 3H̄12 − 6RH̄1 − 6S, (28)
d ∂ ∂ dX ∂ ∂
H̄ = H̄ + H̄ = H̄ + ε1/2 H̄. (31)
dX ∂X ∂X dX ∂X ∂X
The derivatives with respect to the slow variable of terms of order O(ε)
are of order O(ε3/2 ) and the derivatives of terms of the order O(ε2 ) are
of the order O(ε5/2 ). Equations for the constants of integration R and S
as functions of the slow variable X will be obtained from the requirement
that all functions are periodic with respect to the fast variable X. Thus we
have to include terms of the order ε3/2 and ε5/2 additionally to the terms of
order ε and ε2 in the expansion of the dependent variables (19). However,
to facilitate the formulation of the periodicity condition with respect to the
fast variable X we introduce a transformed fast variable ξ
1
ξ= Ω(X ), with X = ε1/2 (X − X0 (ε)), (32)
ε1/2
such that all function have period 1 with respect to the new fast variable ξ.
We have
dξ = ω(Ω) dX, with ω = Ω . (33)
Thus 1/ω is the wave length in terms of the original fast variable X and
part of the solution.
Representative for all dependent state variables the expansion of the
pressure is given by:
Terms of order O(ε) Using the Y -momentum equation (5b) and the
dynamic boundary condition (8) we obtain
P̄1 (ξ, Ω, Y ) = H̄1 (ξ, Ω). (35)
The X-momentum equation (5a) yields
Ū1,ξ = −P̄1,ξ , (36)
with the solution
√
Ū1 (ξ, Ω, Y ) = −H̄1 (ξ, Ω) + AŪD (Y, Ω). (37)
Note that the asymptotic analysis allows the velocity ’defect’ to depend on
the slowly varying variable Ω. From the continuity equation (4) we get the
Y -component of the velocity:
V̄1 (ξ, Ω, Y ) = ω H̄1,ξ (ξ, Ω)Y. (38)
From the dynamic boundary condition (8) and the relation for the surface
velocity (13) the perturbation of the Reynolds shear stress is obtained:
(U V )1 (ξ, Ω, 1) = −H̄1 (ξ, Ω), (U V )1 (ξ, Ω, 0) = 2H̄1 (ξ, Ω). (39)
The comparison of terms of order O(ε) leaves, however, the perturbation
H̄1 of the height of the free surface undetermined. An equation for H̄1 has
to be derived from equations for the higher–order terms.
Terms of order O(ε3/2 ) The analysis of the O(ε3/2 ) terms follows the
same lines as the analysis of the O(ε) terms. Formally in equations (35)-
(39) the subscript ’1’ has to be replaced by the subscript ’3/2’. We note
that in order to avoid secular terms ΔŪ1,Ω has to vanish. Thus we have
ŪD (Y, Ω) = ŪD (Y ).
Up to now we have not used the kinematic boundary condition at all.
Inspection shows that it is already satisfied.
In what follows the discussion of the terms of order O(ε2 ) and O(ε5/2 ),
respectively, will run along the same lines, with the exception that the
kinematic boundary condition will provide solvability conditions that will
serve as equations to determine the perturbation quantities H̄1 and H̄3/2 ,
respectively.
Terms of order O(ε2 ) Using the Y -momentum equation (5b) the O(ε2 )
pressure perturbation P̄2 can be expressed in terms of H̄2 as
9
P̄2 (ξ, Ω, Y ) = H̄2 + H̄1,ξξ (1 − Y 2 )ω 2 − A(V 2 )0 . (40)
2
Multiple Scales Analysis of the Turbulent Hydraulic Jump 207
Inserting P̄2 into the X-momentum equation (5a), using the continuity equa-
tion (4) and integrating with respect to Y gives:
√ √ 1
V̄2 (X, Ω, 1) = −3 + H̄1 + AŪD − 2 A ŪD dY ω H̄1,ξ
0
(41)
+3ω 3 H̄1,ξξξ + ω H̄2,ξ + ω H̄3/2,Ω
However, V̄2 has to satisfy the kinematic boundary condition as well, i.e.
which is a third order differential equation for H̄1 as a function of the fast
variable ξ. It is essentially the same differential equation (26) which we have
obtained in our first approach. Integrating (43) with respect to ξ we obtain
1
ω 2 H̄1,ξξ + H̄12 − H̄1 = R, (44)
2
Terms of order O(ε5/2 ) Using the Y -momentum equation (5b) and the
dynamic boundary condition (8) we obtain the O(ε5/2 ) pressure perturba-
tion P̄5/2 :
9 2 2 dω
P̄5/2 = H̄5/2 + ω H̄3/2,ξξ + ω H̄1,ξΩ + H̄1,ξ ω 1−Y2 . (46)
2 dΩ
208 H. Steinrück
Using the continuity equation (4), the dynamic boundary condition (8) and
equation (39) we obtain
Finally, with the help of the kinematic boundary condition for the O(ε5/2 )
terms, i.e.
with
A a dω
ωrξ = H̄1 − − H̄1,ξξΩ ω 3 − H̄1,ξξ ω 2 − RΩ . (51)
3 3 dΩ
where we have used (U V )0,Y =0 = 1 and (U V )1,Y =0 −(U V )1,Y =1 = 3H̄1 .
In the following we derive differential equations for the slowly varying
’constants’ R and S. Integrating (51) with respect to ξ over one period
gives the following differential equation for R:
1
a A
ω RΩ = − + H̄1 (ξ, Ω) dξ. (52)
3 3 0
Multiple Scales Analysis of the Turbulent Hydraulic Jump 209
Multiplying equation (44) with H̄3/2,ξ and adding (50) multiplied by H̄1,ξ
we get 2
ω
H̄3/2,ξξ + H̄3/2 H̄1 − H̄3/2− r H̄1,ξ +
1 (53)
+ ω H̄1,ξξ + H̄12 − H̄1 − R H̄3/2,ξ = 0.
2
Integrating (53) with respect to ξ over one period, and integrating the result
by parts yields
1
rξ H̄1 dξ = 0. (54)
0
Taking the derivative of (44) with respect to Ω and using (51) and (54) we
obtain the following differential equation for S:
a 1 A 1 2
ω SΩ = H̄1 dξ − H̄ dξ. (55)
3 0 3 0 1
Thus we have derived a set of differential equations for the slowly varying
constants of integration R and S. As we will discuss later the wave length
and amplitude of the undulated surface can be determined from R and S.
The main result of the multiple scales analysis can now be summarized
as follows: Let H(X; R, S) be the solution of
2 1
HX + H3 − H2 = 2R H + 2S, HX (0, R, S) = 0, HXX (0, R, S) > 0,
3
(56)
where R = R(Ω) and S = S(Ω) are the solutions of (52), (55). If H̄1 (0),
H̄1 (0) and H̄1 (0) is given initial conditions for R and S are obtained from
(27) and (28). Then the first-order perturbation of the free surface, i.e. H̄1 ,
is given by:
Ω
ξ − 12 dΩ Ω
H̄1 (X) = H ; R, S , X = , ξ=√ , (57)
ω 0 ω ε
with
1
X(ξ, ε) = √ X + X0 , (58)
ε
where X0 is chosen such that H(−X0 , R(0), S(0)) = H̄1 (0).
has been derived to describe the perturbation of the free surface of an open
channel flow near critical flow conditions. The equations is written here
in a form such that the parameters a, A and ε are the same as defined in
(15), (16, (17), respectively. In the derivation of (59) terms resulting from
the unbalanced external forces, which are of the order O(ε5/2 ) have been
included in the balance of the order O(ε2 ). As a result the perturbation
parameter ε appears in the equation for the leading order terms. However,
equation (59) can be taken as a the starting point of a multiple scales
analysis. We want to show here that the leading order terms for the fast
and slowly varying quantities agree with the results presents in the previous
section. Since equation (59) contains all the information of the multiple
scales analysis we call it a uniformly valid differential equation.
Equation (59) is equivalent to the first order system of differential equa-
tion
2
3ĤX = p(Ĥ, R̂, Ŝ), (60a)
√ √
ε ε
R̂X = −a + AĤ , ŜX = − −a + AĤ Ĥ, (60b)
3 3
where p(Ĥ, R̂, Ŝ) is the third order polynomial defined in (28). In order to
study the asymptotic behavior of the solution (59) in the limit ε → 0 we
can perform a multiple scale expansion according to the same lines as in
preceeding subsection
ˆ Ω̂) + . . . ,
Ĥ ∼ Ĥ1 (ξ, ˆ Ω̂) + . . . ,
R̂ ∼ R̂1 (ξ, ˆ Ω̂) + . . . ,
Ŝ ∼ Ŝ1 (ξ, (61)
where Ĥ1 , R̂1 , Ŝ1 are assumed to be periodic function with period 1 of the
fast variable ξ. ˆ The slow variable Ω̂ and the fast variable ξˆ are defined
analogously to (32). Now it is easy to see that the leading order terms Ĥ1 ,
R̂1 , Ŝ1 satisfy exactly the same differential equations as H̄, R and S. In
particular R̂1 and Ŝ1 are functions of the slow variable Ω̂ only.
We remark that one of the two perturbations terms on the right hand
side of (59) can be eliminated by setting
a √
Ĥ(X) = 1 + b (G(X/c) − 1) , b = 1 − c = 1/ b. (62)
A
We obtain √
ε c
G ± G (G − 1) = ÃG, Ã = , (63)
3 b
provided that A = a and A = 0. The plus sign holds for A > a and the
minus sign for A < a. √
If A = a we set b = A2/3 , c = 1/ b and obtain
√
ε c
G ±G G= ÃG, Ã = . (64)
3 b
Multiple Scales Analysis of the Turbulent Hydraulic Jump 211
Proof
(i) It is sufficient to consider the case A > 0. If a = 0 the transform
a √
H̄1 (ξ, Ω), = 1 + b Ḡ1 (ξ/c, Ω/c), b=1− , c = 1/ b (67)
A
can be used to eliminate a.
(ii) We consider the polynomial p(H) √ defined in (28). If R > 0 it has a
∗
local maximum√ at H max = 1 − 1 + 2R < 0 and a local minimum at
∗ ∗
H min = 1 − 1 + 2R > 0. √ If p(h, R, S) has three roots: h1 < H max <
∗
0 and h3 > H min = 1 − 1 + 2R > 0. Thus h2 = 6S/(h1 h3 ) > 0.
(iii) Since h2 and h3 are positive the integrals Ij , j = 0, 1, 2 are also posi-
tive. With A > 0 it follows that RΩ > 0 and SΩ < 0. Thus zero is a
lower bound for R and an upper bound for S. It can be shown that
the integrals Ij can be estimated by
√
0 < Ij < CR (1 + R)j , j = 0, 1, 2, (68)
212 H. Steinrück
Thus the amplitude of the oscillation decays and the mean value over an
oscillation period increases linearly with increasing X .
In this case the shift of the origin X0 has to depend on ε. It is chosen such
that the first wave crest is at X = 0. We have
1/2
1 1
X(ξ, ε) = √ X + X0 (ε), X0 (ε) = − √ dξ. (73)
ε 0 ω( εξ)
214 H. Steinrück
h3
multiple
scales
sol.
H̄1
h2
homo-clinic orbit
ï ï
X
Figure 3. Multiple scales solution (57) for A = 3, ε = 10−4 and the homo-
clinic orbit (A = 0) according to eq. (74). The solution for the slowly
varying variables is represented by h2 and h3 , i.e. the local minima and
maxima, respectively.
We will see later that X (0) = 0 and X0 is finite. Thus the multiple scales
approximation (57) fails for X → −∞. However, an asymptotic analysis
using the original coordinate X yields the inviscid solution
H̄1 (X) = H(X; 0, 0), (74)
which is a homo-clinic orbit in the phase plane. It is the limiting solution
for the first elevation and describes the behavior for X → −∞ correctly,
but fails to approximate the undular behavior.
In Figure 3 the multiples scales solution (57), evaluated for ε = 10−4 ,
A = 3 and the homo-clinic orbit are plotted. Moreover the solution of the
equations for the slowly varying variables represented by the zeros h2 and
h3 of the polynomial p(h; R, S) are also shown in Figure 3. In Figure 4 the
multiple scales solution (57) and the homo-clinic orbit (74) are shown in the
phase plane.
dH̄1
dX
ï
ï
homo-clinic orbit
ï
H̄1
Figure 4. Phase portrait of the multiple scales solution (57) for A = 3,
ε = 10−4 and the homo-clinic orbit (A = 0) according to eq. (74).
Using
R ∼ RΩ (0)Ω = 4A Ω, S ∼ SΩ (0)Ω = −8A Ω (76)
√ √ 8
h1 ∼ −4 AΩ, h2 ∼ 4 AΩ, h3 ∼ 3 + AΩ + ... . (77)
3
Expanding I0 for Ω 1 gives
1 1
∼ ln + 2 ln 6. (78)
ω AΩ
With
Ω
1 1
X (Ω) = dΩ ∼ Ω ln + 1 + 2 ln 6 (79)
0 ω AΩ
h̄C /h̄r #3
#1 #4
#5
data as in table 1
#2 0.95 ... 1.05 V̇
asymptotic result, eq. (81)
Fr
Figure 5. Comparison of calculated and measured amplitudes h C /h r of
the first elevation. The dotted lines indicate how the non-dimensionalized
height of the first crest changes for a variation of the volume flow rate within
±5% around the value given in table 1.
#3 #4
#1
calculated result (82)
#5
#2
# Exp. V̇ α h̄r w
[ m2 /s] [rad] [ m] [ m]
1 HCUJ8e 0.028 4.10×10−3 0.041 0.25
2 HCUJ10c 0.080 3.99×10−3 0.084 0.25
3 HCUJ4b 0.119 4.99×10−3 0.105 0.25
4 HCUH1b 0.0416 3.7×10−3 0.053 0.25
5 HCUJ4a 0.120 4.33×10−3 0.109 0.25
λas
# Exp. λ h̄C Fr β
λex
[ m] [ m] [−] [−] [−]
1 HCUJ8e 0.310 0.052 1.077 0.115 0.94
2 HCUJ10c 0.600 0.100 1.049 0.225 1.07
3 HCUJ4b 0.750 0.144 1.119 0.075 0.87
4 HCUH1b 0.450 0.068 1.088 0.086 0.83
5 HCUJ4a 0.660 0.138 1.062 0.171 1.20
4 Conclusions
The multiple scaling approach as proposed in this paper permits a self-
consistent asymptotic analysis of the undular hydraulic jump in turbulent
flow. Results for the perturbation of the height of the fluid are obtained
without making use of a turbulence model. Ordinary differential equations
have been derived for the slowly varying quantities, i.e. the amplitude and
the wave-length, as well as for the rapidly varying height of the free surface.
The price for the rigorous analysis is a limited range of parameters where
the requirements for the analysis are satisfied. In view of the high sensitivity
of the the wave length and the amplitude, the agreement with experimental
data appears reasonable.
Bibliography
V.M. Andersen. Undular hydraulic jump. J. Hydraulics Div., 104(HY8):
1185–1188, 1978.
Multiple Scales Analysis of the Turbulent Hydraulic Jump 219
T.B. Benjamin and M.J. Lighthill. On cnoidal waves and bores. Proc. Roy.
Soc., A224:448–460, 1954.
P. Böß. Berechnung der Wasserspiegellage. Forschungsarbeiten VDI, 284,
1927.
H. Chanson. Characteristics of undular hydraulic jumps. University of
Queensland, Department of Civil Engineering Research Report Series,
CE146, 1993.
H. Chanson and J.S. Montes. Characteristics of undular hydraulic jumps:
Experimental apparatus and flow patterns. J. Hydraulic Eng., 121:129–
144, 1995.
V.T. Chow. Open-Channel Hydraulics. McGraw-Hill, 1959.
W. Grillhofer. Der wellige Wassersprung in einer turbulenten Kanal-
strömung mit freier Oberfläche. Dissertation, Technische Universität
Wien, 2002.
W. Grilllhofer and W. Schneider. The undular hydraulic jump in turbulent
open channel flow at large reynolds numbers. Phys. Fluids, 15:730–735,
2003.
W. Hager. Energy Dissipators and Hydraulic Jump. Kluwer, 1992.
W.H. Hager and K. Hutter. On pseudo-uniform flow in open channel hy-
draulics. Acta Mechanica, 53:183–200, 1984.
F.M. Henderson. Open Channel Flow. Macmillan Company, 1966.
R.S. Johnson. Shallow water waves on a viscous fluid - the undular bore.
Phys. Fluids, 15:1693–1699, 1972.
R. Jurisits, W. Schneider, and Y.S. Bae. A multiple scales solution to the
undular hydraulic jump problem. PAMM, 7:4120007 – 4120008, 2007.
K. Kaufmann. Hydromechanik II. Springer, 1934.
H. Lauffer. Wassersprung bei kleinen Sprunghöhen. Wasserwirtschaft und
Technik, pages 137–140, 1935.
I. Ohtsu, Y. Yasuda, H. Gotoh, and M. Iahr. Hydraulic condition for
undular-jump formations. J. Hydraulic Res., 39:203–209, 2001.
R. Reinauer and W. Hager. Non-breaking undular hydraulic jump. J.
Hydraulic Res., 33:683–698, 1995.
H. Steinrück, W. Schneider, and W. Grillhofer. A multiple scales analysis of
the undular hydraulic jump in turbulent open channel flow. Fluid Dyn.
Res., 33:41–55, 2003.
Modern Aspects of High-Reynolds-Number
Asymptotics of Turbulent Boundary Layers
From Fully Attached to Marginally Separated Flows
Bernhard Scheichl†‡∗
†
Institute of Fluid Mechanics and Heat Transfer, Vienna University of
Technology, Vienna, Austria
‡
AC2 T research, Austrian Center of Competence for Tribology,
Wiener Neustadt, Austria
1 Introduction
Despite the rapid increase of computer power in the recent past, the calcula-
tion of turbulent wall-bounded flows still represents an extremely challeng-
ing and sometimes insolvable task. Direct-Numerical-Simulation computa-
tions based on the full Navier–Stokes equations are feasible for moderately
large Reynolds numbers only. Flows characterised by much higher Reynolds
numbers can be investigated if one resorts to turbulence models for the
small scales, as accomplished by the method of Large Eddy Simulation, or
for all scales, as in computer codes designed to solve the Reynolds-averaged
Navier–Stokes equations. Even then, however, the numerical efforts rapidly
∗
This material represents in some places a slight modification and in some an extension
of its original form published by Kluwick and Scheichl (2009) and is by courtesy of
these authors. Special thanks are due to AIC Androsch International Management
Consulting GmbH for generous financial support.
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
222 B. Scheichl
increase with increasing Reynolds number. This strongly contrasts the use
of asymptotic theories, the performance of which improves as the values of
the Reynolds number become larger and, therefore, may be considered to
complement purely numerically based work.
With a few exceptions (e.g. Deriat, 1986; Walker, 1998; Scheichl and
Kluwick, 2008a), studies dealing with the high-Reynolds-number properties
of turbulent boundary layers start from the time- or, equivalently, Reynolds-
averaged equations. By defining non-dimensional variables in terms of a
representative length L̃ and flow speed Ũ and assuming incompressible nom-
inally steady two-dimensional flow they take on the form
∂u ∂v
+ =0, (1a)
∂x ∂y
∂u ∂u ∂p 1 2 ∂u 2 ∂u v
u +v =− + ∇ u− − , (1b)
∂x ∂y ∂x Re ∂x ∂y
∂v ∂v ∂p 1 2 ∂u v ∂v 2
u +v =− + ∇ v− − . (1c)
∂x ∂y ∂y Re ∂x ∂y
Herein ∇2 = ∂ 2 /∂x2 + ∂ 2 /∂y 2 , and (x, y), (u, v), (u , v ), −u2 , −u v , −v 2 ,
and p are Cartesian coordinates measuring the distance along and perpen-
dicular to the wall, the corresponding time mean velocity components, the
corresponding velocity fluctuations, the components of the Reynolds stress
tensor, and the pressure, respectively. The Reynolds number is defined by
Re := Ũ L̃/ν̃, where ν̃ is the (constant) kinematic viscosity. Equations (1)
describe flows past flat impermeable walls when supplemented with the
usual no-slip and no-penetration conditions u = v = u = v = 0. Effects of
wall curvature can be incorporated without difficulty but are beyond the
scope of the present analysis.
When it comes down to the solution of the simplified version of these
equations provided by asymptotic theory in the limit Re → ∞, one is, of
course, again faced with the closure problem. The point, however, is that
these equations and the underlying structure represent closure-independent
basic physical mechanisms characterising various flow regions, identified by
asymptotic reasoning. This has been shown first in the seminal studies
by Yajnik (1970), Bush (1972), Fendell (1972), Mellor (1972), and more
recently and in considerable more depth and breath by Walker (1998) and
Schlichting and Gersten (2000) for the case of small-defect boundary layers,
which are subject of Section 2. Those exhibiting a slightly larger, i.e. a
moderately large, velocity defect are treated in Section 3. Finally, Section 4
deals with situations where the velocity defect is of O(1) rather than small.
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 223
2.1 Preliminaries
Based on dimensional reasoning put forward by L. Prandtl and Th. von
Kármán, a self-consistent time-mean description of firmly attached fully
developed turbulent boundary layers holding in the limit of large Reynolds
numbers Re, i.e. for Re → ∞, has been proposed first in the aforementioned
early asymptotic work (Yajnik, 1970; Bush, 1972; Fendell, 1972; Mellor,
1972). One of the main goals of the present investigation is to show that
this rational formulation can be derived from a minimum of assumptions:
(A) all the velocity fluctuations are of the same order of magnitude in
the limit Re → ∞, so that all Reynolds stress components are equally
scaled by a single velocity scale uref , non-dimensional with a global
reference velocity Ũ (hypothesis of locally isotropic turbulence);
(B) the pressure gradient does not enter the flow description of the viscous
wall layer to leading order (assumption of firmly attached flow);
(C) the results for the outer predominantly inviscid flow region can be
matched directly with those obtained for the viscous wall layer (as-
sumption of “simplest possible” flow structure).
In those outstanding papers (Yajnik, 1970; Bush, 1972; Fendell, 1972; Mel-
lor, 1972), uref is taken to be of the same order of magnitude in the fully
turbulent main portion of the boundary layer and in the viscous wall layer
and, hence, equal to the skin-friction velocity
1/2
uτ := Re −1 (∂u/∂y)y=0 . (2)
This in turn leads to the classical picture, according to which (i) the stream-
wise velocity defect with respect to the external impressed flow is small and
of O(uτ ) across most of the boundary layer, while (ii) the streamwise veloc-
ity is itself small and of O(uτ ) inside the (exponentially thin) wall layer, and,
finally, (iii) uτ /Ue = O(1/ ln Re). Furthermore, then (iv) the celebrated uni-
versal logarithmic velocity distribution
is found to hold in the overlap of the outer (small-defect) and inner (viscous
wall) layer. Here κ denotes the von Kármán constant; in this connection
we note the currently accepted empirical values κ ≈ 0.384, C + ≈ 4.1, which
224 B. Scheichl
refer to the case of a perfectly smooth surface (see Österlund et al., 2000)
and have been obtained for a zero pressure gradient.
This might be considered to yield a stringent derivation of the logarith-
mic law of the wall (3), anticipating the existence of an asymptotic state
and universality of the wall layer flow as Re → ∞; a view which, however,
has been increasingly challenged in more recent publications (e.g. Baren-
blatt and Goldenfeld, 1995; Barenblatt et al., 2000; Barenblatt and Chorin,
2000). It thus appears that – as expressed by Walker (1998) – “. . . although
many arguments have been put forward over the years to justify the log-
arithmic behaviour, non are entirely satisfactory as a proof, . . . ”. As a
result, one has to accept that matching (of inner and outer expansions),
while ensuring self-consistency, is not sufficient to uniquely determine (3).
In the following, from the viewpoint of the time-averaged flow description
the logarithmic behaviour (3), therefore, will be taken to represent an exper-
imentally rather than strictly theoretically based result holding in situations
where the assumption (B) applies. The description of the boundary layer
in the limit Re → ∞ can then readily be formalised.
In passing, we mention that in the classical derivations (see Yajnik,
1970; Bush, 1972; Fendell, 1972; Mellor, 1972) the assumption (B) is not
adopted, in favour of rather heuristic dimensional arguments already put
forward by (among other authors) Clauser (1956) that constitute the scal-
ings of the thicknesses of both the entire boundary and the wall layer. In
view of assumption (A) and dimensional notation, these are given by L̃ uref
and ν̃/(Ũ uref ), respectively. As demonstrated in detail by Mellor (1972),
then (3) results from matching, which decisively contrasts the present study
where it is imposed. However, this procedure additionally requires to an-
ticipate a priori that in the outer layer u ∼ u0 + O(uref ), u0 = O(1). On
balance, we thus render that classical approach less generic than that pro-
posed here and elucidated in the following.
Matching of the expansions (4) and (6) by taking into account (5) forces
a logarithmic behaviour of F1 ,
yields p0 (x) = pe (x), and is achieved provided γ := uτ /Ue satisfies the skin-
friction relationship
From substituting (4) into the x-component (1b) of the Reynolds equa-
tions (1) one obtains the well-known result that the total stress inside the
226 B. Scheichl
The necessary balance with the gradient of the Reynolds shear stress then
determines the magnitude of the thickness of the outer layer, i.e. δo = O(uτ ).
As a consequence, the expansions (6) are supplemented with
δo ∼ γ Δ1 (x) + · · · , (11)
The last equality is due to the usual boundary or patching conditions im-
posed, T1 (x, 1) = F1 (x, 1) = 0. The boundary layer equation (12a) is un-
closed, and in order to complete the flow description to leading order, tur-
bulence models for t+ and T1 have to be adopted. Integration of (12) then
provides the velocity distribution in the outer layer and determines the yet
unknown function C0 (x) entering (7) and the skin-friction law (8).
As a main result, inversion of (8) with the aid of (11) yields
Finally, the skin-friction law (13) implies the scaling law (iii), already men-
tioned in Subsection 2.1, which is characteristic of classical small-defect
flows.
the slow variation of γ, cf. (13), one obtains upon integration (cf. Walker,
1998),
y+
1 ∂u dUe /dx + γ dUe /dx 2
+ τ ∼ γ 2 Ue2 − y + u+ dy + · · · . (14)
Re ∂y γ Re Re 0
Herein the second and the third term on the right-hand side account for,
respectively, the effects of the (imposed) pressure gradient, cf. (10), and the
convective terms, which have been neglected so far. By using (5) and (12),
one can easily derive the asymptotic behaviour of τ as y + → ∞ (e.g. Walker,
1998). When rewritten in terms of the outer-layer variable η, this is found
to be described by
On the basis of these second order results, here matching with the wall
layer is achieved if
where
Herein x = xv denotes the virtual origin of the boundary layer flow. In the
present context flows associated with large values of β0 are of most interest.
1/2
By introducing suitably (re)scaled quantities in the form F1 = β0 F̂ (η̂),
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 229
1/2
T1 = β0 T̂ (η̂), η = β0 η̂, the momentum equation (19) reduces to
Q ∼ Q1 + Q2 + · · · , Q := F, P, Rx , Ry , T, W , (30a)
2
δo ∼ Δ1 + Δ2 + · · · , (30b)
β/βv ∼ B0 (x) + B1 (x) + · · · , βv → ∞ , (30c)
A graph of the relationship (37a) which represents the key result of the
analysis dealing with boundary layers that are in quasi-equilibrium (i.e. self-
similar to first and second order) having a moderately large velocity defect
is shown in Figure 2(b). Most interesting, it is found that solutions describ-
ing flows of this type exist for μ̂ ≥ μ̂∗ = 21/3 /6 only and form two branches,
associated with non-uniqueness of the quantity D̂, which serves as a mea-
sure of velocity defect, for a specific value of the pressure gradient. Along
the lower branch, D̂ ≤ D̂ ∗ = 21/3 and decreases with increasing values of μ̂,
so that the classical small-defect theory is recovered in the limit μ̂ → ∞,
where D̂ ∼ (9μ̂)−1/2 . In contrast, this limit leads to an unbounded growth
of values D̂ ≥ D̂∗ associated with the upper branch: D̂ ∼ 9μ̂ as μ̂ → ∞.
This immediately raises the question if it is possible to formulate a gen-
eral necessarily nonlinear theory which describes turbulent boundary layers
having a finite velocity defect in the limit of infinite Reynolds number. We
also note that the early experimental observations made by Clauser (1956)
seem to strongly point to this type of non-uniqueness.
The non-uniqueness is intrinsically tied to the nonlinearity of the inertia
terms in (1b). For small-defect boundary layers, these come into play as
inhomogeneities in the second-order flow description, as reflected by the
term D̂3 in (37a). It is, therefore, instructive to seek for double-valued self-
preserving flows for various values of m by starting from an ad-hoc boundary
layer approximation of the governing equations (1). We hence assume that
234 B. Scheichl
107
8
10
9
10
ï ï ï ï ï ï m μ̂
(c) (d)
Figure 2: Non-unique quasi-equilibrium flows; (a) F̂ (η), T̂ (η),
dashed : asymptotes found from (31b), (31c); (b) canonical representa-
tion (37a), dashed : asymptotes (see text) and osculating parabola in the
point (μ∗ , D̂∗ ); (c), (d) solid : solutions of (38) for various values of Re;
(c) circles: experimental data; (d) dashed : asymptote (37a) for Re → ∞.
The boundary conditions (38b) and (38c) account for the behaviour of F1 ,
F2 , T1 , T2 entering the expansions (16), even in case of a moderately small
defect still formally valid in the entire small-defect region, near the base
and the outer edge of the latter, respectively. When supplemented with an
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 235
asymptotically correct closure for t, i.e. one consistent with these conditions,
for sufficiently small values of γ and external flows represented by m the
quantities g, t, a represent an asymptotically accurate description of small-
defect flows that are in equilibrium up to second order with respect to the
expansions (16).
We again adopt (25) by setting t = (lg )2 , with l satisfying (35). In
order to detect non-uniqueness, (38) was solved numerically for g(η), t(η),
and m by prescribing values of Re and a. Then γ is computed from (8),
and increasing the value of a means increasing the boundary layer thickness
at some position x. The approximation β0 ∼ −ma[1 − g(1)]/γ 2 + O(γ 2 ) in-
1/2 1/2
ferred from (22) is used to evaluate ∼ γβ0 , Γ 1/3 = γ 1/3 β0 , and, in
turn, μ̂ and D̂ from (37b). Here we set R̂x = R̂y = 0, owing to the lack
of reliable closures for these stress components but in agreement with the
boundary layer approximation adopted, so that r is identified with the so-
called shape factor Ĝ (cf. Schlichting and Gersten, 2000). Also, the esti-
mates F̂ ∼ (1 − g )/ + O(2 ), F̂e ∼ [1 − g(1)]/ + O(2 ) are employed.
Figure 2(c) shows the comparison of the resulting relationship m(Ĝ; Re)
with data extracted from the measurements by Simpson et al. (1981) of a
massively separating boundary layer under the action of an adverse pres-
sure gradient (Schlichting and Gersten, 2000, p. 590): here Ue is found to
obey a power law according to (20) with m being a slowly varying function
of x, and the value of Re is estimated roughly as 3 × 106 . Notwithstand-
ing this uncertainty, the neglect of effects due to Reynolds normal stresses,
and the fact that the experimental flow is in equilibrium only locally, the
agreement with the results that are consistent with the asymptotic theory
is encouraging. Figure 2(d) uncovers the rather slow convergence of the
function μ̂(D̂; Re) towards the canonical relationship (37), cf. Figure 2(b),
attained for Re → ∞, which originates from the logarithmic dependence
of γ on Re.
layer. One expects that this effect will become more pronounced as the
velocity defect increases further, suggesting in turn that the outer part of
the boundary layer, having a velocity defect of O(1), essentially behaves
as a turbulent free shear layer. An attractive strategy then is to combine
the asymptotic treatment of such flows, (see Schneider, 1991) in which the
experimentally observed slenderness is enforced through the introduction
of a Reynolds-number-independent parameter α 1 with the asymptotic
theory of turbulent wall bounded flows.
which leads to
T̄ = Λ0 Y . (43)
Furthermore, T̄ and ψ̄ are subject to the boundary conditions
The solution of the inner wake problem posed by (43), (44) can be obtained
in closed form. It exhibits the expected square-root behaviour of ψ̄Y ,
(Λ0 Y )1/2
ψ̄Y ∼ Ūs (x) + 2 , L̄ ∼ χ(x)Ȳ , Y →0. (45)
χ(x)
Here Ūs (x) denotes the correction of the slip velocity Us (x) caused by the
inner wake region,
Us Δ0
Δ0,M
Δ0,G
Us
Δ0
xG xM
x
Figure 3: Solutions of (40) for |x − xM | 1, |k − kM | 1, dashed : asymp-
totes expressed by (48b) and (49).
∂ ψ̂ ∂ 2 ψ̂ ∂ ψ̂ ∂ 2 ψ̂ dP̂ ∂ T̂
− = −1 − Λ̂(Γ̂ ) + , (50a)
∂ Ŷ ∂ Ŷ ∂ X̂ ∂ X̂ ∂ Ŷ 2 dX̂ ∂ Ŷ
∂ 2 ψ̂ ∂ 2ψ̂
T̂ = , (50b)
∂ Ŷ 2 ∂ Ŷ 2
1 ∞ Â (Ŝ)
P̂ (X̂) = − dŜ , (50c)
π −∞ X̂ − Ŝ
Ŷ = 0 : ψ̂ = T̂ = 0 , (50d)
Ŷ → ∞ : T̂ − Ŷ → Â(X̂) , (50e)
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 241
UD y
y ∼ αΔ0 (x) O(α3/5 )
OW MD−
MD O(α6/5 ) MD+
3/2
O(α )
IW LD− LD+
LD
x
3/5
O(α )
Figure 4: Triple-deck structure, for captions see text, subscripts “−” and
“+” refer to the continuation of flow regions up- and downstream of the
local interaction zone, dashed line indicates inner wake.
ï ï X̂
Â,
Ûs Ŷ
P̂
Ûs
P̂ ï
Γ̂ Â
X̂D X̂R ï
ï
ï ï ï X̂ ï X̂Dï ï ï X̂ ï ï X̂Rï
Figure 5: Specific solution of (50); (a) key quantities, bottom abscissa: X̂-
values for Â(X̂), Ûs (X̂), top abscissa: X̂-values for P̂ (X̂), dashed : asymp-
totes; (b) streamlines, bold : separating streamline ψ̂ = 0.
The streamlines inside the lower-deck region (constant values of ψ̂) are
displayed in Figure 5(b), which clearly shows the formation of a recirculat-
ing eddy. Also, we draw attention to the increasing density of streamlines
further away from the wall and downstream of reattachment, associated
with the strong acceleration of the fluid there as evident from the rapid
increase of Ûs .
The interaction process outlined so far describes the (local) behaviour of
marginally separated turbulent flows in the formal limit 1/Re = 0. As in the
case of conventional, i.e. hierarchical, boundary layers having a velocity of
defect of O(1), additional sublayers form closer to the wall if 1/Re 1 but
finite. Their analysis, outlined in detail by Scheichl and Kluwick (2007b),
provides the skin-friction relationship in generalised form to include the
effects of vanishing and negative wall shear – treated first in a systematic
way by Schlichting and Gersten (2000) – but also shows that these layers
behave passively insofar as the lower deck problem (50) remains intact.
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 243
Bibliography
N. Afzal. Scaling of Power Law Velocity Profile in Wall-Bounded Turbulent
Shear Flows. Meeting paper 2005-109, AIAA, 2005.
G. I. Barenblatt and A. J. Chorin. A note on the intermediate region in
turbulent boundary layers. Phys. Fluids, 12(9):2159–2161, 2000. Letter.
G. I. Barenblatt, A. J. Chorin, and V. M. Prototoskin. Self-similar interme-
diate structures in turbulent boundary layers at large Reynolds numbers.
J. Fluid Mech., 410:263–283, 2000.
G. I. Barenblatt and N. Goldenfeld. Does fully developed turbulence exist?
Reynolds number independence versus asymptotic covariance. Phys. Flu-
ids, 7(12):3078–3082, 1995.
W. B. Bush and F. E. Fendell. Asymptotic analysis of turbulent channel
and boundary-layer flow. J. Fluid Mech., 56(4):657–681, 1972.
F. H. Clauser. The Turbulent Boundary Layer. In H. L. Dryden and
Th. von Kármán, editors, Advances in Applied Mechanics, volume 4,
pages 1–51. Academic Press, New York, 1956.
E. Deriat and J.-P. Guiraud. On the asymptotic description of turbulent
boundary layers. J. Theor. Appl. Mech., 109–140, 1986. Special Issue on
Asymptotic Modelling of Fluid Flows. Original French article in J. Méc.
Théor. Appl.
F. E. Fendell. Singular Perturbation and Turbulent Shear Flow near Walls.
J. Astronaut. Sci., 20(11):129–165, 1972.
P. S. Klebanoff. Characteristics of turbulence in a boundary layer with
zero pressure gradient. NACA Report 1247, NASA – Langley Research
Center, Hampton, VA, 1955. See also NACA Technical Note 3178, 1954.
A. Kluwick and B. Scheichl. High-Reynolds-Number Asymptotics of Tur-
bulent Boundary Layers: From Fully Attached to Marginally Separated
Flows. Invited Paper. In A. F. Hegarty, N. Kopteva, E. O’Riordan,
and M. Stynes, editors, BAIL 2008 – Boundary and Interior Layers.
High-Reynolds-Number Asymptotics of Turbulent Boundary Layers 245
P.-Y. Lagrée
CNRS & UPMC Univ Paris 06, UMR 7190,
Institut Jean Le Rond d’Alembert, Boı̂te 162, F-75005 Paris, France
[email protected] ; www.lmm.jussieu.fr/∼lagree
Abstract We present here the Interacting Boundary Layer Equa-
tions. It is called Inviscid-Viscous Interactions as well. This is a way
to solve an approximation of the Navier Stokes equations at large
Reynolds number using the Ideal Fluid / Boundary Layer decom-
position. But, instead of solving first the ideal Fluid and second the
Boundary Layer, both are solved together. This ”strong coupling”
(or this Viscous-Inviscid Interaction) allows to compute separated
flows. This was impossible in the classical Boundary Layer frame-
work, because in this framework, the boundary layer is constrained
by the Ideal Fluid which imposes its slip velocity at the wall. This
coupling is justified in the Triple Deck theory which is the rational
explanation of IBL. We present some numerical experiments show-
ing some simple academic examples of interactions such as flows
over bumps or wedges in subsonic, supersonic, subcritical and su-
percritical external flows and in pipes. Some examples from the
literature are then presented.
1 Introduction
The concept of Ideal Fluid/ Boundary Layer decomposition is classical (see
Schlichting books (42), (41) (17) or Prandtl (37)). The rational technique
of expansion has been presented by Van Dyke (48) and is explained in
its book (51). The procedure is as follows, starting from Navier Stokes
equations we put first 1/Re = 0; this gives the Euler description (called
”outer problem” see (51)). In this non viscous description, the flow slips
at the wall. This gives an ”outer velocity” at the wall, parallel to the wall.
This singular behavior √ is removed by the introduction of a thin layer of
relative thickness 1/ Re, the boundary layer (called ”inner problem” see
(51)). The velocity at the upper bound of this layer (at infinity in local
inner boundary layer variables) is by matching the ideal fluid velocity at
the wall. In this boundary layer, viscous effects act in order to decrease this
slip velocity to full fit the no slip condition.
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
248 P.-Y. Lagrée
)
7@501;-=2-5?
:@/417585?1
Figure 1. The typical problem a plane plate (neglect curvature) with a small
bump on it. The position L is used to scale x and y, the velocity U∞ is used to
scale the velocities
or ”dominant balance” (Van Dyke (51), Darrozès & François (11)): we want
the convective terms and at least re hook one diffusive term (as ȳ = ỹδ/L):
∂ ũ 1 ∂ 2 ũ
ũ ∝ ,
∂ x̄ Re(δ/L) ∂ ỹ 2
2
at the first position in x̄, a initial profile of ũ should be given to start the
resolution which is marching in x̄. With a given external velocity, those
equations are in principle solvable.
ideal fluid as a Taylor expansion near the wall for small ȳ (and taken into
account the incompressibility of the ideal fluid):
∂v̄ ∂ ūe
v̄ = v̄(x̄, 0) + ȳ + ... = v̄(x̄, 0) − ȳ + ...
∂ ȳ ∂ x̄
matching this velocity and the boundary layer velocity 2 shows that:
d
v̄(x̄, 0) = Re−1/2 (ūe δ̃1 ).
dx
Hence the boundary layer disturbs the ideal fluid at order Re−1/2 . It is
called the ”blowing velocity”. So the velocity in the ideal fluid (called
transpiration boundary condition as well) induces perturbations at the order
Re−1/2 in the fluid:
ū = ū1 + Re−1/2 ū2 , v̄ = v̄1 + Re−1/2 v̄2 , p̄ = p̄1 + Re−1/2 p̄2 ...
with ū1 (x, 0) = ūe (x). The classical asymptotic sequence (as described by
Van Dyke (51)) is then: the ideal fluid at order O(1) drives the boundary
layer at order O(1). In turn the boundary layer disturbs the ideal fluid at
order O(Re−1/2 ), then this perturbation creates a corrective boundary layer
at this O(Re−1/2 ) order, and so on. There is a cascade of disturbances at
increasing order (see figure from Van Dyke 10 left for this classical sequence).
The analysis was presented for a flat plate, but we will present in section
6 examples of flows over a flat plate with a small bump. The bump is
defined by the function f¯(x̄). Before going further, we have to present the
”Prandtl transform” which consists to change the transverse velocity and
variable: ṽ → ṽ − f¯ (x̄)ũ and ỹ → ỹ − f¯(x̄) and keep x̄ and ũ. With this
transformation, the Prandtl system is invariant, and now the position on
the wall is 0 again. The blowing velocity is then corrected by f¯ (x̄)ūe . This
is a trick which allows to change the bumpy wall in a flat one.
β 2n
ūe = x̄n with n = , β= .
2−β n+1
Interactive Boundary Layers 251
This is the so called Falkner Skan problem (41). There is a self similar
n
in the boundary layer, the longitudinal velocity is f (η)x̄ with
solution
ỹ
η = ( n+1
2 ) x̄(1−n)/2 . The selfsimilar stream function equation f (η) is from
Prandt Equation 1:
The Blasius solution is the solution of the flow √ over a∞flat plate, it corre-
sponds to n = 0 and β = 0, and f (0) = 0.332 2 and 0 (1−f )dη = 1.732.
The Hiemenz stagnation point solution is the solution of the flow against
a ∞flat plate, it corresponds to n = 1 and β = 1 and f (0) = 0.92 and
0
(1 − f )dη = 0.8. In principle, for a given geometry (β), one solves the
ideal fluid and obtains n. Then the equation is solved, one finds the value of
f (0) which allows to suit all the boundary conditions. See figure 6 where
f (0) is plotted as a function of β (we will see section 4.1 that the f (0)
function of β is in fact ∞ multivalued in β < 0. One other result is the dis-
placement thickness 0 (1 − f )dη. In fact, this naive direct resolution gives
a very stiff problem, and in practice, for a given β and a negative f (0) it
is impossible to solve the system (eq. 3) in the case of separated flows.
∂ ∂ ūe ∂ ∂ 2 ũ
(ũūe − ũ2 ) + (ūe − ũ) − (ṽ(ũ − ūe )) = − 2 .
∂ x̄ ∂ x̄ ∂ ỹ ∂ ỹ
Using the displacement thickness (eq. 2), and defining the momentum thick-
ness δ̃2 and the shape factor H:
∞
ũ ũ δ̃1
δ̃2 = (1 − )dỹ and H = ,
0 ūe ūe δ̃2
and defining a function f2 linked to the skin friction as: ∂∂ũỹ = f2 Hδū1 e gives
the following equation where the ideal fluid promotes the boundary layer:
Initial condition is for example δ̃1 (0) = 0 (but the Hiemenz value may be
a good first guess) and ūe (0) = 1. In the classical approach, δ̃1 is obtained
through the knowledge of ūe , which we write formaly δ̃1 = F (ūe ).
To solve this boundary layer equation, a closure relationship linking H
and f2 to the velocity and the displacement thickness is needed. This is of
course a strong hypothesis. Defining Λ1 = δ̃12 dū e
dx̄ , the system is closed from
the resolution of the Falkner Skan system using the following fit (fig. 2):
,
2.5905e−0.37098Λ1 if Λ1 < 0.6 1 4
H= , f2 = 1.05(− + 2 ). (5)
2.074 if Λ1 > 0.6 H H
It means that we suppose that each profile remains a Falkner Skan one in
the boundary layer. We used this crude solution in exponential with the
value of the sink solution H = 2.074 as a limiting value (Lorthois & Lagrée
(32)). We tested it to be enough good, other closures may be found in the
literature. Some closures use the concept of entrainment. The closure may
be done with other families of profiles, and Pohlhausen profiles are good
candidates (the solution is part of a polynomia). With those profiles the
reverse flow is over estimated compared to Falkner Skan.
I
+
Figure 3. A sketch from Prandtl (37) of the flow near the point of vanishing
shear stress.
can not compute numerically (here by finite difference) the boundary layer.
Using the Von Kármán equation (Eq. 4 and closure 5) gives the same be-
havior. It fails nearly at the same point (not exactly, but not so bad).
A simple way to see the problem is to observe the Von Kármán equation
linearised around a decelerating flow. Linearizing the velocity near the point
of separation is say ūe = sin(x̄s )−a(x̄− x̄s ) with a = cos(x̄s ) and linearizing
around small Λ1 (which is not true but is a enough good approximation)
gives H = H0 − Hp Λ1 where H0 = 2.59 and Hp −0.96, so the variation
of δ1 /H with respect to x̄ is:
d
therefore dx̄ δ̃1 is infinite, we can not march in x̄ anymore. This crude
estimation shows that the separation point is impossible to cross, but direct
numerical finite difference solution of the boundary layer equations gives
the same result (figure 4).
This difficult problem has been examined, among others by Landau (in
the classical text book (25)) and by Goldstein (18). Landau (25) noticed
that as in the boundary layer v u, so the transverse velocity must increase
a lot to be as large as the longitudinal one. It is apparently the case when the
flow is separated (stream lines are ejected from the wall). In boundary layer
variables Landau infer that v = ∞ and ∂v/∂y = ∞ so that ∂u/∂x = −∞.
The velocity is strongly decelerated near the point of separation xs . So
∂x
he proposes to work with the inverse of the function ( ∂u ) which is zero at
separation and proposes a reciprocal expansion of x in u near xs as (where
Interactive Boundary Layers 255
δ1 τ
x x
Figure 4. Boundary layer separation on a cylinder, the outer velocity is
ūe = sin(x̄), points are numerical finite difference solution of the Boundary
Layer equations, line is the integration of Von Kármán equation with the
proposed closure (Eq. 4 and 5). Separation occurs for an angle of 104o .
∂x 1 ∂2x 1
x−xs = (u−us )+ (u−us )2 +... = 0(u−us )+ 2 (u−us )2 +...
∂u 2 ∂u2 4β (y)
so that one may write the velocity u and by the continuity equation v as:
√ β(y)
u = us (y) + 2β (y) xs − x + ... and v = √ + ...
xs − x
Injecting it in the momentum equation, neglecting the viscosity and writing
the total derivative as u2 (∂y ( uv )) shows that v/u does not depend on y, an
hint for the profile near separation may be deduced as :
∂us ∂A √
u = us (y) + A(x) v = − us with A(x) ∝ xs − x.
∂y ∂x
Unfortunately this description does not fit the good boundary conditions at
the wall....
Goldstein showed that for a given external flow, one can not compute the
boundary layer if the skin friction vanishes (which is consistent to Landau
approach), the skin friction behaves as:
τp = 48a4 (xs − x)),
One other strange problem appeared in the 50’ at the time of the super-
sonic conquest: the problem of ”Upstream Influence”. A model configu-
ration for a supersonic wing was the aligned flat plate in a compressible
supersonic flow. In various experiments in supersonic flows (Ackeret Chap-
man and others, see (45)), it was observed that an impinging shock wave on
a boundary layer produces perturbations far upstream the point of reflex-
ion of the wave. The boundary layer deviates from its basic state upstream
of the impinging shock, see photos on figure 5 from Stewartson 64 book
(45). On this figure we even see that three different accidents (an imping-
ing shock, a forward facing step and a wedge) produce the same upstream
flow. The deviation occurs far away (in boundary layer thickness units)
from the accident.
In the classical supersonic framework this is impossible (figure 5 lower
left). First the ideal fluid is supersonic (hyperbolic) so perturbations move
downstream in the Mach cone. Second, the boundary layer is parabolic,
so perturbations move downstream and across the boundary layer. No dis-
turbance can theoretically move upstream against the flow. This is the
upstream influence paradox.
Is it a dead end? No!
Interactive Boundary Layers 257
along a wedge of half angle βπ/2 (see figure 6 left), the so called Falkner
Skan problem (41). A naive direct resolution gives a very stiff problem, and
in practice, for a given β it is impossible to find the f (0) which solves the
system 3 in the direct way when there is separation. We can nevertheless
obtain flow separation for some values ofβ. To obtain it, we have to solve in
∞
an inverse way, we impose the thickness 0 (1−f )dη, and we find the value
of β associated. Hence, a simple way to feel that the boundary layer must be
solved in inverse way is really the Falkner Skan case. It is representative by
many aspects of the boundary layer behavior: for a given external velocity
one has a given β and one computes the corresponding profile. But, we see
on figure 6 that if the external velocity is with a β to much negative, there
is no solution. Only for an ad hoc external velocity we have solution(s).
η
f )d
∞
(1 −
π 0
β
2
Β
Figure 6. Left the Falkner Skan problem: self similar flow on a wedge. Right,
the numerical solution, value of f (0) as a function of β. Not every external
velocity is compatible with the boundary layer, for example in the Falkner Skan
case, too small β (less than -0.199) are not relevant (small dashing). A larger
value of the outer velocity gradient (large dashing) gives solutions.
layer, so that the retroaction can travel back in this subsonic layer. In
fact it is not the good mechanism as the upstream influence would be of
same length than this subsonic layer is thick. But on the experiments, the
longitudinal scale is far larger than the boundary layer thickness.
• Garvine (16) proposed a simplified boundary layer model linearising around
u = 1 the supersonic boundary layer (neglecting thermal effects):
ỹ
2
∂x̄ ũ = −∂x̄ p̄ + ∂ỹ ũ, ṽ = − ∂x̄ ũdỹ
0
and writing the Ackeret formula (linking the pressure and the blowing ve-
locity) as:
1
p̄ = √ √ ṽ(δ̃)
Re M 2 − 1
he obtains after claiming δ̃ = cst (yes he did!) that the pressure gradient is
δ̃
− √Re√1M 2 −1 0 ũx̄x̄ dỹ so that a model equation of the interaction is:
δ
1
∂x̄ ũ = √ √ ux̄x̄ dỹ + ∂ỹ2 ũ.
Re M 2 − 1 0
He pointed out the come back of ellipticity due to this ux̄x̄ term. He then
obtained a set of eigen solutions with Laplace transform, in fact the expo-
nentially growing one of those solutions can be obtained in looking to eK x̄
260 P.-Y. Lagrée
So the coupling of the two equations produces self induced explosive solu-
tions.
• Numerically those explosive solutions were obtained by Werle Dwoyer,
and Hankey (54) (among others). On figure 8 we have a clear example of
what happens when solving in a marching way the coupled system. Starting
from a given initial location they solved the coupled boundary layer system
with the so called tangent wedge law (valid for stronger shocks than the
linearised Ackeret formula). They showed that changing a bit one parameter
may cause different solutions. Those are called ”branching solutions”.
Figure 8. Branching solutions (54): changing a bit one parameter may cause
different solutions while solving the equations with a marching scheme.
• One may consider the most simple argument, see Le Balleur (26). He
considers the strong coupling of the boundary layer in Von Kármán form
(neglecting again thermal effects) with the Ackeret formula (linking the
perturbation of pressure at M > 1 due to the variations of the effective wall
(represented by δ1 ) as:
so that, supposing that ūe is nearly one and ∂x̄ ūe = −∂x̄ p̄
this equation is ”not so far” from from the basic flow with subscript 0 and
ūe ∼ 1, so after linearisation.
d δ̃10 1 d2 δ̃1
δ̃1 = (H0 + 2) √ √ + ....
dx̄ ūe Re M 2 − 1 dx̄2
where we forget the contribution of the skin friction. So again, we obtain
exponential solutions (called supercritical by Crocco and Lees in 52) for the
disturbance of the displacement thickness δ1 :
√ √
Re M 2 −1
δ̃10 (H0 +2)
x̄
e .
It is nearly the same result than Garvine and than the one obtained numer-
ically by Werle et al.
• In fact, Lighthill in 53 (30) proposed a pre-theory of triple deck explaining
most of the mechanism (see in Stewartson 64 book as well) based on steady
perturbations of the Orr Sommerfeld equation.
• The real definitive theory is the Triple Deck (see Ruban’s contribution).
In this framework, those kind of explosive solutions are called ”self induced
solution” (see Neiland (35), Messiter (34) and Stewartson (46)).
Figure 9. Left, the Triple Deck scales. Right, ”triple decker ship of the line”
from HMS victory brochure Porthmouth (”vaisseau de ligne à trois ponts”). In
german ”Dreierdeck-Theorie”, a french translation of Triple Deck Theory may be
”Triple Pont” instead of ”Triple Couche”.
∂ p̃
ũ = UB (ỹ) + εA(x)UB (ỹ) ṽ = −ε2 A (x)UB (ỹ), and = 0.
∂ ỹ
Note that this function −A is reminiscent to the Landau analysis of section
3. With this description, the velocity is not zero but εA(x)UB (0) on the
wall, so we have to introduce a new layer to full fit the no slip condition.
We reobtain Prandtl equations but with a new transverse scale δ3 = ε5
associated to the longitudinal one x3 = ε3 , in which the longitudinal velocity
is scalled by ε. In this new layer, the Lower Deck, the final system is then:
∂u ∂v ∂u ∂u dp ∂2u
+ = 0, u +v =− + 2. (7)
∂x ∂y ∂x ∂y dx ∂y
With no slip condition at the wall (u = v = 0), the entrance velocity profile
u(x → −∞, y) = UB (0)y, and the matching condition with the Main Deck:
u(x, y → ∞) = (y + A)UB (0). Note, that the system is parabolic, there is
no output condition needed to solve it.
Going back in the Main Deck, the disturbed velocity at the top of the
Main Deck, for ỹ → ∞:
ũ = 1; ṽ = −ε2 A (x),
with φ(0) = φ (0) = 0 and say φ (∞) = UB (0) so that A1 = eKx ; as the
incompressibility is fulfilled, the momentum gives (see Stewartson (47) or
Sychev et al. (49) for details):
∂ 2 φ (y)
= UB (0)Kyφ (y), with φ (0) = KP,
∂y 2
and p1 = − √KA 1
M 2 −1
so that we deduce that the supersonic case allows then
an eigen solution
K = (−3Ai (0)( M 2 − 1)UB (0)2 )3/4 .
This exponential behavior at the longitudinal Triple Deck scale is the ra-
tional explanation of the observed self induced separation. This upstream
influence is then understood as a not well posed problem. In fact, even if
each part of the flow seems hyperbolic/ parabolic, due to the interaction
one recovers the output influence.
This is the case in the supersonic flows, in shallow water flows at small
Froude number, in mixed convection. But there exist flows with no up-
stream influence: for example in the symmetrical pipe flows.
∂ ũ ∂ṽ ∂ ũ ∂ ũ dūe ∂ 2 ũ
+ = 0, ũ + ṽ = ūe + 2,
∂ x̄ ∂ ỹ ∂ x̄ ∂ ỹ dx̄ ∂ ỹ
d(δ̃1 ūe )
v̄e = Re−1/2 .
dx̄
Hence, the outer flow is no more only given by the wall f¯(x̄) (so that the
blowing velocity is f¯ ūe ) but, the wall is ”thickened” by the boundary layer
thickness (or ”blowing velocity”, or ”transpiration boundary condition”),
so that for a subsonic flow:
¯
1 f (x̄)ūe + Re−1/2 d(δ̃dx̄
1 ūe )
ūe = 1 + dξ
π x−ξ
or in a supersonic flow
1 d d(δ̃1 ūe )
ūe = 1 − √ [ f¯(x̄)ūe + Re−1/2 ].
M 2 − 1 dx̄ dx̄
Instead of the usual weak coupling with the hierarchy (figure 10 left), the
boundary layer retroacts on the ideal fluid (figure 10 right). So even if ūe
appears in the definition of himself through d(δ̃dx
1 ūe )
, it is not an issue because
of the iterations involved in the solution. The boundary layer thickness δ̃1
acts as a fictive wall, it disturbs the ideal fluid, the pressure (pressure and
velocity ūe (x̄) are linked) develops the boundary layer itself. It is a strong
interaction. The two layers are coupled. It explains the term ”Interactive
Boundary Layer”, or ”Viscous Inviscid Interaction”.
we have to decompose it into two parts as we cross the Lower and the
Main Decks. Let us introduce Ỹ so that we cut the integral in two parts
266 P.-Y. Lagrée
Figure 10. Left the Classical sequence, image taken from Van Dyke’s book (51).
Right the Interactive Boundary Layer, we do not follow the classical asymptotic
sequence (from left): the ideal fluid at order O(1) drives the boundary layer
at order O(1), in turn the boundary layer disturbs at the ideal fluid at order
O(Re−1/2 ), then this perturbation creates a corrective boundary layer at this
O(Re−1/2 ) order, etc. But, right, we couple the boundary layer and the ideal
fluid at first order.
Ỹ ∞
( 0 + Ỹ ). The first integral is estimated near the wall, so the Lower Deck
description (ỹ = εy) is valid there, but a good idea is to write the velocity
u(x, y) = UB (0)(y + A) + uc where uc is a correction, so the first integral:
Ỹ Ỹ /ε Ỹ /ε
( (1 − ũ(x̄, ỹ))dỹ) = ε( (1 − ε(UB (0)(y + A)))dy − εuc dy)
0 0 0
Re summing the two integrals and changing the order of the terms allows
to recognise :
∞ ∞ Ỹ /ε
−1/2
δ1 = (Re ){ (1−UB (ỹ))dỹ+ (−εA(x)UB (ỹ))dỹ−ε2 uc dy)}.
0 0 0
or ∞
δ1 = (Re−1/2 ){ (1 − UB (ỹ))dỹ − εA(x) − O(ε2 )}.
0
the −εA contribution
∞ of the Triple Deck is the perturbation of the displace-
ment thickness 0 (1 − UB (ỹ))dỹ. The IBL equations (based on δ1 ) even if
they seem to be ill posed as they mix different order of magnitude may be
justified by the Triple Deck analysis (based on −A).
Interactive Boundary Layers 267
∂ ∂ r̃ṽ ∂ ũ ∂ ũ ∂ p̃ ∂ ∂ ũ ∂ p̃
ũ + = 0, ũ + ṽ =− + (r̃ ), 0=− . (8)
∂ x̃ r̃∂ r̃ ∂ x̃ ∂r̃ ∂ x̃ r̃∂ r̃ ∂ r̃ ∂ r̃
slip velocity.
The second ”box” is of course the boundary layer box, given an outer
velocity, it computes the displacement thickness. The equations may be
laminar or turbulent with any turbulent model. It may be full finite differ-
ences resolution or Von Kármán integral method. This box may be used in
standard direct way: for a given slip velocities, it computes a displacement
thickness. This box may be used in reverse, given a displacement thickness
it computes what outer velocity produces it.
Coupling Now, we couple the boxes and present the various possibilities.
In fact we will use δ1 and u in the following figures (we forget all the tildes
in this section). We may use dδ 1
dx instead of δ1 , and instead of u we may use
dp
−p (by Bernoulli linearised) or we may use dx . There is no real influence of
the choice of δ1 instead of his slope, nor in u, p or his gradient (as we deal
with small perturbations).
• Now, having those boxes, we have to branch them. First, the classical
boundary layer theory may be represented as a ideal fluid box followed by
a boundary layer box, figure 11.
yw
ue
δ1
Figure 11. Classical Boundary layer, the geometry gives the velocity which gives
the boundary layer.
yw + δ1
ue
δ1
Figure 12. ”Direct method”: the geometry gives the velocity which gives the
boundary layer, the rebranching will give the second order effects.
• The good way to solve the boundary layer, is to solve it in inverse, we can
imagine that we solve the ideal fluid in inverse as well. This is the ”inverse
270 P.-Y. Lagrée
y w + δ1
ue
δ1
Figure 13. ”Inverse method”, the total geometry (boundary layer thickness and
effective geometry) give the velocity which gives a total geometry, and so on.
• The good way to solve the boundary layer, is to solve it in inverse, the
good way to solve the ideal fluid is in the direct way. So we have to relax the
input depending on the difference of the outputs. This is the semi-inverse
coupling by Le Balleur (figure 14).
Figure 14. ”Semi Inverse method”, inverse boundary layer, direct ideal fluid.
The difference of the two output velocities is used to update the displacement
thickness, and so on.
One has to notice that by the Bernoulli relation variation of velocity are
opposite of variation of pressure so that we can write as well:
δ
8H
τ
[
Figure 15. Incompressible flow. Top the velocity field ũ, ṽ (Prandtl trans-
form), bottom the wall, here a bump, the displacement thickness δ̃1 (start-
ing from Blasius value 1.7 in x̄ = 1), the skin friction (starting from Blasius
value 0.3 in x̄ = 1) and the outer velocity starting from Ideal Fluid value
1 in x̄ = 1. A positive disturbance of the wall increases the velocity and
decreases the displacement. Separation occurs after the bump.
Le Balleur (see (27) and Wigton and Holt (55)) analysis consist to lin-
earize the equation. He defines two operators, one for each box, first B ∗
defined as δ n = B ∗ pnBL and for the ideal fluid, he defines in the same vein
a linear response δ n = Bpne . Then the update is as:
To make it clear, we use Fourier analysis for all the frequencies between
π/L and π/Δx (the smallest linked to the domain size, and the highest
linked to the discretisation step). Furthermore, the B operator may be
obtained in subsonic flow we have B = −1/k. The analysis is then very
simple, defining a ”gain” G = δ n+1 /δ n :
1
G = 1 − λ( + k),
B∗
272 P.-Y. Lagrée
Δδ
SH
τ
[
Figure 16. Supersonic flow on a flat plate with a bump. Top, the velocity
field ũ, ṽ (Prandtl transform), bottom the wall, here a bump, the perturba-
tion of displacement thickness from Blasius Δδ̃1 (starting from 0 in x̄ = 1),
the skin friction (starting from Blasius value 0.3 in x̄ = 1) and the outer
pressure starting from Ideal Fluid value 0 in x̄ = 1. Note the pressure
plateau (here tiny) associated to separation.
we want |G| < 1 for π/L < k < π/Δx. Often ((27), (55)), it was considered
that B ∗ was real (which is not true), so we can find an optimal λ.
For a supersonic flow we have B = (ik/(M 2 − 1))−1 . It is easy to show
that in this case it is impossible to find an optimal λ. The coupling is always
unstable. The good coupling is in fact with the derivative of the pressure:
d n d n
δ n+1 = δ n − μ( p − p )
dx BL dx e
then again we have:
which allows to define a ”gain” G = δ n+1 /δ n . We want |G| < 1 for all the
space frequencies π/L < k < π/Δx. We can find an optimal μ.
Interactive Boundary Layers 273
δ
8H
τ
[
Figure 17. Subcritical flow on a flat plate. Top, the velocity field ũ, ṽ
(Prandtl transform), bottom the wall, here a bump, the displacement thick-
ness δ̃1 (starting from Blasius value 1.7 in x̄ = 1), the skin friction (starting
from Blasius value 0.3 in x̄ = 1) and the outer velocity starting from Ideal
Fluid value 1 in x̄ = 1. A positive disturbance of the wall increases the
velocity and decreases the displacement. Separation takes place after the
bump. There is no upstream influence.
6 Examples
δ
8H
τ
[
Figure 18. Supercritical flow on a flat plate. Top, the velocity field ũ, ṽ
(Prandtl transform), bottom the wall, here a bump, the displacement thick-
ness δ̃1 (starting from Blasius value 1.7 in x̄ = 1), the skin friction (starting
from Blasius value 0.3 in x̄ = 1) and the outer velocity starting from Ideal
Fluid value 1 in x̄ = 1. A positive disturbance of the wall decreases the
velocity and decreases the displacement. Separation occurs far before the
bump, note the long upstream influence and the large increase of δ̃1 .
ūe = 1 + dξ
π x−ξ
Before the bump there is a small decrease of the ūe velocity. In a pure
Hilbert case, the response in ūe is perfectly symmetrical, but here, due to the
boundary layer, the velocity is no more symmetrical. Due to the acceleration
on the bump, the displacement thickness first decreases and increases again
after the bump. It increases more. So, there is a small overshoot of the
thickness associated with the boundary layer separation. This makes the
outer velocity non symmetrical. The skin friction increases before the crest,
and decreases after. This is consistent with the fact that, for instance, before
the crest, the velocity increases, and the boundary layer thickness decreases,
Interactive Boundary Layers 275
Δδ
SH
τ
[
Figure 19. Supersonic flow on a flat plate with a wedge. Top, the velocity
field ũ, ṽ (Prandtl transform), bottom the wall, here a wedge in x̄ = 3.5, the
perturbation of displacement thickness Δδ̃1 (starting from 0 in x̄ = 1), the
skin friction (starting from Blasius value 0.3 in x̄ = 1) and the outer pressure
starting from Ideal Fluid value 0 in x̄ = 1. Note the plateau pressure: the
pressure is nearly constant in the separated bulb before the wedge, and note
the separation occurs far upstream of the wedge.
so the slope of the velocity in the boundary layer increases (it is more or
less the ratio of ūe and δ̃1 ), the reverse happens after. We notice that the
maximum of the skin friction is before the crest, after the inflexion point of
the bump, the velocity increases less, but the boundary layer continues to
decrease because of the inertia of the fluid, so the maximum of skin friction
is between the inflexion point of the bump and the crest. There is eventually
a separated bulb with negative skin friction.
1 d d(δ̃1 ūe )
ūe = 1 − √ [ f¯(x̄) + Re−1/2 ].
2
M −1 dx̄ dx̄
The bump creates upstream influence and a separated bulb far upstream.
The skin friction reincreases and then redecreases to create a second sepa-
rated bulb.
1
ūe = 1 + [f¯(x̄) + δ̃1 Re−1/2 ]
1−F
It means that the velocity increases and decreases after the crest (see figure
17). The skin friction is extremal just before the crest, and there may be
flow separation on the lee side. The behaviour is nearly the same than in
the incompressible case but there is no influence of the bump before the
beginning of it. In the incompressible case there was some small effect due
to the Hilbert integral, but here there is no effect before the bump.
∂ ũ/∂ ỹ
x̃
Figure 20. Longitudinal evolution of the wall shear stress near the incipient
separation case, in an axis symmetrical flow with a stenosis (a constriction). Re-
duced Navier Stokes Prandtl model, integral IBL, full IBL resolution (in RNSP
variables, the bump is located in x = 0.02, and its width is 0.00125), and Triple
Deck resolution. All the curves are rescaled in Triple Deck scales.
that the result from system (8) compares to IBL with integral resolution
and with full equations and triple Deck (case p = A).
The set of RNSP equations in 2D may be solved and compared with
Interactive Boundary Layer equations. This has been done for example in
(23). On figure 21 is the 2D symmetrical flow between two plates with a
constriction. We compare Navier Stokes solved with Castem2000 (5), IBL
(Integral resolution) and RNSP (finite differences). Pressures are nearly the
same.
p̃(x̃)
P̃m P̃t
P̃g
(x − xg )/Rg
(h0/U0)∂(u∗)/∂y∗
p∗/(ρU02 )
x x
Figure 22. Left Right Comparison of integral IBL and NS (with Castem 2000)
pressures. The IBL approach well predicts the over pressure on the flat wall and
the positions of the minima of of the pressures after the throat. Skin friction,
comparison of integral IBL and NS. The integral IBL over predicts the maximum
of skin friction but well predicts the position of the point of separation. The
incipient separation before the bump is well predicted.
Interactive Boundary Layers 279
Τ
Figure 23. A bump in a Poiseuille flow at the lower will disturb the core flow,
the pressure changes across the core flow, perturbations are induced at the upper
wall. Linear perturbation of skin friction τ1 (left) and pressure p1 (right) over a
bump f1 (x) = cos(πx/2)2 , for −1 < x < 1 in the Triple Deck framework. The
A = 0 is in plain line and the A case is in dashed line.
(on the top) decreases before the bump, and pressure on the bumpy wall
increases before the bump).
Note that we recover a result that looks like Smith (43) (or Sobey (44))
result in pipe flow, the transversal perturbation of pressure in a perturbed
Poiseuille flow is ph − pb = A /30 where −A is a displacement of the stream
lines as δ1b − δ1h is. Of course the two configurations are very different. On
figure 23 we plot the perturbation of a Poiseuille skin friction in the linear
case for the A = 0 symmetrical case and the non symmetrical A case (see
Smith (43)). We see the that the case A = 0 presents no upstream influence
as already mentioned, but we clearly see that the case with A promotes
upstream response of the flow (before the first position of the bump, the
pressure has increased and the skin friction has decreased).
Figure 24. Example of comparison of IBL computation, Drela & Giles (12)
Figure 26. Example of comparison of IBL computation, Lock & Williams (31)
2#0
7 Conclusion
So we now know that the Boundary Layer equations are more than
useful. They can handle flow separation and compute reverse flow bubbles.
The strong coupling between the Ideal Fluid and the Boundary Layer allows
this. The explanation of this lies in the Triple Deck theory which couples a
thin wall layer (the Lower Deck) with the Ideal Fluid (Upper Deck) trough
a displacement of the stream lines in the Boundary Layer (the Main Deck).
Further more, the IBL allows as well some upstream influence. It means
that in some special régimes such as supersonic, hypersonic, supercritical
flows, in non-symmetrical 2D pipe flows, disturbances in the Boundary
Layer influence the flow pattern far upstream the position of the distur-
bance. In axisymmetrical pipe flows, in symmetrical 2D pipe flows, in sub-
critical flows, there is no upstream influence at all. In the subsonic case,
there is only a small influence due to the elliptic character of the flow (seen
with the Hilbert integral which is global).
The methodology of IBL, or Inviscid Viscous Interactions, may be sum-
marized in the figure 28 extracted from Jameson (20). Even if this paper
was written in 1983, it seems that most of the flying aircraft have been
defined since by Viscous Inviscid interactions. The Airbus A380 is one of
the first aircraft designed with ”full Navier Stokes” (in fact certainly crude
RANS models). In the late 90’, before the end of the century, a large effort
has been done on Navier Stokes solvers. Lot of people are working on this
equation. Tremendous progress have been done, and with Navier Stokes,
the complexity of the geometry is a problem with lot of solutions. So Navier
Stokes solvers are very promising, and give a lot of practical results.
To a certain extent, IBL-IVI methods are less versatile and require spe-
Interactive Boundary Layers 283
cific methods, they need a kind of savoir faire (Aftosmis et al. (1) point
some difficulties of the IBL). But when used, they are very good. For ex-
ample Le Balleur has codes which may compute even large stall on wings,
giving results very close to experiments. NS solver have difficulties to re-
produce those results. XFOIL free code from Drela allows now everybody
to do quick computations of flows over airfoils.
The review of Piomelli & Balaras (36), shows that up to now only very
simple models are used for boundary layer near the wall. They suggest a
coupling of a Large Eddy Simulation Navier Stokes with a boundary layer
code near the wall.
Bibliography
[1] Michael J. Aftosmis Marsha J. Berger, Juan J. Alonso, (2006) ”Applica-
tions of a Cartesian Mesh Boundary-Layer Approach for Complex Con-
figurations”, 44th AIAA Aerospace Sciences Meeting, Reno NV, January
9-12, 2006 AIAA 2006-0652
[2] G.R. Barrenechea, F. Chouly (2009) ”A finite element method for the
resolution of the Reduced Navier-Stokes/Prandtl equations” ZAMM -
Volume 89 Issue 1, Pages 54 - 68
[3] P. Bradshaw, T. Cebecci & J.H. Whitelaw (1981): ”Engineering calcu-
lation methods for turbulent flow” Academic Press.
[4] J.E. Carter (1979) ”A new boundary layer inviscid iteration technique
for separated flow”. AIAA-79-1450, Jul 1979.
[5] Castem 2000 http://www-cast3m.cea.fr/cast3m/index.jsp
[6] D. Catherall and K.W. Mangler (1966) ”The integration of the two-
dimensional laminar boundary-layer equations past the point of vanish-
ing skin friction. ”J. Fluid Mech. 26 (1966) pp.163182.
[7] Cebeci T. & Cousteix J. (2005): “Modeling and computation of bound-
ary layer flows“, Springer Verlag 502p.
[8] F. Chouly, A. Van Hirtum, X. Pelorson, Y. Payan, and P.-Y. Lagre
(2008): ”Numerical and experimental study of expiratory flow in the
case of major upper airway obstructions with fluid-structure interaction”
Journal of Fluids and Structures 24 (2008) pp. 250 -269
[9] Cousteix J., & Mauss J. (2007): ”Asymptotic analysis and boundary
layers”, Scientific Computations Springer, 432 p.
[10] A. Dechaume, J. Cousteix, J. Mauss, An interactive boundary layer
model compared to the triple deck theory, Eur. J. Mech. B Fluids 24
(4)(2005) 439 447
284 P.-Y. Lagrée
Herbert Steinrück*
*
Vienna University of Technology, Institute of Fluid Mechanics and Heat
Transfer, Vienna, Austria
Abstract Two different interaction mechanisms arise in the asymp-
totic analysis of mixed convection flow past a horizontal plate in the
limit of large Reynolds and Grashof numbers. A global interaction
mechanism between the wake flow and the potential flow and a local
triple deck interaction mechanism at the trailing edge. Both inter-
action mechanisms will be analyzed in the framework of matched
asymptotic expansions.
1 Introduction
Two different interaction mechanisms will be investigated by considering
the flow past a horizontal heated plate which is aligned under a small angle
of attack φ to the oncoming parallel flow with velocity Ũ∞ in a distin-
guished limit of large Reynolds Re and large Grashof number Gr. The
global structure of the flow field is shown in figure 1. According to the
method of matched asymptotic expansions in the limit of large Reynolds
numbers the flow field can be decomposed into the outer inviscid potential
flow, the boundary-layer flow along the plate, the wake behind the plate
and several sub-layers near the trailing edge of the plate.
However, the asymptotic approximation in the different layers cannot be
determined in a hierarchical order. The following two different interaction
mechanism can be identified:
Global interaction: The wake and the potential flow have to be de-
termined simultaneously since the temperature perturbation in the wake
causes a pressure difference across the wake which influences the global flow
field. On the other hand the potential flow determines the inclination of the
wake and thus the velocity and temperature distribution in the wake.
Local interaction: Near the trailing edge of the plate the boundary-
layer interacts with the potential flow according to the well known triple-
deck mechanism, see Stewartson (1969). Here the influence of buoyancy
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
288 H. Steinrück
g
y
U∞ = 1, θ∞ = 0 K
−1 Re−1/2 x
θp = 1 0
The starting point of the analysis are the Navier Stokes equations for
an incompressible fluid using Boussinesq’s approximation to take buoyancy
forces into account and the energy equation. We introduce a Cartesian co-
ordinate system such that its x-axis is horizontal and its origin is at the
trailing edge of the plate. In the following we will use dimensionless vari-
ables. All lengths (if not stated otherwise) are scaled with the plate length
L̃. Velocities are scaled with the velocity Ũ∞ of the unperturbed parallel
flow. The dimensionless velocity components u, v in x and y direction and
the dimensionless temperature perturbation θ = (T̃ − T̃∞ )/(T̃p − T̃∞ ) satisfy
the equations of motion, the energy equation and the continuity condition
1
uux + vuy = −px + (uxx + uyy ), (1a)
Re
1 Gr
uvx + vvy = −py + (vxx + vyy ) + θ, (1b)
Re Re2
1
uθx + vθy = (θxx + θyy ), (1c)
Re P r
ux + uy = 0, (1d)
subject to the asymptotic boundary conditions
u = 1, v = φ, θ = 0 as x2 + y 2 → ∞ (2)
g̃β̃ΔT̃ L̃
κ2 = K = Gr2 Re−9/2 . (4)
ũ2∞
Thus the velocity profile and the temperature profile in the wake depend on
the potential flow, namely on the inclination of the streamline emanating
from the trailing edge. As a consequence the wake flow problem and the
potential flow problem form an interaction problem and thus both problems
have to be solved simultaneously. This problem has been first formulated
by Savić and Steinrück (2005) and solved numerically. However, solutions
of the potential flow problem exist only if the pressure perturbation across
the wake decay downstream. This is the case if the oncoming flow has a
positive angle of attack φ which is of the order of the
√ induced inclination of
the wake. Thus an inclination parameter λ = φK Re is introduced. The
global interaction problem is formulated in section 2 and and numerical
290 H. Steinrück
The first two terms on the right side of equation (5) describe the potential
flow past a horizontal plate of a free stream with an angle φ to the horizontal
axis. The third term on the right side of (5) takes the buoyancy effects into
Interaction Mechanisms in Mixed Convection Flow… 291
account. Along the plate the vertical velocity component v1 has to vanish.
From the view point of the potential flow the scaled pressure has a jump
discontinuity of size γw across the wake. Using the linearized Bernoulli
equation we have
where γw is the dimensionless pressure jump across the wake, see (11).
If γw (x) is given, following Savić and Steinrück (2005), we obtain for the
dimensionless inclination of the wake
x
yw (x) = φ + Kv1 (x, 0), (7)
x+1
with
∞
1 x ξ + 1 γw (ξ) dξ
v1 (x, 0) = . (8)
2π 0 ξ x+1 x−ξ
Ux + VY = 0, (9a)
Thus in the far field the velocity and temperature profiles of the wake
flow tend to the velocity and temperature profiles of a two-dimensional
Interaction Mechanisms in Mixed Convection Flow… 293
laminar plume. Since the flow and temperature profile of the wake flow
is symmetric with respect to the centerline, it is sufficient to integrate the
enthalpy flux only over one half of the wake.
For the numerical solution of the coupled wake (9), (10), (11), and po-
tential flow equations (8) and the interaction or inclination condition (12)
an iterative method is proposed.
(0) x
i) First a suitable wake centerline Yw = λ x+1 is chosen.
ii) The wake equations are integrated for a velocity U (i) and temperature
(i−1)
field Θ(i) by a marching technique for a prescribed inclination Yw
of the wake.
(i) ∞
iii) Then the pressure jump γw = 2 0 θ(i) dy across the wake is deter-
mined
(i)
iv) Evaluating (8) a new centerline Yw of the wake is determined and
steps ii) -iv) are repeated until convergence is obtained.
We note that for κ = 0 no iterations are necessary. In the following we
keep the inclination parameter λ = 1 and the Prandtl number P r = 0.71
fixed. The interaction parameter κ will be increased starting from zero.
In figure 2 the velocity at the centerline of the wake and the pressure
jump across the wake are shown. For κ = 0 the shape of the wake is given
by the well known 2-d potential flow solution of the flow past an inclined
plate Schneider (1978). The centerline velocity increases from u = 0 at the
trailing edge due to viscosity. Then buoyancy leads to further acceleration
and a velocity overshoot forms. Accordingly the vortex distribution γw (x)
(or the pressure jump across the wake) decreases.
Evaluating the integral (8) shows that the induced vertical velocity com-
ponent v1 is negative. Thus for κ sufficiently large the wake turns down-
wards about a plate length behind the trailing edge. After attaining a min-
imum the wake turns upwards again. Accordingly the graph of centerline
velocity first becomes flat. Increasing κ further a minimum forms. When
κ attains a critical value κ = κc this minimum becomes zero. Since this
solution is singular at the zero of the centerline velocity a further increase
of κ is not possible. The physical mechanism which causes the singularity is
the following: In the parts of the wake with downward inclination the wake
flow is decelerated. The deceleration of the wake causes the wake to broaden
there. The increase of the wake thickness causes finally an increase of the
hydro-static pressure jump across the wake. In the limiting case κ = κc ,
the wake thickness becomes infinite in wake coordinates and thus γw also
tends to infinity.
In order to compute solutions with κ close to the critical value κc a
different strategy has to be employed. First a value Umin for the minimum
294 H. Steinrück
with Ŷw (s) = Yw (x) − Yw (x0 ). The local behavior of the centerline Ŷw is
not known a priorily and thus a corresponding term in the expansion of W
is added. Inserting into the differential equation (18) we obtain
W1 = 2Θ̃0 , W2 = W0 W0 , Θ̃1 = ( W0 Θ̃0 ) . (21)
Interaction Mechanisms in Mixed Convection Flow… 295
κ=0
U (x, 0)
κ = 0.5
κ = 0.9141
κ = 0.9173
κ = 0.9174
Umin = 0.13
Umin = 0
x
a) Centerline velocity in the wake
Umin = 0
κ = 0.9141
κ = 0.5
κ=0
x
b) Pressure jump across the wake
Using the local asymptotic expansion we can determine the local be-
havior of γw . We choose some value ψ ∗ > 0 and approximate W ∼
ε + W0 ψ/2 + Ŷw (s)W1 (0) and Θ̃ ∼ Θ̃0 (0) for ψ < ψ ∗ and obtain
∞ ψ∗
Θ̃ Θ̃0 (0)
γw (s) = 2 √ dψ ∼ 2 dψ
0 W 0 ε + W0 ψ 2 /2 + Ŷw (s)W1 (0) (22)
√
2Θ̃0 (0)
∼ − ln |ε + Ŷw (s)W1 |.
W0 (0)
It can be shown that the singular part of γ(s) is independent of the choice
of ψ∗ .
Considering that u1 − iv1 is a potential flow, using the complex valued
function theory and u1 = −γw /2 we conclude
Θ̃0 (0)
u1 − i v1 = ln F (z, ε), (23)
2W0 (0)
Θ̃0 (0)
|F (s)| = |ε + Ŷw (s) W1 |, − arg F (s) = Ŷw for s real, (24)
2W0 (0)
and Ŷw (0) = 0, Ŷw (0) = 0. This constitutes a problem for finding F (z, ε)
and Ŷw (s, ε) simultaneously.
We can express the solution F (z, ε) = εF̃ (z/ε), Ŷw (s, ε) = εỸw (s/ε) of
(24) for arbitrary values of ε by the solution F̃ and Ỹw of (24) for ε = 1. In
the limiting case ε = 0 we can guess the solution F (z, 0) = −iπ √Θ̃0 (0)
z.
2 2W0 (0)
Thus in that case the centerline of the wake has a corner of size
Θ̃0 π
[Yw ] = . (25)
2W0 (0)
∂ Ū ∂ Ū κ2 ∂ΔU ∂ΔU
Ū + V̄ + ΔU + ΔV =
∂x ∂Y Re1/2 ∂x ∂Y
(28a)
∂ P̄ ∂ 2 Ū 1 ∂ 2 Ū
− + + ,
∂x ∂Y 2 Re ∂x2
1 ∂ V̄ ∂ V̄ κ2 ∂ΔV ∂ΔV
Ū + V̄ + ΔU + ΔV =
Re ∂x ∂Y Re3/2 ∂x ∂Y
(28b)
∂ P̄ κ2 1 ∂ 2 V̄ 1 ∂ 2 V̄
− + Δθ + + .
∂Y Re1/2 Re ∂Y 2
Re2 ∂x2
Thus in the equations for the symmetric part the reduced buoyancy
parameter κ appears in the terms of order Re−1/2 . However, these terms
do not influence the equations for leading order terms of the triple deck
analysis. For the anti-symmetric parts we obtain
∂ΔU ∂ Ū ∂ΔU ∂ Ū
Ū + ΔU + V̄ + ΔV =
∂x ∂x ∂Y ∂Y
(29a)
∂ΔP ∂ 2 ΔU 1 ∂ 2 ΔU
− + + ,
∂x ∂Y 2 Re ∂x2
298 H. Steinrück
1 ∂ΔV ∂ V̄ ∂ΔV ∂ V̄
Ū + ΔU + V̄ + ΔV =
Re ∂x ∂x ∂Y ∂Y
(29b)
∂ΔP 1 ∂ 2 V̄ 1 ∂ 2 ΔV
− + θ̄ + + .
∂Y Re ∂Y 2 Re2 ∂x2
If the symmetric parts of the flow and pressure field are known the
equations for the anti-symmetric part are linear and independent of κ.
In the limit of large Reynolds numbers Re the flow structure near the
trailing edge can be described by a triple deck problem, cf. Stewartson
(1969); Messiter (1970). The scaling of the different layers is sketched in
Figure 13 in Ruban (2010). We define the independent variables in the
different layers (lower, main and upper deck) as
with the constant A0 given in table 2. In analogy to the velocity profile of the
symmetric part in the main deck the temperature profile of the symmetric
Interaction Mechanisms in Mixed Convection Flow… 299
part is given as
θ ∼ ΘB (Y ) + Re−1/8 Ā(x∗ )ΘB (Y ). (33)
In the following we will discuss the interaction problem for the anti-
symmetric part of the solution.
The expansions for the velocity components Δu and Δv follow the same
lines as Stewartson (1969) and the solution of the equations of the leading
order terms can be expressed in terms of an yet undetermined function
ΔA of x∗ , which can be interpreted as the scaled difference of the negative
displacement thicknesses on the upper and lower side of the plate. The
leading order terms of the anti-symmetric part of the velocity components
are given as
c1
Δu = ln Re U (Y ) + ΔA(x∗ )UB (Y ) + C1 (Y ) + ..., (37a)
UB (0) B
Δu − iΔv = Δu0 (0, 0) + Re−1/8 (Δu∗,1 (x∗ , y∗ ) − iΔv∗,1 (x∗ , y∗ )) + ..., (38)
where Δu0 (0, 0) = κ1 u1 (0, 0+), see (5), and Δu∗,1 (x∗ , 0) = −Δp∗,1 (x∗ , 0),
Δv∗,1 (x∗ , 0) = −ΔA (x∗ ) holds, the pressure Δp∗,1 (x∗ , 0) and the negative
displacement thickness ΔA (x∗ ) can be interpreted as the real and imaginary
part of a complex analytical function ΔΦ1 evaluated on the real axis. We
have
Considering ΔP1 (x∗ , 0) = 0 for x∗ > 0 and using the asymptotic behav-
ior of Ā for x∗ → ∞ we conclude that
√ ΔA → 0 for
holds. The constants a and b are determined by using that
x(3) → −∞. They turn out to be a = A0 and b = − 3A0 . Thus the
asymptotic behavior of ΔP1 and ΔA is given by
√
ΔA (x∗ ) ∼ − 3A0 |x∗ |1/3 , for x∗ → ∞. (42)
0
1 ΔP1 (ξ, 0) + 2A0 |ξ|1/3
− dξ (43)
π −∞ x∗ − ξ
∞
1 Ā(ξ) − A0 h(ξ)|ξ|1/3
+ dξ,
π −∞ x∗ − ξ
where h(x) denotes the Heaviside function with h(x) = 1 for x > 0 and
h(x) = 0 for x < 0.
We have written the interaction law in a form such that the singular parts
are separated and the integrand in the Hilbert integral decays sufficiently
fast to zero for x∗ → ±∞.
The asymptotic behavior of the vertical velocity component v for x∗ →
∞ and Y → ±∞ is given by
v = ±v̄ + κRe−1/4 Δv ∼
∼ − κRe−3/8 ΔA UB (Y ) ± Re−2/8 Ā UB (Y ) = (44)
√ 1/3 1 −2/3
= κRe−3/8 3A0 (x∗ ) UB (Y ) ± Re−2/8 A0 x∗ UB (Y ).
3
Rewriting (44) in the outer variables we have
√ 1
v(x, y) ∼ κRe−2/8 3A0 x1/3 ± Re−4/8 A0 x−2/3 . (45)
3
Applying the matching principle we conclude for the asymptotic behaviour
for of v1 given in (8)
√
v1 (x, 0) ∼ 3A0 x1/3 for x → 0 + . (46)
We remark that changes of the temperature profile in the lower deck are
too small to influence the leading order terms of the hydrostatic pressure
distribution and thus a discussion of the energy equation is not necessary.
Integrating (48) we obtain the pressure difference in the lower deck
which matches with the pressure difference in the main deck (36). Thus
ΔP1 (x∗ , 0) = ΔP∗,1 (x∗ , 0) and in the interaction law (43) ΔP1 can be re-
placed by ΔP∗,1 .
It remains to specify the asymptotic behavior of the velocity profile for
x∗ → −∞ and Y∗ → ∞.
Considering the asymptotic behavior of the pressure ΔP∗,1 and of Ū∗,1 ∼
UB (0)Y∗ for x∗ → −∞ we conclude that the asymptotic behavior of the flow
field ΔU∗,0 , ΔV∗,0 in the lower deck is self-similar. Using a scaled stream
function E defined by
Y∗
ΔU∗,0 ∼ E (η), with η= , (51)
|x∗ |1/3
a20 ΔA
a20 ΔA
ï
a0 ΔP∗,1
ï
3/4
3/4
a0 ΔP∗,1
ï
ï
ï ï
5/4
x∗ a 0
1/3
a0 ΔP∗,1
x ∗|
a20 ΔA 2 A 0|
− ï
3/4
a20 ΔA
ï
(− c1
3 ln |
ï 3/4 x
a0 ΔP∗,1 ∗| + ï
c1 +
ï c2 ) 2
a0
ï
ï ï ï ï ï ï ï ï
5/4
x∗ a0
X = 0+
X = 10−5 X = 10−4X = 10−3
X = 0.163 X = 0.057
X = 0−
X = −0.163
a0 ΔU∗
X = −3.24
ï
ï
X = −∞
ï
ï
η̃ = Y∗ / x2∗ + 1
−5/4
Figure 4. Velocity profiles ΔU∗,1 at different locations x∗ = a0 X
3.4 The local behavior of the lower deck velocity field near the
trailing-edge
It turns out that the interaction pressure ΔP∗,1 has a jump disconti-
nuity at the trailing edge x∗ = 0. As a consequence the derivative of the
displacement thickness ΔA has a logarithmic singularity at x∗ = 0. To dis-
cuss the behavior of the velocity profile ΔU∗ we integrate the momentum
equation (47) across the jump discontinuity at x∗ . We use the fact that the
symmetric part of the flow field Ū∗,0 , V̄∗,0 is continuous. Let
denote the jump in the Δu-component, the jump in the interaction pressure
ΔP∗,1 and the integral of the ΔV∗ component across the jump discontinuity,
306 H. Steinrück
where B is a constant and Ū∗s (Y∗ ) = Ū∗,1 (0, Y∗ ) . The jump in ΔU∗,0 is
therefore given by
[ΔP∗ ]
Δu∗,1 − iΔv∗,1 ∼ Ā(0) − i ln z∗ =
π
(66)
[ΔP∗ ] y∗
Ā(0) + arctan 2
− i ln x∗ + y∗ .2
π x∗
Thus the pressure and the derivative of the displacement thickness behave
locally like
[ΔP∗ ] y∗ [Δp]
Δp∗,1 ∼ −Ā(0) − arctan , ΔA (x∗ ) ∼ − ln |x∗ |. (67)
π x∗ π
In order to resolve the discontinuity in the main deck we introduce the
sub-layer with
X = Re1/2 x, Y = Re1/2 y. (68)
The velocity profile has to match with the (x∗ , Y )-region (37a),(37b). Thus
we use the following expansion of the anti-symmetric part
[ΔP∗ ]
Δv ∼ −Re−1/8 ln Re UB (Y ) + Re−1/8 Ṽ1 (X, Y ) + · · · , (69b)
8π
Δp ∼ ΔP0 (0, Y ) + Re−1/8 ΔP̃1 (X, Y ) + · · · . (69c)
The term of order Re−1/8 ln Re in the vertical velocity component Δv arises
from matching with the main deck solution and the logarithmic behavior of
ΔA as x∗ → 0. As a consequence a term of the same magnitude must be
present in the expansion of the horizontal velocity component Δu. However,
in the expansion of the pressure Δp these “logarithmic” terms are missing.
The constant c1 has been introduced in (53).
We obtain the following equations for the leading order terms
∂ΔŨ1 ∂ΔP1
UB + ΔṼ1 UB = − , (70a)
∂X ∂X
∂ΔṼ1 ∂ΔP̃1
UB =− + Θ̄1 , (70b)
∂X ∂Y
308 H. Steinrück
ΔPh
ï Y
ï
X
Figure 5. Local behavior of the interaction pressure ΔP1 near the trailing-
edge. The solution ΔPh of the homogenous problem , cf. (73), is shown
∂ΔŨ1 ∂ΔṼ1
+ = 0. (70c)
∂X ∂Y
The flow in the (X, Y )-sub-layer is inviscid. But in contrast to the main
deck the y-momentum equation is not degenerate. Eliminating ΔŨ1 and
ΔṼ1 an elliptic equation for ΔP̃1 can be derived,
∂ 2 ΔP̃1 ∂ 2 ΔP̃1 ∂ Θ̄1 ∂ΔP̃1
UB + − + 2UB Θ̄1 − = 0. (71)
∂X 2 ∂Y 2 ∂Y ∂Y
|ΔP∗ |
ΔP̃1 = − ΔPh (X, Y ) + Ā(0) (ΘB (Y ) − 1) , (73)
π
with ΔPh ∼ arctan Y /X for X 2 + Y 2 → ∞ and ΔP̃h (X, 0) = π for X < 0
and ΔPh (X, 0) = 0 for X > 0.
The local behavior near the singularity can be discussed by transforming
the equation (71) to polar coordinates R, ϕ. Expanding ΔPh ∼ ΔPh,0 (ϕ) +
O(R) for R 1 we obtain
sin ϕ ΔPh,0 − 2 cos ϕΔPh,0 = 0, ΔPh,0 (0) = 0, ΔPh,0 (π) = π, (74)
However in both approaches the wake and the potential flow have to be
considered simultaneously. It turns out that if the interaction parameter
approaches a critical value a singularity in the wake forms. The nature of
this singularity is essentially an inviscid phenomenon. The author believes
that also in the case when the plate is placed in a horizontal channel such
a singularity can occur, but it has not been investigated up to now.
To resolve the singularity in the flow field at the trailing edge a triple deck
problem has been formulated. This problem constitutes a local interaction
problem between the a sub-layer of the boundary-layer (lower deck) and the
potential flow (upper deck). Here the influence of the hydrostatic pressure
on the interaction mechanism has been studied. Surprisingly the interaction
pressure at the trailing edge turns out to have a jump discontinuity which
can be resolved be introducing additional sub-layers.
Bibliography
R. Chow and E. Melnik. Numerical solutions of triple-deck equations for
laminar trailing edge stall. Technical Report RE-526J, Grumman Re-
search Dept, 1976.
P.-Y. Lagree. Thermal mixed convection induced locally by a step change in
surface temperature in a Poiseuille flow in the framework of triple deck
theory. Int. J. Heat and Mass Transfer, 42:2509–2524, 1999.
A. F. Messiter. Boundary layer flow near the trailing edge of flat plate.
SIAM J. Appl. Math, 18:241–257, 1970.
A. Ruban. Asymptotic theory of separated flows. In H. Steinrück, editor,
Asymptotic Methods in Fluid Mechanics-Recent Advances, CISM Lecture
Notes 523, pages 311–408. Springer, 2010.
Lj. Savić and H. Steinrück. Mixed convection flow past a horizontal plate.
Theor. Appl. Mech., 32:1–19, 2005.
Lj. Savić and H. Steinrück. The trailing-edge problem for mixed-convection
flow past a horizontal plate. J. Fluid Mech., 588:309–330, 2007.
W. Schneider. Lift, thrust and heat transfer due to mixed convection flow
past a horizontal plate. J. Fluid Mech., 529:51–69, 2005.
W. Schneider. Mathematische Methoden der Strömungslehre. Vieweg, 1978.
K. Stewartson. On the flow near the trailing edge of a flat plate. Mathe-
matica, 16:106–121, 1969.
V. V. Sychev, A. I. Ruban, Vic. V. Sychev, and G. L. Korolev. Asymptotic
theory of separated flows. Cambridge University Press, 1998.
Asymptotic Theory of Separated Flows
Anatoly I. Ruban
Department of Mathematics, Imperial College London
1 Introduction
Separation is a fluid dynamic phenomenon that influences the behaviour of
a wide variety of liquid and gas flows. The difference between an attached
flow and its separated counterpart is demonstrated in Figure 1 where the
theoretical streamline pattern, given by the classical solution of the inviscid
flow theory1
a2
ϕ + iψ = V∞ z + , (1)
z
is compared with a real flow visualisation for a circular cylinder in a water
tank (Taneda, 1956). In the theoretically predicted form of the flow the
fluid particles follow closely the cylinder contour from the front stagnation
point all the way to the rear stagnation point. Contrary to that in the
experimentally observed flow the fluid particles brake away from the cylinder
1
See, for example, Lamb (1932).
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
312 A. Ruban
surface at a separation point and form a pair of eddies in the wake behind
the cylinder.
The irony of the situation is that the theoretical flow in Figure 1(a)
has been constructed based on the Euler equations. These are intended
for describing the motion of fluids with extremely small viscosity. While
the Euler equations admit solutions in the form of attached flows, such
flows can not be observed in practice except for some special cases. In
fact, experiments clearly indicate that the attached form of fluid motion
past rigid bodies is characteristic for relatively small values of the Reynolds
number. In particular, the flow past a circular cylinder assumes an attached
form only if the Reynolds number Re = V∞ a/ν, with a being the cylinder
radius, is smaller than a critical value of about six. The actual flow shown in
Figure 1(b) corresponds to Re = 26, and we see that the separation eddies
are already well developed. Further increase of the Reynolds number results
in an extention of the eddies, and then the flow looses its symmetry and
becomes unstable, but it never returns to an attached form (see Figure 2).
Figure 2. Visualisation of the cylinder flow by Werlé & Gallon (1972); the
Reynolds number Re = 2000.
Since many ”common” gases and liquids, such as air and water, have
extremely small viscosity, their flows are characterised by very large values of
the Reynolds number. This explains why most liquid and gas flows observed
in nature and encountered in engineering applications involve separation.
The difference between a separated flow and its theoretical unseparated
counterpart (constructed on the basis of inviscid flow analysis) concerns
not only the form of trajectories of fluid particles, but also the magnitudes
of aerodynamic forces acting on the body. For example, for bluff bodies
in an incompressible flow, it is known from experimental observations that
the drag force is never zero; furthermore, it does not approach zero as the
Asymptotic Theory of Separated Flows 313
Reynolds number becomes large. On the other hand, one of the most famous
results of the inviscid flow theory is d’Alembert’s paradox which states that
a rigid body does not experience any drag in incompressible steady flow.
This contradiction is associated with the assumption of an attached form of
the flow.
Separation imposes a considerable limitation on the operating character-
istics of aircraft wings, helicopter and turbine blades leading to a significant
degradation of their performance. It is well known that the separation is
normally accompanied by a loss of the lift force, sharp increase of the drag,
increase of the heat transfer at the reattachment point, development of flow
oscillations, etc.
It is hardly surprising that the problem of flow separation has attracted
considerable interest amongst researchers. The first theoretical model of a
separated flow was due to Helmholtz (1868) and Kirchhoff (1869) in the
framework of the classical theory of inviscid fluid flows. This model was
originally applied to the flow past a flat plate perpendicular to the free
stream when the separation is known to take place at the plate edges. The
Kirchhoff model may be, of course, applied to other body shapes. In partic-
ular, Levi-Civita (1907) used this model for the flow past a circular cylinder
(see Figure 3). A major conclusion that may be drawn from this theory is
that Euler equations allow for a family of solutions where the position of
separation point S remains a free parameter.
S
Indeed with τw being positive upstream of this point, all the fluid particles
in the boundary layer move downstream along the wall (see Figure 4) and
the flow appears to be attached to the body surface. However, as soon as the
skin friction τw turns negative a layer of reversed flow (u < 0) emerges near
the wall, giving rise to a region of recirculation which, obviously, originates
from point S where condition (2) holds.
y
u
2
A review of these efforts may be found, for example, in Chapman et al. (1956).
Early theoretical models to explain the phenomenon are reviewed in a resent paper
by Lighthill (2000).
316 A. Ruban
incident shock
V∞
A
real behaviour of the flow. The experiments invariably showed (see, for
example, Figure 6) that, unless the incident shock was very weak, the flow
separated from the plate surface some distance upstream of the incident
shock (see Figure 7). It was also established that the boundary layer was
perturbed even upstream of the separation point S, and the distance over
which the pressure perturbations were able to propagate upstream of point
S through the boundary layer proved to be significantly larger then the
boundary-layer thickness. An increase of the pressure in the boundary layer
prior to separation and, even more so, the separation of the boundary layer
cause the streamlines at the bottom of the inviscid flow region to deviate
from the wall giving rise to a secondary shock as shown in Figure 7. Together
with the primary shock they form a characteristic shock structure called the
λ-structure.
primary shock
secondary shock
boundary layer with supersonic inviscid flow was termed by Chapman et al.
(1956) the free-interaction.
free-interaction
region
z }| {
2 Self-Induced Separation
2.1 Formulation of the Problem
The theory that will be discussed here applies to boundary-layer sep-
aration in a wide variety of supersonic flow. However, for the purpose of
describing the theory it is convenient to choose a particular flow layout,
for example, the flow past a flat plate surface; see Figure 9. We shall as-
sume that the plate is aligned with oncoming flow which is supersonic. We
shall further assume that the boundary layer on the plate surface separates
due to some downstream disturbance. The position of the separation point
S depends on the nature of this perturbations, and in this study will be
treated as known.
In what follows we shall assume that the gas considered may be treated
as perfect. Its motion will be assumed steady and two-dimensional, in which
Asymptotic Theory of Separated Flows 319
V∞
We denote the distance from the leading edge to the separation point S
by L; the velocity, density, viscosity and pressure in the unperturbed flow
upstream of the plate are denoted by V∞ , ρ∞ , μ∞ and p∞ respectively. To
study the flow we shall use Cartesian coordinate system x̂Oŷ with x̂ mea-
sured parallel to the plate from its leading edge, and ŷ in the perpendicular
direction. The velocity components in these coordinates are denoted by û
and v̂. Other quantities used in the Navier-Stokes equations are the gas
density ρ̂, pressure p̂, enthalpy ĥ and viscosity μ̂. As before, the “hat” de-
notes dimensional variables. The non-dimensional variables are introduced
as
û = V∞ u, v̂ = V∞ v, ρ̂ = ρ∞ ρ,
2 2
p̂ = p∞ + ρ∞ V∞ p, ĥ = V∞ h, μ̂ = μ∞ μ,
x̂ = Lx, ŷ = Ly.
320 A. Ruban
role of the background on which the interaction develops. The flow out-
side the boundary layer is almost uniform with small perturbations of order
O(Re−1/2 ) caused by the presence of the boundary layer.
Asymptotic analysis of the boundary layer is based on the limit proce-
dure
x = O(1), Y = Re1/2 y = O(1), Re → ∞,
and the solution of the Navier-Stokes equations may may be sought in the
form of the asymptotic expansions
u(x, y; Re) = U0 (x, Y ) + · · · , v(x, y; Re) = Re−1/2 V0 (x, Y ) + · · · ,
ρ(x, y; Re) = ρ0 (x, Y ) + · · · , p(x, y; Re) = Re−1/2 P1 (x, Y ) + · · · , (4)
h(x, y; Re) = h0 (x, Y ) + · · · , μ(x, y; Re) = μ0 (x, Y ) + · · · .
Substitution of (4) into the Navier-Stokes (3) equations leads to the classical
boundary-layer equations
∂U0 ∂U0 ∂ ∂U0
ρ0 U0 + ρ0 V0 = μ0 , (5a)
∂x ∂Y ∂Y ∂Y
2
∂h0 ∂h0 1 ∂ ∂h0 ∂U0
ρ0 U0 + ρ0 V0 = μ0 + μ0 , (5b)
∂x ∂Y P r ∂Y ∂Y ∂Y
∂(ρ0 U0 ) ∂(ρ0 V0 )
+ = 0, (5c)
∂x ∂Y
1 1
h0 = 2
. (5d)
(γ − 1)M∞ ρ0
They should be solved with the free-stream conditions at the leading edge
of the flat plate
1
U0 = 1, h0 = 2
at x = 0, Y ∈ [0, ∞) (6)
(γ − 1)M∞
∂h0
= 0 at Y = 0, x ∈ [0, 1]. (9b)
∂Y
Boundary-value problem (5)–(9) admits a self-similar solution for ther-
mally isolated wall (9b) as well as in the case when the wall temperature
is known to be constant, i.e. function F (x) in (9a) does not depend on x.
However, we do not need to restrict ourselves to these particular flow condi-
tions. If they are not satisfied then the boundary-value problem (5)–(9) may
be solved numerically. To proceed further we only need to know that the
solution remains smooth when the trailing edge of the plate is approached.
Therefore we shall assume that the sought functions U0 , h0 , ρ0 and μ0 may
be represented in the form of Taylor expansions
⎫
U0 (x, Y ) = U00 (Y ) + (−s)U01 (Y ) + · · · ,⎪
⎪
⎪
h (x, Y ) = h (Y ) + (−s)h (Y ) + · · · , ⎬
0 00 01
as s = x − 1 → 0− . (10)
ρ0 (x, Y ) = ρ00 (Y ) + (−s)ρ01 (Y ) + · · · , ⎪
⎪
⎪
⎭
μ0 (x, Y ) = μ00 (Y ) + (−s)μ01 (Y ) + · · ·
The leading order terms in (10) exhibit the following behaviour near the
plate surface
⎫
U00 (Y ) = λY + · · · ,⎪ ⎪
⎪
h00 (Y ) = hw + · · · , ⎬
as Y → 0, (11)
ρ00 (Y ) = ρw + · · · , ⎪
⎪
⎪
⎭
μ00 (Y ) = μw + · · ·
where λ, hw , ρw , μw are positive constants representing the dimensionless
skin friction, enthalpy, density and viscosity on the wall surface.
∂u ∂p
ρu ∼ . (12)
∂x ∂x
Asymptotic Theory of Separated Flows 323
Δp > 0
Taking into account that the perturbations are small, and therefore the
velocity u and density ρ may be represented by their initial profiles given
by the leading order terms in (10), we have
∂u ∂p
ρ00 U00 ∼ .
∂x ∂x
Approximating the derivatives by finite differences, we have
Δu Δp
ρ00 U00 ∼ ,
Δx Δx
and it follows that
Δp
Δu ∼ . (13)
ρ00 U00
Since everywhere in the boundary layer, except near the wall, both ρ00 and
U00 are order one quantities, we can finally write
Δu ∼ Δp. (14)
Applying the same arguments to the energy equation (3c), we find
Δh ∼ Δp,
and then it follows from the state equation (3e) that
Δρ ∼ Δp. (15)
Let us now we consider a small filament in the boundary layer confined
between two neighbouring streamlines with δ being the initial distance be-
tween them, as shown in Figure 11. Using the mass conservation law, we
can write
ρ00 U00 δi = (ρ00 + Δρ)(U00 + Δu)(δi + Δδi ).
324 A. Ruban
Since both ρ00 and U00 are order one quantities and their variations are
given by (14) and (15), we can conclude that the thickness of the filament
increases by the value
Δδi ∼ δi Δp.
The integral effect of the thickening of all the filaments in boundary layer
is
Δδ = Δδi ∼ δi Δp ∼ Δp δi ∼ Re−1/2 Δp. (16)
i i i
δi + Δδi
Re−1/2
δi
Δx
Figure 11. Thickening of a stream filament in the main part of the bound-
ary layer.
The above analysis, obviously, is invalid near the bottom of the boundary
layer. Indeed, according to (11) the initial velocity U00 tends to zero as
Y → 0, and equation (13) predicts unbounded growth of the perturbation
velocity. Of course, before it happens non-linear effects take over. In a thin
sublayer near the wall, where Δu ∼ U00 , equation (13) may be written as
Δu ∼ U00 ∼ Δp. (17)
Combining (17) with the formula for U00 in (11), we can deduce that
Y ∼ Δp,
To estimate the displacement effect of the sublayer we again use the mass
conservation law. Treating the sublayer as one filament (see Figure 12) and
Asymptotic Theory of Separated Flows 325
taking into account that along this filament Δu ∼ U00 , we have to conclude
that the variation of the filament thickness is of the same order as its initial
value given by (18), i.e.
Δδ ∼ Re−1/2 Δp. (19)
Re−1/2
y y + Δδ
Δx
Comparing (19) with (16), we see that for any Δp 1 the contribution
of the sublayer into the displacement effect of the boundary layer is signifi-
cantly larger than that of the main part of the boundary layer. Hence, the
slope angle θ of the streamlines at the outer edge of the boundary layer
should be estimated based on the displacement effect of the sublayer (19).
We have √
Δδ Re−1/2 Δp
θ∼ ∼ ,
Δx Δx
Using further the Ackeret formula, we find that the pressure perturbations
in the interaction region
√
Re−1/2 Δp
Δp ∼ θ ∼ . (20)
Δx
It remains to recall that the sublayer is adjacent to the wall and therefore
should be viscous,
∂u 1 ∂ ∂u
ρu ∼ μ . (21)
∂x Re ∂y ∂y
Indeed, if the flow were inviscid, then the Bernoulli equation could be used
along each streamline. Near the wall the velocity small, and therefore, we
can write this equation in the incompressible form
u2 p U2
+ = 00 . (22)
2 ρ 2
Here it is taken into account that p denotes the perturbation of the pressure
with respect to its value in the oncoming flow; everywhere upstream of the
interaction region p = 0.
326 A. Ruban
Writing equation (22) in the form
u2 U2 p
= 00 − , (23)
2 2 ρ
we can see that for any increase of pressure, no matter how small, we can
always find a streamline close enough to the wall for which the right hand
side of (23) appears to be negative. Since this is impossible, we have to
conclude that the flow in the sublayer should be viscous.
Taking into account that the density ρ and viscosity μ are order one
quantities everywhere in the boundary layer, and representing equation (21)
in the finite differences form, we have
Δu 1 Δu
U00 ∼ .
Δx Re (Δy)2
Since in the sublayer Δy ∼ y, this equation may be written as
U00 1 1
∼ . (24)
Δx Re y 2
Let us now summarise the results of the above analysis. Taking into
account that in the sublayer the velocity u is same order as initial velocity
U00 at the bottom of the boundary layer approaching the interaction region,
we will write equation (17) as
u ∼ Δp. (25)
Equation (18) for the thickness of the viscous sublayer is written as
y ∼ Re−1/2 Δp. (26)
From equation (20), representing the pressure induced in the interaction
region, it follows that
Re−1/2
Δp ∼ . (27)
Δx
In order to close this set of equations we need to add equation (24) which
we will now write as
u 1
∼ . (28)
Δx Re y 2
As a result we have four equations (25) – (28) for four unknowns, the velocity
u in the sublayer, characteristic thickness of the sublayer y, induced pressure
Δp and longitudinal extent Δx of the interaction region. These equations
may be treated as algebraic equations. They are easily solved to give
u ∼ Re−1/8 , y ∼ Re−5/8 , Δp ∼ Re−1/4 , Δx ∼ Re−3/8 . (29)
Asymptotic Theory of Separated Flows 327
3
Re −5/8
Re−1/2
1
Re−3/8
Since the flow in the viscous sublayer is slow, it should behave as in-
compressible. To confirm this proposition, we need to consider the energy
equation (32c). It is a parabolic equation, and requires an initial condition
(x∗ → −∞), condition on the wall (Y∗ = 0) and condition at the outer edge
of the sublayer (Y∗ → ∞). We start with the initial condition. It may be
formulated by matching with the solution in the boundary layer upstream
of the interaction region. This solution may be termed the outer solution
for our purposes. According to (4), the outer asymptotic expansion for the
enthalpy has the form
x = 1 + Re−3/8 x∗ .
Since x − 1 is small, we can use the Taylor expansion (10) for the enthalpy.
Being substituted into (33), it gives
y = Re−1/2 Y,
330 A. Ruban
while in region 1
y = Re−5/8 Y∗ .
Comparing these formulae, we see that
Y = Re−1/8 Y∗ ,
which means that Y is small, and we can use the asymptotic formula (11)
for h00 in (33). We have
h(x, y; Re) = hw + · · · .
h∗ → hw as x∗ → −∞. (35)
h∗ = hw at Y∗ = 0, (36a)
∂h∗
=0 at Y∗ = 0. (36b)
∂Y∗
Concerning condition (36) it should be noted that even when the wall tem-
perature is not constant, its variation over a short distance occupied by the
interaction region is small in normal circumstances, and therefore, should
be neglected in the leading order approximation.
It remains to consider the outer edge of the viscous sublayer. At large
values of Y∗ the energy equation (32c) turns into the inviscid the form
∂h∗ ∂h∗
U∗ +V∗ = 0, (37)
∂x∗ ∂Y∗
which shows that the enthalpy h∗ does not chance along streamlines. All
the streamlines, except in the separation region, originate from an upstream
location, where condition (35) holds. Integrating (37) with (35), we arrive
at a conclusion that the sought boundary condition may be written as
h∗ → hw as Y∗ → ∞. (38)
Asymptotic Theory of Separated Flows 331
Therefore, the convective terms on the left hand side of equation (41a) and
the viscous term on the right hand side are written as
∂U ∗ dA0 2α−2
ρw U ∗ = ρw αA0 Y + ··· ,
∂x∗ dx∗ ∗
∂U ∗ dA0 2α−2
ρw V ∗ = −ρw α(α − 1)A0 Y + ··· ,
∂Y∗ dx∗ ∗
∂2U ∗
μw = α(α − 1)(α − 2)A0 Y∗α−3 + · · · .
∂Y∗2
We see that if we assume, subject to subsequent confirmation, that α > 1,
then the convective terms will dominate not only over the viscous term, but
also over the pressure gradient, which remains finite as Y∗ → ∞. We have
dA0
O(Y∗2α−2 ) : A0 = 0.
dx∗
The initial condition for this equation may be obtained by substituting
the first of formulae (48) into (45). We find
λ/α if α = 2,
A0 (−∞) =
0 if α
= 2.
Asymptotic Theory of Separated Flows 333
Hence, a non-trivial solution exists only if α = 2, in which case A0 = 12 λ,
and (47) turns into
1
Ψ∗ (x∗ , Y∗ ) = λY 2 + · · · as Y∗ → ∞.
2 ∗
Now we shall try to find the next order term in this expansion:
1
Ψ∗ (x∗ , Y∗ ) = λY 2 + A1 (x∗ )Y∗α + · · · as Y∗ → ∞. (49)
2 ∗
In order to ensure that the second term in (49) is small as compared with
the first one, we have to assume that α < 2. Substitution of (49) into (46)
yields
Therefore, the convective terms on the left hand side of equation (41a) and
the viscous term on the right hand side are written as
∂U ∗ dA1 α
ρw U ∗ = ρw λα Y + ··· ,
∂x∗ dx∗ ∗
∂U ∗ dA1 α
ρw V ∗ = −ρw λ Y + ··· ,
∂Y∗ dx∗ ∗
∂ 2U ∗
μw = α(α − 1)(α − 2)A1 Y∗α−3 + · · · .
∂Y∗2
Hence, the convective terms remain dominant provided that α > 0, in which
case the momentum equation (41a) reduces at
dA1
O(Y∗α ) : (α − 1) = 0.
dx∗
Since the initial condition (45) does not contain any terms except the
one which matches with the leading order term in (49), we have to conclude
that
A1 (−∞) = 0.
We see that a non-trivial solution for A1 exists only if α = 1. Function
A1 (x∗ ) remains arbitrary in the framework of the asymptotic analysis of
equations (41). We, of course, expect that this function will be found as a
part of the solution of the problem as a whole.
Redenoting A1 (x∗ ) as A(x∗ ) renders (49) in the form
1
Ψ∗ = λY 2 + A(x∗ )Y∗ + · · · as Y∗ → ∞. (50)
2 ∗
334 A. Ruban
Substitution of (51) into (46) shows that at the outer edge of the viscous
sublayer
dA
U ∗ (x∗ , Y∗ ) = λY∗ + A(x∗ ) + · · · , V ∗ (x∗ , Y∗ ) = − Y∗ + · · · , (52)
dx∗
and we can see that the streamline slope angle ϑ, which is zero on the wall,
reaches at the outer edge of the viscous sublayer the following value
∗
1 dA
v −1/4 V −1/4
ϑ = arctan = Re + · · · = Re − + · · · . (53)
u U ∗ Y∗ →∞ λ dx∗
Main part of the boundary layer Region 2, the middle tier of the
triple-deck structure (see Figure 13), is a continuation of the conventional
boundary layer developing on the plate surface before the interaction. Its
thickness is estimated as y = O(Re−1/2 ). The longitudinal extent of region
2 coincides with that of the entire interaction region and is estimated as
|x − 1| = O(Re−3/8 ). Consequently, the asymptotic analysis of the Navier-
Stokes equations (3) in the main part of the boundary layer should be based
on the limit procedure
This suggests that the solution in region 2 should be written in the form of
asymptotic expansions
The leading order term U00 (Y ) in the expansion for u(x, y; Re) coincides
with the velocity profile (10) in the boundary layer immediately before the
interaction region. According to (11)
U00 = λY + · · · as Y → 0. (58)
It further follows from (57) that the perturbation terms U1 (x∗ , Y ) and
V1 (x∗ , Y ) in (57) satisfy the following boundary conditions at the bottom
of region 2,
⎫
U1 = A(x∗ ) + · · · , ⎬
dA as Y → 0. (59)
V1 = − Y + · · · .⎭
dx∗
Finally, taking into account that in view of (32b) the pressure perturbations
in region 2 should be same order as in region 1, we write
Substitution of (57), (60) and (61) into the Navier-Stokes equations (3)
336 A. Ruban
results in
∂ U1
U00 (Y ) + V1 U00 (Y ) = 0, (62a)
∂x∗
∂ P1
= 0, (62b)
∂Y
∂ h1
U00 (Y ) + V1 h00 (Y ) = 0, (62c)
∂x∗
∂ U1 ∂ ρ1 ∂ V1
ρ00 (Y ) + U00 (Y ) + ρ00 (Y ) + V1 ρ00 (Y ) = 0, (62d)
∂x∗ ∂x∗ ∂Y
1 1 1 ρ1
h00 = , h1 = − . (62e)
2
(γ − 1)M∞ ρ00 (γ − 1)M∞ ρ200
2
These equations are easily solved using the following elimination process.
We first substitute (62e) into the energy equation (62c). This leads to
∂ ρ1
U00 (Y ) + V1 ρ00 (Y ) = 0,
∂x∗
showing that the continuity equation (62d) may be written as
∂ U1 ∂ V1
+ = 0. (63)
∂x∗ ∂Y
Now, using (63), we can eliminate ∂ U1 /∂x∗ from the longitudinal mo-
mentum equation (62a). This results in
∂ V1
U00 (Y ) − V1 U00 (Y ) = 0. (64)
∂Y
2
Dividing both terms in (64) by U00 , we have
1 ∂ V1 U
− V1 00
2 = 0,
U00 (Y ) ∂Y U00
or equivalently
∂ V1
= 0.
∂Y U00
We see that the ratio V1 /U00 is a function of x∗ only, say G(x∗ ), i.e.
V1
= G(x∗ ).
U00
Asymptotic Theory of Separated Flows 337
This function may be found by making use of the behaviour of V1 and U00
at the bottom of region 2, as given by (58), (59). We find that
1 dA
G(x∗ ) = − .
λ dx∗
Hence, everywhere across region 2
V1 1 dA
=− .
U00 λ dx∗
Let us now return to the asymptotic expansions (55) of the velocity
components, and calculate the streamline slope angle in region 2 :
v V1 1 dA
ϑ = arctan = Re−1/4 + · · · = Re−1/4 − + ··· . (65)
u U00 λ dx∗
We see that ϑ does not depend on Y , i.e. stays unchanged across the
main part of the boundary layer. This confirms an important result of
the inspection analysis that the displacement effect of the main part of the
boundary layer is relatively small, and may be neglected to the leading
order. The streamline slope angle (53), produced by the viscous sublayer,
is simply transported by the main part of the boundary layer towards the
bottom of the the upper tier of the triple-deck structure, shown as region 3
in Figure 13
2 ϑ
p̂ = p∞ + ρ∞ V∞ (66)
2 −1
M∞
known as the Ackeret formula (see, for example, Siebert (1948)). Since
the pressure does not change across regions 2 and 1, we can apply formula
(66) directly to the pressure in the viscous sublayer. Converting (66) into
non-dimensional form and using (65), we have
1 dA
P∗ = − . (67)
2
λ M∞ − 1 dx∗
This equation establishes a relationship between the the displacement effect
of the boundary layer and the pressure induced in the inviscid from. This
is why it is referred to as the interaction law.
338 A. Ruban
∂U ∗ ∂U ∗ dP ∗ ∂ 2U ∗
ρw U ∗ + ρw V ∗ =− + μw , (68)
∂x∗ ∂Y∗ dx∗ ∂Y∗2
∂U ∗ ∂V ∗
+ = 0, (69)
∂x∗ ∂Y∗
U∗ = V ∗ = 0 at Y∗ = 0, (70)
1 dA
P∗ = − . (72)
λ M∞2 − 1 dx∗
Affine transformations
1
1.5
1 0.5
P̄
τ
0.5
0
0 −0.2
−12 −7 −2 3 8 13 −12 −7 −2 3 8 13
X̄ X̄
(a) Pressure distribution (b) Skin friction
Figure 14. Results of the numerical solution of the interaction problem
(74).
We see that far upstream, where the interaction region matches with
the unperturbed boundary layer, the skin friction τ = 1 and the pressure
perturbation function P̄ = 0. As the pressure starts to rise in the inter-
action region, it cause the flow in the viscous sublayer to decelerate. This
340 A. Ruban
is revealed by the observed decrease of the skin friction. The skin friction
decreases crossing zero at the separation point. It should be noted that the
interaction problem (74) is invariant with respect to arbitrary shift in the
direction parallel to the body surface,
X̄ −→ X̄ + C,
P̄separation = 1.046,
Exercise 1. Using (75) and the Ackeret formula (66), find the shape of
the boundary between the separation region and the main flow above it.
Asymptotic Theory of Separated Flows 341
M
S
(a)
S
(b)
S
(c)
Figure 15. The flow past a thin aerofoil: (a) with a short separation bubble;
(b) with an extended separation region that forms after the bubble bursting;
(c) with a long separation bubble.
Experiments show that for thick aerofoils (with thickness to cord ratio
larger than 15%) the boundary layer first separates near the trailing edge.
For thin aerofoils (with thickness to cord ratio larger than 12%) this is the
leading edge separation that causes the stall of the aerodynamic character-
istics. For the stall to take place the angle of attack α should exceed a
critical value αc . When α is small, the flow over the aerofoil remains fully
attached, and the pressure has its maximum at the front stagnation point
S; see Fig. 15(a). As one moves from this point around the aerofoil nose,
the pressure first drops dramatically reaching a minimum at some point M
on the upper side of the aerofoil, and then starts to recover, so that down-
stream of point M the boundary layer finds itself under an adverse pressure
gradient. Its magnitude increases with increase of the angle of attack, and
342 A. Ruban
With this choice the radius of the curvature of the leading edge appears to
be r = c ε2 .
Asymptotic Theory of Separated Flows 343
y
1 x
V∞
O
α
In what follows the flow past the aerofoil will be investigated using the
asymptotic analysis of the Navier-Stokes equations with a limit procedure,
where the Reynolds number, based on the radius of the leading edge of the
profile, tends to infinity, and the thickness of the aerofoil tends to zero, i.e.
V∞ r
Re = → ∞, ε → 0.
ν
α = εα∗ ,
with α∗ = O(1), then the flow in this region can be described by the classical
thin aerofoil theory. The asymptotic expansions for the velocity components
(u, c) and pressure p are written as.
u = 1 + ε u1 (x, y) + · · · , v = ε v1 (x, y) + · · · ,
p = ε p1 (x, y) + · · · .
(77)
Substituting (77) into the Navier-Stokes equations, and assuming that ε → 0
and Re → ∞, one can find that the pressure p1 satisfies the Laplace equation
∂ 2 p1 ∂ 2 p1
2
+ = 0.
∂x ∂y 2
written as
1
1 Y+ (ζ) − Y− (ζ)
p1 (x, 0±) = dζ±
2π ζ −x
0
(78)
1
1−x 1 ζ Y+ (ζ) + Y− (ζ)
± − α∗ + dζ .
x 2π 1−ζ ζ −x
0
Here “plus” corresponds to the upper surface of the aerofoil and “minus”
to the lower.
The asymptotic expansion for the pressure in (77) shows that the pres-
sure perturbations in the main inviscid-flow region are too small to the
cause the boundary-layer separation. However, the solution in this region
develops a singularity near the leading edge. Indeed, it follows from (78)
that
k
p1 (x, 0±) = ± √ ,
2x
where
1
√ 1 G(x ) 1 dY+ dY−
k = 2 α∗ + dx , G=− + . (79)
π x (1 − x ) 2 dx dx
0
x = ε2 X , y = ε2 Y ,
with X and Y being order one quantities. In this new region (in what
follows we shall call it region
√ 1) the aerofoil contour is represented by the
infinite parabola Y = ± 2X ; see Figure 17. The tangential component of
the velocity vector on the surface of the parabola is given by
Y+k
Ue = √ . (80)
Y 2 + 1
Here Y is the distance from the point on the surface of the parabola where
Ue is calculated to the X -axis; parameter k measures the degree of non-
symmetry of the flow. It is related to the angle of attack and aerofoil shape
through equation (79). It is easily seen from (80) that at the stagnation
point Y = −k. Differentiating (80) and setting the derivative to zero, one
can find that the maximum of the velocity is achieved at the point where
Y = 1/k.
Asymptotic Theory of Separated Flows 345
Y
y
x
M 1
k
X
O
k
O
Figure 17. The flow near the leading edge of a thin aerofoil.
Vτ ∂Vτ ∂Vτ κVτ Vn 1 ∂p 1 1 ∂ 1 ∂Vτ
+ Vn + =− + +
H1 ∂x ∂y H1 H1 ∂x Re H1 ∂x H1 ∂x
(81a)
∂ 2 Vτ ∂ Vτ κ ∂Vn 1 ∂ κVn
+ +κ + 2 + ,
∂y 2 ∂y H1 H1 ∂x H1 ∂x H1
Vτ ∂Vn ∂Vn κVτ2 ∂p 1 1 ∂ 1 ∂Vn
+ Vn − =− + +
H1 ∂x ∂y H1 ∂y Re H1 ∂x H1 ∂x
(81b)
∂ 2 Vn ∂ Vn κ ∂Vτ 1 ∂ κVτ
+ + κ − − ,
∂y 2 ∂y H1 H12 ∂x H1 ∂x H1
1 ∂Vτ ∂Vn κVn
+ + = 0. (81c)
H1 ∂x ∂y H1
346 A. Ruban
Here κ is the local curvature of the body contour, and H1 is the Lamé
coefficient,
H1 = 1 + κ(x)y.
The velocity components are related to the stream function through the
equations
∂ψ ∂ψ
= −H1 Vn , = Vτ . (82)
∂x ∂y
In the boundary layer the asymptotic expansion of the stream function
has the form
Substitution of (83) into (82) and then into the Navier-Stokes equations
(81), results in the classical boundary-layer equation
∂Ψ ∂ 2 Ψ ∂Ψ ∂ 2 Ψ dpe ∂ 3Ψ
− 2
=− + . (84a)
∂Y ∂x∂Y ∂x ∂Y dx ∂Y 3
Ψ=0 at x = 0, (84b)
∂Ψ
Ψ= = 0 at Y = 0, (84c)
∂Y
and the matching condition with the solution in the inviscid flow region,
∂Ψ
= Ue (x) at Y = ∞. (84d)
∂Y
The pressure gradient dpe /dx on the right hand side of equation (84a)
does not depend on Y , and may be calculated with the help of the Bernoulli
equation,
pe + 12 Ue2 = 12 . (85)
Differentiating (85), we have
dpe dUe
= −Ue .
dx dx
Remind that the velocity at the outer edge of the boundary layer, Ue (x),
is given by equation (80). It is zero at the front stagnation point O, but
Asymptotic Theory of Separated Flows 347
rapidly grows as the fluid flows around√ the leading edge of the aerofoil. It
then reaches a maximum value of 1 + k 2 at point M on the upper side
of the aerofoil, for which Y = 1/k; see Figure 17. Downstream from this
point Ue (x) shows monotonic decay, and tends to unity as x → ∞.
The results of numerical solution of problem (84) are shown in Figures 18
in the form of the skin friction distribution along the aerofoil surface. The
skin friction is calculated as
∂ 2 Ψ
τ= . (86)
∂Y 2 Y =0
It appears that there exists a critical value of parameter k = k0 = 1.1575.
If k > k0 , then the solution terminates at a finite position where the skin
friction turns zero and the Goldstein (1948) singularity develops in the so-
lution. The larger the parameter k, the earlier on the aerofoil surface the
singularity is encountered.
k = 1.3
k = 1.15
τ × 100
τ
k = k0
x x
a) Skin friction distribution on b) Local behaviour of the solu-
the aerofoil nose for k = 1.15 tion near point x = x0 ; the
and k = 1.3 graphs of
Figure 18. Results of the numerical solution of problem (84).
If, on the other hand, k < k0 , then the solution exists for all values
of x. Interestingly enough, in this case the skin friction develops a mini-
mum, which tends to zero as k → k0 − 0. We denote the coordinate of the
point where τ first becomes zero by x0 . The calculations show that for the
boundary layer on the parabola surface x0 = 8.265.
the position of zero skin friction on the body surface gives the position of flow
separation. It is therefore of considerable interest to study the behaviour of
the solution of the problem (84) in the vicinity of this point.
We shall denote the point of zero skin friction as xs . For k > k0 it
lies upstream of x0 , and tends to x0 as k → k0 + 0. We first consider the
region lying before the line x = xs ; our task will be to find an asymptotic
expansion of the stream function Ψ as x → xs −0. At any point (x, Y ) in the
region upstream of x = xs , the longitudinal velocity component u = ∂Ψ/∂Y
is positive. Consequently, the boundary-layer equation (84a) possesses the
standard properties of equations of parabolic type. Its solution Ψ(x, Y ) near
the line x = xs depends upon the velocity distribution Ue (x) at the outer
edge of the boundary layer throughout the whole range of values of x from
the stagnation point O to the point x = xs of zero skin friction. On the
other hand, the asymptotic procedure utilised below is restricted to analysis
of a small vicinity of the line x = xs only, and therefore, does not take into
account all the boundary conditions affecting the solution for Ψ(x, Y ). This
means that the sought asymptotic expansion of Ψ(x, Y ) must be based on
the eigensolutions of the local problem. The coefficients multiplying the
eigenfunctions remain arbitrary from the viewpoint of the local analysis.
At the same time, they can be determined uniquely if the solution of the
problem (84) is constructed in the entire region x ∈ [0, xs ).
We start by noticing that the pressure gradient dpe /dx is a smooth
function. Near point x = xs it may be represented in the form of Taylor
expansion
dpe
= λ0 + λ1 s + · · · as s → ∞. (87)
dx
Here s = x − xs is the distance from the point x = xs . The leading order
term, λ0 in (87) coincides with the pressure gradient at point x = xs . For
all k ≤ k0 , when the zero skin point exists, λ0 > 0.
Setting Y = 0 in (84a) and using the no-slip conditions (84c), it is easy
to find that at any point on the aerofoil surface
∂ 3Ψ dpe
= . (88)
∂Y 3 dx
At the point of zero skin friction, in addition to the no-slip conditions (84a)
we also know that
∂ 2 Ψ
τ (xs ) = = 0. (89)
∂Y 2 Y =0
Substituting (87) into (88), and integrating the resulting equation with
(84a) and (89), we find that the leading order term of the asymptotic rep-
Asymptotic Theory of Separated Flows 349
Ψ = 16 λ0 Y 3 + · · · . (90)
We can now determine the thickness of the viscous flow region that forms
inside the boundary layer upstream of the point x = xs . In this region the
viscous term on the right hand side of equation (84a) should be comparable
with either term on the left hand side of (84a). Using, for example, the first
of these, we can write
∂Ψ ∂ 2 Ψ ∂ 3Ψ
∼ . (91)
∂Y ∂x∂Y ∂Y 3
It follows from (90) that the coefficient ∂Ψ/∂Y in the convective term on
the left hand side of (91) may be estimated as an order O(Y 2 ) quantity,
which allows to express (91) in a more simple form,
∂2Ψ ∂ 3Ψ
Y2 ∼ . (92)
∂x∂Y ∂Y 3
Approximating the derivatives in (92) by finite differences, we arrive at a
simple algebraic equation
Ψ Ψ
Y2 ∼ 3
(x − xs )Y Y
which, being solved for Y , shows that the thickness of the viscous region
decreases, as the point x = xs is approached, according to the law
! "
Y = O (−s)1/4 . (93)
Substituting (94) together with (95) into (84a), and setting s → −0, we
find that function f1 (η) satisfies the following ordinary differential equation
1 1 1 2
f1 − λ0 η 3 f1 + λ0 α + η f1 − λ0 αηf1 = 0. (96)
8 2 4
As the region considered adjoins the aerofoil surface, we pose the no-slip
conditions
f1 (0) = f1 (0) = 0, (97)
which are deduced by substituting (94), (95) into (84c).
Equation (96) is linear and homogeneous. Three complementary solu-
tions to this equation, f11 (η), f12 (η) and f13 (η), can be chosen such that
f11 (0) = 1, f11 (0) = 0, f11 (0) = 0,
f12 (0) = 0, f12 (0) = 1, f12 (0) = 0, (98)
f13 (0) = 0, f13 (0) = 0, f13 (0) = 1.
The first two solutions do not satisfy the boundary conditions (97) and
must be rejected. As for the third solution, it can easily be verified by
direct substitution into (96) to be simply f13 = 12 η 2 . Hence, a non-trivial
solution of equation (96) with boundary conditions (97) exists for all α, and
may be written in the form
f1 (η) = 12 a0 η 2 , (99)
a20
f2p (η) = η,
2λ0
Asymptotic Theory of Separated Flows 351
where c0 = 1.
A recurrent equation for the coefficients cn of the series is obtained by
substituting cn η 4n+1 into the right hand side of the equation (104) and
cn+1 η 4n+5 into its left hand side. We find that
! "
λ0 n − (2α − 1) n − 14
cn+1 = cn .
32 n + 34 n + 54 (n + 1)
Γ(a + n)
(a)n = .
Γ(a)
Therefore,
n
Γ(5/4) λ0 Γ(n + 1 − 2α)
cn = − . (109)
Γ(1 − 2α) 32 Γ(n + 5/4) (4n − 1) n!
Now our task will be to determine the asymptotic behaviour of f22 (η)
as η → ∞. For this purpose we shall express f22 (η) through the Kummer’s
3
Here we use a well known property of the Gamma function, zΓ(z) = Γ(z + 1).
Asymptotic Theory of Separated Flows 353
function M (a, b, z), whose properties are well known4 . Remind that the
Kummer’s function is a solution of the confluent hypergeometric equation
d2 w dw
z 2
+ (b − z) − aw = 0,
dz dz
which remains regular at point z = 0. In fact, it may be represented by the
Taylor series
a (a)2 2 (a)n n
M (a, b, z) = 1 + z + z + ···+ z + ··· =
b (b)2 2! (b)n n!
∞ (111)
Γ(b) Γ(a + n) n
= z
Γ(a) n=0 Γ(b + n) n!
Γ(b) z a−b
M (a, b, z) = e z + ··· (113)
Γ(a)
as z tends to infinity along a ray that lies in the right half (z > 0) of the
complex plane z. Through (112) this makes function f22 (η) also to grow
exponentially as η → ∞. However, if
a = −m, m = 0, 1, 2 . . . , (114)
then Γ(a) = 0, and formula (113) cannot be used. In this case (a)m+1
and all the subsequent members of the sequence (108) vanish, reducing the
Taylor expansion (111) to a polynomial of degree m.
It is important to notice that regardless of the choice of parameter a, the
solution (94), (99), (103), (112) for the viscous region does not satisfy the
boundary condition (84d) at the outer edge of the boundary layer. There-
fore, in addition to viscous region (which is shown as region 2a in Figure 19),
4
See, for example, Abramowitz & Stegun (1965).
354 A. Ruban
Y
1
2c Re−1/2
2b
2d
2 2a x
xs
Figure 19. Asymptotic regions’ layout near the point of zero skin friction.
it is necessary to consider the main part of the boundary layer (region 2b),
where the asymptotic analysis of the boundary-layer equations (84) is based
on the limit
Y = O(1), s = x − xs → 0 − .
Since the boundary layer, we are dealing with, is exposed to a finite
pressure gradient, the fluid velocity in region 2b should remain finite. It
cannot be matched with the exponentially growing solution in region 2a.
Hence, an acceptable solution for region 2a may only be obtained under the
condition (114). Setting a = 1 − 2α in (114), leads to a conclusion that the
sought eigenvalues are
m+1
α= , m = 0, 1, 2 . . . . (115)
2
α = 1.
In addition to the terms shown, the expansion (116) contains the sum of
an infinite number of successive eigenfunctions, and also includes additional
terms produced by the higher order terms in the expansion of the pressure
gradient (87). All of these, however, are small as compared to (−s)5/4 f2 (η)
and can be disregarded.
Let us now turn to region 2b, where Y = O(1); see Figure 19. The form
of the asymptotic expansion of the stream function Ψ(x, Y ) in this region
may be determined with the help of the following standard procedure based
on the principle of matched asymptotic expansions. We perform the change
of variables η = Y /(−s)1/4 in (116) and, assuming Y = O(1), collect terms
of the same order as s → 0−. We find
1 a2 1 ! "
Ψ(x, Y ) = λ0 Y 3 − 0 Y 5 + · · · + (−s)1/2 a0 Y 2 + · · · + O (−s)3/4 .
6 240 2
This suggests that the solution in region 2b should be sought in the form
∂Ψ ∂ 2 Ψ dpe
= − 21 (−s)−1/2 Ψ00 Ψ01 + · · · , = λ0 + · · · ,
∂Y ∂x∂Y dx
∂Ψ ∂ 2 Ψ ∂ 3Ψ
= − 21 (−s)−1/2 Ψ01 Ψ00 + · · · , = Ψ
00 + · · · .
∂x ∂Y 2 ∂Y 3
Clearly, the pressure gradient and the viscous term lose their significance as
s → 0−, and equation (84a) reduces to
with the exponent μ being unknown in advance. The functions fˆ1 (ξ) and
fˆ2 (ξ) are determined in the same way as the functions f1 (η) and f2 (η) in
the solution (94) for region 2a. We have
Finally, we substitute (123) and (125) into (122), and recall that ξ = Y /s1/4 .
As a result we find that at the outer edge of region 2d
2μ−1 8μ−3
â2 Γ(5/4) λ0 Y
Ψ = 16 λ0 Y 3 + 0 + ··· . (126)
2λ0 Γ(1/4 + 2μ) 32 8μ − 5
358 A. Ruban
a20 5
Ψ = 16 λ0 Y 3 − Y + ··· . (127)
240
It is obtained by using the first of formulae (118) in the solution (121) in
region 2c.
According to the principle of matched asymptotic expansions, expres-
sions (126) and (127) should coincide with one another. This is only possible
if μ = 1 and5
â20 = −a20 . (128)
While the constant a0 is found in the process of solving the boundary-layer
equation (84a) from the stagnation point x = 0 to the point of zero skin
friction x = xs , in order to find constant â0 one has to use equation (128).
As this equation does not allow for a real solution, we can conclude that
the solution of the boundary-layer equation (84a) only exists upstream of
the line x = xs and cannot be extended beyond this line.
Ψ0 (x, Y ). In view of equation (128), the latter is possible only if the coeffi-
cient a0 in the first eigenfunction in the expansion (116) is zero.
This conjecture is confirmed by the numerical solution of boundary-layer
equation (84a). We have seen that for all k > k0 , Goldstein’s singularity
develops in the solution at the point of zero skin friction x = xs . However, as
the parameter k decreases, the point of zero friction xs moves downstream,
approaching the point x0 . Simultaneously, the singularity becomes weaker,
i.e. the coefficient a0 decreases. At k = k0 it becomes equal to zero, and
the first eigenfunction in (116) disappears.
The second eigenvalue
3
α= (129)
2
corresponds to m = 2 in (115). With (129) the solution in region 2a (see
Figure 19) turns into
or, equivalently,
Ψ01 λ0 − Ψ00
= 2 .
Ψ00 Ψ00
360 A. Ruban
Y
Ψ00 (Y ) − λ0
Ψ01 (Y ) = Ψ00 (Y ) C− ! "2 dY
. (132)
Ψ00 (Y )
0
Matching the solution (131), (132) in region 2b with the solution (130)
in region 2a shows, firstly, that the constant C in (132) is
a0
C= , (133)
λ0
and secondly, that
2λ0 λ1 7 λ0 a20 9
Ψ00 (Y ) = 16 λ0 Y 3 + Y + Y + ··· as Y → 0. (134)
7! 8!
This completes the construction of the solution before the point of zero
skin friction, x = x0 . It is interesting to notice that substitution of (133) into
(132) and then into (131) yields an asymptotic representation of Ψ0 (x, Y ),
Y
Ψ00 (Y ) − λ0 a0
Ψ0 (x, Y ) = Ψ00(Y ) + sΨ00 (Y ) ! "2 dY
− + · · · , (135)
Ψ00 (Y ) λ0
0
which proves to be valid both in region 2b and region 2a; see Exercise 2.
Let us now show that this solution can be continued downstream of point
x = x0 . We start with region 2c (see Figure 19), where Ψ0 (x, Y ) is sought
in the form
# 01 (Y ) + · · ·
Ψ0 (x, Y ) = Ψ00 (Y ) + sΨ as s → 0 + . (136)
Ψ0 (x, Y ) = s3/4 16 λ0 ξ 3 + s3/2 fˆ1 (ξ) + s7/4 F#1 (ξ) + s9/4 fˆ2 (ξ) + · · ·
(138a)
as s → 0+,
Asymptotic Theory of Separated Flows 361
# = â0 .
C (140)
λ0
Substitution of (140) back into (137) and then into (136) results in the
formula
Y
Ψ00 (Y ) − λ0 â0
Ψ0 (x, Y ) = Ψ00 (Y ) + sΨ00 (Y ) !
"2 dY + + · · · , (141)
Ψ00 (Y ) λ0
0
which, similar to (135), may be used not only in the main part of the
boundary layer (region 2c) but also in the viscous sublayer (region 2d).
Equation (138) shows that the solution of the boundary-layer problem
(84), which is unique before the point of zero friction, can be continued
downstream of this point in two ways. The first one is given by
â0 = −a0 .
In this case the formulae (135) and (141) may be written together
Y
Ψ00 (Y ) − λ0 a0
Ψ0 (x, Y ) = Ψ00 (Y ) + sΨ00 (Y ) !
"2 dY − + O(s2 ), (142)
Ψ00 (Y ) λ0
0
showing that the solution is smooth in the vicinity of the point of zero skin
friction, x = x0 . It follows from (142) that the skin friction changes sign at
362 A. Ruban
this point, namely,
∂ 2 Ψ0
τ= = −a0 s + O(s2 ), (143)
∂Y 2 Y =0
â0 = a0 .
Y
a0 Ψ00 (Y ) − λ0
Ψ0 = Ψ00 (Y ) + Ψ00 (Y ) |s| + s !
2
"2 dY + O(s ). (144)
λ0 Ψ00 (Y )
0
θ = Re−1/2 Θ, (145a)
where
Y
∂Ψ0 /∂x a0 λ0 − Ψ
00 (Y )
Θ=− = − sign(s) + ! "2 dY + O(s). (145b)
∂Ψ0 /∂Y λ0 Ψ00 (Y )
0
one can then conclude that the pressure in the boundary layer, pe (x), is also
representable by the Taylor expansion
Here ! "
1 2
p0 = 2 1 − Ue,0 , p1 = −Ue,0 Ue,1 .
Being guided by (147) and (148), we seek a solution of the boundary-
layer problem (84) in the form
The leading order term Ψ0 (x, Y ) has been studied in the preceding section.
We shall now consider the function Ψ1 (x, Y ). Substitution of (149) together
with (147) and (148) into (84) yields
Ψ1 = 0 at x = 0, (150b)
∂Ψ1
Ψ1 = =0 at Y = 0, (150c)
∂Y
∂Ψ1
= Ue,1 (x) at Y = ∞. (150d)
∂Y
The solution of (150) near point x = x0 may be constructed in the same
way as it was done for the leading order problem. We start with region
364 A. Ruban
2a (see Figure 19), where the asymptotic expansion of function Ψ1 (x, Y ) is
sought in the form
with
Y
η= (152)
(−s)1/4
assumed an order one quantity. The leading order term in (151) represents
an eigenfunction of the local solution; constant β is the eigenvalue to be
determined.
Substitution of (151) together with (130) into (150a) results in the fol-
lowing equation for g1 (η):
1 1 1 2
g1 − λ0 η3 g1 + λ0 β + η g1 − λ0 βηg1 = 0.
8 2 4
For any β, its solution satisfying the no-slip conditions
is written as
g1 = 12 a1 η 2 , (153)
where a1 is an arbitrary constant. In order to find the eigenvalue β one
needs to consider the second term in (151). Function g2 (η) satisfies the
equation
1 1 3 β 1
g2 − λ0 η 3 g2 + λ0 β+1 η2 g2 −λ0 β+ ηg2 = − + a0 a1 η 2 . (154)
8 2 4 2 4
It should be solved with the no-slip conditions on the aerofoil surface
The solution to (154) may be constructed in the same way as it was done
with equation (100). We find that
a0 a1
g2 (η) = (η − g22 ) + 12 b1 η 2 , (155)
λ0
where function g22 (η) allows for the power series representation
∞
n
λ0 (−β − 1/2)n
g22 (η) = − η4n+1 , (156)
n=0
32 (5/4)n (4n − 1) n!
Asymptotic Theory of Separated Flows 365
It follows from (157) that g22 (η) grows exponentially as η → ∞ for all
values of β except
β = m − 12 , m = 0, 1, 2 . . . . (158)
Equation (158) gives the sought sequence of the eigenvalues; the first of
these being
β = − 12 . (159)
With (159) all the coefficients in (156), except the first one (n = 0), are
zeros, and we have
g22 = η. (160)
Using (160) in (155), and substituting (155) together with (159) and (153)
into (151), we arrive at a conclusion that the solution of equation (150a) in
region 2a has the form
In order to predict the form of the solution in region 2b (see Figure 19),
we rearrange (161) with the help of (152). We have
which suggests that in region 2b, where Y = O(1), the asymptotic expansion
of Ψ1 (x, Y ) should be written in the form
! "
Ψ1 (x, Y ) = (−s)−1 Ψ11 (Y ) + O (−s)−1/4 as s → 0−, (162)
Ψ11 (Y ) = 12 a1 Y 2 + · · · as Y → 0. (163)
The solution of this equation, satisfying the boundary condition (163), has
the form
a1
Ψ11 = Ψ (Y ). (164)
λ0 00
366 A. Ruban
It remains to substitute (164) back into (162), and we can conclude that in
region 2b
a1 ! "
Ψ1 (x, Y ) = (−s)−1 Ψ00 (Y ) + O (−s)−3/4 as s → 0−, (165)
λ0
Similar to (135), equation (165) proves to be valid not only in region 2b
but also in region 2a; see Exercise 2. This statement is easily verified by
substituting (134) into (165) and comparing the resulting expression with
(161).
Let us now return to the expansion (149). As any other asymptotic
expansion it is expected to have a proper ordering of the terms, namely,
each subsequent term should be much smaller than the previous one. The
assumption that the expansion (149) satisfies this requirement was used
when deriving equation (150a), and became the basis of the entire proce-
dure employed for analysing the behaviour of the function Ψ1 (x, Y ). For
each point (x, Y ) situated upstream of the point x = x0 this assumption
is indeed satisfied thanks to the smallness of Δk. However, as x → x0 − 0
the function Ψ1 (x, Y ) increases without bound, leading to violation of the
supposed relationship between the terms in (149). Consequently, we have
to examine the vicinity of the point x = x0 separately.
In order to find the size of a new region that has to be introduced near
x = x0 , we notice that the x-dependence of both Ψ0 (x, Y ) and Ψ1 (x, Y ) is
determined by the eigenfunctions in the solutions (130) and (161) for region
2a. Comparing the contribution of the eigenfunctions in the asymptotic
expansion (149),
(−s)3/2 12 a0 η2 ∼ Δk (−s)−1/2 a1 η 2 .
we find that the longitudinal extent of the new region is estimated as
√
|x − x0 | ∼ ,
where = |Δk|.
The asymptotic structure of this region is shown in Figure 20. It is
composed of√two
layers. The first one (region 3) is a continuation of region 2a
into the O vicinity of the point x = x0 . Since η = Y /(−s)1/4 is an
order one quantity in region 2a, the thickness of region 3 may be estimated
as
Y ∼ |x − x0 |1/4 ∼ 1/8 .
We see that in region 3 the asymptotic analysis of the boundary-layer equa-
tions (84) is based on the limit
x − x0 Y
x∗ = = O(1), Y∗ = = O(1), → 0. (166)
1/2 1/8
Asymptotic Theory of Separated Flows 367
1
2c
4
2b 2d
3
2 2a
√
√
Figure 20. The O vicinity of the point x = x0 .
This suggests that in region 3 the solution of the boundary layer equations
(84) should be sought in the form
1
Ψ =3/8 λ0 Y∗3 + 6/8 Ψ∗1 (x∗ , Y∗ )+
6
(172)
7/8 1 3 2λ0 λ1 7
+ λ1 x∗ Y∗ + Y∗ + 9/8 Ψ∗2 (x∗ , Y∗ ) + · · · .
6 7!
It also follows from (171) that functions Ψ∗1 (x∗ , Y∗ ) and Ψ∗2 (x∗ , Y∗ ) satisfy
the following matching conditions with the solution in region 2a,
1$ %
Ψ∗1 = a0 (−x∗ ) + sign(Δk)a1 (−x∗ )−1 Y∗2 + · · · as x∗ → −∞
2
(173a)
and
a20 λ0 a20 9
Ψ∗2 = x∗ Y∗5 + Y +
5! 8! ∗
1$ %
+ b0 (−x∗ )7/4 + sign(Δk)b1 (−x∗ )−1/4 Y∗2 + · · · as x∗ → −∞.
2
(173b)
dΨ̆1
Ψ̆1 = = 0 at Y∗ = 0. (179)
dY∗
Since the ordinary differential equation (178) is linear and homogeneous, its
general solution may be written as
where ψ11 , ψ12 and ψ13 are the three complementary solutions of (178).
They can be chosen such that
ψ11 (0) = 1, ψ11 (0) = 0, ψ11 (0) = 0,
ψ12 (0) = 0, ψ12 (0) = 1, ψ12 (0) = 0,
ψ13 (0) = 0, ψ13 (0) = 0, ψ13 (0) = 1.
The first two solutions do not satisfy the conditions (179), and therefore, we
have to set C1 = C2 = 0 in (180). As far as the third solution is concerned,
it is written as
ψ13 = 12 Y∗2 ,
which is easily verified by direct substitution into (178). This reduces (180)
to
Ψ̆1 = 12 C3 Y∗2 . (181)
370 A. Ruban
1 2 dΨ̆2 d3 Ψ̆2
2 ikλ0 Y∗ − ikλ0 Y∗ Ψ̆2 = , (188a)
dY∗ dY∗3
dΨ̆2
Ψ̆2 = 0, = −Ğ(k) at Y∗ = 0. (188b)
dY∗
Here Ψ̆2 (k, Y∗ ) and Ğ(k) are Fourier Transforms of functions Ψ2 and G∗ ,
respectively.
The general solution of equation (188a) is written as
where ψ21 , ψ22 and ψ23 are the three complementary solutions of the equa-
tion (188a). We shall choose them using the initial conditions
ψ21 (0) = 1, ψ21 (0) = 0, ψ21 (0) = 0,
ψ22 (0) = 0, ψ22 (0) = 1, ψ22 (0) = 0, (190)
ψ23 (0) = 0, ψ23 (0) = 0, ψ23 (0) = 1.
C1 = 0, C2 = −Ğ(k). (191)
Our task now will be to determine the function ψ22 . The power series
presentation of this function may be constructed in the same way as it was
done with function f22 (η) representing the second complementary solution
of equation (100). We start with the conditions (190). They show that the
first term in the series is
ψ22 = Y∗ + · · · . (194)
Using (194) on the left hand side of (188a) yields
d3 ψ22
= − 21 ikλ0 Y∗2 .
dY∗3
372 A. Ruban
Using (195) on the left hand side of the equation (188a) and (196) on the
right hand side, we find that the coefficients of the series (195) satisfy the
following recurrent equation
ikλ0 n − 14
cn+1 = cn , (197a)
32 n + 34 n + 54 (n + 1)
c0 = 1. (197b)
By direct substitution one can easily verify that the solution of (197) is
n
ikλ0 Γ(5/4)
cn = − . (198)
32 Γ(n + 5/4)(4n − 1)n!
Substituting (198) back into (195), we have
∞
n
5 ikλ0 Y∗4n+1
ψ22 (Y∗ ) = −Γ 4 . (199)
n=0
32 Γ(n + 5/4)(4n − 1)n!
We shall show now that the function ψ22 (Y∗ ) may be expressed through
the Bessel function. It is known (see, for example Abramowitz & Stegun,
1965) that the Bessel function of the first kind and order ν may be repre-
sented by the power series
∞
2n+ν
(−1)n z
Jν (z) = , (200)
n=0
Γ(n + ν + 1) n! 2
Asymptotic Theory of Separated Flows 373
Therefore
∞
2n+1/2
(−1)n z
=
n=0
Γ(n + 5/4)(4n − 1)n! 2
z
3/2
1/2 (203)
z 2 J1/4 (z) 1 1 z
= 1/4
− dz − .
8 z (z/2) Γ(5/4) Γ(5/4) 2
0
We substitute (201) into (203), and restrict our attention to the leading
order exponentially growing term. Using the integration by parts with
9/4
2
u= , dv = cos z − 38 π dz,
z
13/4
9 2
du = − dz, v = sin z − 38 π ,
8 z
we find that
∞
2n+1/2
(−1)n z
=
n=0
Γ(n + 5/4)(4n − 1)n! 2
5/4
1 2
= √ sin z − 38 π + · · · as |z| → ∞.
4 π z
374 A. Ruban
It remains to set z = 12 iΩ1/2 Y∗2 , where Ω = 12 ikλ0 , and we will see that
function ψ22 , given by (199), grows exponentially as Y∗ → ∞, namely,
5/4
Γ(5/4) 4 1 1/2 2
ψ22 (Y∗ ) = − √ iπ/4 1/4 sin 2 iΩ Y∗ − 38 π + · · · .
2 πe Ω iΩ1/2 Y∗2
(204)
This means that the exponential growth of the solution (193) of the equation
(187a) can only be suppressed by setting
Ğ(k) = 0. (205)
G∗ (x∗ ) = 0.
Ψ = Ψ00 (Y )+
Y −1
1/2 Ψ00 − λ0 a0 (−x∗ ) + a1 sign(Δk)(−x∗ )
+ Ψ00 (Y ) x∗ 2 dY + +· · · .
Ψ00 λ0
0
Asymptotic Theory of Separated Flows 375
This suggests, firstly, that the asymptotic expansion of the stream function
Ψ(x, Y ) in region 4 should be sought in the form
Secondly, we see that the matching condition with the solution in region 2b
reads
Y
Ψ00 − λ0 A∗ (x∗ )
Ψ1 = Ψ00 (Y ) x∗
2 dY + +··· as x∗ → −∞. (211)
Ψ00 λ0
0
Here we use the fact that function A∗ (x∗ ) is represented at large negative
x∗ by (183).
We will also need the matching condition with the solution (208) in
region 3. It can be formulated with the help of a usual routine. We start
by expressing the solution in region 3 in terms of the variables of region 4.
Since the scaled tangential coordinate x∗ is common for regions 3 and 4, we
just need to recall that the relationship between the normal coordinates is
given by (166), i.e.
Y
Y∗ = 1/8 . (212)
Substitution of (212) into (208) yields
Now we turn our attention to the asymptotic expansion (210) of the stream
function in region 4. At the “bottom” of region 4 the leading order term in
(210) may be simplified with the help of (134), which turns (210) into
It remains to compare (213) with (214), and we can conclude that the sought
matching condition is written as
∂ 2 Ψ1 ∂ Ψ1
Ψ00 − Ψ00 = −λ0 + Ψ
00 , (216)
∂x∗ ∂Y ∂x∗
376 A. Ruban
or, equivalently,
∂2 Ψ1 Ψ − λ0
= 00 2 . (217)
∂x∗ ∂Y Ψ00 Ψ00
It follows from (215) and (134) that
Ψ1 A∗ (x∗ )
= .
Ψ00 Y =0 λ0
Y
∂ Ψ1 Ψ
00 − λ0 A∗ (x∗ )
= 2 dY + . (218)
∂x∗ Ψ00 Ψ00 λ0
0
Y
Ψ00 − λ0 A∗ (x∗ )
Ψ1 (x∗ , Y∗ ) = Ψ00 (Y ) x∗
2 dY + , (219)
Ψ00 λ0
0
Through making use of (134) one can verify that in region 3, where Y =
1/2 Y∗ , the expansion (220) reduces to (208). This means that we can use
(220) not only in region 4 but also in region 3. Substituting (207) into (220),
and returning to the original variable x = x0 + 1/2 x∗ , we have
Ψ(x, Y ) = Ψ00 +
Y
a0 a1 Ψ00 − λ0
(221)
+ Ψ00 (x − x0 ) + 2 Δk + (x − x0 )
2
2 dY + · · · .
λ0 a0 Ψ00
0
sign of the argument of the square root in (221). If a1 Δk > 0, then the
solution exists for all x. The skin friction
∂ 2 Ψ a1
τ= = a0 (x − x0 )2 + 2 Δk (222)
∂Y 2 Y =0 a0
remains positive everywhere and at the point x = x0 it reaches the mini-
mum, whose value decreases as Δk → 0 according to the rule
τmin = 2a0 a1 Δk. (223)
If, however, a1 Δk < 0, then the solution exists only up to the point of zero
friction,
|a1 Δk|
xs = x0 − 2 ;
a0
downstream of this point the argument of the square root becomes nega-
tive. On approach to the point x = xs the solution develops Goldstein’s
singularity. Indeed, it follows from (222) that
1/4 √
τ = 8a30 |a1 Δk| xs − x + · · · as x → xs − 0.
Notice that the singularity becomes progressively weaker as Δk → 0.
According to the numerical calculations (see Figure 18) the first of the
situations described is realized for Δk < 0 and the second for Δk < 0. This
means that the constant a1 is negative.
Remind that in the case when Δk = 0, the boundary-layer equations (84)
admit two solutions. One of them, given by (142), passes smoothly through
the point x = x0 , while the second solution, given by (144), develops a
singularity at this point. Comparing (142) and (144) with (221), one can see
that this is the singular solution (144) that represents the limiting solution
of the boundary-layer equations as k → k0 − 0.
In order to determine the constants a0 and a1 we need to compare the
analytical solution with the results of the numerical calculations of the
boundary-layer equations. Firstly, the skin friction distribution in Figure 18
calculated for k = k0 should exhibit near x = x0 the behaviour predicted
by formula (146). Using this fact, we found that the the flow at the leading
edge of the aerofoil, a0 = 0.0085. Then constant a1 was found using formula
(223). It appeared that a1 = −1.24.
region where the solution of the Euler equations was constructed using the
impermeability conditions on the aerofoil surface, i.e. the existence of the
boundary layer was completely ignored. As a result we found the veloc-
ity distribution (80) on the aerofoil surface, while the pressure distribution
was calculated using the Bernoulli equation. Then, as a second step, the
boundary layer was analysed. For this purpose the classical Prandtl’s equa-
tions (84) were used, with the pressure, pe , and the velocity at the outer
edge of the boundary layer, Ue , assumed uninfluenced by the presence of
the boundary layer. This assumption is based on the observation that the
boundary layer is only capable of causing an O(Re−1/2 ) displacement of the
streamlines from the aerofoil surface. Consequently, as long as the solution
in the boundary layer remains regular, its influence on the pressure field
remains weak.
The situation changes as the parameter k approaches its critical value k0 ,
and the singularity forms in the boundary layer. Remind that when k = k0 ,
the streamlines in the boundary layer develop a corner. The deflection angle
θ may be calculated using (145). We have
2a0
θ = Re−1/2 Θ − Θ = −Re−1/2 . (224)
x=x0 +0 x=x0 −0 λ0
Then it follows from the small perturbation inviscid flow theory (see Exer-
cise 3) that the pressure induced by the boundary layer
2
−1/2 2a0 U0
p = Re ln |x − x0 | + O(1) as x → x0 . (225)
πλ0
Here U0 = Ue (x0 ), and λ0 is the leading order term in the pressure gradient
(87); for an aerofoil with parabolic nose U0 = 1.286 and λ0 = 0.024.
We see that the induced pressure gradient exhibits an unbounded growth
as the singularity is approached, namely,
The fourth term in the above expansion is a part of equation (184a). When
comparing it with the induced pressure gradient (226),
Re−1/2
∼ 3/4 , (227)
x − x0
k = k0 + Re−2/5 k1 , (231)
1 5
2c
4
2b 2d
3
2 2a
−1/5
Re
Remind that U0 denotes the value of the tangential inviscid flow velocity at
point x = x0 on the aerofoil surface; Pe0 is the corresponding value of the
pressure. These are related to one another through the Bernoulli equation,
Pe0 = 12 (1 − U02 ).
Setting Y = 0 in (232c), yields the pressure at the outer edge of the
boundary layer in the form
pe (x) = p = Pe0 + c1 (x − x0 ) + c3 (x − x0 )2 + · · · . (233)
y=0
c1 = λ0 , c3 = 12 λ1 . (234)
2a0 U0
Vτ = −Re−1/2 ln (x − x0 )2 + y 2 + · · · , (235a)
πλ0
2a0 U0 y
Vn = −Re−1/2 π − arctan + ··· , (235b)
πλ0 x − x0
2a0 U02
p = Re−1/2 ln (x − x0 )2 + y 2 + · · · . (235c)
πλ0
Let us now turn our attention to region 5; see Figure 21. The longitudinal
extent of this region is given by (230). We shall see that the flow in region 5 is
described by the Laplace equation (241). The principle of least degeneration
requires the lateral size of region 5 to be comparable with its longitudinal
size,
y ∼ |x − x0 | ∼ Re−1/5 .
This means that the asymptotic analysis of the Navier-Stokes equations (81)
in region 5 has to be based on the limit procedure
x − x0 y
x∗ = = O(1), y∗ = = O(1), Re → ∞. (236)
Re−1/5 Re−1/5
In order to predict the form of the asymptotic expansions for the velocity
components and pressure in region 5, we shall express (232) and (235) in
terms of the new variables (236) and combine these together. Being applied
to the tangential velocity component Vτ this procedure yields
& ' & '
Vτ = U0 +Re−1/5 a1 x∗ + a2 y∗ + Re−2/5 a3 x2∗ + a4 x∗ y∗ + a5 y∗2 +
2a0 U0 2a0 U0 2
+ Re−1/2 ln Re − Re−1/2 ln x∗ + y∗2 + · · · .
5πλ0 πλ
This suggests that the solution in region 5 should be sought in the form
& '
Vτ =U0 + Re−1/5 a1 x∗ + a2 y∗ +
& '
+ Re−2/5 a3 x2∗ + a4 x∗ y∗ + a5 y∗2 + (237a)
2a 0 U0
+ Re−1/2 ln Re + Re−1/2 u∗1 (x∗ , y∗ ) + · · · ,
5πλ0
& '
Vn =Re−1/5 b2 y∗ + Re−2/5 b4 x∗ y∗ + b5 y∗2 +
(237b)
+ Re−1/2 v1∗ (x∗ , y∗ ) + · · · ,
& '
p =Pe0 + Re−1/5 c1 x∗ + c2 y∗ +
& '
+ Re−2/5 c3 x2∗ + c4 x∗ y∗ + c5 y∗2 − (237c)
2
2a 0 U 0
− Re−1/2 ln Re + Re−1/2 p∗1 (x∗ , y∗ ) + · · · ,
5πλ0
where functions u∗1 , v1∗ and p∗1 satisfy the following matching conditions with
the solution in region 1,
⎫
∗ 2a0 U0 2 ⎪
⎪
u1 = − ln x∗ + y∗ + · · · ,
2
⎪
⎪
πλ0 ⎪
⎪
⎪
⎬
∗ 2a0 U0 y∗
v1 = − π − arctan + ··· , as x2∗ + y∗2 → ∞. (238)
πλ0 x∗ ⎪
⎪
⎪
⎪
2a0 U02 2 ⎪
⎪
p∗1 = ln x∗ + y∗2 + · · · ⎪
⎭
πλ0
The equations for these functions are deduced by substituting (237) into
the Navier-Stokes equations (81), and working with the O(Re−3/10 ) terms.
We find
∂u∗1 ∂p∗ ∂v1∗ ∂p∗ ∂u∗1 ∂v ∗
U0 = − 1, U0 = − 1, + 1 = 0. (239)
∂x∗ ∂x∗ ∂x∗ ∂y∗ ∂x∗ ∂y∗
The set of equations (239) is easily reduced to a single equation for
function p∗1 . We start by eliminating u∗1 . This is done by solving the last of
equations (239) and substituting the result into the first equation. We have
∂v1∗ ∂p∗1 ∂v1∗ ∂p∗
U0 = , U0 = − 1. (240)
∂y∗ ∂x∗ ∂x∗ ∂y∗
Now we eliminate v1∗ by cross-differentiating equations (240). We find that
the pressure p∗1 satisfies the Laplace equation
∂ 2 p∗1 ∂ 2 p∗1
+ = 0. (241)
∂x2∗ ∂y∗2
Asymptotic Theory of Separated Flows 383
When solving this equation we will use the method of Fourier Trans-
forms. The later is applicable to functions that tend to zero as |x∗ | → ∞.
It follows from (238) that p∗1 does not belong to this category. Therefore,
we shall differentiate (241) with respect to x∗
∂2 ∂p∗1 ∂2 ∂p∗1
+ = 0, (242)
∂x2∗ ∂x∗ ∂y∗2 ∂x∗
and treat ∂p∗1 /∂x∗ as the sought function. The “far-field” boundary condi-
tion for this function is written as
x − x0 y
x∗ = , Y∗ = . (244)
Re−1/5 Re−11/20
Y∗
(−s) = Re−1/5 (−x∗ ), η= , Δk = k − k0 = Re−2/5 k1 ,
(−x∗ )1/4
384 A. Ruban
which, being substituted into (167), yield
1
ψ = Re−1/2 Ψ = Re−13/20 λ0 Y∗3 +
6
−16/20 1 1
+ Re (−x∗ ) a0 Y∗2 + (−x∗ )−1 k1 a1 Y∗2 +
2 2
1 2λ λ
0 1 7
+ Re−17/20 λ1 x∗ Y∗3 + Y∗ +
6 7!
2
a λ0 a20 9 1
+ Re−19/20 0 x∗ Y∗5 + Y + (−x∗ )7/4 b0 Y∗2 +
5! 8! ∗ 2
1
(−x∗ )−1/4 k1 b1 Y∗2 + · · · .
2
and
a20 λ0 a20 9
Ψ∗2 = x∗ Y∗5 + Y +
5! 8! ∗
1 $ %
+ b0 (−x∗ )7/4 + k1 b1 (−x∗ )−1/4 Y∗2 + · · · as x∗ → −∞. (246b)
2
We shall represent the pressure in region 3 by the asymptotic expansion
1 ∂Ψ∗1
Vτ =Re−2/20 λ0 Y∗2 + Re−5/20 +
2 ∂Y∗
2λ0 λ1 6 ∂Ψ∗2
(248a)
+ Re−6/20 12 λ1 x∗ Y∗2 + Y∗ + Re−8/20 + ··· ,
6! ∂Y∗
∂Ψ∗1 ∂Ψ∗2
Vn = −Re−12/20 − Re−13/20 61 λ1 Y∗3 − Re−15/20 + ··· . (248b)
∂x∗ ∂x∗
We then substitute (248) together with (247) into the Navier-Stokes equa-
tions (81). We find from the x-momentum equation (81a) that the equation
(175) for Ψ∗1 retains its form
2 ∗
1 2 ∂ Ψ1 ∂Ψ∗1 ∂ 3 Ψ∗1
2 λ0 Y∗ ∂x ∂Y − λ0 Y∗ = .
∗ ∗ ∂x∗ ∂Y∗3
∂Ψ∗1
Ψ∗1 = = 0 at Y∗ = 0,
∂Y∗
is written as
Ψ∗1 = 12 A∗ (x∗ )Y∗2 . (249)
At this stage, the function A∗ (x∗ ) remains arbitrary; we can only claim that
in view of (246a),
The equation (184a) for function Ψ∗2 now acquires an additional term,
the induced pressure gradient,
∂P ∗
= 0. (252)
∂Y∗
386 A. Ruban
Main part of the boundary layer (region 4) When the parameter k
belongs to the range (229), the asymptotic expansion (210) of the stream
function in region 4 is written as
Here
x − x0 y
x∗ = , Y = .
Re−1/5 Re−1/2
By analogy with (247) we shall seek the asymptotic expansion for the pres-
sure in region 4 in the form
∂ Ψ1 ∂ Ψ1
Vτ = Ψ00 (Y ) + Re−1/5 + ··· , Vn = −Re−1/2 + · · · . (255)
∂Y ∂x∗
We then substitute (255) and (254) into the x-momentum equation (81a).
We find that the equation (216) retains its form6 . Since the boundary
conditions (211) and (215) also remain unchanged, we can use for Ψ1 the
solution given by (219),
Y
Ψ00 − λ0 A∗ (x∗ )
Ψ1 (x∗ , Y∗ ) = Ψ00 (Y ) x∗
2 dY + . (256)
Ψ00 λ0
0
It follows from (255) and (256) that the angle made by the streamlines
with the aerofoil contour is calculated as
Y
Vn λ0 − Ψ
00 1 dA∗
ϑ= = Re−1/2
2 dY − + ··· .
Vτ Ψ00 λ0 dx∗
0
∂ϑ 1 d2 A∗
= −Re−3/10 + ··· (257)
∂x λ0 dx2∗
6
The induced pressure gradient ∂P ∗ /∂x∗ is simply too weak to affect the flow in region 4.
Asymptotic Theory of Separated Flows 387
stays constant across region 4.
To complete the flow analysis in region 4 we substitute (255) and (254)
into the y-momentum equation (81b). Restricting our attention to the lead-
ing order terms, we find
∂P ! "2
= κ(x0 ) Ψ00 (Y ) .
∂Y
Here κ(x0 ) is the curvature of the aerofoil contour at point x = x0 . We see
that while the pressure P changes with Y , the pressure gradient ∂ P /∂x∗
remains constant across region 4:
∂ ∂P
= 0.
∂Y ∂x∗
Comparing (265) with (263), we can conclude that the sought matching
condition is written as
∂v1∗ U 0 d2 A ∗
=− .
∂x∗ y∗ =0 λ0 dx2∗
It remains to reformulate this condition for the pressure gradient. For this
purpose the second equation in (240) is used. Differentiating the equation
with respect to x∗ and setting y∗ = 0 we find
∗
∂ ∂p1 U 2 d3 A ∗
= 0 at y∗ = 0. (266)
∂y∗ ∂x∗ λ0 dx3∗
λ0 a20 9 a20
Ψ∗2 = Ψ2 + Y + x∗ Y∗5 + 12 B∗ (x∗ )Y∗2 + G∗ (x∗ )Y∗ ,
8! ∗ 5!
with
A2∗ − a20 x2∗ − 2k1 a0 a1
G∗ (x∗ ) = ,
2λ0
Asymptotic Theory of Separated Flows 389
1 2 ∂ 2 Ψ̄ ∂ Ψ̄ ∂ 3 Ψ̄
2 Ȳ − Ȳ = − R(X), (267a)
∂X∂ Ȳ ∂X ∂ Ȳ 3
∂ Ψ̄
Ψ̄ = 0, = −G(X) at Ȳ = 0, (267b)
∂ Ȳ
Ψ̄ → 0 as X → −∞. (267c)
Here
G(X) = 1
2 A2 − X 2 + 2a , (268)
with parameter a defined as
2/5
(−a1 )λ0
a = k1 1/5 8/5
. (269)
a 0 U0
In the new variables the condition (250) is written as
Notice that being written in the new variables, the equations (267), (271)
describing the flow in the interaction region, involve a single controlling
parameter (269), which measures the deviation of the angle of attack from
its critical value; as a1 < 0, the parameter a increases with k1 , and hence,
with the angle of attack.
We shall now try to solve the equations of viscous-inviscid interaction
with the help of Fourier Transforms. We start with the inviscid flow in
region 5. It is easily seen that the Fourier transformation renders (271) in
the form
d2 r̆
− k 2 r̆ = 0, (273a)
dȳ 2
r̆ = 0 at ȳ = ∞, (273b)
dr̆
= ik Q̆ at ȳ = 0. (273c)
dȳ
Here r̆(k, ȳ) stands for the Fourier Transform of the function r(X, ȳ) defined
as
∞
r̆(k, ȳ) = r(X, ȳ)e−ikX dX,
−∞
It remains to substitute (276) back into (275), and we will have the solution
of the problem (273) in the form
ik
r̆ = − Q̆e−|k|ȳ . (277)
|k|
Let us now consider the flow in the viscous sublayer (region 3). The
Fourier transformation converts the problem (267) for this region into
1 2 dΨ̆ d3 Ψ̆
2 ik Ȳ − ik Ȳ Ψ̆ = − R̆, (278a)
dȲ dȲ 3
dΨ̆
Ψ̆ = 0, = −Ğ(k) at Ȳ = 0. (278b)
dȲ
Here Ψ̆(k, Ȳ ), Ğ(k) and R̆(k) are the Fourier Transforms of functions Ψ̄,
G(X) and R(X), respectively.
Unlike (188a) the equation (278a) is not homogeneous. Therefore, its
general solution is written as
where ψ2p is a particular solution of the equation (278a) and ψ21 , ψ22 and
ψ23 are the complementary solutions of the homogeneous part of (278a).
These may be chosen such that
ψ21 (0) = 1, ψ21 (0) = 0, ψ21 (0) = 0,
ψ22 (0) = 0, ψ22 (0) = 1, ψ22 (0) = 0,
ψ23 (0) = 0, ψ23 (0) = 0, ψ23 (0) = 1,
ψ2p (0) = 0, ψ2p (0) = 0, ψ2p (0) = 0. (280)
with Ω = 12 ik.
Now our task will be to study the particular solution ψ2p (Ȳ ). Function
ψ2p (Ȳ ) satisfies the equation (278a). If we set Ȳ = 0 in this equation, we
will find that
d3 ψ2p
= R̆. (283)
dȲ 3
Integration of (283) with the boundary conditions (280) yields
ψ2p (Ȳ ) = 16 R̆Ȳ 3 . (284)
We then use (284) to calculate the left hand side in (278a), and perform the
integration again. We find
2
ψ2p (Ȳ ) = 16 R̆Ȳ 3 +
ik R̆Ȳ 7 .
7!
The above procedure may be repeated as many times as one wishes, leading
to a conclusion that the power series of function ψ2p should be sought in
the form
∞
ψ2p (Ȳ ) = 16 R̆Ȳ 3 + cn+1 Ȳ 7+4n . (285)
n=0
Alternatively, we can write (285) as
∞
ψ2p (Ȳ ) = cn Ȳ 3+4n . (286)
n=0
Using (286) on the left hand side of the equation (278a), and (285) on the
right hand side, results in the following recurrent equation for the coefficients
cn ,
ik n + 14
cn+1 = cn ,
32 n + 4 n + 54 n + 32
7
We shall now show that the function ψ2p (Ȳ ) may be expressed in terms
of the Struve function Hν (z). It is known that the Struve function may be
represented by the power series
∞
2n+ν+1
(−1)n z
Hν (z) = , (289)
n=0
Γ(n + 3/2)Γ(n + ν + 3/2) 2
This means that as Ȳ tends to infinity the corresponding point in the com-
plex z-plane runs to infinity along one of the rays shown in Figure 22(b);
which one depends on the sign of k. On both rays {iz} < 0. Consequently,
k z
k
k>0 k<0
r
ϑ
⎫
ei(z− 38 π) − e−i(z− 38 π) ⎪
1 3 ⎪
⎪
sin z − 38 π = = − e−i(z− 8 π ) + · · · ,⎬
2i 2i (296)
ei(z− 38 π) + e−i(z− 38 π) ⎪
⎪
3
cos z − 8 π =
1 3
= e−i(z− 8 π) + · · · ⎪
⎭
2 2
as Ȳ → ∞.
Asymptotic Theory of Separated Flows 395
Substituting (296) into (295) and (282), and then into (281), we find
that at the outer edge of region 3 (see Figure 21),
5/4
1 Γ(5/4) Γ(3/4) 4 3
Ψ̆ = − i 3 π √ Ğ(k)+ 1/2
R̆(k) 1/2 2
e−i(z− 8 π) +· · · .
4e 4 Ω 1/4 π 4Ω iΩ Ȳ
In order to ensure that the solution in region 3 can be matched with the
solution in region 4, we have to suppress the exponential growth of the
function Ψ̆, which is done by setting
Γ(5/4) Γ(3/4)
√ Ğ(k) + R̆(k) = 0. (297)
π 4Ω1/2
The above equation relates (in the Fourier space) the pressure gradient
in region 3 with the function A(X). The second relationship between these
functions is given by the solution (277) in region 5. The solutions in regions 3
and 5 are linked to one another through the equation (272). Writing this
equation in terms of the Fourier Transforms, and using (277), we find that
ik
R̆(k) = r̆ = − Q̆(k). (298)
ȳ=0 |k|
The Fourier Transform of the pressure gradient, R̆(k), may be easily elimi-
nated from (298) and (297), leading to
√
π (ik)1/2
Ğ(k) = Λ Q̆(k). (299)
2 |k|
Here Λ is a constant given by
Γ(3/4)
Λ= √ .
2Γ(5/4)
We shall now try to express (299) in physical variables. Applying the
inverse Fourier transformation to the equation (299), renders it in the form8
√ ∞
1
2 2
π 1 (ik)1/2
2 A − X + 2a = Λ Q̆(k)eikX dk. (300)
2 2π |k|
−∞
8 1
` ´
Recall that Ğ(k) is the Fourier Transform of the function G(X) = 2
A2 − X 2 + 2a .
396 A. Ruban
If we substitute (301) into the integral on the right hand side of (300) and
change the order of integration, we will find that
∞
2 2 Λ
A − X + 2a = √ A (ξ)I(X, ξ)dξ, (302)
2 π
−∞
where
∞
(ik)1/2 ik(X−ξ)
I(X, ξ) = e dk. (303)
|k|
−∞
This equation does not allow for further simplification, and should be
solved numerically. When performing the calculations one needs to use
appropriate boundary conditions. The first of these is given by (270),
A(X) = (−X) − a(−X)−1 + · · · as X → −∞, (306)
and represents the condition of matching with the solution in the boundary
layer upstream of the interaction region. It should be noted, however, that
this condition alone does not make the solution of (305) unique. In addition
to solutions with short separation bubbles, we are interested in, it also allows
for the solutions with semi-infinite separation regions. Indeed, if, to make
it simple, we choose a = 0, then the equation (305) with the boundary
condition (306) will admit the solution A = −X. The latter matches with
the smooth branch (143) of the solution to the boundary-layer equation
outside the interaction region. In order to ensure the matching with the
singular branch (146), we shall require
A(X) = X + · · · as X → ∞. (307)
For more detailed analysis of the asymptotic behaviour of A(X) at large
values of |X| see Exercise 5.
Asymptotic Theory of Separated Flows 397
1
2
A(X)
3
6
5
4
X
Figure 23. Solutions of the equation (305) on the upper branch of the
fundamental curve: 1) a = −0.5; 2) a = 0.0; 3) a = 0.5; 4) a = 1.0; 5)
a = as = 1.139; 6) a = ac = 1.330.
minimal skin friction is lifted, and the skin friction appears to be positive
for all values of X ∈ (−∞, ∞). For graph 5 the parameter a has been
adjusted in such a way that the minimal skin friction returns back to zero
to capture the incipience of the separation. We found that this happens
at point X = 0.406 when the parameter a reaches the value as = 1.139.
Finally, graph 6 is plotted for the critical value of the parameter ac = 1.330.
It shows a region of negative A between X = −0.566 and X = 1.605,
occupied by the separation bubble. Interestingly enough, the solution does
not exist beyond a = ac .
This important result is illustrated by Figure 23, where the so called fun-
damental curve is displayed. This curve shows the entire set of admissible
solutions of the Marginal Separation theory. It is constructed in the follow-
ing way. Given a, the solution of the boundary value problem (305), (306),
(307) yields the distribution of the shear stress A(X) along the aerofoil sur-
face. Each such solution is represented by a point on the fundamental curve,
which is obtained by taking the value of A(X) at X = 0, and plotting A(0)
against the parameter a. The numbered circles on the fundamental curve
1 A(0)
1.5
2
1
3
0.5
4
−0.5 0 0.5 1 5 ac a
as
9
8
−0.5 6
−1 7
A(X)
5
7
X
Figure 25. Comparison of the solutions on the upper and lower branches
for a = as = 1.139.
in the second the separation region is already well developed. In fact, the
length of the separation region grows monotonically as an “observer” follows
the fundamental curve from the point 5 towards the critical point 6 and then
all the way along the lower branch. This trend is demonstrated by Figure 26,
where in addition to the solution at point 7 the solutions at points 8 and 9,
that lie on a small loop on the lower branch, are shown.
Despite the parameter a is still rather large on the loop, the solution
already shows an asymptotic behaviour characteristic of small a. When
a = 0 the equation (305) admits two solutions
A = −X and A = X. (308)
Neither of these satisfy both boundary conditions (306), (307), but they
are clearly visible in Figure 26 as two major fragments of the curves 8 and
9, being connected to one another through a sharp jump in a region that
becomes progressively shorter and moves to the right as a → 0−.
Summarising the results of the above analysis, we can conclude that
according to the Marginal Separation theory the flow near the leading edge
of a thin aerofoil exhibits a hysteresis behaviour. It should be noted that
hysteresis is routinely observed in experiments with aerofoils. These are
normally conducted in such a way that the angle of attack is at first increased
slowly enough to keep the flow quasi-steady, and then, after achieving the
aerofoil stall, it is gradually decreased. What one usually observes is that
the angle of attack at which the separation region forms on the upper surface
of the aerofoil does not coincide with the angle of attack at which the flow
400 A. Ruban
9
8
A(X)
X
Figure 26. Solutions on the lower branch of the fundamental curve: 7)
a = as = 1.139; 8) a = 0.600; 9) a = 0.680.
returns back to attached form. As a result the graph of the lift force versus
the angle of attack assumes the shape of a hysteresis curve. Within the
hysteresis loop, for each value of the angle of attack two flow states become
possible. The choice between them depends on the flow history of the
development of the flow.
Still experimental observations of the short separation bubbles show that
these are formed in a smooth manner without abrupt change of the flow
field, which is only possible if the solution remains on the upper branch of
the fundamental curve when the parameter a passes through as . Of course,
the flow cannot continue to change smoothly when the parameter a passes
through the critical value ac . The non-existence of the solution to (305),
(306), (307) for a > ac suggests that the flow has to undergo a sudden
change, known from experiment as the “bubble bursting”.
In order to calculate the values of the angle of attack at the incipience of
the short separation bubble and at the bubble bursting, we need to return
to the equation (79). Combining it with (231) and (269), we find that the
scaled angle of attack α∗ = α/ε is given by
1/5 8/5 1
1 a U0 1 G(x )
α∗ = √ k0 + Re−2/5 0 2/5
a − dx . (309)
2 π x (1 − x )
(−a1 )λ0
0
Acknowledgment: The author would like to thank S. Braun for his help
in preparing the manuscript.
θ
O x
0 ∞
(ik)1/2 ik(X−ξ) (ik)1/2 ik(X−ξ)
I(X, ξ) = e dk + e dk, (316)
−k k
−∞ 0
and assume, first, that ξ < X. In this case, when calculating the first
integral, I1 ,in (316) change the contour of integration from the negative
real semi-axis (contour C1 in Figure 28a) to the positive imaginary semi-
axis, C1 . Observe that, according to Jordan’s lemma, the integral along the
k N
CR1 CR2
C1 k C2
s
C1 ϑ C2
a) Contour change for the first b) Contour change for the second
integral in (316). integral in (316).
from a point on C1 to the coordinate origin, and confirm that the adopted
rule of calculating (ik)1/2 gives on C1 ,
(ik)1/2 = −is1/2 .
For the case when ξ > X, change the contour of integration for the two
integrals in (316) as indicated in Figure 29, and show that
√
π
I1 = I2 = √ .
ξ−X
C3 C4
C3 C4
CR3 CR4
Exercise 5. Notice that the right hand side of the equation (305) tends
to zero as X → ∞. This means that the first two terms of the asymptotic
expansion of A(X) at large positive values of X may be written as
a
A(X) = X − + ··· as X → ∞. (317)
X
Differentiate (317) twice, and substitute the result into the integral on the
right hand side of (305). Perform the integration using the following chain
of substitutions: ξ = Xs, then s = t2 + 1 and, finally, t = tan θ. Conclude
that
A2 − X 2 + 2a = − 43 aΛπX −5/2 + · · · as X → ∞. (318)
Now, assume that X → −∞, and split the integral on the right hand
side of (305) as
∞ −Δ Δ ∞
A (ξ) A (ξ) A (ξ) A (ξ)
√ dξ = √ dξ + √ dξ + √ dξ, (319)
ξ−X ξ−X ξ−X ξ−X
X X −Δ Δ
Consider, first, the middle integral. Notice that thanks to the first in-
equality in (320) it may be approximated as
Δ Δ )
A (ξ) 1 1 (
I2 ≈ √ dξ = √ A (ξ)dξ = √ A (Δ) − A (−Δ) .
−X −X −X
−Δ −Δ
Keeping in mind that Δ is large, calculate A (Δ) and A (−Δ) using (317)
and (306), respectively. Conclude that
2
I2 = √ + ··· . (321)
−X
When evaluating the first integral in (319), calculate A (ξ) with the help
of (306). Then use the fact that (−ξ) ≤ Δ everywhere in the integration
interval, and show that
−Δ
2a dξ 4a
|I1 | ≤ √ = 3 (−X) − Δ.
Δ3 ξ−X Δ
X
406 A. Ruban
Use the asymptotic expansion (317) to calculate A (ξ) in (322), and intro-
duce a new integration variable t through the substitution ξ = (−X)(t2 −1).
This leads to
∞
4a dt
I3 = − .
(−X)5/2 q (t − 1)3 (t + 1)3
Δ
1+ (−X)
Confirm that under the conditions (320), integrals I1 and I3 are much
smaller than I2 . Substitute (321) into (319), and conclude that
2Λ
A2 − X 2 + 2a = √ + ··· as X → −∞.
−X
Exercise 6. Show that the equation (305) may be inverted to take the
form
∞
1 A2 (ξ) − ξ 2 + 2a
A (X) = 1 − √ dξ.
πΛ ξ−X
X
For this purpose introduce a new function S(X) such that S = G (X), and
deduce from the equation (299) that
2 |k|
Q̆ = √ S̆. (323)
πΛ (ik)3/2
Apply the inverse Fourier transformation to (323), and using the technique
described in Exercise 4, show that
∞
1 S(ξ)
A (X) = − √ dξ. (324)
2πΛ ξ−X
X
Asymptotic Theory of Separated Flows 407
Recall that the function G(X) was introduced through the equation (268),
and it follows from (318) that G(X) = O(X −5/2 ) as X → ∞. Keeping this
in mind, apply the integration by parts to (325), and show that
∞
F (X) = −2 ξ − XS(ξ)dξ. (326)
X
Finally, substitute (327) into (324) and integrate the resulting equation
using the fact that, according to (317), A = 1 at X = ∞.
Bibliography
Abramowitz, M. & Stegun, I. A. 1965 Handbook of Mathematical Functions,
3rd edn. Dover Publications.
Chapman, D. R., Kuehn, D. M. & Larson, H. K. 1956 Investigation of
separated flows in supersonic and subsonic streams with emphasis on
the effect of transition NACA Rep. 1356.
Goldstein, S. 1948 On laminar boundary-layer flow near a position of sep-
aration. Q. J. Mech. Appl. Math. 1 (1), 43–69.
Helmholtz, H. 1868 Über diskontinuirliche Flüssigkeitsbewegungen. Monats-
bericht Akad. Wiss. Berlin pp. 215–228.
Hsiao, C. T. & Pauley, L. L. 1994 Comparison of the triple-deck theory,
interactive boundary layer method, and Navier-Stokes computations for
marginal separation. Trans. ASME J. Fluids Eng. 116, 22–28.
Jones, B. M. 1934 Stalling. J. Roy. Aero. Soc. 38, 753–770.
Kirchhoff, G. 1869 Zur Theorie Freier Flüssigkeitsstrahlen. J. für die Reine
und Angew. Math. 70 (4), 289–298.
Lamb, H. 1932 Hydrodynamics, 6th edn. Cambridge University Press.
Landau, L. D. & Lifshitz, E. M. 1944 Mechanics of Continuous Media.
Gostekhizdat, Moscow.
408 A. Ruban
1 Problem formulation
The viscous inviscid interactions of steady weakly three-dimensional tran-
sonic flows in narrow channels are considered which are triggered, for ex-
ample, by a shallow deformation of the channel walls. Using asymptotic
analysis for large Reynolds number Re = ũr L̃/ν̃ 1 Kluwick and Git-
tler, assuming two dimensional steady flows of a perfect gas, showed that
a consistent interaction theory can be formulated in which the flow inside
the inviscid core region is almost one-dimensional, A. Kluwick [2001]. The
former theory can be extended to the weakly three dimensional case if the
heights H̃ and h̃ of the channel and the surface mounted obstacle are of
orders 3 L̃ and 7 L̃ and if the length ΔX and width ΔZ of the obstacle are
of orders 3 L̃ and 2 L̃ with = Re−1/12 1. Here ũr , L̃ and ν̃ denote
the flow velocity in the core region just upstream of the local interaction
region, a characteristic length associated with the unperturbed boundary
layer adjacent to the channel wall and a reference value of the kinematic
viscosity. As in A. Kluwick [2001] the field quantities inside the inviscid
core region do not depend on the distance measured perpendicular to the
channel wall in leading order resulting, however, in a two-dimensional rather
than one-dimensional flow behavior.
The interaction region exhibits a triple deck structure, where as in the clas-
sical triple deck theory, e.g. see Stewartson [1974], the role of the main deck
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
410 A. Kluwick and M. Kornfeld
∂U ∂V
+ = 0,
∂X ∂Y
∂U ∂U ∂P ∂2U
U +V =− + , (1)
∂X ∂Y ∂X ∂Y 2
∂W ∂W ∂P ∂ 2W
U +V =− + ,
∂X ∂Y ∂Z ∂Y 2
where (X, Y, Z), (U, V, W ) and P denote Cartesian coordinates parallel and
perpendicular to the channel wall and the lateral direction, the correspond-
ing velocity components and the pressure. All quantities are suitable scaled.
The boundary conditions include the no slip condition on the channel walls,
the requirement that the unperturbed velocity profile is recovered in the
limit X → −∞ and a matching condition between the lower and main deck
for large Y
Y = S(X, Z) : U = V = W = 0, X → −∞ : U = Y,
1 X ∂P (2)
Y → ∞ : U = Y + A(X, Z), W =− dζ = 0,
Y −∞ ∂Z
4.5
Y 3.5
upper deck
3
H ∼
3
Y
2.5
S(X,Z) −Z 6 2
7 1.5
2
1
main deck
0.5
lower deck
X 0
-1.5 -1 -0.5
X
0 0.5 1 1.5
L=1 3
Figure 2. Stream lines and separa-
Figure 1. Triple deck structure of
tion bubble through the centerline
interaction region; := Re−1/12 .
of the hump
2 Results
For the numerical solution of the boundary layer equations (1) to (3) a
pseudo spectral method is used. Figure 3 shows the streamlines (dashed
lines) as well as the separation stream line (solid line) in a cross section
through the √ center line of the surface mounted hump (Z = 0) with S(X, Z) =
h · cos2 ( π2 X 2 + Z 2 ) and a height h = 2.25 for Λ = 2.5, K = 1 (subsonic
flow). The analytical separation angle ϑ (dashed dotted line), obtained by
Oswatitsch [1980], at the lee side of the hump matches the angle of the
separation streamline at the wall quite well.
For a constriction with h = 1 (to avoid the above mentioned effect of flow
separation) the pressure perturbation increases with decreasing absolute
values |K| under both, sub- and supersonic flow conditions, see Figure 4.
In the subsonic case, if |K| is small enough, i.e. the flow is close enough to
the point of transition P = −1, the flow exhibits a local supersonic region in
the upper deck. There is no corresponding local subsonic region in the core
region, i.e a pressure perturbation greater than P = 1, under supersonic
flow conditions for this surface perturbation.
Another remarkable effect is the phenomenon of upstream influence. For su-
personic flow conditions a first mathematical explanation of this effect was
given by Lighthill [1953]. For the two-dimensional subsonic case A. Kluwick
[to appear 2010] showed, that there is strictly no upstream influence. In
contrast to these results one finds an upstream influence in the three-
412 A. Kluwick and M. Kornfeld
0.2
0.4
0.18 analytical K = −1 K = −2
separation angle 0.2
0.16 K = −10
0.14
0
0.12
separation streamline -0.2
(numerical result)
Y
P
0.1
-0.4
0.08
-0.6
K = −0.1
K = −0.001
0.06
ϑ -0.8
0.04
0.02 -1 K →0
0 -1.2
0.55 0.56 0.57 0.58 0.59 0.6 0.61 0.62 0.63 0.64 0.65 -10 -5 -1 0 1 5 10 15 20
X X
also under subsonic flow conditions. A deeper insight into the effect of
upstream influence is gained through studying the linearized problem in
spectral space. The pressure perturbation in spectral space is found to be
sign(K) Λ2 k 2
P ∗∗ = 1 2
S ∗∗ (4)
−k 2 − sign(K) |K| l + sign(K) Λ2 γ −4/3 (ik)7/3
with γ −4/3 = − 3Ai1 (0) , where Ai (0) denotes the first derivative of the
Airy function and k, l the spectral variables corresponding to the physical
variables X and Z, respectively. Investigation of the poles of the pressure
perturbation (4) shows that there is at least one pole causing an upstream
influence even in the 3D subsonic case. Consequently, the upstream behav-
γ4
ior for −X 1 and Z = 0 is found to be P ∼ − X1 e8 Λ3 X for the supersonic
case K < 0 and P ∼ X12 for the subsonic case K > 0. In the supersonic
case the pressure decays algebraic-exponentially with the same exponent as
in the two-dimensional case, which is not surprising since a weakly three-
dimensional problem is studied and solutions deviating only slightly from
strictly two-dimensional solutions are expected. In the subsonic case an al-
gebraic decay, which is typical for the decay of perturbations under subsonic
conditions, is found.
Both analytical results match the numerical results, found upstream of a
perturbation caused by a surface mounted hump, cf. Figure 6 and Figure
5, quite well. Following the idea of Lighthill to formulate a free interaction
theory under supersonic conditions we are looking for solutions of the form
1 8 γ 43 X 1 γ4
U ∼Y + e Λ cos(β(X)Z)Ũ (Y ), V ∼ − e8 Λ3 X cos(β(X)Z)Ṽ (Y ),
X X
1 8 γ 43 X 1 γ4
P ∼ − e Λ cos(β(X)Z), A ∼ + e8 Λ3 X cos(β(X)Z)Ũ (∞),
X X
(5)
Weakly 3D Effects in Transonic Flows 413
-2
10 -2
10
-3
10
10
-3 2D; K < 0
-4
10
-4
10 -5
10
log(P)
3D; K > 0
log(P)
-6
10
-5
10
3D; K < 0
-7
10
-6
10
-8
10
-9
10
-7
10
1 0
-40 -35 -30 -25 -20 -15 -10 -5 -1 0
-10 -10
log(X) X
Figure 5. Decay of the effect of the Figure 6. Decay of the effect of the
upstream influence for K > 0 upstream influence for K < 0
1 γ 4 |K|
β(X) ∼ √ 8 3 . (6)
3 Λ |X|
Consequently, the three-dimensional structure of the perturbations is get-
ting weaker as −X increases, which is confirmed by numerical results. In
the limiting case |K| → 0 a two-dimensional behavior with no upstream
influence is obtained.
With the above discussed results of the effect of upstream influence there
arises the question if there exists a weakly three-dimensional regularized
shock profile similar to the two-dimensional case, e.g. a regularized shock
with a curved shock front. Since the weakly three-dimensional problem
covers the two-dimensional shock solution each additional weakly three-
dimensional solution would cause a non-uniqueness. In this case the as-
pects of the stability of the solutions as well as the transition between these
solutions are of interest.
Acknowledgement: This work has been financed by the Austrian Sci-
ence Fund in the framework of the WK Differential Equations.
Bibliography
G. Meyer A. Kluwick. Shock regularisation in dense gases by viscous inviscid
preparation. J. Fluid Mech., 664:473–507, 2010.
Ph. Gittler A. Kluwick. Transonic laminar interacting boundary layers in
narrow channels. ZAMM, 81:473–474, 2001.
A. Kluwick. Transonic nozzle flow of dense gases. J. Fluid Mech., 247:
661–699, 1993.
M.J. Lighthill. On boundary layers and upstream influence. Proc. Roy. Soc.
A, 217, pages 344–356, 1953.
414 A. Kluwick and M. Kornfeld
1 Fundamental Equations
For a comprehensive description of the underlying theory based on matched
asymptotic expansions the reader is referred to Ruban (2010) of this mono-
graph and to Sychev et al. (1998) for the case of steady planar flows. Ex-
tensions of the theory to incorporate transient and three-dimensional effects
can be found in Smith (1982), Ruban (1983), Duck (1990) and Braun &
Kluwick (2004), respectively. Our main concern is the investigation of the
Cauchy problem associated with the evolution of the displacement function
or equivalently the local wall shear stress A = A(x, t) and A = A(x, z, t)
governed by the fundamental equations
A2 −x2 +Γ = − 2π
λ
JK[ΔA]−γI[∂t A]+g(x, z, t) for (x, z) ∈ R2 , t ≥ 0, (2)
H. Steinrück (ed.), Asymptotic Methods in Fluid Mechanics: Survey and Recent Advances
© CISM, Udine 2010
416 M. Aigner and S. Braun
stream- and spanwise direction, the time, a control parameter and positive
constants. The functions g in (1) and (2), which remain bounded ∀t ≥ 0,
account for flow control devices which provide physically meaningful initial
conditions for A at t = 0, see Braun & Kluwick (2004). Also, we have
formally introduced the following integral operators
∞
∗ 1
J [f ](x, ·) = f (ξ, ·)dξ,
(ξ − x)1/2
x
x x
1 1
J[f ](x, ·) = f (ξ, ·)dξ, I[f ](x, ·) = f (ξ, ·)dξ,
(x − ξ)1/2 (x − ξ)1/4
−∞ −∞
x−ξ
K[f ](x, z) = f (ξ, η)dξdη,
((x − ξ) + (z − η)2 )3/2
2
R2
(3)
sometimes referred to as fractional derivatives, Weyl operators or Riesz
potentials.
Remark 1.1. If we consider problem (1) in the stationary case (i.e. ∂t A ≡ 0
and g ≡ 0) the resulting equation coincides exactly with the fundamental
problem derived in Ruban (2010) of this monograph (cf. equation (305)
therein, where x = X, Γ = 2a and λ = Λ).
It is well known (see Smith (1982) and Duck (1990)) that solutions to equa-
tions (1) and (2) may blow up at a finite time t = ts > 0. In this short
treatise we focus on the terminal structure of these solutions as t → ts (see
section 2).
Remark 1.2. Matching to the classical boundary layer flow requires the
far-field behaviour for equations (1) and (2) to be
A(x, ·) ∼ |x| as x → ±∞ and A(x, z, ·) ∼ |x| as x → ±∞ (4)
whereas A(x, z, ·) stays at least bounded as z → ±∞.
Remark 1.3. Duck (1990) considered a fully three-dimensional incompress-
ible marginally separated boundary layer flow in a line of symmetry which
led to a different z-dependent left hand side than in (2). As in the usual ap-
proach use was made of the interaction law, which relates the pressure P to
the displacement function A via P = K[∂x A]. In this case the fundamental
equation reads (omitting positive coefficients)
x
2 2 2 2
A + γ(z, t) − μ (βx + z ) = J[∂x P + ∂z2 P dξ] − I[∂t A] (5)
−∞
Self Similar Blow-up Structures in Unsteady Flows 417
bounded. Thus, by matching  with A at |(x̄, z̄)| = O(1), and using polar
coordinates (x̂, ẑ) → (r̂, φ̂) the far-field condition for (8)
Remark 2.2. Let  = Â(x̂) be independent of ẑ, then equations (1) and
(2) are equivalent (with the same far-field condition). In other words, the
fundamental equation for a planar flow can be considered as a special case
of the locally three-dimensional problem. This also holds for the blow-up
profile (cf. equations (7) and (8)).
3 Numerical Solution
A first attempt to gain insight into the behaviour of equations (1) and (7)
can be found in Smith (1982), which has been extended by Scheichl et al.
(2008) to a more detailed survey. Therein a finite differencing scheme to
solve equations (1) and (7) is presented. Performed numerical experiments
give strong evidence for the existence and uniqueness of a non-trivial solu-
tion of (7).
To solve the singular, nonlinear, homogeneous, partial integro-differential
equation (8) numerically we follow the approach of Scheichl et al. (2008)
for a discretization of the singular integrals appearing in (3). A nontrivial so-
lution is sought in the form Â(x̂, ẑ) = Â(x̂∗ , ẑ ∗ )φ(x̂, ẑ), where A(x̂∗ , ẑ ∗ )
= 0
for a fixed (x̂∗ , ẑ ∗ ) (and since we want  to be smooth in some sense, this
is true for a whole neighbourhood). Thus φ
≡ 0 with φ(x̂∗ , ẑ ∗ ) = 1.
Another key for a stable scheme is to find an appropriate, bijective map
ψ : C → R2 , where C ⊂ R2 is compact, such that one can impose bound-
ary conditions Â(ψ(u)) = 0 for u ∈ ∂C. The applied methods lead to an
inhomogeneous system (in the sense, that the null vector is not a possible
Self Similar Blow-up Structures in Unsteady Flows 419
4
3
2
1
x̂
Remark 3.1. The convergence of the used Newton method depends quite
sensitively on the map ψ, due to which the x̂-ẑ grid on the whole R2 is non-
equidistant. As a consequence, increasing the grid points does not directly
imply a more accurate solution. One additionally has to use an accordingly
modified ψ.
Despite all the faced difficulties, the numerical solution still provides a good
first insight into equation (8) and gives further ideas for some analytical
investigation.
420 M. Aigner and S. Braun
Â
4
3
2
1
ẑ
Bibliography
Braun, S., Kluwick, A. Unsteady three-dimensional marginal separation
caused by surface-mounted obstacles and/or local suction, J. Fluid Mech.
514, 121-152, 2004.
Duck, P.W. Unsteady three-dimensional marginal separation, including
breakdown, J. Fluid Mech. 220, 85-98, 1990.
Ruban, A.I. Stability of preseparation boundary layer on the leading edge
of a thin airfoil, Izv. Akad. Nauk SSSR: Mekh. Zhidk. Gaza 6, 55-63,
(Engl. transl. Fluid Dyn. 17, 835-843), 1983.
Ruban, A.I. Asymptotic Theory of Separated Flows in Steinrück, ed. Asymp-
totic Methods in Fluid Mechanics: Survey and Recent Advances, CISM
Courses and Lectures Vol. 523, Springer, 311–408, 2010.
Scheichl, S., Braun, S., Kluwick, A. On a similarity solution in the theory
of unsteady marginal separation, Acta Mech. 201, 153-170, 2008.
Smith, F.T. Concerning dynamic stall, Aeron. Quart. 33, 331-352, 1982.
Sychev, V.V., Ruban, A.I., Sychev, Vik.V., Korolev, G.L. Asymptotic The-
ory of Separated Flows, Cambridge University Press, Cambridge, 1998.