Tải bản đầy đủ (.pdf) (50 trang)

Vorticity and Vortex Dynamics 2011 Part 15 ppsx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (388.8 KB, 50 trang )

698 A Vectors, Tensors, and Their Operations
The particular importance of the permutation tensor in vorticity and
vortex dynamics lies in the fact that
ω
i
= 
ijk
u
k,j
= 
ijk

jk
, (A.11a)
due to (A.3) and (A.6). Inversely, by using (A.9b) it is easily seen that

jk
=
1
2

ijk
ω
i
. (A.11b)
This pair of intimate relations between vorticity vector and spin tensor show
that they have the same nonzero components and hence can represent, or are
dual to, each other. Note that (A.11b) also indicates that the inner product of
a vector and an antisymmetric tensor can always be conveniently expressed as
the cross-product of the former and the dual vector of the latter; for instance,
a ·Ω =


1
2
ω × a, Ω ·a =
1
2
a ×ω. (A.12a,b)
Obviously, we also have
∇·Ω = −
1
2
∇×ω. (A.13)
A.2 Integral Theorems and Derivative Moment
Transformation
The key result of tensor integrations is the extension of the fundamental the-
orem of calculus in one dimension

b
a
f

(x)dx =

b
a
df(x)=f(b) −f(a)
to multidimensional space. We state two classic theorems without giving proof
(e.g., Milne-Thomson 1968), which are then used to derive useful integral
transformations.
A.2.1 Generalized Gauss Theorem and Stokes Theorem
First, let V be a volume, having closed boundary surface ∂V with outward

unit normal vector n and ◦ denote any permissible differential operation of the
gradient operator ∇ on a tensor F of any rank. Then the generalized Gauss
theorem states that ∇◦FdV must be a total differentiation, and its integral
can be cast to the surface integral of n ◦FdS over the boundary surface ∂V
of V , where n is the unit outward normal vector:

V
∇◦FdV =

∂V
n ◦FdS. (A.14)
A.2 Integral Theorems and Derivative Moment Transformation 699
In particular, if F is constant, (A.14) yields a well-known result

∂V
n dS =

∂V
dS = 0, (A.15)
i.e., the integral of vectorial surface element dS = n dS over a closed surface
must vanish. Moreover, for the total dilatation and vorticity in V we have

V
ϑ dV =

∂V
n ·u dS, (A.16a)

V
ωdV =


∂V
n ×u dS. (A.16b)
One often needs to consider integrals in two-dimensional flow. In this case
the volume V can be considered as a deck on the flow plane of unit thickness.
Then (A.14) and some of the volume-integral formulas below remain the same
in both two and three dimensions, but care is necessary since in n-dimensions
δ
ii
= n is n-dependent. Some formulas for n = 3 need to be revised or do not
exist at all; see Sect. A.2.4 for issues special in two dimensions.
Next, let S be a surface with unit normal n, then without leaving S only
tangential derivatives can be performed and have chance to be integrated out,
expressed by line integrals over the boundary loop ∂S. The tangent differential
operator is naturally n ×∇, and the line element of ∂S has an intrinsic
direction along its tangent, dx = t ds, where t is the tangent unit vector
and ds the arc element. The directions of the normal n of S and t obey the
right-hand rule. Then, as the counterpart of (A.14), the generalized Stokes
theorem states that on any open surface S any (n ×∇) ◦FdS must be the
total differentiation, and its surface integral can be cast to the line integral of
dx ◦Falong ∂S:

S
(n ×∇) ◦FdS =

∂S
dx ◦F. (A.17)
Thus, if F is constant, (A.17) shows that the integral of element dx over a
closed line must vanish


t ds =

dx = 0; (A.18)
and if F = x, since
[(n ×∇) ×x]
i
= 
ijk

jml
n
l
x
k,m
= −2n
i
,
we obtain the well-known formula for the integral of a vectorial surface

S
n dS =
1
2

∂S
x ×dx (A.19)
700 A Vectors, Tensors, and Their Operations
with (A.15) as its special case since a closed surface has no boundary. In
general, (A.17) implies


S
(n ×∇) ◦FdS = 0 on closed S. (A.20)
The most familiar application of (A.17) to fluid mechanics is the relation
between total vorticity flux through a surface and circulation along the boun-
dary of the surface. Since (n ×∇) · u = n · (∇×u), there is

S
ω · n dS =

∂S
u ·dx. (A.21)
A.2.2 Derivative Moment Transformation on Volume
The Gauss and Stokes theorems permit the construction of useful identities for
integration by parts. In particular, we need to generalize the one-dimensional
formula

b
a
f(x)dx = bf(b) − af(a) −

b
a
xf

(x)dx,
which expresses the integration of f(x)bythex-moment of its derivative,
to various integrals of a vector f over a volume or surface, so that they are
cast to the integrals of proper moments of the derivatives of f plus bound-
ary integrals. We call this type of transformations the derivative moment
transformation.

We first use the generalized Gauss theorem (A.14) to cast the volume
integral of f to the moments of its divergence and curl. Since
(f
i
x
j
)
,i
= f
j
+ f
i,i
x
j
,

ijk

jlm
(f
m
x
k
)
,l
= 
ijk
(
jlm
f

m,l
)x
k
+ 
ijk

jkm
f
m
,
where x is the position vector, by (A.14) we find a pair of vector identities

V
f dV = −

V
x(∇·f)dV +

∂V
x(n ·f)dS, (A.22)

V
f dV =
1
n −1

V
x ×(∇×f)dV −
1
n −1


∂V
x ×(n ×f)dS, (A.23)
where n =2, 3 is the space dimension. The factor difference comes from the
use of (A.9c) for n = 3 and (A.10b) for n = 2. Note that for two-dimensional
flow, (A.22) still holds if f is on the plane (e.g., velocity), but becomes trivial
if f is normal to the plane (e.g., vorticity).
2
2
This can be verified by considering a deck-like volume of unit thickness. When
the vector is normal to the deck plane, one finds 0 = 0 from (A.22).
A.2 Integral Theorems and Derivative Moment Transformation 701
Then, we need to cast the first vector moment x×f to the second moments
of its curl, say F = ∇×f. When n =3,F has three second moments x
2
F ,
x ×(x ×F ), and x(x ·F ), related by
x ×(x ×F )=x(x · F ) −x
2
F .
Note that x(x ·F )=0 for n = 2. Then one finds
2

V
x ×f dV = −

V
x
2
F dV +


∂V
x
2
n ×f dS, n =2, 3, (A.24a)

V
x ×f dV =

V
x(x ·F )dV −

∂V
x(n ×f) ·x dS, n =3, (A.24b)
3

V
x×f dV =

V
x×(x×F )dV −

∂V
x×x×(n×f )dS, n =3. (A.24c)
Here, (A.24c) is the sum of (A.24a) and (A.24b). The trick of proving the first
two is using (A.14) to cast the surface integrals therein to volume integrals
first.Insodoingn becomes operator ∇ which then has to act on both f and
x.
If we make a Helmholtz–Hodge decomposition f = f


+f

, see (2.87) and
associated boundary conditions (2.98a) or (2.98b), then we can replace f by f

on the right-hand side of (A.22). Namely, the integral of a vector is expressible
solely by the derivative-moment integrals of its longitudinal part. However,
(A.23) is not simply a counterpart of this result in terms of the transverse
part of the vector. Rather, as long as n ×f

= 0 on ∂V , a boundary coupling
with the longitudinal part must appear. For some relevant discussions see Wu
and Wu (1993).
A.2.3 Derivative Moment Transformation on Surface
By similar procedure, we may use the Stokes theorem (A.17) to cast surface
integrals of a vector to that of its corresponding derivative moments plus
boundary line integrals. To this end we first decompose the vector to a normal
vector φn and a tangent vector n×A, since they obey different transformation
rules. Then for the normal vector we find a surface-integral identity effective
for n =2, 3

S
φn dS = −
1
n −1

S
x ×(n ×∇φ)dS +
1
n −1


∂S
φx ×dx. (A.25)
And, for n =3only, the integral of tangent vector can be cast to

S
n ×A dS = −

S
x ×[(n ×∇) ×A]dS +

∂S
x ×(dx ×A). (A.26)
In deriving these identities the operator n ×∇should be taken as a whole for
the application of (A.17). Setting φ = 1 in (A.25) returns to (A.19). In fact,
702 A Vectors, Tensors, and Their Operations
(A.25) is also a special case of (A.23) with f = ∇φ. Note that the cross-
product on the right-hand side of (A.26) can be replaced by inner product

S
n ×A dS =

S
x(n ×∇) ·A dS −

∂S
x(A ·dx), (A.27)
where (n ×∇) · A = n · (∇×A).
Then, for both n = 2 and 3, the integral of the first moment x × nφ can
be transformed to the following alternative forms:


S
x ×nφ dS =
1
2

S
x
2
n ×∇φ dS −
1
2

∂S
x
2
φ dx (A.28a)
= −

S
x[x ·(n ×∇φ)] dS +

∂S
φx(x ·dx) (A.28b)
= −
1
3

S
x ×[x ×(n ×∇φ)] dS +

1
3

∂S
φx ×(x ×dx). (A.28c)
Finally, to cast the surface integral of
x ×(n ×A)=n(x ·A) −A(x ·n)
to the second-moment of the derivatives of A, we start from two total deriv-
atives

ijk
(n ×∇)
j
(x
2
A
k
)=x
2

ijk
(n ×∇)
j
A
k
+ A
k

ijk
(n ×∇x

2
)
j
= x
2

ijk
(n ×∇)
j
A
k
+2(x
i
n
k
A
k
− n
i
x
k
A
k
),

ljk
(n ×∇)
j
(x
i

x
l
A
k
)=x
i
x
l

ljk
(n ×∇)
j
A
k
+ A
k

ljk
(n ×∇)
j
(x
i
x
l
)
= x
i
x
l


ljk
(n ×∇)
j
A
k
+3x
i
n
k
A
k
− A
i
x
k
n
k
.
Here, since what matters in x ×(n ×A) is only the tangent components of A,
we may well drop its normal component n
k
A
k
. Hence, subtracting the second
identity from 1/2 times the first yields an integral of x×(n×A). Using (A.17)
to cast the left-hand side to line integral then leads to the desired identity

S
x ×(n ×A)dS =


S
S ·[(n ×∇) ×A]dS −

∂S
S ·(dx ×A), (A.29a)
where
S =
1
2
x
2
I −xx or S
ij
=
1
2
x
2
δ
ij
− x
i
x
j
. (A.29b)
is a tensor depending on x only.
As a general comment of derivative moments, we note that, since in
(A.22),(A.23), and (A.25)–(A.27) the left-hand side is independent of the
choice of the origin of x, so must be the right-hand side. In general, if I
A.2 Integral Theorems and Derivative Moment Transformation 703

represents any integral operator (over volume or surface or a sum of both),
than the above independence requires
I{(x
0
+ x) ◦F}= I{x ◦F}
for any constant vector x
0
.Thus,wehave
x
0
◦ I{F} + I{x ◦F}= I{x ◦F},
which implies that, due to the arbitrariness of x
0
, there must be
I{F} =0. (A.30)
Namely, if we remove x from the right-hand side of (A.22), (A.23), and (A.25)–
(A.27), the remaining integrals must vanish. It is easily seen that this condition
is precisely the Gauss and Stokes theorem themselves.
A.2.4 Special Issues in Two Dimensions
The preceding integral theorems and identities are mainly for three-dimensional
domain, with some of them also applicable to two-dimensional domain. A few
special issues in two dimensions are worth discussing separately.
In many two-dimensional problems it is convenient to convert a plane
vector ae
x
+be
y
to a complex number z = x+iy by replacing e
z
× by i =


−1
(Milne-Thomson 1968), so that
e
y
= e
z
× e
x
=⇒ ie
x
(A.31a)
and hence
ae
x
+ be
y
=(a + be
z
×)e
x
=⇒ e
x
(a +ib). (A.31b)
Then the immaterial e
x
can be dropped. Thus, denoting the complex conju-
gate of z by ¯z = x − iy, for derivatives there is

x

= ∂
z
+ ∂
¯z
,∂
y
=i(∂
z
− ∂
¯z
), (A.32a)
2∂
z
= ∂
x
− i∂
y
, 2∂
¯z
= ∂
x
+i∂
y
, (A.32b)
so by (A.31)
∇ =⇒ 2e
x

¯z
, ∇

2
=⇒ 4∂
z

¯z
. (A.33)
The replacement rule (A.31) cannot be extended to tensors of higher ranks.
If in a vector equation one encounters the inner product of a tensor S and a
vector a that yields a vector b, then (A.31) can be applied after b is obtained
by common real operations. For example, consider the inner product of a
trace-free symmetric tensor
S = e
x
e
x
S
11
+
1
2
(e
x
e
y
+ e
y
e
x
)S
12

+ e
y
e
y
S
22
,S
11
+ S
22
=0,
704 A Vectors, Tensors, and Their Operations
and a vector a = e
x
a
1
+ e
y
a
2
=⇒ e
x
(a
1
+ia
2
). After obtaining a · S = S ·a
by real algebra, we use (A.31) to obtain
2a ·S =⇒ e
x

(a
1
− ia
2
)(2S
11
+iS
12
),S
11
+ S
22
=0. (A.34)
In this case S appears as a complex number S
11
+iS
12
/2 but a appears as its
complex conjugate a
1
− ia
2
.
Now, if F = f(x, y) is a scalar function, (A.17) is reduced to
e
z
×

S


e
x
∂f
∂x
+ e
y
∂f
∂y

dS =

S

e
y
∂f
∂x
− e
x
∂f
∂y

dS =

∂S
fdx.
Hence, by (A.31) this formula becomes
i

S


∂f
∂x
+i
∂f
∂y

dS =

∂S
f dz,
which by (A.32) is further converted to

∂S
f(z,z)dz =2i

S
∂f
∂z
dS, (A.35a)

∂S
f(z,z)dz = −2i

S
∂f
∂z
dS, (A.35b)
the second formula being the complex conjugate of the first. Milne-Thomson
(1968) calls this result the area theorem.

The two-dimensional version of the derivative-moment transformation on
surface, i.e., the counterpart of (A.25) and (A.26), also needs special care.
We proceed on the real (x, y)-plane. Let C be an open plane curve with end
points a and b,ande
s
and n be the unit tangent and normal vectors along
C so that (n, e
s
.e
z
) form a right-hand orthonormal triad. Then since
∂x
∂s
= e
s
, n ×∇φ = n ×

e
s

∂s
+ n

∂n

φ = e
z
∂φ
∂s
for any scalar φ and tangent vector tA, there is


C
nφ ds = −e
z
× (xφ)|
b
a


C
x ×(n ×∇φ)ds, (A.36)

C
tA ds =(xA)|
b
a


C
x
∂A
∂s
ds. (A.37)
The transformation of the first-moment integral of a normal vector nφ has
been given by (A.28). But that of a tangent vector, say x × e
s
A, cannot be
similarly transformed at all, because
x ×e
s

A =

∂s
(x ×xA) −e
s
× xA −x ×x
∂A
∂s
A.3 Curvilinear Frames on Lines and Surfaces 705
simply leads to a trivial result. What we can find is only a scalar moment

b
a
x ·At ds = −
1
2

b
a
x
2
∂A
∂s
ds +
1
2
x
2
A




b
a
. (A.38)
A.3 Curvilinear Frames on Lines and Surfaces
In the above development we only encountered Cartesian components of
vectors and tensors. In some situations curvilinear coordinates are more conve-
nient, especially when they are orthonormal. Basic knowledge of vector analy-
sis in a three-dimensional orthonormal curvilinear coordinate system can be
found in most relevant text books (e.g., Batchelor 1967), where the coordi-
nate lines are the intersections of a set of triply orthogonal surfaces. But,
the Dupin theorem (e.g., Weatherburn 1961) of differential geometry requires
that in such a system the curves of intersection of every two surfaces must
be the lines of principal curvature on each. While the concept of principal
curvatures of a surface will be explained later, here we just notice that the
theorem excludes the possibility of studying flow quantities on an arbitrary
curved line or surface and in its neighborhood by a three-dimensional ortho-
normal curvilinear coordinate system. These lines and surfaces, however, are
our main concern. Therefore, in what follows we construct local coordinate
frames along a single line or surface only and as intrinsic as possible, with an
arbitrarily moving origin thereon.
A.3.1 Intrinsic Line Frame
If we are interested in the flow behavior along a smooth line C with length
element ds, say a streamline or a vorticity line, the intrinsic coordinate frame
with origin O(x)onC has three orthonormal basis vectors: the tangent vector
t = ∂x/∂s,theprincipal normal n (toward the center of curvature), and the
binormal b = t × n, see Fig. A.2. This (t, n, b) frame can continuously move
along C and is known as intrinsic line frame. The key of using this frame is to
know how the basis vectors change their directions as s varies. This is given

by the Frenet–Serret formulas, which form the entire basis of spatial curve
theory in classical differential geometry (e.g., Aris 1962):
∂t
∂s
= κn,
∂n
∂s
= −κt + τb,
∂b
∂s
= −τn, (A.39a,b,c)
where κ and τ are the curvature and torsion of C, respectively. The curvature
radius is r = −1/κ with dr = −dn. The torsion of C measures how much a
curve deviates from a plane curve, i.e., it is the curvature of the projection of
C onto the (n, b) plane. For a plane curve τ =0andwehavea(t, n) frame
as already used in deriving (A.36)–(A.38).
706 A Vectors, Tensors, and Their Operations
b
S
n
t
O
Fig. A.2. Intrinsic triad along a curve
Now, let the differential distances from O along the directions of n and b
be dn and db, respectively. Then
∇ = t

∂s
+ n


∂n
+ b

∂b
, (A.40)
which involves curves along n and b directions that have their own curvature
and torsion. Then one might apply the Frenet–Serret formulas to these curves
as well to complete the gradient operation. But due to the Dupin theorem we
prefer to leave the two curves orthogonal to C undetermined.
For example, if C is a streamline such that u = qt, then the continuity
equation for incompressible flow reads
∇·u =
∂q
∂s
+ q∇·t =0, (A.41)
where, by using (A.39),
∇·t = n ·
∂t
∂n
+ b
∂t
∂b
. (A.42)
Similarly, there is
∇×t = κb +

n ×
∂t
∂n
+ b ×

∂t
∂b

.
Here, since |t| = 1, it follows that:
n ·

n ×
∂t
∂n
+ b ×
∂t
∂b

= t ·
∂t
∂b
=
1
2

∂b
|t|
2
=0,
b ·

n ×
∂t
∂n

+ b ×
∂t
∂b

= −t ·
∂t
∂n
= −
1
2

∂n
|t|
2
=0.
A.3 Curvilinear Frames on Lines and Surfaces 707
Therefore, the second term of ∇×t must be along the t direction, with the
magnitude
ξ ≡ t · (∇×t)=b ·
∂t
∂n
− n ·
∂t
∂b
. (A.43)
The scalar ξ is known as the torsion of neighboring vector lines (Truesdell
1954). Thus, using this notation we obtain
∇×t = ξt + κb. (A.44)
This result enables us to derive the vorticity expression in the streamline
intrinsic frame

ω = ∇×(qt)=∇q × t + q∇×t = ∇q × t + ξqt + κqb.
Thefirsttermofis
∇q × t =
∂q
∂b
n −
∂q
∂n
b,
so we obtain (Serrin 1959)
ω = ξqt +
∂q
∂b
n +

κq −
∂q
∂n

b. (A.45)
Thus, ξ0ifω · u = 0. Note that in a three-dimensional orthonormal frame
there must be ξ ≡ 0, so by the Dupin theorem a curve with ξ = 0 cannot be
the principal curvature line of any orthogonally intersecting surfaces.
A.3.2 Intrinsic operation with surface frame
Derivatives of tensors along a curved surface S can be made simple by an
intrinsic use of an intrinsic surface frame, which is more complicated than the
intrinsic line frame since now there are two independent tangential directions
on S.
Covariant Frame
At a given time, a two-dimensional surface S in a three-dimensional space is

described by the position vector x of all points on S, which is a function of
two independent variables, say u
α
with α =1, 2. Then
r
α
(u
1
,u
2
) ≡
∂x
∂u
α
,α=1, 2, (A.46)
define two nonparallel tangent vectors (not necessarily orthonormal) at each
point x ∈ S, see Fig. A.3. Note that by convention when an upper index
appears in the denominator it implies a lower index in the numerator, and
708 A Vectors, Tensors, and Their Operations
n
r
1
r
2
r
1
r
2
Fig. A.3. Covariant and contravariant frames on a surface
vise versa. r

α
areusedasthecovariant tangent basis vectors characterized
by lower indices. Their inner products
g
αβ
≡ r
α
· r
β
,α,β=1, 2, (A.47)
form a 2 × 2 matrix which gives the covariant components of the so-called
metric tensor and completely determine the feature of r
α
. Note that the area
of the parallelogram spanned by r
1
and r
2
is |r
1
× r
2
| =

g = det{g
αβ
}.
From r
α
one can obtain the unit normal vector

n(u
1
,u
2
)=
r
1
× r
2
|r
1
× r
2
|
=
r
1
× r
2

g
. (A.48)
Then the set (r
1
, r
2
, n) form a covariant right-handed surface frame. We stress
that only n is the intrinsic feature of the surface, but r
α
created by u

α
are ar-
tificially chosen. If we introduce a variable u
3
along the n direction, then since
for the given surface r
α
and n depends only on u
1
,u
2
, there is ∂r
α
/∂u
3
=0
and ∂n/∂u
3
=0.
Contravariant Frame
A vector f can be decomposed in terms of (r
α
, n)
f · r
α
= f
α
, f · n = f
3
.

However, to recover the vector from its components, one does not have
f = r
α
f
α
+ f
3
n, since then f · r
β
= f
β
would equal r
α
· r
β
f
α
= g
αβ
f
α
,
which is possible only if g
αβ
= δ
αβ
, but in general this is not the case. Thus,
a covariant frame alone is insufficient; one has to construct another frame
conjugate to it. This is similar to the situation in complex domain, where
a complex basis-vector set, say a

i
, needs be complemented by its complex
conjugate a

i
with a
i
· a

j
= δ
ij
, such that if f · a
i
= f
i
then there should
be f = f
i
a

i
.
A.3 Curvilinear Frames on Lines and Surfaces 709
We therefore introduce the second pair of tangent vectors r
α
with upper
index, by requiring the hybrid components of the metric tensor be a unit
matrix
r

α
· r
β
≡ g
α
β
= δ
α
β
. (A.49)
To this end, we set (see Fig. A.3)
r
1
=
r
2
× n

g
, r
2
=
n ×r
1

g
. (A.50)
Then obviously r
α
·r

β
=0forα = β, and by (A.48) r
α
·r
β
=1forα = β as
desired. Moreover, by using identity
(a ×b) ×(c ×d)=b[a ·(c ×d)] −a[b ·(c ×d)] (A.51)
there is
r
1
× r
2
=
1
g
(r
2
× n) ×(n ×r
1
)=
1
g
n[n ·(r
1
× r
2
)] =
n


g
,
so we have
n =

gr
1
× r
2
, |r
1
× r
2
| =
1

g
. (A.52)
Thus, we have a contravariant right-handed frame (r
1
, r
2
, n). Similar to
(A.46), r
α
can be created by a pair of covariant variables u
1
,u
2
:

r
α
(u
1
,u
2
)=
∂x
∂u
α
,α=1, 2. (A.53)
With
g
αβ
= r
α
· r
β
,α,β=1, 2, (A.54)
being the contravariant components of the metric tensor. Substituting (A.50)
into (A.54), and using (A.47) and identity
(a ×b) ·(c ×d)=(a ·c)(b ·d) −(a ·d)(b ·c), (A.55)
we find
g
11
=
g
11
g
,g

22
=
g
22
g
,g
12
= g
21
= −
g
12
g
.
Namely, g
αβ
are the cofactors of g
αβ
, or the two matrices are inverse of each
other:
g
αβ
g
βγ
= δ
α
γ
. (A.56)
Since in both frames n is the same and normal to any tangent basis vectors,
the covariant and contravariant component of a vector f along n coincide and

will be denoted by f
n
. We thus write
f = r
α
f
α
+ f
n
n = r
β
f
β
+ f
n
n. (A.57)
710 A Vectors, Tensors, and Their Operations
This time the result is correct
f · r
γ
= f
α
δ
γ
α
= f
α
, f · r
γ
= f

β
δ
β
γ
= f
γ
.
Here we see a hybrid use of the two frames. Accordingly, the summation
convention over the repeated index is implied only if one index is upper and
the other is lower.
Moreover, by (A.57) we have
f
α
= r
α
· f = r
α
· r
β
f
β
= g
αβ
f
β
,
f
β
= r
β

· f = r
β
· r
α
f
α
= g
βα
f
α
.
(A.58)
Thus, raising or lowering the index of a component can be achieved by using
the metric tensor g
αβ
or g
αβ
, respectively. Actually, in (A.56) we have done
this for the metric tensor itself.
A tensor of higher rank can be similarly decomposed by its repeated inner
products with basis vectors. Depending on the choice of basis vectors, the
components of a tensor of rank 2 can be covariant, contravariant, or hybrid.
For example, if a tensor T has only tangent component, we can write
T = T
αβ
r
α
r
β
= T

·β
α·
r
α
r
β
= T
α·
·β
r
α
r
β
= T
αβ
r
α
r
β
,
where the dots indicate the order of components which cannot be exchanged
unless the tensor is symmetric with respect to relevant indices.
Tangent Derivatives of a Vector
The most important differential operations on a surface S is the tangent deriv-
atives of vectors and tensors. We split the gradient vector into
∇ = ∇
π
+ n

∂n

, ∇
π
(·)=r
α
∂(·)
∂u
α
= r
α
(·)

, (A.59)
where a hybrid use has been made: r
α
is not created from u
α
but from u
α
.
Now, for the tangent derivatives of a vector f , there is

π
f = r
α
(r
β
f
β
+ f
n

n)

= r
α
(r
β

f
β
+ r
β
f
β,α
+ n

f
n
+ nf
n,α
).
Here, the derivatives of f
β
and f
n
are merely that of numbers; the key is the
derivatives of basis vectors r
α
, r
β
,andn. These vectors are not coplaner, and

hence their derivatives with respect to u
α
can be expressed by their linear
combinations. The result is given by the classic Gauss formulas
r
β,α
= Γ
λ
αβ
r
λ
+ b
αβ
n, r
β

= −Γ
β
αλ
r
λ
+ b
β
α
n,Γ
λ
αβ
= Γ
λ
βα

, (A.60a,b,c)
A.3 Curvilinear Frames on Lines and Surfaces 711
and Weingarten formulas
n

= −b
β
α
r
β
= −b
αβ
r
β
,b
αβ
= b
βα
= b
γ
β
g
αγ
. (A.61a,b)
The six coefficients Γ
λ
αβ
= Γ
λ
βα

in (A.60) are called Christoffel symbols of the
second kind. They do not form a tensor since r
α
are artificially chosen. In
contrast, three coefficients b
αβ
= g
γα
b
α
β
= b
βα
are covariant components of
the symmetric curvature tensor, defined by the intrinsic operation
K ≡−∇
π
n = −r
α
n

= r
α
r
β
b
β
α
= r
α

r
β
b
αβ
. (A.62)
Intrinsic Operation on Surface
In view of the complicated involvement of nontensorial Γ
λ
αβ
in the derivatives
of r
α
or r
β
, it is desired to bypass these operations. We call this kind of oper-
ation intrinsic operation. To start, we look at some more intrinsic properties
of the normal vector n.First,themean curvature κ is defined as half of the
two-dimensional divergence
2κ ≡−∇
π
· n = b
β
α
δ
α
β
= g
αβ
b
αβ

= b
α
α
. (A.63)
This result directly comes from the contraction of (A.62. Second, by (A.61a,b)
there is

π
× n = r
α
× n

= −b
αβ
r
α
× r
β
= 0. (A.64)
Finally, (A.61a) implies
n ·n

=0. (A.65)
From the derivation of (A.64 we can already see the main feature of the
intrinsic operation: (1) The operation is in vector form rather than component
form; and (2) the tangent basis vectors are carried along in the operation but
their derivatives do not appear. The final result will be automatically free from
these vectors. The following examples further show the strategy.
First, we compute the jump of normal vorticity across a vortex sheet
[[ ω

n
]] = n ·[[ ∇×u]] = ( n ×∇) ·(n ×γ),
where γ is the sheet strength. Using the surface frame, we have
[[ ω
n
]] = ( n × r
α
) ·(n ×γ

+ n

× γ),
which amounts to common vector algebra. Thus, by (A.55),
[[ ω
n
]] = ( n · n)(r
α
· γ

) −(n ·γ

)(r
α
· n)
+(n ·n

)(r
α
· γ) −(n ·γ)(r
α

· n

),
712 A Vectors, Tensors, and Their Operations
where n · γ =0,r
α
· n = 0, and we have (A.65). Thus, we simply obtain
(Saffman 1992; where no proof is given)
[[ ω
n
]] = r
α
· γ

= ∇
π
· γ. (A.66)
Second, we show that for any vectors A and B satisfying n · A =0and
B = n × A, there is

π
· A =(n ×∇) ·B = n · (∇×B). (A.67)
Indeed, since n ×∇= n ×∇
π
,wehave
(n ×∇) ·B = n ·(∇
π
× B)=n · [r
α
× (n


× A + n × A

)]
= n · n

(r
α
· A) −n ·A(r
α
· n

)+r
α
· A

,
in which the first two terms vanish due to (A.65) and assumed feature of A.
The last term gives (A.67).
Third, (n ×∇) ×A is a tangent-derivative operation and we develop it to
several fundamental constituents. There is
(n ×∇) ×A =(n ×r
α
) ×A

= r
α
(A

· n) −n(r

α
· A

)
=(∇
π
A) ·n −n∇
π
· A.
(A.68)
The second term comes from the tangent divergence of A, while the first term
must be related to the normal component of A; but even if A
n
= 0, the surface
curvature may cause ∇
π
A to have normal components. In fact, we can further
split the first term of (A.68) to
(∇
π
A) ·n = ∇
π
(A ·n) −A ·∇
π
n = ∇
π
A
n
+ A · K,
where K has only tangent components. Thus, it follows that

(n ×∇) ×A = ∇
π
A
n
+ A
π
· K − n(∇
π
· A). (A.69)
In all the above examples r
α
are merely a temporary scaffold for the
convenience of operation. However, if the final result is to be written in terms
of components, then the derivatives of r
α
become inevitable. For example, in
(A.69) and (A.68) we have

π
· A =
1

g
(

gA
α
)

, n ·(∇

π
× B)=
1

g
(B
2,1
− B
1,2
), (A.70a,b)
of which the proof requires the knowledge of Γ
λ
αβ
that is beyond our present
concern.
A.3 Curvilinear Frames on Lines and Surfaces 713
Orthonormal Surface Frame
If in the above surface-moving frame the covariant tangent basis vectors r
1
and r
2
are already orthogonal, then by (A.50) there is r
α
= r
α
/

g, implying
that there is no need to distinguish covariant and contravariant tangent basis
vectors. Thus, like Cartesian tensors, it suffices to use one pair of tangent

basis vectors and denote the components by subscripts only.
The orthogonality of basis vector implies that
g
ii
= h
2
i
δ
ii
(no summation with respect to i),
where
h
i
=

g
ii
=

r
i
· r
i
(no summation with respect to i) (A.71)
is the length of r
i
, called scale coefficients or Lam´e coefficients. They are still
functions of the moving point x. Nevertheless, we can now introduce a set of
orthonormal basis vectors
e

α
=
r
α
h
α
(no summation with respect to α), e
3
= n, (A.72)
which form an orthonormal surface frame. For a given curved surface S, h
α
depends on its intrinsic geometric feature as well as the orientation of r
α
,the
latter can be arbitrarily chosen, not restricted by the Dupin theorem as long
as the frame is only defined on a single surface S.
With the e
i
frame and using the notation

i
=
1
h
i

∂x
i
,i=1, 2, 3, (A.73)
the Gauss formulas (A.60) and Weingarten formulas (A.61) can be cast to


1
e
1
= −
h
1,2
h
1
h
2
e
2
+ b
11
e
3

1
e
2
=
h
1,2
h
1
h
2
e
1

+ b
12
e
3

1
e
3
= −b
11
e
1
− b
12
e
2

2
e
1
=
h
2,1
h
1
h
2
e
2
+ b

12
e
3

2
e
2
= −
h
2,1
h
1
h
2
e
1
+ b
22
e
3

2
e
3
= −b
12
e
1
− b
22

e
2




































, (A.74)
and, of course,

i
e
3
=0,i=1, 2, 3.
714 A Vectors, Tensors, and Their Operations
Here, b
αβ
= b
βα
(α, β =1, 2) are redefined by
b
αβ
= e
α
· K · e
β
or K = b
αβ
e
α

e
β
(A.75)
instead of (A.62). From b
αβ
one can construct two intrinsic scalar curvatures
which describe the wall geometry and are independent of the choice of e
1
and
e
2
. One is
b
11
+ b
22
= κ
B
= −∇
π
· n, (A.76)
twice of the mean curvature as already seen in (A.63); the other is
b
11
b
22
− b
2
12
= det{b

αβ
}≡K, (A.77)
the total curvature. Moreover, except some isolated points a curved surface
has a pair of orthogonal principal directions.Ife
1
and e
2
coincide with these
directions then b
αβ
=0forα = β. In this case b
11
= K
1
and b
22
= K
2
are the principal curvatures, which are the greatest and least b
αβ
among all
orientations of the tangent vectors. The total curvature is simply K = K
1
K
2
.
On a sphere any tangent direction is a principal direction.
It is convenient to express b
αβ
by principal curvatures since the latter are

independent of the choice of (e
1
, e
2
). Denote the unit tangent vectors along
the principal directions by p
1
and p
2
(they define the curvature lines of the
surface), then by (A.75) there is
K = p
1
p
1
K
1
+ p
2
p
2
K
2
.
Thus, if (p
1
, p
2
, e
3

) form a right-handed frame and β is the angle by which
the (p
1
, p
2
) pair rotates to the (e
1
, e
2
) pair in counterclockwise sense, there
is
b
11
= K
1
cos
2
β + K
2
sin
2
β, (A.78a)
b
12
= −
1
2
(K
1
− K

2
)sin2β, (A.78b)
b
22
= K
1
sin
2
β + K
2
cos
2
β. (A.78c)
Hence, for a given surface, b
αβ
depend solely on a single parameter β. Because
reversing the direction of (p
1
, p
2
) does not affect K, without loss of generality
we set β ∈ [0, π/2].
On the other hand
κ
1
=(∂
1
e
1
) ·e

2
= −(∂
1
e
2
) ·e
1
= −
h
1,2
h
1
h
2
, (A.79a)
κ
2
=(∂
2
e
2
) ·e
1
= −(∂
2
e
1
) ·e
2
= −

h
2,1
h
1
h
2
(A.79b)
define a pair of on-surface curvatures of coordinate lines x
1
and x
2
, respec-
tively (they are the geometric curvatures of these lines if K = 0). Therefore,
A.3 Curvilinear Frames on Lines and Surfaces 715
(A.74) can be written in a geometrically clearer form:

1
e
1
= κ
1
e
2
+ b
11
e
3

1
e

2
= −κ
1
e
1
+ b
12
e
3

1
e
3
= −b
11
e
1
− b
12
e
2

2
e
1
= −κ
2
e
2
+ b

12
e
3

2
e
2
= κ
2
e
1
+ b
22
e
3

2
e
3
= −b
12
e
1
− b
22
e
2
























(A.80)
Note that since in (A.71) h
α
(α =1, 2) are functions of x
1
and x
2
, operators

α

and ∂
β
for α = β are not commutative. Instead, there is
(∂
1
− κ
2
)∂
2
=(∂
2
− κ
1
)∂
1
or ∂
1

2
− ∂
2

1
= κ
1

1
− κ
2


2
, (A.81)
but we still have ∂
α

3
= ∂
3

α
.
The orthogonality of x
1
-lines and x
2
-lines implies that the variation of κ
1
and κ
2
are not independent. In fact, by (A.81) and using (A.80), there is

2
κ
1
= ∂
2
[(∂
1
e
1

) ·e
2
]
=(∂
1

2
e
1
) ·e
2
+(∂
1
e
1
) ·(∂
2
e
2
)+κ
2
1
+ κ
2
2
= −∂
1
κ
2
+ b

11
b
22
− b
2
12
+ κ
2
1
+ κ
2
2
,
from which it follows a differential identity

2
κ
1
+ ∂
1
κ
2
= κ
2
1
+ κ
2
2
+ K. (A.82)
Finally, by using (A.80), the components of the gradient of any vector

f = e
i
f
i
in the (x
1
,x
2
,x
3
)-frame read
(∂
i
f) ·e
j
=



c
11
c
12
c
13
c
21
c
22
c

23

3
f
1

3
f
2

3
f
3



, (A.83a)
where
c
11
= ∂
1
f
1
− κ
1
f
2
− b
11

f
3
c
12
= ∂
1
f
2
+ κ
1
f
1
− b
12
f
3
c
13
= ∂
1
f
3
+ b
11
f
1
+ b
12
f
2

c
21
= ∂
2
f
1
+ κ
2
f
2
− b
12
f
3
c
22
= ∂
2
f
2
− κ
2
f
1
− b
22
f
3
c
23

= ∂
2
f
3
+ b
12
f
1
+ b
22
f
2
























. (A.83b)
Thus, b
αβ
and κ
α
appear in all tangent derivatives.
716 A Vectors, Tensors, and Their Operations
A.4 Applications in Lagrangian Description
Tensor analysis is necessary in studying the transformation of physical quan-
tities and equations between the physical space and the reference space (in
Lagrangian description). In this section, we present some materials which are
directly cited in the main text.
A.4.1 Deformation Gradient Tensor and its Inverse
Consider the deformation gradient tensor and associated Jacobian in the ref-
erence space, defined by (2.3) and (2.4)
F = ∇
X
x or F
αi
= x
i,α
, (A.84)
J =
∂(x
1

,x
2
,x
3
)
∂(X
1
,X
2
,X
3
)
= detF. (A.85)
First, J has a few explicit forms exactly the same as the determinant of a
matrix
J = 
αβγ
x
1,α
x
2,β
x
3,γ
, (A.86a)

ijk
J = 
αβγ
x
i,α

x
j,β
x
k,γ
. (A.86b)
Next, keeping the labels of particles, any variation of J can only be caused by
that of x. Using (A.86), an infinitesimal change of J is then given by
δJ = 
αβγ
(δx
1,α
x
2,β
x
3,γ
+ x
1,α
δx
2,β
x
3,γ
+ x
1,α
x
2,β
δx
3,γ
)
= 
αβγ

x
1,α
x
2,β
x
3,γ
(δx
1,1
x
2,2
x
3,3
+ x
1,1
δx
2,2
x
3,3
+ x
1,1
x
2,2
δx
3,3
)
= Jδx
l,l
.
Namely,
δJ = J∇·δx. (A.87)

Then, owing to (2.8), (2.3) is invertible, and hence F has inverse tensor
defined in the physical space
F
−1
= ∇X, or F
−1

= X
α,i
, (A.88)
which satisfies F ·F
−1
= F
−1
· F = I, i.e.,
x
i,α
X
β,i
= δ
αβ
,X
α,i
x
j,α
= δ
ij
. (A.89)
The Jacobian of F
−1

is
J
−1
= detF
−1
=
∂(X
1
,X
2
,X
3
)
∂(x
1
,x
2
,x
3
)
=
1
J
, (A.90)
A.4 Applications in Lagrangian Description 717
which has expressions symmetrical to (A.86a,b). For example, we have

ijk
J
−1

= 
αβγ
X
α,i
X
β,j
X
γ,k
. (A.91)
Similar to (A.87), there is
δJ
−1
= J
−1
δX
α,α
= J
−1
δX
α,i
x
i,α
= J
−1

X
· δX. (A.92)
The deformation tensor F and its inverse F
−1
have one index in physical

space and one in the reference space. This property can be used to transform
a physical-space vector to its dual or image in the reference space, or vise
versa.
A.4.2 Images of Physical Vectors in Reference Space
We are particularly interested in seeking the image of vorticity ω in the X-
space. To this end consider a general vector f first. Multiply both sides of
(A.86b) by f
k,j
/J and notice that
f
k,j
x
j,β
= f
k,β
,
αβγ
x
k,βγ
=0.
It follows that:

ijk
f
k,j
=
1
J

αβγ

x
i,α
x
k,γ
f
k,β
=
1
J

αβγ
(x
k,γ
f
k
)

x
i,α
,
of which the vector form is
∇×f =
1
J
[∇
X
× (F ·f)] ·F. (A.93)
Taking inner product of this result with F
−1
, the inverse of (A.93) reads


X
× (F ·f)=J(∇×f) · F
−1
. (A.94)
The mapping between ∇×f and ∇
X
× (F · f) is one-to-one, and has three
features: (1) they are identical at t = τ = 0; (2) if one vanishes, so must the
other; and (3) they are divergenceless in their respective spaces. We therefore
identify the latter as the image of the former in X-space. Besides, for two-
dimensional vector field f =(f
1
,f
2
, 0) it can be shown that (A.94) is simplified
to

X
× (F ·f)=J∇×f. (A.95)
Note that the above identification naturally leads us to define the image of f
itself in the X-space as F · f .
Now let f = u = ∂x/∂τ be the velocity field. Its image in X-space is
U
α
= x
i,α
u
i
or U = F · u. (A.96)

718 A Vectors, Tensors, and Their Operations
Then by (A.94), the image of ω is
Ω ≡∇
X
× U, (A.97)
which we call the Lagrangian vorticity. By using (A.89) and mass conservation
(2.40), the transformation between ω and Ω is given by
Ω = Jω ·F
−1
= ω
ρ
0
ρ
· F
−1
, (A.98a)
ω =
1
J
Ω · F =
ρ
ρ
0
Ω · F. (A.98b)
For two-dimensional flow there is
Ω = Jω =
ρ
0
ρ
ω (A.99)

and, if in addition the flow is incompressible we simply have Ω = ω.
Note that the images of u and ω have opposite structures. In the former F
is used but in the latter, in addition to the factor J,itisF
−1
. This is related
to the fact that u is a true vector (polar vector) but ω is a pseudo-vector
(axial vector). One might switch the use of F and F
−1
for true and pseudo
vectors; which however does not lead to useful result unless U is to be written
as the curl of another vector, which will be used once below.
Once we introduced the Lagrangian vorticity, we may study vorticity kine-
matics in physical space by that of Ω in reference space. First, since

∂τ
(x
k,γ
u
k
)=x
k,γ
a
k
+
1
2
q
2

,

where
A
α
≡ x
i,α
a
i
= A
α
or bF · a = A (A.100)
is the X-space image of acceleration, there is
∂U
∂τ
= A + ∇
X

1
2
q
2

. (A.101)
This is the X-space image of a =Du/Dt. Its curl gives an elegant equation
at once
∂Ω
∂τ
= ∇
X
× A, (A.102a)
where by A.94,


X
× A = J(∇×a) ·F
−1
(A.102b)
is the image of ∇×a. Therefore, the rate of change of the image of the curl of
velocity (Lagrangian vorticity) equals the image of the curl of acceleration. We
stress that this result is not true for Lagrangian velocity U and acceleration
A.4 Applications in Lagrangian Description 719
A: in (A.101) there is an extra gradient term. The implication of this difference
has been made clear in Sect. 3.6.
As an application of utilizing the vector images in reference space, we
extend the content of Sect. 3.6.2 by showing that the circulation preserving is
sufficient but not necessary for having a Bernoulli integral. Rewrite (A.101)
as
∂U
∂τ
= A

+ ∇
X
Ψ, (A.103)
where Ψ is defined in (3.152) and A

is the rotational part of A. Similar to
the approach leading to (3.148), assume there exists a family of material sur-
faces defined by two parameters, say ξ
1
(X)andξ
2

(X), such that the normal

X
ξ
1
×∇
X
ξ
2
is along A

A

× (∇
X
ξ
1
×∇
X
ξ
2
)=0. (A.104)
Then a Bernoulli integral like (3.155a) along these surfaces can be obtained.
An important example is inviscid baroclinic flow without shock waves, for
which Ds/Dt =0ors = s(X)intheX-space. Hence, any material sur-
faces are isotropic. Note that the existence of these surfaces is ensured by
the Crocco–Vazsonyi equation (2.163), in which the rotational term T ∇s is a
complex-lamellar vector field according to the definition in Sect. 3.3.1. Then,
by using (A.96) there is
A


(X,τ)=F · T ∇s,
which is normal to isentropic surfaces. Therefore, on these surfaces we have
Bernoulli integral (3.155). Correspondingly, those conservation theorems
involves material integrals can survive along these surfaces. For example,
instead of volume integrals (3.132) and (3.134) we may have a similar sur-
face integrals; and the Kelvin circulation theorem (3.130c) will hold if the
loop C is on one of such surfaces.
References
Abramowitz M., Stegun A. (1972) Handbook of Mathematical Functions.
Dover, Mineola New York.
Acarlar M.S., Smith C.R. (1987a) A study of hairpin vortices in a laminar
boundary layer. Part.1, Hairpin vortices generated by a hemisphere protu-
berance. J. Fluid Mech. 175: 1–41.
Acarlar M.S., Smith C.R. (1987b) A study of hairpin vortices in a laminar
boundary layer. Part.2, Hairpin vortices generated by fluid injection. J.
Fluid Mech. 175: 43–83.
Adem J. (1956) A series solution for the barotropic vorticity equation and its
application in the study of atmospheric vortices. Tel lus, 8: 364–372.
Adrian, R.J., Meinhart, C.D., Tomkins, C.D. (2000) Vortex organization in
the outer region of the turbulent boundary layer. J. Fluid Mech. 422: 1–54.
Ahlborn B., Lefrancois M., King D.H. (1998) The clockwork of vortex shed-
ding. Phys. Essays, 11: 144–154.
Ahlborn B., Seto M.L., Noack B.R. (2002) On drag, Strouhal number and
vortex-street structure. Fluid Dyn. Res. 30: 379–399.
Althaus W., Brucker C., Weimer M. (1995) Breakdown of slender vortices. In:
Green S.I. (ed.) Fluid Vortices, Chap. IX, Kluwer, Dordrecht Netherlands
317–372.
Anderson C.R., Reider M.B. (1996) A high order explicit method for the
computation of flow about a circular cylinder. J. Comp. Phys. 125: 207–

224.
Anderson J.D. Jr. (1991) Fundamentals of Aerodynamics. McGraw-Hill, New
York.
Antkowiak A., Brancher P. (2004) Transient energy growth for the lamb-oseen
vortex. Phys. Fluids, 16(1): L1–L4.
Aref H. (1980) Vortex dynamics of the two-dimensional turbulent shear layer.
J. Fluid Mech. 100: 705.
Aref H. (1983) Integrable, chaotic and turbulent vortex motion in two-
dimensional flows. Ann. Rev. Fluid Mech. 15: 345–389.
722 References
Aref H., Flinchem E.P. (1984) Dynamics of a vortex filament in a shear flow.
J. Fluid Mech. 148: 477–497.
Aref H., Pomphrey N. (1982) Integratable and chaotic motions of four vortices.
Phys. Lett. 78A(4): 297–300.
Aref H., Zawadzki I. (1990) Comment of vortex ring reconnections. In: Moffatt
H.K., Trinsber A. (eds.) Topological Fluid Mechnics, Proc. IUTAM Symp.
Combridge, Cambridge University Press Cambridge, 535–539.
Aref H., Zawadzki I. (1991) Linking of vortex rings. Nature, 354: 50–53.
Aris R. (1962) Vectors, Tensors and the Basic Equations of Fluid Mechanics.
Dover, Mineala New York.
Arnold V.I. (1965a) Sur la totologie des ´ecoulments stationaires des fluid par-
faits. C. R. Acad. Paries A261: 17–20.
Arnold V.I. (1965b) Conditions for nonlinear stability of stationary plane
curvilinear flows of an ideal fluid, Dokl Akad. Nauk SSSR, 162: 975–978.
(Soviet Math. 6: 773–777.)
Arnold V.I. (1965c) Variational principle for three-dimensional steady-state
flows of an ideal fluid. Prikl Mat. Mekh. 29: 846–851.
Arnold V.I. (1969) On an apriori estimate in the theory of hydrodynamical
stability. Amer. Math. Soc. Transl. 19: 267–269.
Arnold V.I. (1989) Mathematical Methods in Classical Mechanics. Springer-

Verlag, Berlin Heidelberg New York.
Arrese J.C., Fernandez-Feria R. (1996) Nuerical solution for the boundary
layer induced by a conical vortex. Proceedings of the CFD96, Ottawa, On-
tario, June 2–6, 1996.
Asai M., Nishioka M. (1995) Boundary-layer transition triggered by hairpin
eddies at subcritical Reynolds numbers. J. Fluid Mech. 297: 101–122.
Asai M., Nishioka M. (1997) Development of wall turbulence structure in
transitional flows. In: Tatsumi T., Watanabe E., Kambe T. (eds.) Theoret-
ical and Applied Mechanics, 19th IUTAM Congress, 1996, Elsevier Kyoto,
Japan, 121–138.
Asai M., Sawada K., Nishioka M. (1996) Development of turbulent patch in
a subcritical boundary-layer transition. Fluid Dyn. Res. 18: 151–164.
Ash R.L., Khorrami M.R. (1995) Vortex stability. In: Green S.I. (ed.) Fluid
Vortices, Chap. VIII, Kluwer, Dordrecht Netherlands 317–372.
Ashley H., Landahl M. (1965) Aerodynamics of Wings and Bodies,Dover
Mineola New York.
Ashurst W.T., Meiburg E. (1988) Three-dimensional shear layer via vortex
dynamics. J. Fluid Mech. 189: 87–111.
Baggett J.S., Trefethen L.N. (1997) Low-dimensional models of subcritical
transition to turbulence. Phys. Fluids, 9(4): 1043–1053.
Baker G.R., Meiron D.I., Orszag S.A. (1982) Generalized vortex methods for
free-surface flow problems. J. Fluid Mech. 123: 477–501.
Baker S.J., Crow S.C. (1977) The motion of two-dimensional vortex pairs in
a ground effect. J. Fluid Mech. 82: 659–671.
References 723
Bakker P.G. (1991) Bifurcations in flow patterns. Kluwer, Dordrecht Nether-
lands.
Barkley D., Henderson B.D. (1996) Three-dimensional floquet stability analy-
sis of the wake of circular cylinder. J. Fluid Mech. 322: 215–241.
Baron A., Quadrio M. (1996) Turbulent drag reduction by spanwise wall

oscillations. Appl. Sci. Res. 55: 311–326.
Basdevant C., Philipovich T. (1994) On the validity of the “Weiss criterion”
in two-dimensional turbulence. Physica D, 73: 17–30.
Batchelor G.K. (1956a) Unsteady laminar flow with closed streamlines at large
Reynolds number. J. Fluid Mech. 1: 177–190.
Batchelor G.K. (1956b) A proposal concerning laminar wakes behind bluff
bodies at large Reynolds number. J. Fluid Mech. 1: 388–398.
Batchelor G.K. (1964) Axial flow in trailing line vortices. J. Fluid Mech. 20:
645–658.
Batchelor G.K. (1967) An Introduction to Fluid Mechanics. Cambridge Uni-
versity Press, Cambridge, UK.
Batchelor G.K. (1969) Computation of the energy spectrum in homogeneous
two-dimensional turbulence. Phys. Fluids Suppl. II 12: 233–239.
Bayly B.J. (1986) Three-dimensional instability of strained vortices in a vis-
cous fluid. Phys. Rev. Lett. 57(17): 2160–2163.
Beale J.T., Greengard C. (1992) Convergence of Euler-Stokes splitting of the
Navier-Stokes equations. Research Report RC 18072, IBM.
Beale J.T., Kato T., Majda A. (1984) Remarks on the breakdown of smooth
solution for the 3-D Euler equations. Comm. Math. Phys. 94: 61–66.
Beckers M., Vlercx H.J.H., van Heijst G.J.F., Verzicco R. (2003) Evolution
and instability of monopolar vortices in a stratified fluid. Phys. Fluids, 15:
1033–1045.
Bellamy-Knights P.G. (1970) An unsteady two-celled vortex solution of the
Navier-Stokes equations. J. Fluid Mech. 41: 673–687.
Bengtsson L., Lighthill J. (eds.) (1982) Intense Atmospheric Vortices,
Springer-Verlag, Berlin Heidelberg New York.
Benjamin T.B. (1962) Theory of the vortex breakdown phenomenon. J. Fluid
Mech. 14: 593–629.
Beran P.S., Culick F.E.C. (1992) The role of nonuniqueness in the develop-
ment of vortex breakdown in tubes. J. Fluid Mech. 242: 491–527.

Berestov A.L. (1979) Solitary rossby waves. Izv. Acad. Sci. USSR Atmos.
Oceanic Phys. 15: 443–447.
Bernal L.P., Roshko A. (1986) Streamwise vortex structure in plane mixing
layers, J. Fluid Mech. 170: 499–525.
Bernard P.S., Thomas J.M., Handler R.A. (1993) Vortex dynamics and the
production of Reynolds stress. J.Fluid Mech. 253: 385–419.
Bers A. (1975) Linear waves and instabilities. In: DeWitt C., Peyraud J. (eds.)
Physique des Plasmas, Gordon & Breach, New York, 117–215.

×