You are on page 1of 20

Proc. R. Soc.

A (2012) 468, 3745–3764


doi:10.1098/rspa.2012.0335
Published online 8 August 2012

Cauchy integral formula for generalized analytic


functions in hydrodynamics
BY MICHAEL ZABARANKIN*
Department of Mathematical Sciences, Stevens Institute of Technology,
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

Castle Point on Hudson, Hoboken, NJ 07030, USA

It is shown that for several classes of generalized analytic functions arising in linearized
equations of hydrodynamics and magnetohydrodynamics, the Cauchy integral formulae
follow from the one for generalized holomorphic vectors in a uniform fashion. If
hydrodynamic fields (velocity, pressure and vorticity) admit representations in terms
of corresponding generalized analytic functions, those representations and the Cauchy
integral formulae form two essential parts of the generalized analytic function approach,
which readily yields either closed-form solutions or boundary integral equations. This
approach is demonstrated for problems of axisymmetric and asymmetric Stokes flows,
two-phase axisymmetric Stokes flows, two-dimensional and axisymmetric Oseen flows.
Keywords: generalized analytic function; Cauchy integral formula; Stokes flow; Oseen flow

1. Introduction

(a) Vector fields in linear hydrodynamics


A vector field U = U(x) and a scalar field J = J(x) related by
curl U + 2[a × U] = −VJ, div U = 0, (1.1)
where a is a constant real-valued vector and x is the position vector (multiplier 2
is introduced for convenience), arise in linearized equations of hydrodynamics
and magnetohydrodynamics (MHD) (Happel & Brenner 1983; Zabarankin &
Krokhmal 2007; Zabarankin 2010, 2011a). For a = 0, (1.1) is known as the
Moisil–Theodorescu system (Moisil & Theodorescu 1931; Bitsadze 1969), whereas
for J = 0 and a = 0, (1.1) simplifies to the classical potential flow equations
(Mises 1944; Bitsadze 1969).
Example 1.1 (Ideal fluid). The velocity field u of an ideal fluid is irrotational
and incompressible (solenoidal), i.e.
curl u = 0, div u = 0, (1.2)
which corresponds to (1.1) with U = u, J ≡ 0 and a = 0.
Example 1.2 (Stokes flows). Under the assumption of negligible inertial and
thermal effects, the time-independent velocity field u of a viscous incompressible
*mzabaran@stevens.edu

Received 1 June 2012


Accepted 13 July 2012 3745 This journal is © 2012 The Royal Society
3746 M. Zabarankin

fluid is governed by the Stokes equations


mDu = Vp, div u = 0, (1.3)
where p is the pressure in the fluid, m is shear viscosity and Du = V(div u) −
curl(curl u) (Happel & Brenner 1983). The Stokes equations (1.3) imply that the
vorticity u = curl u and pressure p are related by
m curl u = −Vp, div u = 0, (1.4)
which corresponds to (1.1) with J = p, U = mu and a = 0.
Example 1.3 (Oseen flows). Suppose a solid body translates with constant
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

velocity v in a quiescent viscous incompressible fluid. If the Reynolds number is


sufficiently small, the time-independent velocity field u with partially accounted
inertial effects can be described by the Oseen equations
m Du + r(v · V)u = Vp, div u = 0, (1.5)
where p is the pressure, and m and r are fluid shear viscosity and density,
respectively (Happel & Brenner 1983). Let v · u = 0 with u = curl u. Then the
Oseen equations (1.5) can be recast in two equivalent forms:
curl (mu + r[v × u]) = −Vp, div (mu + r[v × u]) = 0 (1.6)
and
m curl u + r[v × u] = −V(p − r(v · u)), div u = 0, (1.7)
which are both particular cases of (1.1): J = p, U = mu + r[v × u] and a = 0
in (1.6), and J = p − r(v · u), U = mu and 2a = rv/m in (1.7).
Let (x, y, z) be a Cartesian coordinate system with basis (i, j, k), and let (r, 4, z)
be a cylindrical coordinate system with basis (er , e4 , k). The both coordinate
systems have the same z-axis and are related in the ordinary way.
Example 1.4 (Magnetohydrodynamics). Let a non-magnetic solid body of
revolution translate at constant velocity in an electrically conducting viscous
incompressible fluid under the presence of an initially constant and uniform
magnetic field. It is assumed that body’s axis of revolution, body’s velocity
and the direction of the undisturbed magnetic field are all parallel to the
z-axis. The fluid velocity u, fluid pressure p and magnetic field disturbances h+
and h− inside and outside the body, respectively, can be described by linearized
dimensionless equations of MHD, provided that u and h− are small:

R(k · V)u = −Vp + Du + M2 [[(u − h− ) × k] × k], div u = 0
(1.8)
and curl h+ = 0, curl h− = Rm [(u − h− ) × k], div h± = 0,
where R is the Reynolds number, M is the Hartmann number and Rm
is the magnetic Reynolds number. In this case, the electric field is zero
everywhere, and u = ur (r, z)er + uz (r, z)k, u = curl 
u = u(r, z)e4 , p = p(r, z) and
 
± ± ±
h = hr (r, z)er + hz (r, z)k. Let l1,2 = R + Rm ± (R − Rm )2 + 4M2 /4. It is
shown in Zabarankin (2011a) that (1.8) can be recast in the form of (1.1)
with Jj = p + 2lj uz + (R − 2lj )hz− , Uj = u + (2lj − R)(ur − hr− )e4 and a = −lj k,
j = 1, 2.

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3747

(b) Generalized analytic functions


A generalized analytic (pseudoanalytic) function G = u + iu with the real
√ and
imaginary parts u = u(x, h) and v = v(x, h), respectively, and with i = −1 is
defined by the Bers–Vekua system (Bers 1953; Vekua 1962):
ux − vh + au + bv = 0 and uh + vx + cu + dv = 0, (1.9)
where a = a(x, h), b = b(x, h), c = c(x, h) and d = d(x, h) are known real-valued
functions. For a ≡ b ≡ c ≡ d ≡ 0, (1.9) simplifies to the Cauchy–Riemann system.
The relationship (1.1) leads to several classes of generalized analytic functions.
It can be readily seen that with the substitution
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

U = e−a·x L and J = e−a·x Y,


(1.1) takes the form
curl L + [a × L] = Ya − VY, div L = a · L, (1.10)
so that DL − a2 L = 0 and DY − a2 Y = 0, where DY = div(VY). The
system (1.10) defines a generalized holomorphic vector (Y, L) introduced by
Obolashvili (1975) (see also Liede (1990)) and is a particular case of the
quaternionic equation (Kravchenko & Shapiro 1996; Kravchenko 2003) related to
time-harmonic Maxwell’s equations. Next two corollaries show that under certain
assumptions on the symmetry of L and Y, the system (1.10) defines h-analytic
and H -analytic functions.
Corollary 1.5 (h-analytic functions). If a = −li, where l is a real-valued
constant, and
L = u(x, y)k and Y = v(x, y), (1.11)
then (1.10) reduces to the system
ux − lu = vy and uy = −vx − lv, (1.12)
which defines an h-analytic function G = u + iv. In this case, both u and v satisfy
   
the modified Helmholtz equation uxx + uyy − l2 u = 0 and vxx + vyy − l2 v = 0.
The system (1.12) was introduced by Duffin (1971) in context of the Yukawan
potential theory and also arises in the two-dimensional Oseen equations.
Corollary 1.6 (H -analytic functions). Let a = −lk and

Y = u(r, z) cos n4
(1.13)
and L = −er v(r, z) sin n4 + e4 v(r, z) cos n4 − ku(r, z) sin n4,
where l is a real-valued constant and n is a nonnegative integer. In this case, the
vectorial relationship (1.10) simplifies to two equations
       
v n v v v n+1
− u= − l v and +l u=− + v, (1.14)
vr r vz vz vr r
which define an nth-order H -analytic function G = u + iv and imply that
v2 1 v v2 n2
Dn u − l2 u = 0, Dn+1 v − l2 v = 0 and Dn = + + − .
vr 2 r vr vz 2 r2

Proc. R. Soc. A (2012)


3748 M. Zabarankin

Obviously, (1.14) is a particular case of (1.9). Also, if u and v satisfy (1.14),


then elz r −n u + ie−lz r n+1 v is a so-called p-analytic function (with p = e−2lz r 2n+1 )
introduced by Polozhii (1973) for arbitrary real-valued characteristic p = p(r, z)
as yet another generalization of ordinary analytic functions.
The system (1.14) has several important cases:

(a) For n = 0 and l = 0, (1.14) defines a zero-order r-analytic function and


arises in axially symmetric problems of Stokes flows (Zabarankin &
Krokhmal 2007; Zabarankin 2008a) and isotropic elastic medium (Polozhii
1973; Alexandrov & Soloviev 1978).
(b) For l = 0, (1.14) defines an nth-order r-analytic function1 and appears in
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

asymmetric Stokes flow problems (Zabarankin 2008b).


(c) For n = 0, (1.14) defines a zero-order H -analytic function and arises
in axially symmetric problems of Oseen flows (Zabarankin 2010) and
linearized MHD (Zabarankin 2011a).

(c) Generalized Cauchy integral formula and its application


The theories of p-analytic functions and generalized analytic functions defined
by (1.9) furnish general forms for the Cauchy integral formula, which often need to
be specialized and refined for particular classes of generalized analytic functions
(Chemeris 1995; Kravchenko 2008; Zabarankin 2008a). For example, the theory
of p-analytic functions facilitated obtaining the Cauchy integral formula for
zero-order H -analytic functions (Zabarankin 2010). However, it does not readily
yield an explicit-form Cauchy kernel for nth-order r-analytic functions, which
was derived via an integral representation involving ordinary analytic functions
similar to the representation for zero-order r-analytic functions (Alexandrov &
Soloviev 1978; Zabarankin 2008a). Section 2 shows that the Cauchy integral
formulae for h-analytic functions and H -analytic functions with particular cases
of n = 0 and l = 0 follow in a uniform fashion from the Cauchy integral formula
for a generalized holomorphic vector defined by (1.10).
In applications, the generalized Cauchy integral formula is indispensable
only if involved physical fields (e.g. velocity, pressure, vorticity, etc.) admit
representations in terms of corresponding generalized analytic functions. In
some cases as for an ideal fluid, the velocity u is already a generalized
analytic function, whereas, for example, for a two-dimensional Stokes flow, the
fields u, p and u are represented by Kolosov’s formulae with two ordinary
analytic functions involving their derivatives. Analogues of Kolosov’s formulae
with generalized analytic functions are available for three-dimensional Stokes
flows (Zabarankin 2008a,b), axisymmetric Oseen flows (Zabarankin 2010) and
axisymmetric linearized MHD (Zabarankin 2011a). However, in contrast to
Kolosov’s formulae, they contain no derivatives of the involved generalized
analytic functions, which considerably simplifies obtaining closed-form solutions
and deriving boundary-integral equations based on the generalized Cauchy
integral formula (Zabarankin 2008a,b, 2010, 2011a). Thus, the Cauchy integral
formula and the analogues of Kolosov’s formulae for u, p, and u form two essential
parts of the generalized analytic function approach to linear hydrodynamics.

1 This function is called n-harmonically analytic in Zabarankin (2008a,b).

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3749

Section 3 demonstrates this approach in problems of three-dimensional Stokes


flows, two-phase axisymmetric Stokes flows, two-dimensional Oseen flows and
axisymmetric Oseen flows.

2. Generalized Cauchy integral formula

This section shows that for several classes of generalized analytic functions, the
Cauchy integral formulae follow from the one for generalized holomorphic vectors
defined by (1.10). The next theorem is the vector-form restatement of the matrix-
form Cauchy integral formula for generalized holomorphic vectors given by (8)
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

and (16) in Obolashvili (1975).


Theorem 2.1 (Cauchy integral formula for generalized holomorphic vectors).
Let D be a bounded open region in R3 with a piecewise smooth boundary vD, and
let Y and L satisfy (1.10) in D, be continuously differentiable in D and continuous
in D ∪ vD. Then Y and L can be represented in D via their boundary values by

Y(x) = − (Y(y)ny + [ny × L(y)]) · (Vy − a)F(y − x)dS (y) (2.1)
vD
and

L(x) = − ((Y(y) ny + [ny × L(y)]) × (Vy + a)F(y − x)
vD
+ (ny · L(y))(Vy − a)F(y − x))dS (y), (2.2)
where dS (y) is the surface area element, ny is the outward normal of vD at
point y, Vy is the gradient with respect to y, (Vy ± a)F(y − x) ≡ Vy F(y − x) ±
F(y − x)a, and
⎧ −a y−x

⎪ e
⎨ in the three-dimensional case,
F(y − x) = 4py − x

⎪ 1
⎩ K0 (a y − x) in the two-dimensional case
2p
is the fundamental solution of the modified Helmholtz equation2 with Km (·) being
the mth-order modified Bessel function of the second kind.3 If a = 0, then in the
two-dimensional case, F(y − x) = −1/(2p) ln y − x.
For brevity, let I1 (Y, L) and I2 (Y, L) denote the right-hand sides of (2.1)
and (2.2), respectively. Two remarks are in order.
Remark 2.2. Let D − be the complement of D ∪ vD in R3 (or in R2 ) and let
(Y, L) be a generalized holomorphic vector continuously differentiable in D − and
continuous in D − ∪ vD. Then the Cauchy integral formula (2.1) and (2.2) takes
the form: Y(x) = −I1 (Y, L), x ∈ D − and L(x) = −I2 (Y, L), x ∈ D − , provided that
Y and L vanish at infinity (Liede 1990).
2 The fundamental solution F(x) solves (D − a2 )F(x) = −d(0), where d(·) is the Dirac

delta function.
3 In the two-dimensional case, D is a region in the xy-plane, and Y and L are represented by (1.11)

with a being a constant vector parallel to the z-axis.

Proc. R. Soc. A (2012)


3750 M. Zabarankin

Remark 2.3. Let f and g be scalar and vector functions, respectively, that are
continuous on vD. Then I1 (f , g) and I2 (f , g) determine a generalized holomorphic
vector for x ∈ vD. If f and g are Hölder continuous4 on vD and x0 ∈ vD,
then the limits of I1 (f , g) and I2 (f , g) as x = x+ and x = x− approach x0 from
inside and outside D, respectively, are determined by ±lim I1 (f , g) = ± 12 f (x0 ) +
x →x0
I1 (f , g)|x=x0 and ±lim I2 (f , g) = ± 12 g(x0 ) + I2 (f , g)|x=x0 , which are analogues of
x →x0
the Sokhotski–Plemelj formulae (Obolashvili 1975, eqn 19).
Corollary 2.4 (Ideal fluid). The equations (1.2) governing the velocity field u of
an ideal fluid correspond to (1.10) with Y = 0, L = u, and a = 0, so that for (1.2) in
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

the three-dimensional case, (2.2) simplifies to the formula (Mises 1944; Bitsadze
1969; Morgunov 1974)
  
1 ny · (y − x) [ny × (y − x)]
u(x) = u(y) + × u(y) dS (y), (2.3)
4p vD y − x3 y − x3
whereas in the two-dimensional case, the components ux and −uy form an ordinary
analytic function ux − iuy , and the linear combination (2.2)·i − i(2.2)·j with F(y −
x) = −1/(2p) ln y − x reduces to the ordinary Cauchy integral formula.
Corollary 2.5 (Stokes flow). The equations (1.4) that relate the pressure p and
vorticity u in a Stokes (creeping) flow correspond to (1.10) with p = Y, L = m u
and a = 0. Thus, for (1.4), the Cauchy integral formula (2.1)–(2.2) yields

1 y−x
p(x) = (p(y) ny + m[ny × u(y)]) · dS (y)
4p vD y − x3
and 
1 p(y)[ny × (y − x)] + m[u(y), y − x, ny ]
u(x) = dS (y),
4pm vD y − x3
where [a, b, c] = (b · c)a + (a · c)b − (a · b)c is the triple product of vectors a,
b and c. These formulae are a vector-form restatement of the matrix-form
Cauchy integral formula for the generalized analytic functions defined by the
Moisil–Theodorescu system (Bitsadze 1969).
The Cauchy integral formula for h-analytic functions defined by (1.12) is stated
in theorem 18 in Duffin (1971) for l ≥ 0. The next corollary shows that it follows
from (2.1) and (2.2) and holds for positive and negative l.
Corollary 2.6 (Cauchy integral formula for h-analytic functions). Let D be a
bounded open region in the xy-plane with a piecewise smooth positively oriented
boundary , and let G be an h-analytic function in D and Hölder continuous on
, then

1 sK1 (s)
G(z) = Ch (G; ) ≡ G(t) dt − lG(t)K0 (s)dt, z ∈ D, (2.4)
2pi  t−z
where Ch denotes the Cauchy operator for h-analytic functions, z = x + iy, t =
x1 + iy1 and s = |l| |t − z|. For l → 0, (2.4) reduces to the Cauchy integral formula
for ordinary analytic functions.
4 A function f (x) is Hölder continuous on vD, if f (x ) − f (x ) ≤ c x − x a for any x , x ∈ vD,
2 1 2 1 1 2
some a ∈ (0, 1] and non-negative constant c.

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3751

Detail. Let L and Y be represented by (1.11) with a = −li, where l is a real-


valued constant, and let x = (x, y) and y = (x1 , y1 ) be vectors in the xy-plane. In
this case, the formula (2.2), projected onto k, and the formula (2.1) simplify to

u(x) = L · k = − (v(y)[k × ny ] + u(y) ny ) · (Vy − li)F(y − x)ds(y) (2.5)


and

v(x) = Y = − (v(y)ny − u(y)[k × ny ]) · (Vy + li)F(y − x)ds(y), (2.6)

Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

respectively, where F(y − x) = K0 (|l| |t − z|)/(2p) and ds(y) is the curve length
element. Let G(z) = u(x) + iv(x). Then with u(x) = (G(z) + G(z))/2 and v(x) =
(G(z) − G(z))/(2i), the expression (2.5)+i(2.6) reduces to

G(z) = − (G(t)(ny − i[k × ny ]) · Vy F(y − x)


− lG(t)(ny + i[k × ny ]) · iF(y − x))ds(y). (2.7)

For positively oriented  with the outward normal ny , we have ny ds(y) = i dy1 −
j dx1 and
(ny ∓ i[k × ny ])ds(y) = (∓ii + j)(dr1 ± i dz1 ). (2.8)

Now, with the identity

|l| (x1 − x)i + (y1 − y)j


Vy F(y − x) = − K1 (|l| |t − z|)
2p |t − z|
and the relationship (2.8), the representation (2.7) takes the form of (2.4).
Finally, the limits lime→0 eK0 (e) = 0 and lime→0 eK1 (e) = 1 imply that (2.4)
reduces to the Cauchy integral formula for ordinary analytic functions as l → 0.

Theorem 2.7 (Cauchy integral formula for H -analytic functions). Let D be a


bounded open region in the rz-half plane (in the cylindrical coordinates (r, 4, z),
r ≥ 0) with a piecewise smooth positively oriented boundary , which is either closed
or an open curve with the endpoints lying on the z-axis (if D contains a segment
of the z-axis). Also, let G be an nth-order H -analytic function in D and Hölder
continuous on . The Cauchy integral formula for G is given by
  
1 U+ (z, t) U− (z, t)
G(z) = C (G; ) ≡ G(t) dt − G(t) dt̄ , z ∈ D, (2.9)
2pi  t−z t̄ + z

where C is the generalized Cauchy operator, z = r + iz, t = r1 + iz1 and U+ (z, t)


and U− (z, t) are determined by
 p/2
U± (z, t) = 2(−1) r1
n
L± (z, t, t) e−|l|9(z,t,t) dt, (2.10)
0

Proc. R. Soc. A (2012)


3752 M. Zabarankin

with

(2n + 1 + il(t − z)) cos t cos(2n + 1)t ⎪
L+ (z, t, t) = + |l| sin t sin(2n + 1)t,⎪


9(z, t, t) ⎪




(2n + 1 − il(t̄ + z)) sin t sin(2n + 1)t
L− (z, t, t) = + |l| cos t cos(2n + 1)t ⎪
9(z, t, t) ⎪



 ⎪



and 9(z, t, t) = (r1 + r) + (z1 − z) − 4r1 r sin t.
2 2 2

(2.11)
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

Proof. Let D be the axially symmetric region obtained by revolving D around


the z-axis in the cylindrical coordinate system (r, 4, z). For nth-order H -analytic
functions, L and Y are represented by (1.13) with a = −lk, where l is a real-
valued constant. Since u(r, z) and v(r, z) in (1.13) do not depend on the angular
coordinate 4, the formulae (2.1) and (2.2) can be considered in the rz-half
plane corresponding to 4 = 0. In this case, x = (r er + z e4 )|4=0 = ri + zk, and the
coordinate 4 will be used to describe the vector y as y = r1 cos 4i + r1 sin 4j + z1 k,
so that
y − x = (r1 cos 4 − r)i + r1 sin 4j + (z1 − z)k. (2.12)
Then in the identified rz-half plane, (2.1) and (2.2) projected onto j take the form
 2p 
u(r, z) = − (v(r1 , z1 )[(i sin(n + 1)4 − j cos(n + 1)4) × ny ]
0 
+ u(r1 , z1 )(ny cos n4 + [k × ny ] sin n4)) · (Vy + lk)Fr1 ds d4 (2.13)
and
v(r, z) = L(x) · j
 2p 
=− (v(r1 , z1 )(ny cos(n + 1)4 − [k × ny ] sin(n + 1)4)
0 
+ u(r1 , z1 )[(i sin n4 + j cos n4) × ny ]) · (Vy + lk)F r1 ds d4, (2.14)
respectively, where F ≡ F(y − x) = F(r, z, r1 , z1 , 4), dS (y) = r1 ds d4 and ds =
ds(r1 , z1 ) is the curve length element.
Let G(z) = u(r, z) + iv(r, z), so that u(r, z) = (G(z) + G(z))/2 and u(r, z) =
(G(z) − G(z))/(2i). In this case,
 
y−x 1 |l|
Vy F = − + e−|l| y−x ,
4p y − x3 y − x2
where y − x is given by (2.12). Then, with the relationship ny ds = (i cos 4 +
j sin 4)dz1 − k dr1 , which holds for the outward normal ny and positively oriented
, and with the identity
(y − x) · ((cos n4 ± cos(n + 1)4)(ii + k) ± i(sin(n + 1)4 ∓ sin n4)j)
= ±i(r1 ∓ iz1 ∓ z̄)(cos n4 ± cos(n + 1)4),

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3753

the combination (2.13)+i(2.14) simplifies to


 
1 2p
G(z) = r1 (G(t)X+ (z, t, 4)dt − G(t)X− (z, t, 4)dt)d4, (2.15)
8pi 0 
where
  
1 |l|
X± (z, t, 4) = (r1 ∓ iz1 ∓ z̄)(cos n4 ± cos(n + 1)4) +
y − x 3 y − x2

cos n4 ∓ cos(n + 1)4 −|l| y−x
± il e .
y − x
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

Changing the variable 4 = p − 2t and using 9(z, t, t) in (2.11), we have


 2p  
1 |l|
(cos n4 + cos(n + 1)4) + e−|l| y−x d4
0 y − x 3 y − x 2
 p/2  
1 |l|
= 8(−1)n sin t sin(2n + 1)t 3 + 2 e−|l|9(z,t,t) dt
0 9 (z, t, t) 9 (z, t, t)
n  p/2  
8(−1) (2n + 1) cos t cos(2n + 1)t
= + |l| sin t sin(2n + 1)t
|t − z|2 0 9(z, t, t)
× e−|l|9(z,t,t) dt (2.16)
and
 2p  
1 |l|
(cos n4 − cos(n + 1)4) + e−|l| y−x d4
0 y − x3 y − x2
 p/2  
1 |l|
= 8(−1) n
cos t cos(2n + 1)t 3 + 2 e−|l|9(z,t,t) dt
0 9 (z, t, t) 9 (z, t, t)
n  p/2  
8(−1) (2n + 1) sin t sin(2n + 1)t
= + |l| cos t cos(2n + 1)t
|t + z̄|2 0 9(z, t, t)
× e−|l|9(z,t,t) dt, (2.17)
where the middle integrals in (2.16) and (2.17) are integrated by parts with the
relationships
   
sin t 1 d cos t cos t 1 d sin t
=− and 3 = ,
93 (z, t, t) |t − z|2 dt 9(z, t, t) 9 (z, t, t) |t + z̄|2 dt 9(z, t, t)

respectively. Now, (2.15) with (2.16) and (2.17) yields (2.9)–(2.11). 


The generalized Cauchy integral formula (2.9)–(2.11) extends the one for
zero-order H -analytic functions derived in Zabarankin (2010, theorem 1) via the
theory of p-analytic functions (Polozhii 1973), whereas the following result was
obtained in Zabarankin (2008a) based on an integral representation of nth-order
r-analytic functions through ordinary analytic functions.

Proc. R. Soc. A (2012)


3754 M. Zabarankin

Corollary 2.8 (Cauchy integral formula for nth-order r-analytic functions).


For l = 0, the kernels (2.10) become real-valued functions determined by
 
[G(n + 3/2)]2 k2n (z, t) 1 1
U± (z, t) = 2r1 F n + , n + 1 ∓ , 2(n + 1), k (z, t) ,
2
(2n + 1)! |z + t̄| 2 2
(2.18)
1
where G(·) is the gamma function, F(a, b, c, k) = G(c)/(G(b)G(c − b)) 0 t b−1
(1 − t)c−b−1 (1 − kt)−a dt is the hypergeometric function, and k2 (z, t) = 4r1 r
|z + t̄|−2 .

Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

Detail. With 9(z, t, t) = |t + z̄| 1 − k2 (z, t) sin2 t, the formula (2.18) follows
from the relationships
 p/2 √  
cos t cos(2n + 1)t pG(n + 1/2)k2n 1 1
I1 =  dt = 2(n+1) F n + , n + , 2(n + 1), k 2
0 1 − k2 sin2 t 2 (−1)n n! 2 2
(2.19)
and
 p/2 √  
sin t sin(2n + 1)t pG(n + 1/2)k2n 3 1
I2 =  dt = 2(n+1) F n + , n + , 2(n + 1), k 2
,
0 1 − k2 sin2 t 2 (−1)n n! 2 2
(2.20)

see the appendix A, and the fact that G(n + 3/2) = p(2n + 1)!/(22n+1 n!).
Corollary 2.9 (Cauchy integral formula for zero-order r-analytic functions).
The case of l = 0 and n = 0 in (1.14) corresponds to zero-order r-analytic
functions, for which (2.18) simplifies to

|t + z̄| ⎪
U+ (z, t) = (E(k(z, t)) − (1 − k (z, t))K(k(z, t)))⎪
2 ⎪

2r
(2.21)
|t + z̄| ⎪



and U− (z, t) = (K(k(z, t)) − E(k(z, t))),
2r
p/2 p/2 
where K(k) = 0 (1 − k2 sin2 t)−1/2 dt and E(k) = 0 1 − k2 sin2 t dt are
complete elliptic integrals of the first and second kinds, respectively. The function
K(k) has a logarithmic singularity as k → 1−: K(k) = −(ln(1 − k2 ))/2 + O(1).
Corollary 2.10 (Cauchy integral formula for first-order r-analytic analytic
functions). The case of l = 0 and n = 1 in (1.14) corresponds to first-order
r-analytic functions, for which (2.18) simplifies to

|t + z̄| ⎪
U+ (z, t) = ((8 − 7k )E(k) − (8 − 3k )(1 − k )K(k))⎪
2 2 2

2r (2.22)
|t + z̄| 2 ⎪

and U− (z, t) = ((k − 8)E(k) + (8 − 5k )K(k)),
2 ⎭
2r
where k2 = 4r1 r|z + t̄|−2 .
Proc. R. Soc. A (2012)
Generalized Cauchy integral formula 3755

3. Application to linear hydrodynamics

This section demonstrates the approach of generalized analytic functions to


problems of linear hydrodynamics. Advantages of this approach are in the
convenience of representations of hydrodynamic fields (velocity, vorticity and
pressure) and key characteristics (drag, torque and lift) in terms of generalized
analytic functions and in the simplicity of obtaining closed-form solutions
and boundary-integral equations via the generalized Cauchy integral formulae;
compare to the methods, e.g. in Pozrikidis (1992) and Bardzokas et al. (2007).
For clarity, C0 , C1 and CH will denote the Cauchy operator C in (2.9) for
zero-order r-analytic, first-order r-analytic and zero-order H -analytic functions,
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

respectively, whereas Ch is the Cauchy operator in (2.4) for h-analytic functions.


(a) Ideal fluid
In the two-dimensional case, equations (1.2) for the velocity field u = ux (x, y)i +
uy (x, y)j of an ideal fluid reduce to the Cauchy–Riemann system for ux and −uy ,
so that g = ux − iuy is an ordinary analytic function of a complex variable z =
x + iy. If a solid infinitely long airfoil is aligned with the z-axis and is immersed
into a uniform flow ux∞ i + uy∞ j, then limz→∞ g(z) = ux∞ − iuy∞ ≡ v∞ and u · n ≡
Re[gvz/vn] = 0 on the boundary  of the airfoil cross section in the  xy-plane,
where n is the outward normal for . In this case, Fx + iFy = −(ir/2)  g 2 (z)dz is
the airfoil lifting force (Blasius–Chaplygin formula), where r is the fluid density.
It simplifies to the Kutta–Joukowski formula Fx + iFy = −ir v∞  g(z)dz.
In the three-dimensional case, u satisfying (1.2) is itself a generalized analytic
function for which the Cauchy integral formula is given by (2.3); see Mises (1944)
and Morgunov (1974) for application of (2.3) to the ideal fluid. In particular, in
an axisymmetric flow with the z-axis of revolution, uz and ur are independent of
the angular coordinate 4 and form a zero-order r-analytic function G = uz + iur
with Im[Gvz/vn] = 0 on any fixed solid boundary, where z = r + iz.
(b) Stokes flows
Under the zero Reynolds number assumption, the time-independent velocity
field u and pressure p of a viscous incompressible fluid are governed by the Stokes
equations (1.3). The relationship (1.4) plays a pivotal role in constructing solution
forms for (1.3) in terms of generalized analytic functions; see example 1.2.

(i) Two-dimensional case


In the two-dimensional case, u and p depend on the Cartesian coordinates
x and y only, i.e. u = ux (x, y)i + uy (x, y)j, u = u(x, y)k and p = p(x, y), and
consequently, the system (1.4) implies that p/m − iu is an ordinary analytic
function. In this case,
p
ux + iuy = g1 (z) − zg1 (z) − g2 (z) and − iu = −4g1 (z), (3.1)
m
where z = x + iy is a complex variable, and g1 (z) and g2 (z) are analytic functions;
see (4.1) and (4.2) in Richardson (1995). In fact, the representation (3.1) is
Kolosov’s formulae (Kolosov 1909) for an incompressible elastic medium (with
Poisson’s ratio 1/2). It is used to reduce two-dimensional Stokes flow problems

Proc. R. Soc. A (2012)


3756 M. Zabarankin

to boundary-integral equations via the Cauchy integral formula (Muskhelishvili


1977, 1992). If a solid infinitely long cylinder
 of arbitrary cross section is aligned
with the z-axis, then Fx + iFy = −4mi  g1 (z)dz is the total flow reaction force per
unit length of cylinder’s span, where  is the cross section’s boundary in the xy-
plane. However, the problem of unbounded two-dimensional Stokes flow with solid
obstacles has no solution bounded at infinity (Happel & Brenner 1983). This
well-known paradox is resolved by the Oseen approximation of the Navier–Stokes
equations.

(ii) Three-dimensional axially symmetric flows


Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

If a flow is axisymmetric with the z-axis of revolution, then in the cylindrical


coordinates (r, 4, z), u and p are independent of the angular coordinate 4: u =
ur (r, z) er + uz (r, z)k, u4 ≡ 0, p = p(r, z) and u = u(r, z) e4 (u = e4 · curl u). In
this case, a complex variable is introduced by z = r + iz, and proposition 7 in
Zabarankin (2008a) states that
 
ir p
uz + iur = z − G1 + G2 and + iu = G1 , (3.2)
2 m
where G1 and G2 are zero-order r-analytic functions. The representation (3.2) is
a three-dimensional analogue of (3.1), but in contrast to the latter, it involves no
derivatives of G1 and G2 .
Suppose a solid axisymmetric finite body with the z-axis of revolution
translates in the quiescent fluid at constant velocity vz k. Then u = vz k on body’s
surface S (no-slip boundary condition), and u and p vanish at infinity. Let open
regions D+ and D− be the interior and exterior of the body’s cross section in
the rz-half plane (r ≥ 0) with common positively oriented boundary  (cross
section of S ). Then (3.2) implies that (z − ir/2)G1 + G2 = vz on  and that G1
and G2 vanish at infinity.5 If G1 on  is known, then G2 = vz − (z − ir/2)G1 on
, and with the Cauchy integral formula (2.9) for zero-order r-analytic functions
(see corollary 2.9), (3.2) yields representations for u, p and u in D− :
p
uz + iur = A (G1 ; ) and + iu = −C0 (G1 ; ), z ∈ D− ,
m
where the operator A is determined by
    
ir ir
A (G1 ; ) = C0 z− G1 ;  − z − C0 (G1 ; ). (3.3)
2 2
Note that A has no Cauchy-type singularity on , and thus, A (G1 ; ) is
continuous as z approaches  from within D− . This fact and the boundary
condition uz + iur = vz on  imply that G1 on  satisfies the boundary-
integral equation
A (G1 ; ) = vz , z ∈ . (3.4)
5A zero-order r-analytic function G vanishing at infinity behaves as G = o(|z|−1 ) as |z| → ∞
(Zabarankin 2008a). Since the second equation in (3.2) implies that G1 vanishes at infinity, i.e.
G1 = o(|z|−1 ) as |z| → ∞, the first one in (3.2) yields G2 = o(1) as |z| → ∞.

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3757

Theorem 10 in Zabarankin (2008a) proves that a homogeneous solution of (3.4)


is any real constant and that G1 = w − c, z ∈ , where w is a solution of (3.4) and
c is the real constant determined by c = w(z)/2 + C0 (w; ), z ∈ .
The operator (3.3) has only a logarithmic-type singularity, and (3.4) is solved as
follows. Let  be parametrized by z = z(t),  t ∈ [t1 , t2 ], then G1 on  is approximated
by a finite functional series G1 (t) = vz m k=1 ak f1k (t) + ibk f2k (t), t ∈ [t1 , t2 ], with
basis functions {f1k (t)}m k=1 , {f 2k (t)} m
k=1 , where none of f1k (t) is constant, and
coefficients ak , bk are found by minimizing the total quadratic error on [t1 , t2 ]:
 m 2
 
 
min  ak A (f1k ; ) + bk A (if2k ; ) − 1 .
ak ,bk  
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

k=1

The representation (3.2) and the boundary condition uz + iur = vz on 


yield vuz /vn + ivur /vn = −(vz̄/vn)Im G1 , z ∈ , and proposition 11 in Zabarankin
(2008a) shows that the drag exerted on the body is given by
 
Fz = 2pmRe r G1 (z)dz .

Note that Fz is unaffected if G1 is added a real constant; so Fz holds for any
solution of (3.4). For example, if in the rz-half plane, a sphere of radius a
is parametrized by z(t) = aepit/2 , t ∈ [−1, 1], then G1 (t) = −3vz i/(2a)e−pit/2 , t ∈
[−1, 1] and Fz = −6pm vz a, which is the well-known Stokes formula for the sphere
drag. Zabarankin (2008a) showed that for prolate and oblate spheroids, biconvex
lens and torus of circular cross section, solutions of the boundary-integral
equation (3.4) coincide with corresponding analytical solutions in Happel &
Brenner (1983), Zabarankin & Krokhmal (2007) and Zabarankin &Ulitko (2006).
Also, Zabarankin (2008a) solved (3.4) for solid bi-spheroids (two separate
spheroids of equal size) and torus of elliptical cross section, whereas Zabarankin &
Molyboha (2010) used (3.4) to find minimum-drag shapes for solid bodies of
revolution subject to constraints on body’s volume and body’s shape.

(iii) Two-phase three-dimensional axially symmetric flow


Suppose an initially spherical liquid drop with radius a is placed into an
extensional flow u∞ = 2(−rer + 2zk) at the origin (r = z = 0), where 2 is the share
rate, and suppose D + and D − are the regions occupied by the deformed drop and
the ambient fluid, respectively, with common boundary S . Let u+ and u− be the
actual velocity in the drop and the velocity disturbance of the extensional flow,
respectively, and let p ± be the pressure in D ± . It is assumed that the drop and the
ambient fluid are incompressible and viscous with the same viscosity m and that
for fixed S , u± and p ± satisfy the Stokes equations (1.3) in D ± . For convenience,
let the linear dimensions, u± , and p± be rescaled by a, a 2 and m 2, respectively,
and let Ca = ma 2/g be the capillary number, where g is the interfacial tension.
This problem is axisymmetric with the z-axis of revolution, and the boundary
conditions on S are given by
u+ = u− + u∞ , p+ − p − = Ca−1 div n and u+ = u− , (3.5)
± ±
where n is the outward normal for S and u = e4 · curl u (Zabarankin & Nir
2011, proposition 3.1). Also, u− and p− vanish at infinity.

Proc. R. Soc. A (2012)


3758 M. Zabarankin

At the steady state, u+ · n = 0 on S (kinematic condition), and in this case, u± ,


u and p ± are time-independent and have a representation similar to that of (3.2):
±

 
ir
uz± + iur± = z − G1± + G2± (3.6)
2
and
p + = Re G1+ + c, p − = Re G1− and u± = Im G1± , (3.7)
where G1± = G1± (z) and G2± = G2± (z) are zero-order r-analytic functions in D ±
with z = r + iz and with G1− and G2− vanishing at infinity, and c is a real constant.
Let open regions D± be the interior of cross sections of D ± in the rz-half
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

plane, and let  be common positively oriented smooth boundary of D± . With


the Cauchy integral formula for zero-order r-analytic functions (corollary 2.9),
the representations (3.6) and (3.7) and the boundary conditions (3.5) yield a
closed-form solution for u± , p ± and u± (Zabarankin & Nir 2011, theorem 3.3):

uz± + iur± = f ± − Ca−1 A (div n; ), z ∈ D± ∪ , ⎪


± ± −1 ±
p + iu = Ca C0 (div n; ), z ∈ D (3.8)



and p ± + iu± = ±(2Ca)−1 div n + Ca−1 C0 (div n; ), z ∈ ,
where f + = 2z − ir, f − = 0 and the operator A is defined by (3.3). Zabarankin &
Nir (2011) used (3.8) to find the drop’s steady shape from the kinematic condition.

(iv) Three-dimensional asymmetric flows


Suppose a solid axisymmetric finite body with the z-axis of revolution either
translates in the quiescent fluid along the x-axis (x-translation) or rotates around
the y-axis (y-rotation). Proposition 2 in Zabarankin (2008b) shows that in this
case, u and p are represented by

uz (r, 4, z) + iur (r, 4, z) = ((r + z)G1 + iG2 + G3 ) cos 4,⎪

u4 (r, 4, z) = Im[−zG1 − iG2 + G3 ] sin 4 (3.9)


and p(r, 4, z) = 2mIm G1 cos 4,
where z = r + iz is a complex variable, G1 = G1 (z) and G2 = G2 (z) are zero-order
r-analytic functions, G3 = G3 (z) is a first-order r-analytic function, and all three
G1 , G2 and G3 vanish at infinity. On the body’s surface S , the boundary conditions
are given by u = vx i for the x-translation and by u = [6y j × (x i + zk)] for the
y-rotation, where vx and 6y are constants.
Let D+ and D− be open regions corresponding to the interior and exterior
of the body’s cross section in the rz-half plane, and let  be common positively
oriented boundary of D± (cross section of body’s surface S ). Then (r + z)G1 +
iG2 + G3 = f1 and Im[zG1 + iG2 − G3 ] = f2 on , where f1 = ivx and f2 = vx for the
x-translation and f1 = −6y z̄ and f2 = 6y z for the y-rotation. If G1 , G2 and G3 on
 are determined, then representing ur , u4 and uz in D− with (3.9) and with the
Cauchy operators C0 and C1 is straightforward (see corollaries 2.9 and 2.10).
Suppose that G1 and Re G3 on  are known, then G2 = f3 + i((r + z)G1 + G3 )
and Im G3 = − 12 rIm G1 on , where f3 = vx for the x-translation and f3 = i6y z̄ for

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3759

the y-rotation. Theorem 3 in Zabarankin (2008b) proves that G1 and Re G3 on 


are found from a system of two boundary-integral equations:

C0 (i(z + r)G1 + iG3 ; ) − i(z + r)C0 (G1 ; ) − iC1 (G3 ; ) = f4 , z ∈ 
(3.10)
2Im G3 = −rIm G1 and G3 + 2C1 (G3 ; ) = 0, z ∈ ,
where f4 = −vx for the x-translation and f4 = −6y (iz̄/2 + C0 (iz̄; )) for the y-
rotation (2Im G3 = −rIm G1 , z ∈ , is retained for brevity of notation). Observe
that G3 = G3 /2 − C1 (G3 ; ), z ∈ , is the Sokhotski–Plemelj formula for G3 on ,
which is equivalent to a single real-valued equation in the sense that if Re G3
on  is known then Im G3 on  is found from this formula, and vice versa.
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

Thus, the system (3.10) is viewed as three real-valued equations with three real-
valued unknowns Re G1 , Im G1 and Re G3 . Zabarankin (2008b) solved (3.10)
with the quadratic error minimization technique for the x-translation of bi-
spheroids and for the y-rotation of torus of elliptical cross section and showed
that the solutions for bi-spheres and torus of circular cross section coincide with
corresponding analytical solutions in Goren & O’Neill (1980), Wakiya (1967) and
Zabarankin (2007).
Propositions 7 and 8 in Zabarankin (2008b) state that the drag exerted on the
body in the x-translation and the torque in the y-rotation are determined by
   
Fx = 4pmRe rG1 (z)dz and Ty = 4pmRe r((2z − ir)G1 (z) + G2 (z))dz ,
 

which compared with (48a) and (54a) in Zabarankin (2008b), respectively, have
additional multiplier 2 because multiplier 1/2 was omitted from the right-hand
side in (3.9). Observe that Fx and Fz in the axisymmetric translation have similar
expressions.

(c) Oseen flows


Suppose a solid body translates in a viscous incompressible fluid with constant
velocity. Under the low Reynolds number assumption, the time-independent
velocity field and pressure are governed by the Oseen equations (1.5). Example 1.3
implies that u and p can be represented by h-analytic functions in the two-
dimensional case and by H -analytic functions in the three-dimensional case.

(i) Two-dimensional case


If the body is a solid infinitely long cylinder of arbitrary cross section, which is
aligned with the z-axis and translates in the quiescent fluid at constant velocity
v = −vx i, then u = ux (x, y)i + uy (x, y)j, u = u(x, y)k and p = p(x, y), and u and
p vanish at infinity in the xy-plane. In this case, (1.6) and (1.7) determine
an ordinary analytic function g and h-analytic function h, respectively, with a
complex variable z = x + iy, and thus, u, p and u have a representation
uy + iux = g + elx h, p = −2lmIm g and u = 2lelx Re h, (3.11)
where l = rvx /(2m)  = 0, and g and h vanish as |z| → ∞.
Remark 3.1. The functions g and h are uniquely determined. Indeed, let pairs
g1 , h1 and g2 , h2 be two different solutions, then g = g2 − g1 and h = h2 − h1 solve
the exterior homogeneous two-dimensional Oseen flow problem, which with u and

Proc. R. Soc. A (2012)


3760 M. Zabarankin

p vanishing as |z| → ∞ has only zero solution (Finn 1965). Thus, (3.11) with p ≡ 0
and u ≡ 0 yields Im g ≡ 0 and Re h ≡ 0, and then (3.11) and u ≡ 0 imply Re g ≡ 0
and Im h ≡ 0.
Let open regions D + and D − be cross sections of the cylindrical body and fluid
in the xy-plane, respectively, with common boundary  (positively oriented Jordan
curve). The boundary condition u = −vx i on  and the representation (3.11) imply
uy + iux = −ivx on . Suppose the value of g on  is known. Then h = −e−lx (g +
ivx ) on , and with the Cauchy integral formulae for analytic and h-analytic
functions, (3.11) yields representations for u, p and u in D − :
  
1 g(t)
dt , z ∈ D − ,
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

uy + iux = Bh (g; ), p = 2lmIm


2pi  t − z
and u = 2lelx Re[Ch (e−lx g; )], z ∈ D − , where the operator Bh is given by

−lx 1 g(t)
Bh (g; ) = e Ch (e g; ) −
lx
dt
2pi  t − z

1 sK1 (s)el(x−x1 ) − 1
≡ g(t)dt − l K0 (s)el(x−x1 ) g(t)dt
2pi  t−z
with s = |l| |t − z| and Km (·) being the mth-order modified Bessel function of the
second kind. Observe that the operator Bh has no Cauchy-type singularity on
. Consequently, Bh (g; ) is continuous as z approaches  from within D − , and
uy + iux = −ivx on  implies that g satisfies the boundary-integral equation
Bh (g; ) = −ivx , z ∈ . (3.12)
Theorem 3.2. Let l = 0, then a homogeneous solution of (3.12) is any imaginary
constant and g = g̃ − c, z ∈ , where g̃ is a solution
 of (3.12), and c is the imaginary
constant determined by c ≡ g̃(z)/2 + 1/(2pi)  (g̃(t)/(t − z))dt, z ∈ .
Proof. The proof follows the approach of Muskhelishvili (1992). The necessary
and sufficient condition for a complex-valued function g on  to be the boundary
value of an analytic function in D− that vanishes at infinity is given by the
Sokhotski–Plemelj formula

g(z) 1 g(t)
g(z) = − dt, z ∈ . (3.13)
2 2pi  t − z
Let g̃(z), z ∈ , be a solution of (3.12), then the Cauchy-type integrals

1 ivx − g̃(t)
g1 (z) = dt, h1 (z) = Ch (e−lx g̃; ), z ∈ D + ,
2pi  t − z
define ordinary analytic and h-analytic functions, respectively, in D + . Then the
boundary-integral equation (3.12) is equivalent to g1 + elx h1 = 0 on  (through
Sokhotski–Plemelj formulae as z approaches  from within D + ). Observe that
similar to (3.11), uy + iux = g1 + elx h1 is a formal solution of the two-dimensional
Oseen equations (1.5) in D + . However, (1.5) with u = 0 on  yields u ≡ 0 in
D + . Thus, g1 ≡ −elx h1 in D + , which holds only if g1 ≡ c1 in D + , where c1 is an
+
imaginary constant. As z approaches  from  within D , the Cauchy-type integral
that defines g1 implies g̃(z)/2 + 1/(2pi)  (g̃(t)/(t − z)) dt = ivx − c1 = c for any

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3761

z ∈ . Since g should satisfy (3.13), the solution g = g̃ − c, z ∈ , follows. Now


if g̃ ∗ is another solution of (3.12) with a corresponding imaginary constant c ∗ ,
then because g is uniquely determined (remark 3.1), g̃ ∗ − c ∗ = g̃ − c, z ∈  or,
equivalently, g̃ ∗ − g̃ = c ∗ − c, z ∈ , and thus, a homogeneous solution of (3.12) is
only an imaginary constant. 
With (3.11), (3.12) is equivalent to the two-dimensional Oseenlet-based
boundary-integral eqn (12) in Williams (1965).
It follows from (3.11) and uy + iux = −vx i on  that p/m − iu = 2lig and
vuy /vn + ivux /vn = −2l(vz̄/vn)Re g, z ∈ , which yields the drag exerted on the
cylinder per unit length of cylinder’s span:
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

    
vu
Fx = i · 2m + m[n × u] − pn ds = −2lmRe g(z)dz , (3.14)
 vn 

where n is the outward normal for  and ds is the curve length element. Since Fx is
unchanged by adding a constant to g, it holds for any solution of (3.12). Observe
that 2lm = rvx and (3.14) bears a close resemblance to the Kutta–Joukowski
formula (see §3a).
As a verification, (3.12) is solved for a circular cylinder with radius 1. In this

case, z(t) = eit and g(t) = vx 10
k=1 ak sin kt + ibk cos kt, t ∈ [−p, p], with ak and bk
found by the quadratic error minimization technique (see §3b(ii)). The obtained
drag coefficient CD = Fx /(2lmvx ) for Reynolds numbers R = 4l = 0.2, 2, 20, 200
and 2000 is 34.6541, 8.0847, 3.4658, 2.5387, 2.3237, respectively, versus 34.655,
8.090, 3.469, 2.545 and 2.324, respectively, computed in Miyagi (1974).

(ii) Translation of solid bodies of revolution


Suppose a solid axisymmetric finite body with the z-axis of revolution
translates in the quiescent fluid at constant velocity −vz k, so that u = −vz k
on the body’s surface S , and u and p vanish at infinity. In this case, u and p
are independent of the angular coordinate 4: u = ur (r, z)er + uz (r, z)k, u4 ≡ 0,
p = p(r, z) and u = u(r, z)e4 , and proposition 1 in Zabarankin (2010) states that

uz + iur = G + elz H , p = −2lmRe G, u = 2lelz Im H , (3.15)


where l = rvz /(2m) = 0, and G and H are zero-order r-analytic and zero-order
H -analytic functions, respectively, that both vanish at infinity.
Let open regions D+ and D− be the interior and exterior of the body’s cross
section in the rz-half plane (r ≥ 0), respectively, with common positively oriented
boundary . Then (3.15) implies that G + elz H = −vz on . If the boundary value
of G on  is known, then H = −e−lz (vz + G) on  and with the Cauchy integral
formula (2.9), (3.15) furnishes representations for u, p and u in D− :
uz + iur = BH (G; ), p = 2lmRe[C0 (G; )], z ∈ D− ,
and u = 2lelz Im[CH (e−lz G; )], z ∈ D− , where BH is determined by
BH (G; ) = elz CH (e−lz G; ) − C0 (G; ).
The operator BH has no Cauchy-type singularity on , so that BH (G; ) is
continuous as z approaches  from within D− , and the boundary condition uz +

Proc. R. Soc. A (2012)


3762 M. Zabarankin

iur = −vz on  implies that G on  satisfies the boundary-integral equation


BH (G; ) = −vz , z ∈ . (3.16)
Theorem 10 in Zabarankin (2010) proves that a homogeneous solution of (3.16)
is any real constant and that G = w − c, z ∈ , where w is a solution of (3.16) and
c is the real constant determined by c = w/2 + C0 (w; ), z ∈ .
The representation (3.15) and uz + iur = −vz on  yield p/m + iu = −2lG
and vuz /vn + ivur /vn = 2l(vz̄/vn)Im G, z ∈ , and proposition 10 in Zabarankin
(2010) shows that the drag exerted on the body is given by
 
Fz = −4plmRe
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

rG(z)dz .


Observe that a real constant added to G has no effect on the drag.


As an illustration, Zabarankin (2010) solved the boundary-integral
equation (3.16) with the quadratic error minimization technique for sphere,
spheroids and bi-spheroids, and showed that the use of the conjugate kernels in the
Cauchy integral formula for zero-order H -analytic functions improves technique’s
running time by several times. Zabarankin & Molyboha (2011) used (3.16) with
the conjugate kernels in an iterative procedure for finding minimum-drag shapes
for solid bodies under the low Reynolds number assumption subject to constraints
on body’s volume and body’s shape.

(d) Magnetohydrodynamics
The approach of generalized analytic functions to the MHD problem
formulated in example 1.4 is demonstrated in Zabarankin (2011a). With the
generalized Cauchy integral formulae, the MHD problem is reduced to boundary-
integral equations, which are then employed in an iterative procedure for finding
minimum-drag shapes for solid nonconducting bodies in the MHD flow under the
assumption of small R, Rm and M; see Zabarankin (2011b).
We are grateful to the anonymous referees for their valuable comments and suggestions, which
greatly helped to improve the quality of the paper.

Appendix A

For |b| < 1 and n ∈ Z+


0 , formula (10) in §2.4 in Bateman & Erdelyi (1953) yields
   2p
1 1 n! cos n4 d4
F , n + , n + 1, b = √
2

2 2 2 pG (n + 1/2) b 0n
1 − 2b cos 4 + b2
 p/2
2(−1)n n! cos(2nt)dt
=√  ,
pG(n + 1/2)b (1 + b) 0
n
1 − 4b/(1 + b)2 sin2 t
whereas formula (24) in §2.1.5 in Bateman & Erdelyi (1953) implies
   
1 1 1 1 4b
F , n + , n + 1, b2 = (1 + b)−2n−1 F n + , n + , 2n + 1, .
2 2 2 2 (1 + b)2

Proc. R. Soc. A (2012)


Generalized Cauchy integral formula 3763

Consequently, √
 p/2
cos(2nt)dt pG(n + 1/2)k2n
 = F1 ,
0 1 − k2 sin2 t 22n+1 (−1)n n!
where F1 = F[n + 1/2, n + 1/2, 2n + 1, k2 ], so that
 p/2 √  
cos(2nt) + cos 2(n + 1)t pG(n + 1/2)k2n (n + 1/2)k2
I1 =  dt = 2(n+1) F1 − F2 ,
0 2 1 − k2 sin2 t 2 (−1)n n! 4(n + 1)
where F2 = F[n + 3/2, n + 3/2, 2n + 3, k2 ]. For any positive a, b and c, formulae
(20) and (22) in §2.8 in Bateman & Erdelyi (1953) yield
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

  c(c − 1)     
F a + 1, b + 1, c + 1, k2 = 2
F a, b, c − 1, k2 − F a, b, c, k2 ,
abk
so that F2 = 4(n + 1)/((n + 1/2)k2 )(F1 − F3 ), where F3 = F[n + 1/2, n + 1/2,
2n + 2, k2 ], and thus, I1 simplifies to (2.19). Similarly,
 p/2 √
cos(2nt) − cos 2(n + 1)t pG(n + 1/2)k2n
I2 =  dt = 2(n+1) (2F1 − F3 ).
0 2 1 − k2 sin2 t 2 (−1)n n!
Now, (34) in §2.8 in Bateman & Erdelyi (1953) implies that
 
3 1 k2
F1 = F n + , n + , 2n + 1, k −
2
F3 ,
2 2 2(1 − k2 )
which, with formula (39) in §2.8 in Bateman & Erdelyi (1953), results in (2.20).

References
Alexandrov, A. Y. & Soloviev, Y. I. 1978 Three-dimensional problems of the theory of elasticity.
Moscow: Nauka. [In Russian.]
Bardzokas, D. I., Filshtinsky, M. L. & Filshtinsky, L. A. 2007 Mathematical methods in electro-
magneto-elasticity. Lecture Notes in Applied and Computational Mechanics. New York,
NY: Springer.
Bateman, H. & Erdelyi, A. 1953 Higher transcendental functions, vol. 1. New York,
NY: McGraw-Hill.
Bers, L. 1953 Theory of pseudo-analytic functions. New York, NY: Institute for Mathematics and
Mechanics.
Bitsadze, A. 1969 Foundations of the theory of analytic functions of complex variable. Moscow:
Nauka. [In Russian.]
Chemeris, V. S. 1995 Construction of Cauchy integrals for one class of nonanalytic functions of
complex variables. J. Math. Sci. 75, 1857–1865. (doi:10.1007/BF02367742)
Duffin, R. J. 1971 Yukawan potential theory. J. Math. Anal. Appl. 35, 105–130. (doi:10.1016/0022-
247X(71)90239-3)
Finn, R. 1965 On the exterior stationary problem for the Navier–Stokes equations, and associated
perturbation problems. Arch. Ration. Mech. An. 19, 363–406. (doi:10.1007/BF00253485)
Goren, S. L. & O’Neill, M. E. 1980 Asymmetric creeping motion of an open torus. J. Fluid Mech.
101, 97–110. (doi:10.1017/S0022112080001553)
Happel, J. & Brenner, H. 1983 Low Reynolds number hydrodynamics. New York, NY: Springer.
Kolosov, G. V. 1909 On the application of complex function theory to a plane problem of the
mathematical theory of elasticity. Tartu, Estonia: Dorpat (Yuriev) University. [In Russian.]

Proc. R. Soc. A (2012)


3764 M. Zabarankin

Kravchenko, V. V. 2003 Applied quaternionic analysis. Research and Exposition in Mathematics.


Lemgo, Germany: Heldermann.
Kravchenko, V. V. 2008 On a transplant operator and explicit construction of Cauchy-type integral
representations for p-analytic functions. J. Math. Anal. Appl. 339, 1103–1111. (doi:10.1016/
j.jmaa.2007.07.068)
Kravchenko, V. V. & Shapiro, M. V. 1996 Integral representations for spatial models of mathematical
physics. Pitman Research Notes in Mathematics Series, vol. 20. London, UK: Longman.
Liede, H. 1990 The existence and uniqueness theorems of the linear and nonlinear Riemann–Hilbert
problems for the generalized holomorphic vector of the second kind. Acta Math. Sci. 10, 185–199.
Mises, R. 1944 Integral theorems in three-dimensional potential flow. Bull. Am. Math. Soc. 50,
599–611. (doi:10.1090/S0002-9904-1944-08175-5)
Miyagi, T. 1974 Oseen flow past a circular cylinder. J. Phys. Soc. Jpn. 37, 1699–1707. (10.1143/JP
Downloaded from https://royalsocietypublishing.org/ on 04 July 2021

SJ.37.1699)
Moisil, G. C. & Theodorescu, N. 1931 Fonctions holomorphes dans l’espace. Mathematica 5,
142–159.
Morgunov, G. M. 1974 Three-dimensional analog of generalized Cauchy formula and its application
to hydromechanics. Fluid Dyn. 9, 287–289. (doi:10.1007/BF01092664)
Muskhelishvili, N. I. 1977 Some basic problems of the mathematical theory of elasticity, 1st edn.
New York, NY: Springer.
Muskhelishvili, N. I. 1992 Singular integral equations: boundary problems of function theory and
their applications to mathematical physics, 2nd edn. New York, NY: Dover Publications.
Obolashvili, E. I. 1975 Three-dimensional generalized holomorphic vectors. Diff. Equ. 11, 82–87.
Polozhii, G. N. 1973 Theory and application of p-analytic and (p, q)-analytic functions, 2nd edn.
Kiev: Naukova Dumka. [In Russian.]
Pozrikidis, C. 1992 Boundary integral and singularity methods for linearized viscous flow. New York,
NY: Cambridge University Press.
Richardson, S. 1995 Optimum profiles in two-dimensional Stokes flow. Proc. R. Soc. Lond. A 450,
603–622. (doi:10.1098/rspa.1995.0103)
Vekua, I. N. 1962 Generalized analytic functions. Oxford, UK: Pergamon Press.
Wakiya, S. 1967 Slow motion of a viscous fluid around two spheres. J. Phys. Soc. Jpn. 22, 1101–1109.
(doi:10.1143/JPSJ.22.1101)
Williams, W. E. 1965 Integral equation formulation of Oseen flow problems. IMA J. Appl. Math.
1, 339–347. (doi:10.1093/imamat/1.4.339)
Zabarankin, M. 2007 Asymmetric 3D Stokes flows about two fused equal spheres. Proc. R. Soc. A
463, 2329–2349. (doi:10.1098/rspa.2007.1872)
Zabarankin, M. 2008a The framework of k-harmonically analytic functions for three-dimensional
Stokes flow problems, part I. SIAM J. Appl. Math. 69, 845–880. (doi:10.1137/080715913)
Zabarankin, M. 2008b The framework of k-harmonically analytic functions for three-dimensional
Stokes flow problems, part II. SIAM J. Appl. Math. 69, 881–907. (doi:10.1137/080715925)
Zabarankin, M. 2010 Generalized analytic functions in axially symmetric Oseen flows. SIAM J.
Appl. Math. 70, 2473–2508. (doi:10.1137/090776937)
Zabarankin, M. 2011a Generalized analytic functions in magnetohydrodynamics. Proc. R. Soc. A
467, 3343–3370. (doi:10.1098/rspa.2011.0223)
Zabarankin, M. 2011b Minimum-drag shapes in magnetohydrodynamics. Proc. R. Soc. A 467,
3371–3392. (doi:10.1098/rspa.2011.0224)
Zabarankin, M. & Krokhmal, P. 2007 Generalized analytic functions in 3D Stokes flows. Quart. J.
Mech. Appl. Math. 60, 99–123. (doi:10.1093/qjmam/hbl026)
Zabarankin, M. & Molyboha, A. 2010 Three-dimensional shape optimization in Stokes flow
problems. SIAM J. Appl. Math. 70, 1788–1809. (doi:10.1137/090751578)
Zabarankin, M. & Molyboha, A. 2011 3D shape optimization in viscous incompressible fluid under
Oseen approximation. SIAM J. Control Optim. 49, 1358–1382. (doi:10.1137/100790033)
Zabarankin, M. & Nir, A. 2011 Generalized analytic functions in an extensional Stokes flow with
a deformable drop. SIAM J. Appl. Math. 71, 925–951. (doi:10.1137/100797370)
Zabarankin, M. & Ulitko, A. F. 2006 Hilbert formulas for r-analytic functions and Stokes flow
about a biconvex lens. Q. Appl. Math. 64, 663–693. (doi:10.1137/050632403)

Proc. R. Soc. A (2012)

You might also like