You are on page 1of 28

Proc. R. Soc.

A (2011) 467, 3343–3370


doi:10.1098/rspa.2011.0223
Published online 12 July 2011

Generalized analytic functions in


magnetohydrodynamics
BY MICHAEL ZABARANKIN*
Department of Mathematical Sciences, Stevens Institute of Technology,
Castle Point on Hudson, Hoboken, NJ 07030, USA

An approach of generalized analytic functions to the magnetohydrodynamic (MHD)


problem of an electrically conducting viscous incompressible flow past a solid non-
magnetic body of revolution is presented. In this problem, the magnetic field and the
body’s axis of revolution are aligned with the flow at infinity, and the fluid and body
are assumed to have the same magnetic permeability. For the linearized MHD equations
with non-zero Hartmann, Reynolds and magnetic Reynolds numbers (M , Re and Rem ,
respectively), the fluid velocity, pressure and magnetic fields in the fluid and body
are represented by four generalized analytic functions from two classes: r-analytic and
H -analytic. The number of the involved functions from each class depends on whether the
Cowling number S = M 2 /(Rem Re) is 1 or is not 1. This corresponds to the well-known
peculiarity of the case S = 1. The MHD problem is proved to have a unique solution
and is reduced to boundary integral equations based on the Cauchy integral formula
for generalized analytic functions. The approach is tested in the MHD problem for a
sphere and is demonstrated in finding the minimum-drag spheroids subject to a volume
constraint for S < 1, S = 1 and S > 1. The analysis shows that as a function of S, the
drag of the minimum-drag spheroids has a minimum at S = 1, but with respect to the
equal-volume sphere, drag reduction is smallest for S = 1 and becomes more significant
for S  1.
Keywords: generalized analytic function; magnetohydrodynamics;
boundary integral equations; drag

1. Introduction

The steady flow of an electrically conducting viscous incompressible fluid in


the presence of a magnetic field and with neglected thermal effects can be
characterized by three independent parameters: the Hartmann number M ,
Reynolds number Re and magnetic Reynolds number Rem . M is the ratio of
the Lorentz force to the viscous force in the Navier–Stokes equations (when
M = 0, the velocity field is uncoupled from the electromagnetic field), whereas
Rem is interpreted as the ratio of magnetic advection to magnetic diffusion (when
Rem = 0, the magnetic field is uncoupled from the velocity field). The ratio of
*mzabaran@stevens.edu

Received 8 April 2011


Accepted 9 June 2011 3343 This journal is © 2011 The Royal Society
3344 M. Zabarankin

(v)
2.0
(iv)
1.8
(iii)

drag
1.6
(ii)
1.4
(i)
1.2
0 0.5 1.0 1.5 2.0
S

Figure 1. Drag for the unit sphere normalized to the Stokes drag (when Re = Rem = M = 0) of the
sphere as a function of S in five cases: (i) Re = Rem = 1; (ii) Re = 1, Rem = 3; (iii) Re = Rem = 2;
(iv) Re = 3, Rem = 1; and (v) Re = Rem = 3.

the magnetic forces to the inertial forces is characterized by the Cowling number
S (Cowling 1957),1 which can be expressed as M 2 /(Rem Re) when Re = 0 and
Rem  = 0.
A magnetohydrodynamic (MHD) problem that has attracted much theoretical
interest is the one of an electrically conducting flow past a non-magnetic sphere
in the presence of a uniform magnetic field being aligned with the undisturbed
flow at infinity (Chester 1957; Blerkom 1960; Gotoh 1960a,b; Yosinobu 1960;
Goldsworthy 1961; Ludford & Singh 1963). For this problem, figure 1 shows the
drag for the unit sphere normalized to the Stokes drag (when Re = Rem = M = 0)
of the sphere as a function of S in five cases: (i) Re = Rem = 1; (ii) Re = 1, Rem = 3;
(iii) Re = Rem = 2; (iv) Re = 3, Rem = 1; and (v) Re = Rem = 3. At S = 1, the drag
attains minimum and is non-smooth (Gotoh 1960a; Goldsworthy 1961). Namely,
the fact that the sphere’s drag at S = 1 is non-smooth is remarkable. It poses
the following research questions. If the sphere is replaced by an arbitrary non-
magnetic body of revolution, what is the body’s shape that has the smallest drag
subject to a volume constraint and how does it depend on S? Are there any
qualitative differences in the minimum-drag shapes for S < 1, S = 1 and S > 1?
How does drag reduction depend on S? Answering these questions is the subject
of this paper and the follow-up work (Zabarankin 2011).
The challenge in addressing the posed questions is that the traditional partial
differential equation (PDE)-constrained optimization approach coupled with the
finite-element method (FEM) is slowly converging and inaccurate in general.
The second deficiency is attributed to the fact that being applied to external
problems, the FEM truncates and discretizes an external domain, and minimum-
drag shapes are found point-wise. Our approach consists of two parts: (i) reducing
the MHD problem to boundary integral equations and (ii) deriving the optimality
condition for the minimum-drag shapes analytically and obtaining minimum-
drag shapes in a function series form (see Zabarankin & Molyboha (2010, 2011)

1 Thisnumber is also known as the ratio of the magnetic pressure to the dynamic pressure and is
denoted by b in Blerkom (1960) and by S in Yosinobu (1960) and Gotoh (1960a).

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3345

for a discussion of boundary integral equation-constrained optimization versus


FEM-based PDE-constrained optimization). This paper addresses the first part,
whereas Zabarankin (2011) addresses the second.
In the proposed approach, the MHD equations are linearized, and the velocity,
pressure and magnetic fields in the fluid and body of revolution are represented
by four generalized analytic functions. Under the assumption that the fluid and
body are both non-magnetic with the same magnetic permeability, the axially
symmetric MHD problem is reduced to integral equations for the boundary values
of the two involved generalized analytic functions based on the generalized Cauchy
integral formula. The Hartmann number is assumed to be non-zero since when
M = 0, the velocity field is uncoupled from the magnetic field. With M = 0, three
cases are studied separately: (i) Rem = 0, Rem Re = M 2 (S = 1); (ii) Rem = 0; and
(iii) Rem Re = M 2 (S = 1). The reason for this is that in each case, solving the
MHD problem involves different numbers of generalized analytic functions from
two classes: r-analytic and H -analytic. This is the mathematical explanation of
why the case S = 1 is special.

(a) r-analytic and H -analytic functions



A function G(x, y) = U (x, y) + i V (x, y) with x, y ∈ R and i = −1 is called
generalized analytic or pseudo-analytic if its real and imaginary parts, U and
V , respectively, satisfy the so-called Carleman or Bers–Vekua system (Bers 1953;
Vekua 1962)
vU vV vU vV
− + aU + bV = 0 and + + cU + dV = 0, (1.1)
vx vy vy vx
where a = a(x, y), b = b(x, y), c = c(x, y) and d = d(x, y) are real-valued
functions. For example, for a ≡ b ≡ c ≡ d ≡ 0, the system (1.1) defines ordinary
analytic functions of a complex variable x + iy. Generalized analytic functions
arise in various areas of applied mathematics including hydrodynamics,
gas dynamics, theory of elasticity, heat transfer, electromagnetism, quantum
mechanics, etc. (Bers 1958; Vekua 1962; Polozhii 1973; Alexandrov & Soloviev
1978). Their theory has been extensively developed since the mid-twentieth
century and extends the majority of results for ordinary analytic functions, e.g.
the formal powers and the Cauchy integral formula (Bers 1953, 1956; Vekua 1962;
Polozhii 1973; Chemeris 1995).
Several important classes of generalized analytic functions arise from the
relationship
curl F + [a × F] = −grad J and div F = 0 (1.2)
for a vector field F and scalar field J, where a is a known vector function.
This relationship is frequently encountered in problems of applied mathematics.
For example, for J = 0 and a = 0, equation (1.2) simplifies to an irrotational
solenoidal field F found in electrostatics and an ideal fluid, whereas for a = 0, it
defines the so-called related potentials J and F, e.g. the pressure and vorticity in
the Stokes equations, electric potential and magnetic field in conductive materials,
etc. (Zabarankin & Krokhmal 2007). For a  = 0 and J = 0, equation (1.2) arises
in the Maxwell equations for quasi-stationary electromagnetic fields, in the Oseen
equations for viscous incompressible fluid (Zabarankin 2010), and in the linearized
MHD equations as will be shown in this paper.

Proc. R. Soc. A (2011)


3346 M. Zabarankin

Let (r, 4, z) be a cylindrical coordinate system with the basis (er , e4 , k). In
the axially symmetric case with the z-axis being the axis of revolution, let F =
Fr (r, z)er + Fz (r, z)k, J = 0 and a = −2lk, where l is a real-valued constant.
Then, for U = e−lz Fz and V = e−lz Fr , equation (1.2) reduces to
     
vU v v v 1
= − l V and +l U =− + V, (1.3)
vr vz vz vr r
which is a special case of equation (1.1) and implies that
(D0 − l2 )U = 0 and (D1 − l2 )V = 0, (1.4)
where Dk denotes the so-called k-harmonic operator, Dk ≡ v2 /vr 2 + r −1 v/vr +
v2 /vz 2 − k 2 /r 2 . The functions U and V satisfying equation (1.3) form an
H -analytic function G = U + iV of a complex variable z = r + i z. For l = 0,
equation (1.3) defines a so-called r-analytic function and is arguably the most
studied system among various classes of generalized analytic functions. It arises in
the axially symmetric theory of elasticity (Polozhii 1973; Alexandrov & Soloviev
1978) and in the axially symmetric Stokes and Oseen flows (Zabarankin & Ulitko
2006; Zabarankin 2008, 2010). Since an r-analytic function satisfies system (1.4)
for l = 0, it is also referred to as the 0-harmonically analytic function (Zabarankin
2008). Both classes of r-analytic and H -analytic functions will be instrumental
in constructing solutions to axially symmetric MHD problems.

(b) The generalized Cauchy integral formula and series representation


Let G + be a generalized analytic function in a bounded open region D+ in
the right-half rz-plane (D+ may contain parts of the z-axis). The boundary of
D+ is assumed to be a piece-wise smooth positively oriented curve , which is
either closed or open with the endpoints lying on the z-axis.2 Let D− be the
complement of D+ ∪  in the right-half rz-plane (D− is unbounded), and let G − be
a generalized analytic function in D− that vanishes at infinity. For convenience, an
arbitrary function f (r, z) will be denoted by f (z) without assuming its analyticity.
Let G ± (z) satisfy the Hölder condition3 on , and let  be the reflection of  over
the z-axis. Then, under the assumption of the symmetry condition G ± (−z̄) =
G ± (z), the Cauchy integral formula for G ± is given by
1
G ± (z) = ±  G ± (t)W(z, t) dt, z ∈ D± , (1.5)
2pi  

where z = r + iz, t = r1 + iz1 and W(z, t) ≡ W(r, z; r1 , z1 ) is a generalized Cauchy


kernel (Vekua 1962; Polozhii 1973; Alexandrov & Soloviev 1978). If the boundary
 is smooth, then on , G ± (z) satisfies the generalized Sokhotski–Plemelj formula,

1 1
G ± (z) = G ± (z) ±  G ± (t)W(z, t) dt, z ∈ . (1.6)
2 2pi  

2 Open segments of the z-axis that D + may contain are not parts of .
3 This condition means that for some parametrization z(t) of , the boundary value f (z(t)) satisfies
|f (z(t2 )) − f (z(t1 ))| ≤ c|t2 − t1 |g for all t1 and t2 , some g ∈ (0, 1] and non-negative constant c.

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3347

If  is piece-wise smooth (has salient points), in particular if the endpoints of 


lie on the z-axis and the angle between  and the z-axis at one of the endpoints is
not p/2 (conic endpoint), then at a salient point of , the coefficient 1/2 at G ± (z)
in equation (1.6) is replaced by a finite function of the angle between tangents at
the salient point (see Alexandrov & Soloviev (1978, systems 31.13a, 31.13b) and
Zabarankin & Nir (2011, system 3.7)).
Let Wr and WH denote the generalized Cauchy kernels for r-analytic and
H -analytic functions, respectively. For r-analytic functions, Wr is given in
Alexandrov & Soloviev (1978), Zabarankin (2008, theorem 2) and Zabarankin &
Nir (2011, theorem 3.2). Theorem 1 and corollary 4 in Zabarankin (2010) present
two forms for WH with l ≥ 0. The next result extends the Cauchy integral formula
for H -analytic functions for arbitrary real-valued l.

Theorem 1.1. Let G ± be an H -analytic function in D± for some real-valued


l (G − vanishes at infinity) and let it satisfy the Hölder condition on . Suppose
G ± (−z̄) = G ± (z). Then, the generalized Cauchy integral formula (1.5) takes the
form

1
G ± (z) = ±  G(t)WH (z, t, l) dt
2pi  

1
=± U ± (t)elz1 dt K1 (z, t, l) + iV ± (t)r1 e−lz1 dt K2 (z, t, l), (1.7)
2pi 

for z ∈ D± , where dt = v/vr1 dr1 + v/vz1 dz1 ,


 p 
−lz1 i(r cos t − r1 )
K1 (z, t, l) = r1 e − cos t
0 z1 − z + f(l)r(z, t, t)

e−|l| r(z,t,t)
× dt + i e−lz C (t, z),
r(z, t, t)
p   −|l| r(z,t,t)
i (r1 cos t − r) e elz
K2 (z, t, l) = e lz1
−1 dt + i C (z, t),
0 z − z1 + f(l)r(z, t, t) r(z, t, t) r
p
C (z, t) = (f(l) + sign(r − r1 )(f(l) − sign(z − z1 ))
2

and r(z, t, t) = r 2 + r12 − 2rr1 cos t + (z − z1 )2 ,

with f(l) being defined as f(l) = 1 for l ≥ 0 and f(l) = −1 for l < 0. The functions
K1 and K2 are continuous provided that there is a branch cut connecting t and −t̄
and arg(z − t) ∈ [0, 2p].

Proof. The proof is similar to that of theorem 1 in Zabarankin (2010). 

For the region exterior to a sphere, example 5 and proposition 2 in Zabarankin


(2010) provide series representations for r-analytic functions and H -analytic
functions with l > 0, respectively. The next proposition generalizes the series
representation for H -analytic functions for any non-zero real-valued l.

Proc. R. Soc. A (2011)


3348 M. Zabarankin

Proposition 1.2 (H -analytic function in the region exterior to a sphere). Let


(R, w, 4) be a system of spherical coordinates. For the region exterior to a sphere
centered at the system’s origin, an H -analytic function satisfying equation (1.3)
for real-valued l = 0 and vanishing at infinity can be represented by

1 
H (R, w) = √ An (Ln (cos w) Kn+ 1 (|l|R) − (sign l)L−n (cos w) Kn− 1 (|l|R)),
2 2
R n=1
(1.8)
(1) (m)
where Ln (t) = nPn (t) − iPn (t) with Pn (t), t ∈ [−1, 1], being the associated
Legendre polynomial of the first kind of order n and rank m (for m = 0, the
superscript is omitted), Kn+ 1 (·) is a modified spherical Bessel function of the
2
third kind (see Bateman & Erdelyi 1953, §7.2.6), and real-valued coefficients An
are such that the series (1.8) converges for all w ∈ [0, p] and R greater than or
equal to the radius of the sphere.
Proof. Proved similarly to proposition 2 in Zabarankin (2010). 
The rest of this work is organized into three sections. Section 2 constructs
representations for the velocity field, pressure and magnetic fields in the fluid
and body in terms of four generalized analytic functions. Section 3 shows that in
all three cases (i)–(iii), the corresponding boundary-value problems for the four
involved functions have unique solutions and can be reduced to boundary integral
equations based on the Cauchy integral formula for generalized analytic functions.
Section 4 demonstrates the approach of the generalized analytic functions in
solving the MHD problem for a sphere and in finding minimum-drag prolate
spheroids subject to a volume constraint.

2. Magnetohydrodynamics

The steady flow of an electrically conducting viscous incompressible fluid is


governed by the (stationary) MHD equations

r(U · grad)U = −grad ℘ + r nDU + m [J × H],⎪

div U = 0, J = s(E + m [U × H]), curl E = 0, (2.1)


div E = q/e, curl H = J and div H = 0,
where DU = grad div U−curl curl U, U is the fluid velocity, ℘ is the pressure
in the fluid, E and H are the electric and magnetic fields, respectively, J is the
current density, q is the excess charge density, n is the kinematic viscosity, r is the
fluid density, m is the magnetic permeability, s is the electric conductivity and e
is the dielectric constant (see Cowling 1957; Cabannes 1970; Chester 1957). All
constants are measured in metre, kilogram and second (MKS) units.
Suppose a non-magnetic solid body of revolution is immersed in a uniform flow
in the presence of a uniform magnetic field. It is assumed that in the cylindrical
coordinate system (r, 4, z), the body’s axis of revolution is parallel to the z-axis,
and that at infinity, the flow and magnetic fields are constant and aligned with the
z-axis, i.e. U|∞ = V∞ k and H|∞ = H∞ k, where V∞ and H∞ are constants. Under
different approximations of the nonlinear equations (2.1), this MHD problem

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3349

was considered in Chester (1957, 1961, 1962), Gotoh (1960a,b), Imai (1960) and
Stewartson (1956). Let u be the disturbance of the velocity field: U = V∞ (k + u),
and let h+ and h− be the disturbances of the magnetic field in the body and fluid,
respectively: H = H∞ (k + h+ ) in the body and H = H∞ (k + h− ) in the fluid. On
the surface S of the body and at infinity, u and ℘ satisfy the conditions
u|S = −k, u|∞ = 0 and ℘|∞ = 0. (2.2)
In the body, the magnetic field satisfies curl H = 0 and div H = 0, or in terms
of the disturbance h+ ,
curl h+ = 0 and div h+ = 0. (2.3)
As in Chester (1957) and Gotoh (1960b), it is assumed that the fluid and body
have the same magnetic permeability. In this case, the magnetic field is continuous
across the surface of the body4 and also h− vanishes at infinity,
h+ |S = h− |S and h− |∞ = 0. (2.4)
The problem (2.1)–(2.4) is axially symmetric. In this case, the velocity, pressure
and electromagnetic field are independent of the angular coordinate 4, i.e.
u = ur (r, z)er + uz (r, z)k, ℘ = ℘(r, z),
h± = hr± (r, z)er + hz± (r, z)k, and E = E4 (r, z)e4 .
The equation curl E = 0 implies rE4 = c, where c is a constant. Since in the fluid,
[U × H] and curl H vanish at infinity, we have c = 0, and thus, E ≡ 0 everywhere.
Rescaling the linear dimensions, velocity, pressure and magnetic field by a, V∞ ,
V∞ rn/a and H∞ , respectively, where a is the half of the diameter of the body,
and assuming u and h− to be small, we can rewrite equations (2.1) without E
and J in the linearized dimensionless form

Re(k · grad)u = −grad ℘ + Du + M 2 [[(u − h− ) × k] × k],⎪

curl h− = Rem [(u − h− ) × k], (2.5)


div u = 0 and div h− = 0,

where Re = V∞ a/n is the Reynolds number, M = mH∞ a s/(rn) is the Hartmann
number and Rem = V∞ ams is the magnetic Reynolds number.
For M = 0, the first equation in (2.5) reduces to the Oseen equations, and
the problem for the velocity and pressure becomes uncoupled from the magnetic
field. For solving the Oseen equations, see Zabarankin (2010). For M = 0, there
are three cases to analyse:

(a) Rem  = 0 and Rem Re = M 2 (S = 1): a solution to (2.5) is represented by


two H -analytic functions and one r-analytic function, whereas a solution
to equations (2.3) is given by a single r-analytic function;
(b) Rem = 0: the magnetic field is constant everywhere, i.e. h± = 0, and a
solution to equations (2.5) is represented by two H -analytic functions (the
case of Re = 0 yields no simplification); and
4 If
the fluid and body have different magnetic permeabilities, then across the body’s surface, the
tangential component of the magnetic field H and the normal component of the magnetic induction
B should be continuous.

Proc. R. Soc. A (2011)


3350 M. Zabarankin

(c) Rem Re = M 2 (S = 1): a solution to equations (2.5) is represented by


one H -analytic function and two r-analytic functions, and a solution to
equations (2.3) is given by a single r-analytic function.

Let D+ and D− be the regions occupied by the body and fluid, respectively,
with S being their common boundary (the surface of the body), and let D± be
the open region corresponding to the interior of the cross section of D± in the
right-half rz-plane (r ≥ 0). In this case,  denotes the positively oriented common
boundary of D+ and D− , i.e.  is the cross section of S in the right-half rz-plane
and is either a closed curve or has the endpoints on the z-axis, and  denotes the
reflection of  over the z-axis.
Theorem 2.1 (solution representation, case (a)). In the axially symmetric case
with M  = 0 and Rem Re = M 2 , a solution to equations (2.5) and (2.3) is given by
1 − 2l2 8 l1 z 1 − 2l1 8 l2 z 1
uz + iur = e G1 − e G2 − G −, (2.6)
l1 − l 2 l1 − l 2 Rem Re − M 2 3
28 1
hz− + i hr− = − (l2 el1 z G1 − l1 el2 z G2 ) − G− (2.7)
l1 − l 2 Rem Re − M 2 3
and hz+ + ihr+ = G3+ , (2.8)
where G1 and G2 are H -analytic functions in D− satisfying system (1.3) with
l = l1,2 = (Re + Rem ± (Re − Rem )2 + 4M 2 )/4, respectively, and vanishing at
infinity; G3+ and G3− are r-analytic functions in D+ an D− , respectively, with
G − vanishing at infinity; and 8 = Rem /(Rem Re − M 2 ).
In this case, the pressure and scalar vortex function u (u = curl u = u e4 is the
vorticity) are determined by

2(Re 8 − 1) Re −
℘ = Re (l2 e G1 − l1 e G2 ) +
l1 z l2 z
G (2.9)
l1 − l 2 Rem Re − M 2 3
and
1
u= Im[(Re − 2l2 )el1 z G1 − (Re − 2l1 )el2 z G2 ]. (2.10)
l1 − l 2
Proof. With u = curl u and the identity (k · grad)u = −curl[k × u], the first
equation in (2.5) can be rewritten in the form
curl(u − Re[k × u]) − M 2 [k × [k × (u − h− )]] + grad ℘ = 0. (2.11)
The continuity equation div u = 0 implies that u and u satisfy the identity
curl[k × u] + grad(k · u) − [k × u] = 0, (2.12)
− −
whereas div h = 0 implies a similar identity for h ,
[k × curl h− ] = curl[k × h− ] + grad(k · h− ),
with which the second equation in (2.5) can be represented in the form
curl[k × h− ] + grad(k · h− ) + Rem [k × [k × (u − h− )]] = 0. (2.13)

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3351

Forming a linear combination of (2.11), (2.12) and (2.13) with constant weights
b1 , b2 and b3 , respectively, we have
curl(b1 u + [k × ((b2 − b1 Re)u + b3 h− )])
− [k × (b2 u + (b1 M 2 − b3 Rem )[k × (u − h− )])]
+ grad(b1 ℘ + b2 (k · u) + b3 (k · h− )) = 0. (2.14)

Let b3 = −(b2 − b1 Re) and (b2 − b1 Re)/b1 = (b1 M 2 − b3 Rem )/b2 , then (b2 /
b1 )2 − (Re + Rem )b2 /b1 + Rem Re − M 2 = 0, hence b2 /b1 = 2lk , k = 1, 2, and
equation (2.14) becomes a particular case of system (1.2),
curl Lk − 2lk [k × Lk ] + grad Jk = 0 and div Lk = 0, k = 1, 2, (2.15)
where 
Lk = u + (2lk − Re)[k × (u − h− )]
(2.16)
and Jk = ℘ + 2lk (k · u) + (Re − 2lk )(k · h− ).

In the axially symmetric case, Lk = Lk e4 , k = 1, 2, and system (2.15) reduces to


system (1.3) determining H -analytic functions Gk ,
Jk + iLk = 2elk z Gk , k = 1, 2, (2.17)
with lk defined in the theorem.
On the other hand, if b2 = 0 and b1 M 2 − b3 Rem = 0, then equation (2.14)
becomes the relationship for related potentials,
curl L3 + grad J3 = 0 and div L3 = 0, (2.18)
where
L3 = Rem u + [k × (M 2 h− − Rem Re u)] and J3 = Rem ℘ + M 2 (k · h− ). (2.19)
In the axially symmetric case, L3 = L3 e4 and system (2.18) determines an
r-analytic function G3− in D− ,
J3 + iL3 = G3− . (2.20)
Equations (2.15)–(2.20) can be reformulated in the complex form
2elk z Gk = ℘ + iu + 2lk (uz + iur ) − iRe ur + (Re − 2lk )(hz− + ihr− ), k = 1, 2,
and
G3− = Rem (℘ + iu) − iRe Rem ur + M 2 (hz− + ihr− ),
from which the representations (2.6), (2.7), (2.9) and (2.10) follow. The condition
M = 0 guarantees that l1 = l2 .
Finally, in the axially symmetric case, h+ , being an irrotational solenoidal field
(see equations (2.3)), can be represented in the form (2.8). 
Corollary 2.2 (solution representation, case (b)). In the axially symmetric case
with M = 0 and Rem = 0, the magnetic disturbance is zero everywhere: h± ≡ 0, and

Proc. R. Soc. A (2011)


3352 M. Zabarankin

the velocity, pressure and vorticity that solve equations (2.5) can be represented by

1 ⎫
uz + iur = (e G1 − e G2 )
l1 z l2 z ⎪

l1 − l 2 ⎬
    (2.21)
Re Re ⎪
and ℘ + iu = el1 z G1 − G 1 + el2 z G2 + G 2 ,⎪

2(l1 − l2 ) 2(l1 − l2 )
where √G1 and G2 are H -analytic functions satisfying system (1.3) with l = l1,2 =
(Re ± Re 2 + 4M 2 )/4, respectively, and vanishing at infinity.
Detail. For Rem = 0, the representation (2.7) reduces to hz− + i hr− = G3− /M 2 ,
which with the representation (2.8) and the boundary condition (2.4) implies
G3− /M 2 = G3+ on . Consequently, as G3− vanishes at infinity, by the Sokhotski–
Plemelj formula (1.6), G3± ≡ 0 in D± , respectively, and the representation (2.21)
follows from the representations (2.6), (2.9) and (2.10).
Corollary 2.3. For M  = 0 and Rem = Re = 0, representation (2.21) simplifies to
1 Mz/2
uz + iur = (e G1 − e−Mz/2 G2 ) and ℘ + iu = eMz/2 G1 + e−Mz/2 G2 ,
M
where G1 and G2 are H -analytic functions satisfying system (1.3) with l = l1,2 =
±M /2, respectively. This representation is similar to the solution form suggested
by Chester (1957) in the case of Re = Rem = 0.
Corollary 2.4. For M = 0 and Re  = 0, representation (2.21) reduces to
representations (22) and (23) in Zabarankin (2010) for the velocity, pressure and
vorticity for the axially symmetric Oseen flow of a non-conducting fluid.
Theorem 2.5 (solution representation, case (c)). In the axially symmetric case
with M = 0 and Rem Re = M 2 , a solution to equations (2.5) and (2.3) is given by
     
1 2Re elz G1 i Re −
uz + iur = + Rem z − r − G2 + Rem G3 ,
Re + Rem Re + Rem 2 Re + Rem
(2.22)
   
− − Rem 2elz G1 i 1 −
hz + i hr = − + z− r+ G2 + G 3
Re + Rem Re + Rem 2 Re + Rem
(2.23)
and hz+ + ihr+ = G3+ , (2.24)
where l = (Re + Rem )/2; G1 is an H -analytic function in D− that satisfies
system (1.3) with l and vanishes at infinity; G2 and G3− are r-analytic functions
in D− that vanish at infinity; and G3+ is an r-analytic function in D+ .
In this case, the pressure and vortex functions are determined by
   
Rem Re 2elz G1 i 1 −
℘ = Re − z− r+ G2 − G 3 + G 2
Re + Rem Re + Rem 2 Re + Rem
(2.25)
and
1
u= Im[2Re elz G1 + Rem G2 ]. (2.26)
Re + Rem

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3353

Proof. The proof is partially based on the proof of theorem 2.1. When Rem Re =
M 2 , relationships (2.15)–(2.17) hold for l1 = (Re + Rem )/2 and l2 = 0. For l1 =
(Re + Rem )/2, equation (2.17) simplifies to

2e(Re+Rem )z/2 G1 = ℘ + iu + (Re + Rem )(uz + iur ) − iRe ur − Rem (hz− + ihr− ),
(2.27)
where G1 is the H -analytic function defined in this theorem. The case of
l2 = 0 means that equation (2.15) reduces to relationship (2.18) for the related
potentials. Consequently, a different approach is required. As Rem Re = M 2 ,
multiplying equation (2.13) by Re and adding to equation (2.11), we obtain
curl(u − Re [k × (u − h− )]) + grad(℘ + Re (k · h− )) = 0,
which with the second equation in equations (2.5) and conditions div u = 0 and
div h− = 0 can be rewritten as
D(u + (Re/Rem )h− ) = grad(℘ + Re(k · h− )) and div(u + (Re/Rem )h− ) = 0.
(2.28)
System (2.28) is similar to the Stokes equations for a viscous incompressible
fluid. In the axially symmetric case, its solution is given by proposition 7 in
Zabarankin (2008):
  ⎫
Re − i −⎪
uz + iur + (h + ihr ) = z − r G2 + G3 ⎬

Rem z 2 (2.29)
− −


and ℘ + iu − iRe ur + Re(hz + i hr ) = G2 ,

where G2 and G3− are the r-analytic functions defined in this theorem.
Observe that the conditions M = 0 and Rem Re = M 2 guarantee that Rem =
0. The representations (2.22), (2.23), (2.25) and (2.26) follow from equations
(2.27) and (2.29), whereas the representation (2.24) remains the same as in
theorem 2.1. 
The advantage of the representations (2.6)–(2.10) and (2.21)–(2.26) compared
with the existing solution forms, e.g. in Chester (1957); Gotoh (1960a,b) and
Yosinobu (1960), is that these representations, being linear combinations of
the generalized analytic functions, involve no derivates of those functions and
simultaneously represent the velocity, vorticity, pressure and magnetic fields in
the fluid and body. This fact considerably simplifies reducing the MHD problem
to boundary integral equations based on the generalized Cauchy integral formula.

3. Boundary-value problems and integral equations

(a) Case (a): M = 0, Rem  = 0 and Rem Re = M 2


For M  = 0, Rem = 0 and Rem Re = M 2 , the MHD problem (2.2)–(2.5) reduces to
the boundary-value problem for the four generalized analytic functions G1 , G2

Proc. R. Soc. A (2011)


3354 M. Zabarankin

and G3± defined in theorem 2.1,



(1 − 2l2 8)el1 z G1 − (1 − 2l1 8)el2 z G2 ⎪


l1 − l 2 ⎬
− G3− = l2 − l1 , z ∈ , (3.1)
Rem Re − M 2 ⎪



+
and e G1 − e G2 = (l2 − l1 )(G3 + 1), z ∈ .
l1 z l2 z

Proposition 3.1. The boundary-value problem (3.1) has a unique solution.


Proof. The proposition is equivalent to the fact that the homogenous
problem (3.1) has only zero solution, which with the representations (2.6)–(2.10)
corresponds to (2.5) with the zero boundary conditions, i.e. u|S = 0.
The first equation in (2.5) can be recast in the form
vu
curl u + grad ℘ + Re − M 2 [k × [k × (u − h− )]] = 0. (3.2)
vz
With the identities
u · curl u = div(u × u) + |u|2
and
u · grad ℘ = div(℘ u) − ℘ div u = div(℘ u),
the scalar product of equation (3.2) and u takes the form
 
Re 2
div [u × u] + ℘u + |u| k + |u|2 + M 2 ur (ur − hr− ) = 0. (3.3)
2
On the other hand, with the identity
 
− − 1 −2 − −
[k × h ] · curl h = div |h | − h (k · h ) , (3.4)
2
the scalar product of the second equation in equations (2.5) and [k × h− ] results in
 
− − 1 1 −2 − −
hr (hr − ur ) − div |h | − h (k · h ) = 0, (3.5)
Rem 2
provided that Rem = 0. The linear combination (3.3) +M 2 (3.5) reduces to
  
Re 2 M2 1 − 2 − −
div [u × u] + ℘u + |u| k − |h | − h (k · h )
2 Rem 2
+ |u|2 + M 2 (ur − hr− )2 = 0. (3.6)
Since curl h+ = 0 in D+ , we can also write
 
+ + 1 +2 + +
[k × h ] · curl h = div |h | − h (k · h ) = 0. (3.7)
2
Let DR be the region bounded by the body’s surface S and by a sphere SR with
large radius R and centre at the origin, so that DR → D− as R → ∞. Applying
the divergence theorem to the linear combination of the volume integrals,

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3355
 
− M 2 /Rem
DR (3.6) dV D+ (3.7) dV , and using the boundary conditions u = 0
+ −
and h = h on S , we obtain

IR + (|u|2 + M 2 (ur − hr− )2 ) dV = 0, (3.8)
DR

where
   
Re 2 M2 1 −2
IR = n · [u × u] + ℘u + |u| k − |h | − h− (k · h− ) dS ,
SR 2 Rem 2
with n being the outward normal. Note that equation (3.8) holds for multiply
connected DR , since u, u, ℘ and h± are all single-valued functions.5
Next we show that IR → 0 as R → ∞. Let CR be the positively oriented cross
section of SR in the right-half rz-plane, i.e. CR is a semicircle, and let v/vs and
v/vn be the tangential and normal derivatives for CR , respectively. With the
identities
vr vz vr vz
= and =− , (3.9)
vs vn vn vn
and the surface element dS = r ds d4, where ds is the differential of the length of
CR , we have
    
Re M2 − − 2
IR = 2pRe (uz + iur ) ℘ + iu + (uz − iur ) + (h + ihr ) r dz .
CR 2 2Rem z
(3.10)
Now with the representations (2.6)–(2.7) and (2.9)–(2.10), the integral (3.10)
reduces to
  
(G3− )2
IR = 2pRe k1 e G1 + k2 e G2 −
2l1 z 2 2l2 z 2
r dz ,
CR 2Rem (Rem Re − M 2 )
where k1 = (Re/2 − 2l2 + 28l22 )/(l1 − l2 )2 and k2 = (Re/2 − 2l1 + 28l12 )/
(l1 − l2 )2 . Observe that l1 = l2 as M  = 0.
The representation (1.8) implies that in the spherical coordinates (R, w, 4)
related to the cylindrical coordinates in the ordinary way, Gi (R, w) =
fi (w)R−1 e−|li |R + O(R−2 e−|li |R ) as R → ∞ for i = 1, 2, where fi (w) is a bounded
complex-valued function with |fi (w)| < K for some constant K and w ∈ [0, p],
whereas example 5 in Zabarankin (2010) implies that G3− = O(R−2 ) as R → ∞.
Thus, with z = Rei w , w ∈ [0, p], we can evaluate
⎛ ⎞
p  2
|IR | ≤ 2pK 2 ⎝ |kj |e−2R(|lj |−lj cos w) ⎠ sin w dw + o(1)
0 j=1

2
|kj |
= 2pK 2 (1 − e−4|lj |R ) + o(1) → 0 as R → ∞,
j=1
2|l j |R

5 If D is multiply connected, we can make crosscuts in D to make D simply connected, and


R R R
since u, u, ℘ and h± are single valued, they have the same values on the banks of a crosscut.

Proc. R. Soc. A (2011)


3356 M. Zabarankin

where li  = 0 provided that Rem Re = M 2 . Consequently, passing R to infinity in


− 2
D− (|u| + M (ur − hr ) ) dV = 0 so that u = 0 and
2 2
equation (3.8), we obtain
ur = hr− in D− , which along with the representations (2.6), (2.7) and (2.10) imply
that Im Gj = 0, j = 1, 2, in D− . It follows from the system (1.3) that Re Gj =
cj e−lj z , i = 1, 2, where ci are real-valued constants. But since G1 and G2 vanish
at infinity, cj = 0, j = 1, 2, and thus, G1 = 0 and G2 = 0 in D− . Then, the first and
second equations of the homogeneous problem (3.1) imply G3− = 0 and G3+ = 0 on
, and it follows from the Cauchy integral formula (1.5) for r-analytic functions
that G3± = 0 in D± , respectively. 
The next proposition is considered from the mathematical point of view only. It
will be used in determining homogeneous solutions to boundary integral equations
that follow from the boundary-value problem (3.1).
Proposition 3.2 (homogeneous conjugate boundary-value problem, case (a)).
Let G1 and G2 be H -analytic functions in D+ , and let G3± be r-analytic functions
in D± , respectively, with G3− vanishing at infinity. Under the assumptions M = 0,
Rem  = 0 and Rem Re = M 2 , the homogeneous conjugate boundary-value problem

l1 − l 2 ⎪
(1 − 2l2 8)e G1 − (1 − 2l1 8)e G2 −
l1 z l2 z
G3 = 0, z ∈ ,⎬
+
Rem Re − M 2
(3.11)



and e G1 − e G2 − (l2 − l1 )G3 = 0, z ∈ ,
l1 z l2 z

has the solution


G1 = c e−l1 z , G2 = c e−l2 z , G3+ = 2Rem c and G3− = 0, (3.12)
where c is an arbitrary real-valued constant, and l1 , l2 are defined in theorem 2.1.
Proof. The formulae (2.6)–(2.10), in which G3+ and G3− are interchanged and
G1 , G2 and G3± are defined in this proposition, represent u, ℘, u and h+ that
satisfy equations (2.5) in D+ and represent h− that satisfy equations (2.3) in
D− . In this case, equations (2.5) in D+ and (2.3) in D− will be called conjugate
equations for equations (2.5) in D− and (2.3) in D+ . The homogeneous boundary-
value problem (3.11) corresponds to the conjugate equations with the boundary
conditions u = 0 and h+ = h− on S , where h− vanishes at infinity. To find a
solution to this problem, we repeat the proof of proposition 3.1 so that

IR + (|u|2 + M 2 (ur − hr+ )2 ) dV = 0, (3.13)
D+
where   
M2 1 −2  
IR = − n· |h | − h− k · h− dS .
Rem SR 2
As in the proof of proposition 3.1, it can be shown that |IR | → as R → ∞,
and passing R to infinity in equation (3.13), we have u = 0 and ur = hr+ in
D+ . Then, the representations (2.6)–(2.10) with interchanged G3+ and G3− imply
that Im G1 = 0 and Im G2 = 0 in D+ , and it follows from system (1.3) that
Re Gj = cj e−lj z , i = 1, 2, where ci are real-valued constants. The second equation
in the problem (3.11) implies G3− = (c1 − c2 )/(l2 − l1 ) on , and by the Cauchy

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3357

integral formula (1.5) for r-analytic functions, we obtain G3− = (c1 − c2 )/(l2 − l1 )
in D− . In this case, G3− vanishes at infinity only if c1 = c2 , and consequently,
G3− = 0 in D− . Similarly, the first equation in (3.11) implies G3+ = 2Rem c on ,
where c = c1 = c2 , and by the generalized Cauchy integral formula (1.5), we have
G3+ = 2Rem c in D+ . 
In general, as in the classic theory of boundary-value problems for ordinary
analytic functions (Muskhelishvili 1992), the number of unknown generalized
analytic functions should be equal to the number of real-valued boundary
conditions, and the problem (3.1) can be reduced to integral equations for the
boundary values of G1 and G2 based on the generalized Cauchy integral formula.
Theorem 3.3 (boundary integral equations, case (a)). Let M = 0, Rem = 0 and
Rem Re  = M 2 . The boundary-value problem (3.1) yields two integral equations for
the boundary values of Fk (z) = elk z Gk (z), k = 1, 2,
1
(Wr (z, t) − elk (z−z1 ) WH (z, t, lk ))Fk (t) dt
2pi ∪
1 − 28lk 1
+ (F1 (z) − F2 (z)) = − , z ∈ , k = 1, 2, (3.14)
28(l1 − l2 ) 28
where l1 , l2 are defined in theorem 2.1. A solution to equations (3.14) is determined
up to a real-valued constant c, i.e. Fk (z) = c, k = 1, 2, is a homogeneous solution to
equations (3.14). Let F̂k (z), k = 1, 2, solve equations (3.14), then G1 (z) and G2 (z)
on  are determined by

Gk (z) = (F̂k (z) − c)e−lk z , k = 1, 2, z ∈ , (3.15)


where
1 1
c = F̂k (z) +  F̂k (t)elk (z−z1 ) WH (z, t, lk ) dt, z ∈ . (3.16)
2 2pi  

Proof. The necessary and sufficient conditions for G1 (z), G2 (z) and G3± (z),
z ∈ , to be boundary values for the corresponding generalized analytic functions
are given by the generalized Sokhotski–Plemelj formulae (1.6)
1 1
Gk (z) + Gk (t)WH (z, t, lk ) dt = 0, k = 1, 2, z ∈ , (3.17a)
2 2pi ∪

and
1 ± 1
G3 (z) ∓ G3± (t)Wr (z, t) dt = 0, z ∈ . (3.17b)
2 2pi ∪

Expressing G3+ (z) and G3− (z) on  from the boundary-value problem (3.1) in
terms of G1 (z) and G2 (z), then substituting them into corresponding equations
in (3.17b) and solving the latter for the integral terms, we have
1 − 28 lk 1 1 1
(F1 (z) − F2 (z)) + Fk (z) + Fk (t)Wr (z, t) dt = − ,
28(l1 − l2 ) 2 2pi ∪ 28

Proc. R. Soc. A (2011)


3358 M. Zabarankin

for k = 1, 2 and z ∈ . Then, equation (3.17a) is multiplied by elk z and is recast in


terms of Fk (z),

1 1
Fk (z) + Fk (t)elk (z−z1 ) WH (z, t, lk ) dt = 0, k = 1, 2, z ∈ .
2 2pi ∪

Subtracting this equation from the previous one for corresponding k, we obtain
equations (3.14).
Now we will show that the solution to the boundary-value problem (3.1) is
determined by the formulae (3.15) and (3.16) and that a homogenous solution
to equations (3.14) is Fk (z) = c, k = 1, 2. It is known that if a boundary-value
problem for ordinary analytic functions is reduced to boundary integral equations,
then a homogeneous solution to those equations is a solution to the homogeneous
conjugate boundary-value problem (Muskhelishvili 1992). Thus, the approach is
to reduce equations (3.14) to the homogeneous conjugate boundary-value problem
(3.11), whose solution is given by the formulae (3.12). This explains the role of
the problem (3.11).
Let F̂k (z), k = 1, 2, and solve equations (3.14). Then, let Q+k (z), k = 1, 2, be
H -analytic functions in D+ determined by the generalized Cauchy-type integral

1
Q+
k (z) = F̂k (t)e−lk z1 WH (z, t, lk ) dt, k = 1, 2, z ∈ D+ , (3.18)
2pi ∪

and let F± (z) be r-analytic functions in D± , respectively, also determined by the


generalized Cauchy-type integrals

1
F+ (z) = ((1 − 2l1 8)F̂2 (t) − (1 − 2l2 8)F̂1 (t) + l2 − l1 )Wr (z, t) dt
2pi ∪

and
1
F− (z) = (F̂2 (t) − F̂1 (t))Wr (z, t) dt,
2pi ∪

where z ∈ D± for F± , respectively. With the introduced functions Q+ +


1 , Q2 and
±
F and the corresponding generalized Sokhotski–Plemelj formulae (1.6) for these
functions, the difference of equation (3.14) for k = 1 and k = 2 takes the form

el2 z Q+
2 (z) − e
l1 z +
Q1 (z) = F− (z), z ∈ .

Similarly, the linear combination (1 − 2l2 8) · (3.14)|k=1 − (1 − 2l1 8) · (3.14)|k=2


reduces to

(1 − 2l1 8)el2 z Q+
2 (z) − (1 − 2l2 8)e
l1 z +
Q1 (z) = F+ (z), z ∈ .

The last two equations form a system equivalent to the homogeneous conjugate
boundary-value problem (3.11) with Q+ +
k = Gk , k = 1, 2, F = −(l1 − l2 )/
+ − −
(Rem Re − M )G3 and F = (l1 − l2 )G3 . Consequently, by proposition 3.2, the
2

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3359

only solution this system has is Q+k =ce


−lk z
, k = 1, 2, F+ = 2(l2 − l1 )8 c and

F = 0, where c is a real-valued constant. In this case, the representation (3.18)
can be rearranged in the form
1
(F̂k (t) − c)e−lk z1 WH (z, t, lk ) dt = 0, k = 1, 2, z ∈ D+ .
2pi ∪

For z approaching  within D+ , the above equation reduces to the generalized


Sokhotski–Plemelj formula
1 1
(F̂k (z) − c)e−lk z + (F̂k (t) − c)e−lk z1 WH (z, t, lk ) dt = 0, (3.19)
2 2pi ∪
for k = 1, 2 and z ∈ , which is the necessary and sufficient condition for
(F̂k (z) − c)e−lk z , z ∈ , to be the boundary value of an H -analytic function in
D− that vanishes at infinity. Thus, the solution (3.15) and (3.16) follows from
equation (3.19).
Finally, let F̃k (z), k = 1, 2, z ∈  be another solution to equations (3.14).
Similarly, we can show that Gk (z) = (F̃k (z) − c̃)e−lk z , k = 1, 2, z ∈ , solve the
boundary-value problem (3.1), where c̃ is a real-valued constant. However, by
proposition 3.1, the boundary-value problem (3.1) has a unique solution, and
consequently, F̂k (z) − F̃k (z) = c − c̃, k = 1, 2, z ∈ , which means that a solution
to the integral equations (3.14) is determined up to a real-valued constant, and
the proof is finished. 
Several remarks are in order.
Remark 3.4 (logarithmic singularity). The kernels in equations (3.14) have
logarithmic singularity.
Remark 3.5 (alternative form). For efficient solving, the integral equation
(3.14) needs to be rewritten in the form similar to eqns (67) and (68) in
Zabarankin (2010).
Remark 3.6 (conic endpoint). If  is piece-wise smooth (has salient points),
in particular has conic endpoints, the integral equation (3.14) holds for all z ∈ ,
except for salient points and conic endpoints, i.e. almost everywhere.
Remark 3.7 (multiply connected body). The integral equation (3.14) holds for
multiply connected regions, e.g. in the MHD problem for a torus.
The boundary integral equations (3.14) can be solved by the quadratic
error minimization method. Let  be parametrized by z = z(t) ∈ C 1 [−1, 1], and
let Fj = Fj (t), t ∈ [−1, 1], j = 1, 2, be unknown boundary values. For brevity,
equations (3.14) are recast as
Aj (F1 , F2 ) = fj , t ∈ [−1, 1], j = 1, 2,
where Aj , j = 1, 2, are corresponding integral operators, and f1 (t) = f2 (t) =
−1/(28). The functions Fj , j = 1, 2, can be approximated by finite function series

n 
n
F1 (t) = (ak + ibk )Tk−1 (t) and F2 (t) = ck Tk (t) + idk Tk−1 (t), (3.20)
k=1 k=1

Proc. R. Soc. A (2011)


3360 M. Zabarankin

where Tk (t) is the Chebyshev polynomial of the first kind, and ak , bk , ck and dk ,
k = 1, . . . , n, are real-valued coefficients. Observe that the real part in the series
approximating F2 contains no constant term, whereas the series for F1 does. This
is because same real-valued constant in place of F1 and F2 is a homogeneous
solution to equations (3.14).
Unknown coefficients ak , bk , ck and dk , k = 1, . . . , n, can be found by minimizing
the quadratic error in satisfying equations (3.14), i.e.

2
min Aj (F1 , F2 ) − fj 2 , (3.21)
ak ,bk ,ck ,dk
j=1

with the inner product and norm for complex-valued functions f (t) and g(t)
introduced by6
 1
f , g = Re f (t) g(t) dt , f  = f , f .
−1

As Aj , j = 1, 2, are linear operators, the problem (3.21) is unconstrained quadratic


optimization and reduces to a system of linear algebraic equations for ak , bk , ck
and dk , k = 1, . . . , n.

(b) Case (b): M = 0 and Rem = 0


For M  = 0 and Rem = 0, the MHD problem (2.2)–(2.5) reduces to the boundary-
value problem for the H -analytic functions G1 and G2 defined in corollary 2.2,

el1 z G1 − el2 z G2 = l2 − l1 , z ∈ . (3.22)


Proposition 3.8. The boundary-value problem (3.22) has a unique solution.
Proof. The proof is similar to the proof of proposition 3.1. In this case, corollary
2.2 shows that the disturbance of the magnetic field in the fluid and body is zero,
i.e. h± = 0. With h− = 0, equation (3.8) and the surface integral (3.10) reduce to

IR + (|u|2 + M 2 ur2 ) dV = 0 (3.23)
DR

and    
Re
IR = 2p Re (uz + iur ) ℘ + iu + (uz − iur ) r dz ,
CR 2

respectively. Using the representation (2.21), we obtain



2p
IR = Re (e G1 − e G2 )r dz ,
2l1 z 2 2l2 z 2
l1 − l 2 CR

and the rest of the proof is analogous to the proof of proposition 3.1. 
6 f (t) and g(t) are viewed as two-dimensional vector functions from the direct sum L2 ([−1, 1]) ⊕

L2 ([−1, 1]).

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3361

Proposition 3.9 (homogeneous conjugate boundary-value problem, case (b)).


Let G1 and G2 be H -analytic functions in D+ . Under the assumption M = 0, the
homogeneous conjugate boundary-value problem
el1 z G1 − el2 z G2 = 0, z ∈ , (3.24)
has the solution
G1 = c e−l1 z and G2 = c e−l2 z , (3.25)
where c is an arbitrary real-valued constant, and l1 , l2 are defined in corollary 2.2.
Proof. Proved similar to proposition 3.2. 
With the generalized Cauchy integral formula for H -analytic functions, the
boundary-value problem (3.22) readily reduces to a boundary integral equation.
Theorem 3.10 (boundary integral equation, case (b)). Let M = 0 and Rem = 0
and let F1 (z) = el1 z G1 (z). The boundary-value problem (3.22) reduces to the
integral equation for the boundary value of F1 ,
1
(el2 (z−z1 ) WH (z, t, l2 ) − el1 (z−z1 ) WH (z, t, l1 ))F1 (t) dt = l2 − l1 , (3.26)
2pi ∪

for z ∈ . A solution to equation (3.26) is determined up to a real-valued constant c.


Let F̂1 (z) solve equation (3.26), then G1 (z) in the problem (3.22) is given by
G1 (z) = (F̂1 (z) − c)e−l1 z , z ∈ , (3.27)
where
1 1
c = F̂1 (z) +  F̂1 (t)el1 (z−z1 ) WH (z, t, l1 ) dt, z ∈ . (3.28)
2 2pi  

Proof. The integral equation (3.26) is derived similar to equations (3.14).


In the second part of the proof that determines a homogeneous solution to
equation (3.26), H -analytic functions Q+ +
k , k = 1, 2, in D are introduced by the
generalized Cauchy-type integrals
1
Q+
1 (z) = F̂1 (t)e−l1 z1 WH (z, t, l1 ) dt, z ∈ D+ ,
2pi ∪

and
1
Q+
2 (z) = (F̂1 (t) + l1 − l2 )e−l2 z1 WH (z, t, l2 ) dt, z ∈ D+ ,
2pi ∪

where F̂1 is a solution to equation (3.26). With these functions and corresponding
generalized Sokhotski–Plemelj formulae (1.6), the integral equation (3.26) reduces
to the homogenous conjugate boundary-value problem (3.24), whose solution
is given by the formulae (3.25). The rest of the proof is analogous to that of
theorem 3.3. 
The integral equation (3.26) can be solved by the quadratic error minimization
method outlined at the end of §3a. Remarks 3.4–3.7 also hold for equation (3.26).

Proc. R. Soc. A (2011)


3362 M. Zabarankin

(c) Case (c): M = 0 and Rem Re = M 2


For M  = 0 and Rem Re = M 2 , the MHD problem (2.2)–(2.5) reduces to the
boundary-value problem for the four generalized analytic functions G1 , G2 and
G3± defined in theorem 2.5,

2elz G1 − G2 + (Re + Rem )(G3+ + 1) = 0, z ∈ ,⎪

  
i (3.29)
and Rem z − r G2 + G3− + 1 − ReG3+ = 0, z ∈ .⎪ ⎭
2

Proposition 3.11. The boundary-value problem (3.29) has a unique solution.

Proof. The assumptions M = 0 and Rem Re = M 2 imply Rem = 0, and


consequently, as in the proof of proposition 3.1, we obtain equation (3.8) with
the surface integral (3.10). With the representations (2.22)–(2.26), the integral
(3.10) reduces to
  
Re
IR = 2p Re (uz + iur )G2 + (2e G1 − G2 ) r dz .
lz 2
CR 2(Re + Rem )

Since in this case, G1 and G2 are H -analytic and r-analytic functions, respectively,
we have G1 (R, w) = f (w)R−1 e−|l|R + O(R−2 e−|l|R ) as R → ∞, where f (w), w ∈
[0, p], is a bounded complex-valued function, and G2 = O(R−2 ) as R → ∞. The
rest of the proof is completely analogous to the proof of proposition 3.1. 

Proposition 3.12 (homogeneous conjugate boundary-value problem, case (c)).


Let G1 and G2 be H -analytic and r-analytic functions in D+ , respectively, and let
G3± be r-analytic functions in D± , respectively, with G3− vanishing at infinity.
Under the assumptions M = 0 and Rem Re = M 2 , the homogeneous conjugate
boundary-value problem

2elz G1 − G2 + (Re + Rem )G3− = 0, z ∈ ,⎪

  
i (3.30)
and Rem z − r G2 + G3+ − ReG3− = 0, z ∈ ,⎪ ⎭
2

has the solution

G1 = c e−lz , G2 = 2c, G3+ = −c(2z − ir) and G3− = 0, (3.31)

where l = (Re + Rem )/2, and c is an arbitrary real-valued constant.

Proof. Proved similarly to proposition 3.2. 

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3363

Theorem 3.13 (boundary integral equations, case (c)). Let M = 0 and


Rem Re = M 2 , and let l = (Re + Rem )/2, then the boundary-value problem (3.29)
reduces to two integral equations for F1 (z) = elz G1 (z) and F2 (z) = G2 (z), z ∈ ,

1 1 ⎪

l(z−z1 )
(e WH (z, t, l) − Wr (z, t))F1 (t) dt + F1 (z) − F2 (z) = 0 ⎪

2pi  2 ⎬
∪
 
1 i Re(2F1 (z) − F2 (z)) Re + Rem ⎪⎪

and z − z1 + (r1 − r) Wr (z, t)F2 (t) dt − = .⎪

2pi ∪ 2 Re m (Re + Re m ) Re m
(3.32)

Equations (3.32) determine F1 and F2 up to c and 2c, respectively, where c is a


real-valued constant. Let F̂k , k = 1, 2, solve equations (3.32), then the solution to
the boundary-value problem (3.29) is given by

G1 (z) = (F̂1 (z) − c)e−lz and G2 (z) = F̂2 (z) − 2c, z ∈ , (3.33)
where
1 1
2c = F̂2 (z) +  F̂2 (t)Wr (z, t) dt, z ∈ . (3.34)
2 2pi  

Proof. The derivation of equations (3.32) is similar to that of equations (3.14)


(see the proof of theorem 3.3). In the second part of the proof that determines
a homogeneous solution to equations (3.32) and also follows closely the proof of
theorem 3.3, an H -analytic function Q+ in D+ , r-analytic functions F1+ and F2+
in D+ , and an r-analytic function F2− in D− are introduced by the generalized
Cauchy-type integrals
1
Q+ (z) =  F̂1 (t)e−lz1 WH (z, t, l) dt, z ∈ D+ ,
2pi  

1
F1+ (z) =  F̂2 (t)Wr (z, t) dt, z ∈ D+ ,
2pi  
  
1 i Re(2F̂1 (t) − F̂2 (t)) Re + Rem
F2+ (z) =  z1 − r1 F̂2 (t) + +
2pi   2 Rem (Re + Rem ) Rem
× Wr (z, t) dt, z ∈ D+ ,
1
and F2− (z) =  (2F̂1 (t) − F̂2 (t))Wr (z, t) dt, z ∈ D− ,
2pi  

where F̂k , k = 1, 2, are a solution to equations (3.32). With these functions


and corresponding generalized Sokhotski–Plemelj formulae (1.6), the integral
equations (3.32) reduce to the homogeneous conjugate boundary-value problem
(3.30), whose solution is given by the formulae (3.31). The rest of the proof
continues as in the proof of theorem 3.3. 
The integral equations (3.32) can be solved by the quadratic error minimization
method outlined at the end of §3a. Remarks 3.4–3.7 hold for equations (3.32)
as well.

Proc. R. Soc. A (2011)


3364 M. Zabarankin

(d) Drag
The drag (force) exerted on the body by the electrically conducting viscous
incompressible fluid in the presence of the magnetic field has the mechanic and
magnetic components corresponding to the viscous stress and Lorentz force,
respectively,  
F= Pn dS − m [J × H] dV ,
S D−
where Pn = rn(2vU/vn + [n × curl U]) − ℘ n, and n is the outward normal to S
(Cabannes 1970).
In the axially symmetric case, div U = 0 and the boundary conditions (2.2)
imply vU/vn = −[n × curl U] on S ; see e.g. the proof of proposition 8 in
Zabarankin (2008). Also, the next proposition shows that the second term in
F vanishes.
Proposition 3.14. If at infinity, the flow and the magnetic field are uniform and
parallel to the body’s axis of revolution, and if also the fluid and body have the same
magnetic permeability, the Lorentz force has no direct contribution to the drag.
Proof. In the axially symmetric case, the resulting mechanic and Lorentz forces
are parallel to the body’s axis of revolution, and consequently, it is sufficient to
show that the projection of the resulting Lorentz force onto k vanishes. Let DR
be the region bounded by the body’s surface S and a sphere SR with large radius
R and centre at the origin. Using the identity (3.4) and the divergence theorem,
we obtain
 
k · [J × H] dV = −H∞ 2
[k × h− ] · curl h− dV = −H∞
2
(IR − I ),
DR DR

where   
1
I= n · (k · h ) h − |h− |2 k
− −
dS ,
S 2
and IR is defined as the integral I in which S is replaced by SR . As in the proof
of proposition 3.1, we can show that in all three cases (i)–(iii), IR → 0 as R → ∞.
On the other hand, as h+ = h− on S and curl h+ = 0 in D+ , using the identity
similar to (3.4) for h+ , we have
   
+ + 1 +2
I= n · (k · h )h − |h | k dS = − [k × h+ ] · curl h+ dV = 0.
S 2 D+

Consequently, DR k · [J × H] dV → 0 as R → ∞. 
Thus, in terms of dimensionless u and ℘, the only non-zero drag component
Fz = k · F takes the form

Fz = −V∞ rna k · ([n × curl u] + ℘n) dS ,
S

which, as in the proof of proposition 11 in Zabarankin (2008), reduces to



Fz = −2p V∞ rna Re r(℘ + iu) dz . (3.35)


Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3365

Let CD = −Fz /(6pV∞ rna) be a dimensionless drag coefficient, where 6pV∞ rna
is the drag of the sphere with radius a in the Stokes flow of a non-conducting fluid.
Theorem 2.1, corollary 2.2, theorem 2.5 and the corresponding boundary-value
problems (3.1), (3.22) and (3.29) imply that

⎪ Re − 2l2 l1 z Re − 2l1 l2 z

⎪ e G − e G2 + Re, z ∈ , in case (a),

⎨ l1 − l 2
1
l1 − l 2
℘ + iu = 2(el1 z G1 + l1 ), z ∈ , in case (b),



⎪ 1
⎩ (2Re elz G1 + Rem G2 ) + Re, z ∈ , in case (c).
Re + Rem

Now with this representation for ℘ + iu and the fact that Re[  r dz] = 0, the
formula (3.35) yields the following result.

Proposition 3.15 (drag). Let M = 0, then the drag coefficient CD is determined


as follows.

(i) If Rem  = 0 and Rem Re = M2 , then


  
Re − 2l2 l1 z Re − 2l1 l2 z
CD = Re r e G1 − e G2 dz , (3.36)
 3(l1 − l2 ) 3(l1 − l2 )
where G1 , G2 , l1 and l2 are defined in theorem 2.1.
(ii) If Rem = 0, then

2
CD = Re l1 z
re G1 dz , (3.37)
3 

where G1 and l1 are defined in corollary 2.2.


(iii) If Rem Re = M 2 , then

1
CD = Re r(2Re e G1 + Rem G2 ) dz ,
lz
(3.38)
3(Re + Rem ) 

where G1 , G2 and l are defined in theorem 2.5.

4. Axially symmetric magnetohydrodynamic problems

(a) Sphere
The purpose of solving the MHD problem for a sphere is threefold: (i) verifying the
presented approach of generalized analytic functions based on the existing results
for a sphere, (ii) testing accuracy of solutions to the derived boundary integral
equations, and (iii) obtaining drag for various Re, Rem and M to compare it with
the drag of minimum-drag shapes in Zabarankin (2011).
Let (R, w, 4) be the spherical coordinates related to the cylindrical coordinates
in the ordinary way, and let Sa be a sphere with radius a and centre at the origin.

Proc. R. Soc. A (2011)


3366 M. Zabarankin

(i) Case (a): M  = 0, Rem = 0 and Rem Re = M 2 (S  = 1)


In the region exterior to Sa , the functions G1 and G2 in theorem 2.1 can be
represented by a series similar to the series (1.8),
 ∞
a 
Gk (R, w) = Ak,n Nn (cos w, R, lk ), k = 1, 2, (4.1)
R n=1

where Ak,n are unknown real-valued coefficients and

Kn+ 1 (|l|R) Kn− 1 (|l|R)


Nn (t, R, l) = Ln (t) 2
− (sign l) L−n (t) 2
,
Kn+ 1 (|l|a) Kn+ 1 (|l|a)
2 2

whereas G3± can be represented by (see example 5 in Zabarankin (2010))



G3± (R, w) = Bn± (R/a)±n−1 L∓n (cos w), (4.2)
n=1

where Bn± are unknown real-valued coefficients.


Let the inner product ·, · be defined as at the end of §3a. With the
orthogonality property

Ln (t), Lm (t) = 2m dmn , (4.3)

for all integer n and m, where dmn is the Kronecker delta, and the representations
(4.1) and (4.2), the inner product of the first equation in (3.1) with L−m−1 (t) for
m ≥ 0 and the inner product of the second equation in (3.1) with Lm (t) for m ≥ 1
yield a real-valued infinite linear system for A1,n and A2,n ,

∞   ⎫
 1 − 2l2 8 1 − 2l1 8 ⎪

xmn (l1 )A1,n − xmn (l2 )A2,n = 2dm0 , m ≥ 0,⎪


l1 − l 2 l1 − l 2 ⎬
n=1
(4.4)

 ⎪

and (hmn (l1 ) A1,n − hmn (l2 )A2,n ) = 0, m ≥ 1, ⎪



n=1

where xmn (l) = elat Nn (t, a, l), L−m−1 (t) and hmn (l) = elat Nn (t, a, l), Lm (t).
System (4.4) can be truncated at some large m and solved numerically.
With the second equation in (4.4), the drag coefficient (3.36) reduces to


2
CD = A1,n h1n (l1 ).
3 n=1

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3367

Figure 1 shows CD as a function of S in five cases: (i) Re = Rem = 1; (ii) Re = 1,


Rem = 3; (iii) Re = Rem = 2; (iv) Re = 3, Rem = 1; and (v) Re = Rem = 3.

(ii) Case (b): M = 0 and Rem = 0


As in the previous case, the functions G1 and G2 in corollary 2.2 can be
represented by the series (4.1). With the orthogonality property (4.3) and
representations (4.1), the inner products of (3.22) with e−l2 at Lm (t) for m ≥ 1
and with e−l1 at Lm (t) for m ≥ 1 yield a real-valued infinite linear system for A1,n
and A2,n ,

 ∞

x̃mn A1,n − 2m A2,m = b2,m and 2m A1,m − h̃mn A2,n = b1,m , m ≥ 1,
n=1 n=1

(4.5)

where x̃mn = el1 at Nn (t, a, l1 ), e−l2 at Lm (t), h̃mn = el2 at Nn (t, a, l2 ), e−l1 at Lm (t) and
bk,m = l2 − l1 , e−lk at Lm (t), k = 1, 2. System (4.5) can be truncated at some large
m and solved numerically.
The drag coefficient (3.37) takes the form

2
CD = A1,n el1 at Nn (t, a, l1 ), L1 (t),
3 n=1

which is in good agreement with Chester (1957): CD = 1 + 3M /8 + 7M 2 /960 −


43M 3 /7680 + O(M 4 ) for small M up to 0.25.

(iii) Case (c): M = 0 and Rem Re = M 2 (S = 1)


In this case, the H -analytic function G1 and the r-analytic function G2 in
theorem 2.5 are represented by the series
 ∞ ∞  a n+1
a  
G1 (R, w) = A1,n Nn (cos w, R, l) and G2 (R, w) = A2,n Ln (cos w),
R n=1 n=1
R
(4.6)
where l = (Re + Rem )/2, and Nn (cos w, R, l) was introduced in §4a(i), whereas
the r-analytic functions G3± are represented by the series (4.2).
With the orthogonality property (4.3) and representations (4.6) and (4.2), the
inner products of the first equation in (3.29) with Lm (t) for m ≥ 1 and with
L−m−1 (t) for m ≥ 0 yield a real-valued infinite linear system


 ⎪

hmn (l)A1,n − mA2,m = 0, m ≥ 1,⎪ ⎪


n=1
(4.7)
∞ ⎪

xmn (l)A1,n − (Re + Rem )(m + 1)Bm+1 = (Re + Rem )dm0 , m ≥ 0,⎪
+ ⎪
and ⎪

n=1

where hmn (l) and xmn (l) are those in §4a(i).

Proc. R. Soc. A (2011)


3368 M. Zabarankin

Similarly, with equations (4.6), (4.2) and (4.3), the inner products of the second
equation in (3.29) with Lm (t) for m ≥ 1 and with L−m−1 (t) for m ≥ 0 yield
  ⎫
m−1 3(m + 1) ⎪
a A2,m−1 + A2,m+1 + 2Bm = 0, m ≥ 1,⎪



2m − 1 2m + 1
(m + 1)(m + 2) ⎪

and Rem a +
A2,m+1 − Re(m + 1)Bm+1 = −Rem dm0 , m ≥ 0.⎪

(2m + 1)(2m + 3)
(4.8)
It follows from system (4.7) and the second equation in (4.8) that
∞ 
 
a Rem (Re + Rem ) (m + 2) (Re + Rem )2
xmn (l) − h(m+1)n (l) A1,n = dm0 ,
n=1
Re(2m + 1)(2m + 3) Re

for m ≥ 0. The latter can be truncated at some large m and solved numerically.
The drag coefficient (3.38) simplifies to the formula

2 
CD = A1,n h1n (l),
3 n=1

which is used to evaluate CD at S = 1 in figure 1 and which agrees


with Gotoh (1960b) for various small Re and Rem : CD ≈ 1 + 3Re/8 −
(19Re 2 /320 + 2Re Rem /15) + 1/7680(213 Re 3 + 256 Re 2 Rem − 704 Re Rem
2
) + 1Re
Rem /30(Re + 4 Rem ).

(b) Minimum-drag spheroids


This section solves the axially symmetric MHD problem for solid non-magnetic
spheroids having the volume of a unit sphere, i.e. 4p/3, and among those finds
the spheroids that have the smallest drag for given Re, Rem and M . In the right-
half rz-plane (r ≥ 0), the cross section  of spheroids is parametrized by r(t) =
a cos(pt/2), z(t) = a −2 sin(pt/2), t ∈ [−1, 1], with a ∈ (0, 1]. Then, for fixed Re,
Rem and M , the spheroid drag becomes a function of a, and the minimum of this
function along with the value of a at which the minimum is attained is found
by a modified bisection method. This is an iterative procedure that requires the
involved boundary integral equations to be solved fast and accurately.
Let CD and CD∗ be the drag coefficients for the minimum-drag spheroid and
unit sphere, respectively, for the same Re, Rem and S, and let k = a −3 be the axes
ratio of the corresponding minimum-drag spheroid. Figures 2 and 3 show the drag
ratio CD /CD∗ and k as functions of S ∈ [0, 2] in three cases: (i) Re = Rem = 1, (ii)
Re = Rem = 2, and (iii) Re = Rem = 3. As a function of S, CD /CD∗ has a maximum
at S = 1 (and is non-smooth at S = 1), and consequently, drag reduction is smallest
for S = 1. Also, figure 2 indicates that drag reduction becomes more significant
for S  1, whereas figure 3 suggests that the shortest minimum-drag spheroids do
not correspond to S = 1. The minimum-drag spheroids found are used as initial
approximations for minimum-drag shapes in Zabarankin (2011).

Proc. R. Soc. A (2011)


Generalized analytic functions in MHD 3369

0.94
0.93
0.92
(iii)

drag ratio
0.91
0.90
(ii)
0.89
0.88
0.87 (i)
0.86
0 0.5 1.0 1.5 2.0
S

Figure 2. The drag ratio CD /CD∗ for the minimum-drag spheroid as a function of S in three cases:
(i) Re = Rem = 1; (ii) Re = Rem = 2; and (iii) Re = Rem = 3.

3.0 (iii)

2.8
axes ratio

(ii)
2.6

2.4 (i)

2.2
0 0.5 1.0 1.5 2.0
S

Figure 3. The axes ratio k of the minimum-drag spheroid as a function of S in three cases: (i) Re =
Rem = 1; (ii) Re = Rem = 2; and (iii) Re = Rem = 3.

I am grateful to the anonymous referees for their valuable comments and suggestions, that helped
to improve the quality of the paper. Research was supported by AFOSR grant FA9550-09-1-0485.

References
Alexandrov, A. Y. & Soloviev, Y. I. 1978 Three-dimensional problems of the theory of elasticity.
Moscow, Russia: Nauka (in Russian).
Bateman, H. & Erdelyi, A. 1953 Higher transcendental functions, vol. II. New York, NY:
McGraw-Hill.
Bers, L. 1953 Theory of pseudo-analytic functions. New York, NY: Institute for Mathematics and
Mechanics.
Bers, L. 1956 Formal powers and power series. Commun. Pure Appl. Math. 9, 693–711. (doi:10.1002/
cpa.3160090405)
Bers, L. 1958 Mathematical aspects of subsonic and transonic gas dynamics. In Surveys in applied
mathematics, vol. 3. New York, NY: John Wiley & Sons, Inc.
Blerkom, R. V. 1960 Magnetohydrodynamic flow of a viscous fluid past a sphere. J. Fluid Mech.
8, 432–441. (doi:10.1017/S0022112060000712)
Cabannes, H. 1970 Theoretical magnetohydrodynamics. New York, NY: Academic Press.

Proc. R. Soc. A (2011)


3370 M. Zabarankin

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)
Chester, W. 1957 The effect of a magnetic field on Stokes flow in a conducting fluid. J. Fluid Mech.
3, 304–308. (doi:10.1017/S0022112057000671)
Chester, W. 1961 The effect of a magnetic field on the flow of a conducting fluid past a body of
revolution. J. Fluid Mech. 10, 459–465. (doi:10.1017/S0022112061001050)
Chester, W. 1962 On Oseen’s approximation. J. Fluid Mech. 13, 557–569. (doi:10.1017/
S0022112062000932)
Cowling, T. G. 1957 Magnetohydrodynamics. New York, NY: Interscience.
Goldsworthy, F. A. 1961 Magnetohydrodynamic flows of a perfectly conducting, viscous fluid.
J. Fluid Mech. 11, 519–528. (doi:10.1017/S0022112061000706)
Gotoh, K. 1960a Magnetohydrodynamic flow past a sphere. J. Phys. Soc. Jpn 15, 189–196.
(doi:10.1143/JPSJ.15.189)
Gotoh, K. 1960b Stokes flow of an electrically conducting fluid in a uniform magnetic field. J. Phys.
Soc. Jpn 15, 696–705. (doi:10.1143/JPSJ.15.696)
Imai, I. 1960 On flows of conducting fluids past bodies. Rev. Mod. Phys. 32, 992–999. (doi:10.1103/
RevModPhys.32.992)
Ludford, G. S. S. & Singh, P. 1963 The motion of a non-conducting sphere through a
conducting fluid in a magnetic cross-field. Proc. Camb. Phil. Soc. 59, 615–624. (doi:10.1017/
S0305004100037294)
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.
Polozhii, G. N. 1973 Theory and application of p-analytic and (p, q)-analytic functions, 2nd edn.
Kiev: Naukova Dumka (in Russian).
Stewartson, K. 1956 Motion of a sphere through a conducting fluid in the presence of a strong
magnetic field. Proc. Camb. Phil. Soc. 52, 301–316. (doi:10.1017/S0305004100031285)
Vekua, I. N. 1962 Generalized analytic functions. Oxford, UK: Pergamon Press.
Yosinobu, H. 1960 A linearized theory of magnetohydrodynamic flow past a fixed body in a parallel
magnetic field. J. Phys. Soc. Jpn 15, 175–188. (doi:10.1143/JPSJ.15.175)
Zabarankin, M. 2008 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. 2010 Generalized analytic functions in axially symmetric Oseen flows. SIAM J.
Appl. Math. 70, 2473–2508. (doi:10.1137/090776937)
Zabarankin, M. 2011 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 in the domain
exterior to spindle. SIAM J. Appl. Math. 66, 1270–1300. (doi:10.1137/050632403)

Proc. R. Soc. A (2011)

You might also like