21.12.2013 Views

Reflections on Magnetohydrodynamics

Reflections on Magnetohydrodynamics

Reflections on Magnetohydrodynamics

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong><br />

H. K. MOFFATT<br />

1 Introducti<strong>on</strong><br />

<strong>Magnetohydrodynamics</strong> (MHD) is c<strong>on</strong>cerned with the dynamics of fluids<br />

that are good c<strong>on</strong>ductors of electricity, and specifically with those effects<br />

that arise through the interacti<strong>on</strong> of the moti<strong>on</strong> of the fluid and any ambient<br />

magnetic field B(x, t) that may be present. Such a field is produced by electric<br />

current sources which may be either external to the fluid (in which case we<br />

may talk of an ‘applied’ magnetic field), or induced within the fluid itself.<br />

The inducti<strong>on</strong> of a current distributi<strong>on</strong> j (x, t) by flow across the field B is the<br />

result of Faraday’s ‘law of inducti<strong>on</strong>’. The resulting Lorentz force distributi<strong>on</strong><br />

F(x, t) = j A B is generally rotati<strong>on</strong>al, i.e. V A F # 0, and therefore generates<br />

vorticity in the fluid. There is thus a fundamental interacti<strong>on</strong> between the<br />

velocity field v and the magnetic field B, an interacti<strong>on</strong> which not <strong>on</strong>ly leads<br />

to modificati<strong>on</strong> of well-understood flows of ‘c<strong>on</strong>venti<strong>on</strong>al’ fluid dynamics,<br />

but also is resp<strong>on</strong>sible for completely new phenomena that simply do not<br />

exist in n<strong>on</strong>-c<strong>on</strong>ducting fluids.<br />

There are three major fields of applicati<strong>on</strong> of magnetohydrodynamics,<br />

which will be discussed in this survey in the following order.<br />

1.1 Liquid-metal magnetohydrodynamics<br />

This is a branch of the subject which has attracted increasing attenti<strong>on</strong> over<br />

the last 30 years, and which has been much stimulated by the experimental<br />

programmes of such laboratories as MADYLAM in Grenoble, and the<br />

Laboratory of <strong>Magnetohydrodynamics</strong> in Riga, Latvia. These programmes<br />

have been motivated by the possibility of using electromagnetic fields in the<br />

processing of liquid metals (and their alloys) in c<strong>on</strong>diti<strong>on</strong>s that frequently<br />

require high temperatures and high purity. These fields can be used to levitate<br />

samples of liquid metal, to c<strong>on</strong>trol their shape, and to induce internal<br />

347


348 H. K. Mofatt<br />

stirring for the purpose of homogenizati<strong>on</strong> of the finished product, all<br />

these being effects that are completely unique to magnetohydrodynamics.<br />

Electromagnetic stirring (exploiting the rotati<strong>on</strong>ality of the Lorentz force)<br />

is used in the process of c<strong>on</strong>tinuous casting of steel and other metals;<br />

and magnetic fields have an important potential use in c<strong>on</strong>trolling the<br />

interface instabilities that currently limit the efficiency of the industrial<br />

process of extracting aluminium from the raw material cryolite. The industrial<br />

exploitati<strong>on</strong> of MHD in such c<strong>on</strong>texts is of relatively recent origin, and great<br />

developments in this area are to be expected over the next few decades.<br />

1.2 Magnetic fields in planetary physics and astrophysics<br />

Nearly all large rotating cosmic bodies, which are either partly or wholly fluid<br />

in compositi<strong>on</strong>, exhibit magnetic fields of predominantly internal origin, i.e.<br />

fields associated with internal rather than external currents. Such fields may<br />

be generated by amplificati<strong>on</strong> of a very weak applied field, a process that may<br />

be limited by the c<strong>on</strong>ductivity of the fluid. More dramatically however, such<br />

fields may be the result of a ‘dynamo instability’ which is entirely of internal<br />

origin; this can occur (as in the liquid core of the Earth) if buoyancy-driven<br />

c<strong>on</strong>vecti<strong>on</strong> interacts with Coriolis forces to produce ‘helicity’ in the flow field.<br />

When the dynamo instability occurs, the field energy grows exp<strong>on</strong>entially<br />

until the Lorentz forces are str<strong>on</strong>g enough to modify the flow; at this stage,<br />

the magnetic energy is at least of the same order as the kinetic energy<br />

of the flow that generates it, and may even (as in the Earth c<strong>on</strong>text) be<br />

much greater. A fundamental understanding of this process in the c<strong>on</strong>text of<br />

turbulent flow has developed slowly over the last 50 years; and our present<br />

understanding of turbulent dynamo acti<strong>on</strong>, although incomplete, must surely<br />

be regarded as <strong>on</strong>e of the great achievements of research in turbulence of<br />

this last half-century.<br />

1.3 Magnetostatic equilibrium, structure and stability<br />

This large area of MHD is also important in astrophysical c<strong>on</strong>texts; but it has<br />

received even greater stimulus from the field of fusi<strong>on</strong> physics and the need<br />

to design fusi<strong>on</strong> reactors c<strong>on</strong>taining hot fully i<strong>on</strong>ized gas (or plasma) isolated<br />

from the c<strong>on</strong>taining solid boundaries by a suitably engineered magnetic field<br />

- such an arrangement being c<strong>on</strong>venti<strong>on</strong>ally described as a ‘magnetic bottle’.<br />

In ideal circumstances, the gas in such a bottle is at rest, in equilibrium<br />

under the mutual acti<strong>on</strong> of Lorentz and pressure forces. Pressure forces of<br />

course tend to make the gas expand; the Lorentz force must thus be such as


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 349<br />

to prevent this expansi<strong>on</strong>. Moreover the equilibrium, if it is to be effective,<br />

must be stable: the energy of the system must be minimal with respect<br />

to variati<strong>on</strong>s associated with the vast family of perturbati<strong>on</strong>s available to<br />

such a system. It is this problem of stability, first recognized and analysed<br />

in the 1950s, which has bedevilled the subject ever since, and which still<br />

stands in the way of development of an energy-producing therm<strong>on</strong>uclear<br />

reactor. Progress has nevertheless been sustained, and there is still a degree<br />

of optimism that the full stability problem can be eventually, if not solved,<br />

at least sufficiently tamed to allow design of commercially viable reactors.<br />

This is truly <strong>on</strong>e of the immense scientific challenges that c<strong>on</strong>tinues to face<br />

mankind at the dawn of the new millennium. Its soluti<strong>on</strong> would provide a<br />

clean source of low-cost energy at least until the next reversal of the Earth’s<br />

magnetic field; by then we may have other problems to worry about!<br />

2 Fundamental principles<br />

The electromagnetic field is governed by Maxwell’s equati<strong>on</strong>s, and it is<br />

legitimate in all the above c<strong>on</strong>texts to adopt the ‘magnetohydrodynamic<br />

approximati<strong>on</strong>’ in which displacement current and all associated relativistic<br />

effects are neglected. The magnetic field is then related to current by Ampere’s<br />

equati<strong>on</strong><br />

V A B = poj with V * B = 0, (2.1)<br />

where po is c<strong>on</strong>stant (471 x lOP7 in SI units). Moreover B evolves according<br />

to Faraday’s law of inducti<strong>on</strong> which may be expressed in the form<br />

2Bl2t = -V A E , (2.2)<br />

where E(x,t) is the electric field in the ‘laboratory’ frame of reference. The<br />

current in this frame is related to E and B by Ohm’s law:<br />

*<br />

j = o(E + U AB), (2.3)<br />

where u(x,t) is the velocity field and cr is the electrical c<strong>on</strong>ductivity of<br />

the fluid. Like viscosity, o is temperature dependent, and will therefore in<br />

general be a functi<strong>on</strong> of x and t in the fluid. We shall, however, neglect such<br />

variati<strong>on</strong>s, and treat o as a given c<strong>on</strong>stant fluid property. Note that the field<br />

E’ = E + U AB appearing in (2.3) is the electric field in a frame of reference<br />

moving with the local fluid velocity U.<br />

From the above equati<strong>on</strong>s, we have immediately<br />

(2.4)


350 H. K. Mofutt<br />

where v] = (poo)-’, the ‘magnetic diffusivity’ (or ‘resistivity’) of the fluid.<br />

Equati<strong>on</strong> (2.4) is the famous ‘inducti<strong>on</strong> equati<strong>on</strong>’ of magnetohydrodynamics,<br />

describing the evoluti<strong>on</strong> of B if u(x,t) is known. The equati<strong>on</strong> has a marvellous<br />

generality : it holds quite independently of the particular dynamical<br />

forces generating the flow (e.g. whether these are of thermal or compositi<strong>on</strong>al<br />

origin, whether the Lorentz force is or is not important, whether Coriolis<br />

forces are present or not); it holds also whether U is incompressible (V.u = 0)<br />

or not. Equati<strong>on</strong> (2.4) may be regarded as the vector analogue of the scalar<br />

advecti<strong>on</strong>-diffusi<strong>on</strong> equati<strong>on</strong><br />

which describes the evoluti<strong>on</strong> of a scalar c<strong>on</strong>taminant O(x, t) subject to<br />

molecular diffusivity K. Clearly, it is desirable to extract from (2.4) as much<br />

informati<strong>on</strong> as we can, before specializing to any particular dynamical c<strong>on</strong>text.<br />

Note first the striking analogy, first pointed out by Batchelor (1950),<br />

between equati<strong>on</strong> (2.4) and the equati<strong>on</strong> for vorticity o = V A U (the use of<br />

U rather than U here is deliberate) in a n<strong>on</strong>-c<strong>on</strong>ducting barotropic fluid of<br />

kinematic viscosity 11 :<br />

do<br />

- = VA(uAo)+vV2o.<br />

;t<br />

The analogy is incomplete in that o is c<strong>on</strong>strained by the relati<strong>on</strong>ship<br />

o = V A U, whereas B and U in (2.4) suffer no such c<strong>on</strong>straint: there is in<br />

effect far more freedom in (2.4) than there is in (2.6)! Nevertheless a number<br />

of results familiar in the c<strong>on</strong>text of (2.6) do carry over to the c<strong>on</strong>text of (2.4).<br />

2.1 The magnetic Reynolds number<br />

First, suppose that the system c<strong>on</strong>sidered is characterized by a length scale<br />

lo and a velocity scale 00. NaYve comparis<strong>on</strong> of the two terms <strong>on</strong> the right<br />

of (2.4) gives<br />

the magnetic Reynolds number, defined by obvious analogy with the<br />

Reynolds number of classical fluid dynamics. The three major areas of<br />

applicati<strong>on</strong> of magnetohydrodynamics specified in the introducti<strong>on</strong> may be<br />

discriminated in terms of this single all-important dimensi<strong>on</strong>less number:


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 351<br />

(a) R, > 1: here we are into the domain of ‘nearly perfect c<strong>on</strong>ductivity’, in<br />

which inductive effects dominate diffusi<strong>on</strong>. The limit R,n + x (or CT -+ x<br />

or q + 0) may be described as the perfect-c<strong>on</strong>ductivity limit. In this formal<br />

limit, B satisfies what is known as the ‘frozen-field equati<strong>on</strong>’<br />

ZB<br />

- = V A (U A B) ,<br />

2t<br />

which implies that the flux @ of B across any material (Lagrangian) surface<br />

S is c<strong>on</strong>served:<br />

1<br />

d@<br />

- = 0 where @ = BqndS.<br />

(2.10)<br />

dt<br />

This is Alfvh’s theorem, the analogue of Kelvin’s circulati<strong>on</strong> theorem, and<br />

<strong>on</strong>e of the results for which the incomplete analogy between (2.4) and (2.6)<br />

is reliable. The fact that, in the perfect-c<strong>on</strong>ductivity limit, magnetic lines of<br />

force are ‘frozen in the fluid’ provides an appealing picture of how a magnetic<br />

field may develop in time. By analogy with vorticity in an incompressible<br />

fluid, any moti<strong>on</strong> that tends to stretch magnetic field lines (or ‘B-lines’ for<br />

short), will also tend to intensify magnetic field. Is this a manifestati<strong>on</strong> of<br />

dynamo acti<strong>on</strong>? Sometimes it is, but by no means invariably, as we shall see.


352 H. K. Moffatt<br />

2.2 Magnetic helicity c<strong>on</strong>servati<strong>on</strong><br />

Since V B = 0, we may always introduce a vector potential A(x, t) with the<br />

properties<br />

The ‘uncurled’ versi<strong>on</strong> of (2.9) is then<br />

?A<br />

-=uAB-V~,<br />

c‘t<br />

where q(x,t) is a scalar field satisfying<br />

B=VAA, V.A=O. (2.11)<br />

(2.12)<br />

V2q = V * (U A B) . (2.13)<br />

Note that the ‘gauge’ of A may be changed by the replacement A + A + Vx<br />

for arbitrary scalar x. The equati<strong>on</strong> V A = 0 implies a particular choice of<br />

gauge. Equati<strong>on</strong>s (2.9) and (2.12) may be written in equivalent Lagrangian<br />

form :<br />

D ( E) =-.vv, B<br />

DAi ;A. 2~p<br />

__ - c . J - - (2.14)<br />

Dt P P Dt 2xl 2xl*<br />

This form is most c<strong>on</strong>venient for the proof of c<strong>on</strong>servati<strong>on</strong>s of magnetic<br />

helicity, defined as follows: let S be any closed ‘magnetic surface’, i.e. a<br />

surface <strong>on</strong> which B-n = 0, and let V be the (material) volume inside S. Then<br />

the magnetic helicity X,v in V is defined by<br />

s<br />

XAv= AaBdV, (2.15)<br />

a pseudo-scalar quantity that may easily be shown to be independent of the<br />

gauge of A. In the limit y~<br />

1<br />

= 0, this quantity is c<strong>on</strong>served (Woltjer 1958); for<br />

/A*BdV = .BdV + / A. - D (-) B pdV, (2.16)<br />

dt Dt P<br />

and, using (2.14), this readily gives<br />

s<br />

AIA-BdV = V*[B(v-A-q)]dV<br />

dt<br />

= i(n.B)(n.A-q)dS=O. (2.17)<br />

Note that this result holds whether the fluid is incompressible or not; it<br />

merely requires that the fluid be perfectly c<strong>on</strong>ducting.<br />

This invariant admits interpretati<strong>on</strong> in terms of linkage of the B-lines<br />

(which are frozen in the fluid) (Moffatt 1969). To see this, c<strong>on</strong>sider the<br />

simplest ‘prototype’ linkage (figure 1) for which B is zero except in two flux


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 353<br />

Figure 1. Linked flux tubes; here the linkage is right-handed and the linking number<br />

n = 1.<br />

tubes of small cross-secti<strong>on</strong> centred <strong>on</strong> (unknotted) curves C1 and C2; the<br />

field lines within each tube are supposed to be unlinked, <strong>on</strong>e with another.<br />

The sole linkage is then that between the two tubes, or equivalently between<br />

C1 and C2. Let @1 and @2 be the fluxes of B within these<br />

integral (2.16) may then be simply evaluated; it degenerates<br />

two tubes. The<br />

to<br />

(2.18)<br />

The two line integrals are zero if C1 and C2 are unlinked (more strictly if<br />

the flux of B across any surface spanning either C1 or C2 is zero); if C1 and<br />

C2 are linked, with linking number n, then, from (2.18)<br />

(2.19)<br />

t<br />

where the + or - is chosen according to whether the linkage is positive or<br />

negative; thus, for example, it is positive in figure 1, in which the sense of<br />

relative rotati<strong>on</strong> is right-handed; if either arrow is reversed then the sign<br />

changes.<br />

In general, B-lines are not closed curves, and the above simple interpretati<strong>on</strong><br />

of SAw is not available. Nevertheless, Xlw does always carry some<br />

limited informati<strong>on</strong> about the topology of the magnetic field (Arnold 1974;<br />

see also Arnold & Khesin 1998). This will be further c<strong>on</strong>sidered in $6<br />

below.


354 H. K. MofSatt<br />

3 The Lorentz force and the equati<strong>on</strong> of moti<strong>on</strong><br />

As indicated in the introducti<strong>on</strong>, the Lorentz force in the fluid (per unit<br />

volume) is F = j A B. With j = p;’V A B, this admits alternative expressi<strong>on</strong><br />

in terms of the Maxwell stress tensor<br />

1 2<br />

F. 1 - -- (BiBj- iB2Sij) .<br />

PO axj<br />

The first c<strong>on</strong>tributi<strong>on</strong>, which may equally be written p;’(B * V)B, represents<br />

a c<strong>on</strong>tributi<strong>on</strong> to the force associated with curvature of B-lines directed<br />

towards the centre of curvature; the sec<strong>on</strong>d c<strong>on</strong>tributi<strong>on</strong>, which may be<br />

written -(2p0)-’VB2, represents (minus) the gradient of ‘magnetic pressure’<br />

pLw = (2Po)-lB2. (3.2)<br />

In an incompressible fluid of c<strong>on</strong>stant density p, the equati<strong>on</strong> of moti<strong>on</strong><br />

including these c<strong>on</strong>tributi<strong>on</strong>s to the Lorentz force may be written in the<br />

form<br />

2U 1<br />

-+v.VU =-V~+-B.VB+~V~~,<br />

at<br />

(3.3)<br />

POP<br />

where<br />

x = (P + PM/r>/P 7 (3.4)<br />

and v is the kinematic viscosity of the fluid. Coupled with the inducti<strong>on</strong><br />

equati<strong>on</strong> (2.4) and appropriate boundary c<strong>on</strong>diti<strong>on</strong>s, this determines the<br />

evoluti<strong>on</strong> of the fields ( ~(x, t), B(x, t)}.<br />

We may illustrate this with reference to a phenomen<strong>on</strong> of central importance<br />

in MHD, namely the ability of the medium to support transverse wave<br />

moti<strong>on</strong>s in the presence of a magnetic field.<br />

3.1 AlfvCn waves<br />

Suppose that the medium is of infinite extent, and that we perturb about a<br />

state of rest in which the fluid is permeated by a uniform magnetic field Bo.<br />

Let U(X, t) and B = Bo + b(x, t) be the perturbed velocity and magnetic fields.<br />

Then with the notati<strong>on</strong><br />

v = (pop)-1/2Bo, h = (,~op)-”~b, (3.5)<br />

the linearized forms of equati<strong>on</strong>s (2.4) and (3.3) (neglecting squares and<br />

products of U and b) are<br />

2v/Zt = -Vx + V * Vh + v V~V,<br />

2h/dt = ( V * V)U + qV2h.<br />

(3.6)


7 Rejecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 355<br />

These equati<strong>on</strong>s admit wave-like soluti<strong>on</strong>s with<br />

= c<strong>on</strong>st. and<br />

provided<br />

(U, h}<br />

Hence, for a n<strong>on</strong>-trivial soluti<strong>on</strong>,<br />

e<br />

i(k*x-ur)<br />

(-ice + vk2)u = i(k * ~)h,<br />

(-im + qk2)h = i(k - V)U .<br />

(io - vk2)(io - qk2) = -(k - v)~,<br />

which is a quadratic equati<strong>on</strong> for CO with roots<br />

(3.7)<br />

CO = -ii(q + v)k2 {4(k V)2 - (q - v)”‘}li2 . (3.10)<br />

In the ideal-fluid limit (q = 0, v = 0), the roots are real:<br />

CO = f(k* V), (3.11)<br />

a dispersi<strong>on</strong> relati<strong>on</strong>ship of remarkable simplicity. Note that the group<br />

velocity associated with these waves is<br />

cg=vkCO=fV, (3.12)<br />

a result evidently independent of the wave-vector k. The associated n<strong>on</strong>dispersive<br />

waves are known, after their discoverer, as Alfven waves, and the<br />

velocity V is the Alfven velocity (Alfven 1950).<br />

When q + v # 0, these waves are invariably damped. In the liquid-metal<br />

c<strong>on</strong>text, the damping is predominantly due to magnetic resistivity (q >> v);<br />

if, moreover, the field Bo is str<strong>on</strong>g in the sense that lk. Vl >> q2k4, then (3.8)<br />

approximates to<br />

CO rn -iiqk2+/i. V, (3.13)<br />

and the nature of the damping is clear.<br />

If, <strong>on</strong> the other hand, lk. Vl


356 H. K. Moffatt<br />

4.1 Two historic experiments<br />

4 Electromagnetic shaping and stirring<br />

Two early papers provide examples of the manner in which applied magnetic<br />

fields and/or currents may c<strong>on</strong>tribute to the shaping of the regi<strong>on</strong><br />

occupied by a c<strong>on</strong>ducting fluid and to the stirring of the fluid within this<br />

regi<strong>on</strong>. The first is that of Northrup (1907) who described an experiment<br />

in which a steady current is passed through a layer of c<strong>on</strong>ducting liquid<br />

(sodium-potassium alloy NaK) covered by a layer of (n<strong>on</strong>-c<strong>on</strong>ducting) oil.<br />

The c<strong>on</strong>figurati<strong>on</strong> described by Northrup is shown, in plan and elevati<strong>on</strong>,<br />

in figure 2. The steady current j(x) passes through a c<strong>on</strong>stricti<strong>on</strong> between<br />

two electrodes <strong>on</strong> the boundary; the increased current density in this regi<strong>on</strong><br />

gives rise to an increased magnetic field encircling the current lines (via<br />

Ampere’s Law) ; the resulting Lorentz force distributi<strong>on</strong> causes depressi<strong>on</strong><br />

of the oil/NaK interface. Northrup attributes this observati<strong>on</strong> to his friend<br />

Car1 Hering who, he says, “jocosely called it the ‘pinch phenomen<strong>on</strong>’ ”. The<br />

famous ‘pinch effect’ is precisely the effect of c<strong>on</strong>tracti<strong>on</strong> of a (compressible)<br />

cylindrical column of fluid carrying an axial current due to the self-induced<br />

radial magnetic pressure gradient, a behaviour of fundamental importance in<br />

plasma c<strong>on</strong>tainment devices (figure 3) ; as Northrup and Hering recognized,<br />

a similar effect occurs even in liquid metals under experimentally realizable<br />

laboratory c<strong>on</strong>diti<strong>on</strong>s. In the complex geometry of figure 2, the n<strong>on</strong>linear<br />

deformati<strong>on</strong> (i.e. shaping) of the interface is determined by a balance between<br />

Lorentz forces, gravity and surface tensi<strong>on</strong>; this is too complex to be<br />

calculated analytically, and numerical techniques would be required to solve<br />

this type of three-dimensi<strong>on</strong>al problem.<br />

Northrup also commented that the liquid <strong>on</strong> the inclined surface of the<br />

interface ‘showed great agitati<strong>on</strong>’. This is a manifestati<strong>on</strong> of the stirring<br />

effect, which, as will be shown below, inevitably accompanies the shaping<br />

influence of the magnetic field.<br />

The sec<strong>on</strong>d historic experiment is that described by Braunbeck (1932) and<br />

illustrated in figure 4. Here, a cylindrical capsule c<strong>on</strong>taining liquid metal is<br />

suspended <strong>on</strong> a torsi<strong>on</strong> wire, and is subjected to a rotating magnetic field, as<br />

indicated. Such a field may be regarded as the superpositi<strong>on</strong> of two alternating<br />

fields in quadrature and at right angles, and may be easily produced<br />

using a three-phase power supply. Braunbeck observed that the capsule<br />

rotated in the directi<strong>on</strong> of rotati<strong>on</strong> of the field, achieving an equilibrium<br />

angle of rotati<strong>on</strong> depending <strong>on</strong> the c<strong>on</strong>ductivity of the liquid; the device<br />

may thus in principle be calibrated to determine the c<strong>on</strong>ductivity of small<br />

samples of liquid metal. What is happening in equilibrium is that the liquid


7 Rejecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 351<br />

.1<br />

Oil<br />

NaK<br />

Figure 2. Sketch of Northrup’s (1907) experiment: current flows through the NaK<br />

under a layer of oil; the c<strong>on</strong>stricti<strong>on</strong> leads to a Lorentz force distributi<strong>on</strong> which<br />

depresses the oil/NaK interface. (a) Plan; (b) elevati<strong>on</strong>.<br />

><br />

4 B ><br />

Figure 3. The pinch effect: axial current in the cylinder generates an azimuthal<br />

magnetic field; the resulting magnetic pressure gradient causes radial c<strong>on</strong>tracti<strong>on</strong> of<br />

the cylinder.<br />

metal rotates within the stati<strong>on</strong>ary c<strong>on</strong>tainer, exerting <strong>on</strong> it a viscous torque<br />

which is balanced by the restoring torque transmitted by the torsi<strong>on</strong> wire.<br />

We shall analyse the details of this flow, and describe certain limitati<strong>on</strong>s of<br />

the descripti<strong>on</strong>, below. Here, we may simply say that the stirring effect of<br />

an alternating or rotating magnetic field is a fundamental mechanism that is


358 H. K. Mofatt<br />

Figure 4. Sketch of the experiment of Braunbeck (1932) : a capsule, c<strong>on</strong>taining liquid<br />

metal and suspended <strong>on</strong> a torsi<strong>on</strong> wire, is subjected to a rotating magnetic field. This<br />

induces rotati<strong>on</strong> in the fluid; the viscous torque <strong>on</strong> the capsule is then in equilibrium<br />

with the torque transmitted by the wire.<br />

widely exploited in liquid-metal technology; thus, for example, it is used in<br />

the c<strong>on</strong>tinuous casting of steel to stir the melt before solidificati<strong>on</strong> in order<br />

to produce a more homogeneous end-product (see, for example, Moffatt &<br />

Proctor 1984).<br />

4.2 Electrically induced stirring<br />

Northrup’s problem as described above is <strong>on</strong>e of a general class of problems<br />

in which a steady current distributi<strong>on</strong> j(x) is established in the fluid in<br />

a domain 9 through prescripti<strong>on</strong> of an electrostatic potential distributi<strong>on</strong><br />

cps(x) <strong>on</strong> the boundary S of this domain. This current, together with any<br />

external current closing the circuit, is the source for the magnetic field<br />

B(x), and flow is driven by the force F = j A B. If R,


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 359<br />

8=0<br />

Figure 5. The weldpool problem: the Lorentz force drives the fluid towards the axis<br />

8 = 0 and a str<strong>on</strong>g jet flow develops al<strong>on</strong>g this axis; the flow is subject to instability<br />

involving swirl about this axis (BojareviEs et al. 1989).<br />

If the boundary S is entirely rigid, then the electromagnetic problem<br />

thus defined is c<strong>on</strong>veniently decoupled from the fluid dynamical problem,<br />

and may be solved, either analytically or numerically, as a preliminary to<br />

determinati<strong>on</strong> of the force field F and the resulting flow field U. If, <strong>on</strong> the<br />

other hand, as in Northrup’s problem, part of the surface S is an interface<br />

with another fluid (usually n<strong>on</strong>-c<strong>on</strong>ducting) then the dynamical problem of<br />

determining the shape of S is coupled with the electromagnetic problem of<br />

determining j and B, an altogether more complex situati<strong>on</strong>. Two problems of<br />

practical importance, described in the following subsecti<strong>on</strong>s, have attracted<br />

much interest.<br />

4.2.1 The weldpool problem<br />

Here, current is injected through a point electrode at the surface S; the<br />

current spreads out radially (figure 5), and the c<strong>on</strong>figurati<strong>on</strong> is locally axisymmetric<br />

about the directi<strong>on</strong> of injecti<strong>on</strong>. Both j and B are singular at<br />

the point of injecti<strong>on</strong>, but of course these singularities are removed if the<br />

finite size of any real electrode is allowed for. Even so, the soluti<strong>on</strong> to this<br />

problem has some curious features that deserve comment here; for extended<br />

discussi<strong>on</strong>, see BojareviEs et al. (1989).<br />

With local spherical polar coordinates (r, 8, cp) as indicated in figure 5, the


360 H. K. Moffutt<br />

current in the fluid is given by<br />

and the corresp<strong>on</strong>ding field B is then given by<br />

,LQJ sin 6<br />

2nr( 1 + cos 6)<br />

Note the singularities at r = 0. The Lorentz force is given by<br />

(4.4)<br />

( 47c2r3 1 + cos6<br />

--p0J2 sin6 ,o) ,<br />

F=jr\B= 0,-<br />

(4.5)<br />

This force, directed towards the axis 6 = 0, tends to drive the fluid towards<br />

this axis; being incompressible, the fluid has no alternative but to flow out<br />

al<strong>on</strong>g the axis in the form of a str<strong>on</strong>g axisymmetric jet, which becomes<br />

increasingly ‘focused’ for decreasing values of the fluid viscosity.<br />

Experimental realizati<strong>on</strong> of this flow indicates a behaviour that is not<br />

yet fully understood: the flow is subject to a str<strong>on</strong>g symmetry-breaking<br />

instability in which the fluid sp<strong>on</strong>taneously rotates, in <strong>on</strong>e directi<strong>on</strong> or the<br />

other, about the axial directi<strong>on</strong> 6 = 0. It is believed that this rotati<strong>on</strong> c<strong>on</strong>trols<br />

the singularity of axial velocity that otherwise occurs if the dimensi<strong>on</strong>less<br />

parameter<br />

K = poJ2/p~j2 (4.6)<br />

exceeds a critical value K, of order 300; a recent discussi<strong>on</strong> of this perplexing<br />

phenomen<strong>on</strong> is given by Davids<strong>on</strong> et al. (1999).<br />

4.2.2 Aluminium smelting<br />

The essential ingredients of the industrial processes by which aluminium<br />

is extracted from the raw material cryolite (an electrolytic salt, sodium<br />

aluminium fluoride) are shown in figure 6. Current passes from the anodes<br />

to the cathode through the cryolite and a layer of molten aluminium grows<br />

from the cathode. There is now a jump of c<strong>on</strong>ductivity across the fluid/fluid<br />

interface so that the current lines are refracted across this interface. The<br />

magnetic field B is again that due to the current distributi<strong>on</strong> in the fluids<br />

and the external circuit. The Lorentz force drives a flow in both fluids, the<br />

shape of the interface being c<strong>on</strong>trolled by gravity (surface tensi<strong>on</strong> effects<br />

being weak <strong>on</strong> the scale of this industrial process). The process is further<br />

complicated by instabilities of the interface to which the steady c<strong>on</strong>figurati<strong>on</strong><br />

may be subject (for detailed discussi<strong>on</strong>, see Moreau 1990). This is a problem<br />

of enormous industrial importance: in effect the efficiency of the process is


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 361<br />

I 1 Anodes<br />

v 1 / t f / f I<br />

.pF<br />

I I I<br />

Cathode<br />

Figure 6. Sketch of the aluminium smelting process (Moreau 1990). The interface<br />

between the liquid cryolite and the molten aluminium is subject to instabilities of<br />

magnetohydrodynamic origin, which limit the efficiency of the process.<br />

limited by instabilities that may lead to c<strong>on</strong>tact between the aluminium and<br />

the anodes, a 'short-circuiting' that would terminate the process. This is a<br />

billi<strong>on</strong>-dollar industry for which an understanding of the fundamentals of<br />

magnetohydrodynamics would appear to be a first essential.<br />

4.3 Inductive stirring<br />

An equally important stirring mechanism is that associated with the applicati<strong>on</strong><br />

of an alternating (AC) magnetic field<br />

B(x,t) = Re B(x)e-i"t ,<br />

[^ 1<br />

(4.71<br />

such a field being produced by AC currents in external circuits (figure 7).<br />

The field diffuses into the c<strong>on</strong>ductor and a Lorentz force distributi<strong>on</strong><br />

F(x, t) = j AB is established. This force distributi<strong>on</strong> has a time-averaged part<br />

P(x) and an additi<strong>on</strong>al periodic part with period 2co and zero mean. The<br />

resulting flow c<strong>on</strong>sists of a steady part driven by P(x), and a time-periodic<br />

part whose amplitude is c<strong>on</strong>trolled by the inertia of the fluid. We shall focus<br />

<strong>on</strong> the steady part in the present discussi<strong>on</strong>, bearing in mind that this steady<br />

part may be unstable, in which case a turbulent resp<strong>on</strong>se to the force P(x)<br />

may be envisaged. We suppose that R,


362 H. K. Moftatt<br />

C’<br />

Figure 7. Stirring induced by AC fields; the sketch indicates the flow pattern that<br />

would be driven by a travelling magnetic field from the source coils C and a<br />

stati<strong>on</strong>ary AC field from the coils C; a wide range of patterns of stirring may be<br />

generated by appropriate engineering of the external coils and appropriate choice<br />

of field frequency U.<br />

thickness of this boundary layer is<br />

and is small when CO is large. Outside the c<strong>on</strong>ductor, the field &x) is given<br />

by<br />

B = V@, V2@ = 0, (4.9)<br />

and a boundary c<strong>on</strong>diti<strong>on</strong>, which in the high-frequency limit is just<br />

2@/2n=O <strong>on</strong> S. (4.10)<br />

In effect, the field is perfectly excluded from the c<strong>on</strong>ductor in this limit. Of<br />

course, the potential @ has ‘prescribed’ singularities at the external coils. Let<br />

us regard this ‘external’ problem as solved; then Bs = (V@)S is known, as a<br />

tangential field <strong>on</strong> the surface S. Note that the surface divergence of Bs is<br />

n<strong>on</strong>-zero in general; in fact<br />

(4.11)<br />

Within the skin inside the c<strong>on</strong>ductor, the field B and hence j^ can be<br />

calculated by standard methods; the mean force f = ;Re(? A h) may then<br />

be found; here, the * denotes the complex c<strong>on</strong>jugate. Details may be found


in Moffatt (198521); the result is<br />

7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 363<br />

(4.12)<br />

where i is a normal coordinate directed into the c<strong>on</strong>ductor. The first term<br />

here is a str<strong>on</strong>g inwardly directed normal comp<strong>on</strong>ent, which is resp<strong>on</strong>sible<br />

for the shaping of any part of S that is free to move in the normal directi<strong>on</strong>.<br />

By c<strong>on</strong>trast, stirring of the fluid is associated with the curl of F, given by<br />

where<br />

(4.13)<br />

(4.14)<br />

a vector field defined <strong>on</strong> the surface S, and having the dimensi<strong>on</strong>s of an<br />

accelerati<strong>on</strong>.<br />

This force distributi<strong>on</strong> will clearly generate a highly sheared flow within<br />

the boundary layer (thickness 0(8~)). We may estimate the net effect, <strong>on</strong> the<br />

reas<strong>on</strong>able assumpti<strong>on</strong> that fluid inertia is negligible compared with viscous<br />

effects within the layer. On this ‘lubricati<strong>on</strong>’ assumpti<strong>on</strong> the velocity field<br />

within the layer, assuming a no-slip c<strong>on</strong>diti<strong>on</strong> at i = 0, is<br />

so that asymptotically<br />

(4.15)<br />

(4.16)<br />

What this means is that, under the influence of the magnetic field, the no-slip<br />

c<strong>on</strong>diti<strong>on</strong> must be replaced by a c<strong>on</strong>diti<strong>on</strong> of ‘prescribed tangential velocity’<br />

as given by (4.16) for the driven flow in the interior of the c<strong>on</strong>ductor.<br />

A corresp<strong>on</strong>ding analysis using a free-surface c<strong>on</strong>diti<strong>on</strong> 2u/2C: = 0 <strong>on</strong><br />

5 = 0 leads to an effective ‘stress’ c<strong>on</strong>diti<strong>on</strong><br />

7s = -PV (2~l2i) I;== = $P~MQ~(.) (4.17)<br />

as a boundary c<strong>on</strong>diti<strong>on</strong> for the interior flow. Note that this effective stress<br />

is independent of kinematic viscosity v.<br />

Thus, in the high-frequency limit, the influence of the magnetic field is<br />

c<strong>on</strong>fined to the magnetic boundary layer, or skin, in such a way as to simply<br />

replace the no-slip (or zero-stress) c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong> S by an effective slip (or<br />

effective stress) distributi<strong>on</strong> <strong>on</strong> S, which is then resp<strong>on</strong>sible for generating an<br />

internal flow, a flow whose topology will clearly be influenced, and indeed


364 H. K. Moflatt<br />

c<strong>on</strong>trolled, by the functi<strong>on</strong> Qs(x), which is determined through (4.14) by the<br />

surface field Bs(x).<br />

As the frequency is decreased, the skin depth JLw increases, and the above<br />

simple descripti<strong>on</strong> ceases to be valid; the force distributi<strong>on</strong> penetrates more<br />

and more into the interior of the fluid, and extends throughout the fluid<br />

when 8-v increases to the scale L of the fluid domain.<br />

4.3.2 The case of a circular cylinder<br />

The prototype example is just as c<strong>on</strong>ceived by Braunbeck (1932): a horiz<strong>on</strong>tal<br />

rotating field applied to a circular cylinder with axis vertical and c<strong>on</strong>taining<br />

c<strong>on</strong>ducting fluid. Here,<br />

Bs = 2Boeieee, (4.18)<br />

where Bo is the value of field magnitude far from the cylinder, and 0 is the<br />

angular coordinate. Hence, in this case, IBs12 is uniform <strong>on</strong> S, and<br />

2i<br />

Vs Bs = -Boeie<br />

U<br />

<strong>on</strong> S<br />

(4.19)<br />

Hence, from (4.14) and (4.15), the effective slip velocity just inside the skin is<br />

21 = (11B;iPofwva) ee. (4.20)<br />

The resulting moti<strong>on</strong> is a rigid body rotati<strong>on</strong> of the fluid inside this skin,<br />

with angular velocity<br />

SZ = YBi/popova2, (4.21)<br />

a result obtained originally by Moffatt (1965). The viscous stress and resulting<br />

couple G (per unit length of cylinder) acting <strong>on</strong> the cylinder are easily<br />

calculated, with the result<br />

Bi<br />

G=2rr(?) -.<br />

POP<br />

(4.22)<br />

Remarkably, this couple, although of viscous origin, is independent of v;<br />

this is because the core angular velocity, given by (4.21), is proporti<strong>on</strong>al<br />

to vl. In Braunbeck’s experiment, it is the torque (4.22) that is ultimately<br />

in equilibrium with the torque transmitted by the torsi<strong>on</strong> wire, which is<br />

proporti<strong>on</strong>al to the net angle of rotati<strong>on</strong>; thus measurement of this angle<br />

provides a means of determinati<strong>on</strong> of 11, and hence of the c<strong>on</strong>ductivity g.<br />

There are of course a number of limitati<strong>on</strong>s of the above type of analysis<br />

that should be borne in mind. First, the high-frequency approximati<strong>on</strong> is<br />

valid <strong>on</strong>ly if 6.v > q/a2. This is not however a serious restricti<strong>on</strong>;


7 Re$ecti<strong>on</strong>s <strong>on</strong> Magnetohydvodynamics 365<br />

the problem can be solved exactly for arbitrary CO in terms of Bessel functi<strong>on</strong>s.<br />

More seriously however, the analysis fails if the field strength Bo becomes<br />

so str<strong>on</strong>g that Q in (4.21) becomes comparable with CO; for it is in fact not<br />

the absolute angular velocity w of the field that is relevant to the inducti<strong>on</strong><br />

of currents, but rather the relative angular velocity w - Q between field and<br />

fluid. This str<strong>on</strong>g field situati<strong>on</strong> requires a major modificati<strong>on</strong> of approach<br />

(see, for example, Moreau 1990, where problems of this type are extensively<br />

treated).<br />

4.3.3 Magnetic lecitati<strong>on</strong><br />

A high-frequency field also offers the possibility of levitating a volume<br />

(usually a small droplet) of c<strong>on</strong>ducting liquid in the complete absence of<br />

any rigid boundary support (figure 8); here, it may be more appropriate to<br />

talk of a magnetic basket rather than a magnetic bottle! The field is again<br />

expelled from the c<strong>on</strong>ductor (except in a thin skin); the first c<strong>on</strong>tributi<strong>on</strong> to<br />

the force F in (4.12) corresp<strong>on</strong>ds to the effect of magnetic pressure p . <strong>on</strong> ~<br />

the surface S; if S can so adjust itself that each vertical column of liquid<br />

of length Az is support by the magnetic pressure difference Ap-w between<br />

bottom and top, i.e.<br />

APM = PgAz , (4.23)<br />

then levitati<strong>on</strong> is possible. This can be achieved by placing the main current<br />

sources of the field B below the liquid sample as in the c<strong>on</strong>figurati<strong>on</strong><br />

envisaged in figure 8(a). In the axisymmetric c<strong>on</strong>figurati<strong>on</strong> of figure 8(b),<br />

the magnetic pressure vanishes <strong>on</strong> the axis of symmetry, and (4.23) cannot<br />

be satisfied; <strong>on</strong>e must rely <strong>on</strong> surface tensi<strong>on</strong> to compensate gravity for the<br />

column of fluid immediately adjacent to this axis.<br />

C<strong>on</strong>figurati<strong>on</strong>s of this kind have been studied in some detail by Mestel<br />

(1982) and by Sneyd & Moffatt (1982). As pointed out in the latter paper, the<br />

effective stress given by (4.17) drives an interior flow with closed streamlines,<br />

whose intensity is limited <strong>on</strong>ly by viscosity. This is because integrati<strong>on</strong> of<br />

the steady equati<strong>on</strong> of moti<strong>on</strong> in the form<br />

O=VACO-V -+-U +F-vVACO (4.24)<br />

(; ?><br />

round a closed streamline C gives<br />

(4.25)<br />

i.e. inertia and pressure forces play no part in the equilibrium that is established.<br />

In practice, the viscosity in liquid metals is small; put another way,


366<br />

H. K. MofSatt<br />

(a)<br />

Axis of<br />

symmetry<br />

Levitated molten<br />

current<br />

Figure 8. Magnetic levitati<strong>on</strong> by high-frequency magnetic field, providing support<br />

via magnetic pressure distributi<strong>on</strong> <strong>on</strong> the surface (Sneyd & Moffatt 1982).<br />

the Reynolds number of the flow that is necessarily driven in the interior of<br />

a levitated droplet of radius of the order of 10 mm or greater is very large.<br />

It seems likely therefore that this interior flow will in these circumstances be<br />

turbulent. The ‘great agitati<strong>on</strong>’ referred to by Northrup (see $4.1 above) also


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 367<br />

presumably indicated a state of turbulent flow just inside the liquid metal -<br />

and for similar reas<strong>on</strong>s.<br />

5 Dynamo theory<br />

Dynamo theory is c<strong>on</strong>cerned with explaining the origin of magnetic fields<br />

in stars, planets and galaxies. Such fields are produced by currents in the<br />

interior regi<strong>on</strong>s, which would in the normal course of events be subject to<br />

ohmic decay, in the same way that current in an electric circuit decays if not<br />

maintained by a battery. This decay can however be arrested, and indeed<br />

reversed, through inductive effects associated with fluid moti<strong>on</strong>; when this<br />

happens, the fluid system acts as a self-exciting dynamo. The magnetic field<br />

grows sp<strong>on</strong>taneously from an arbitrary weak initial level, in much the same<br />

way as any other perturbati<strong>on</strong> of an intrinsically unstable situati<strong>on</strong>.<br />

The possibility of such ‘dynamo’ instabilities may be understood with<br />

reference to the inducti<strong>on</strong> equati<strong>on</strong> (2.4), which governs field evoluti<strong>on</strong> if<br />

the velocity field v(x,t) in the fluid is regarded as ‘given’. Let us suppose<br />

for simplicity that the fluid fills all space, and that the velocity field is<br />

steady, i.e. U = u(x). We must suppose also that B has no ‘sources at<br />

infinity’ which would mitigate against the c<strong>on</strong>cept of an internally generated<br />

dynamo.<br />

We may seek ‘normal mode’ soluti<strong>on</strong>s of (2.4) of the form<br />

where, by substituti<strong>on</strong>,<br />

pB = v A (U AB) + yv2B. (5.2)<br />

When coupled with the requirement that &x) should be either a localized<br />

field of finite energy or, for example, a space-periodic field if U is<br />

space-periodic, this c<strong>on</strong>stitutes an eigenvalue problem which (in principle)<br />

determines a sequence of possibly complex eigenvalues p1, p2,. . . , which may<br />

be ordered so that<br />

Repl 3 Rep2 3 Rep3 3 ... , (5.3)<br />

and corresp<strong>on</strong>ding ‘eigenfields’ B~(x), &(x), .,.. If Repl > 0, then the corresp<strong>on</strong>ding<br />

field<br />

B(x, t) = Re [B~(x) ePlf] (5.4)<br />

exhibits dynamo behaviour: it grows exp<strong>on</strong>entially in intensity, the growth


368 H. K. Mofatt<br />

being oscillatory or n<strong>on</strong>-oscillatory according to whether Imp1 # 0 or = 0.<br />

The mode of maximum growth-rate is clearly the <strong>on</strong>e that will emerge from<br />

an arbitrary initial c<strong>on</strong>diti<strong>on</strong> in which all modes may be present.<br />

If this exp<strong>on</strong>ential growth occurs, then it can persist <strong>on</strong>ly for as l<strong>on</strong>g as<br />

the velocity field u(x) remains unaffected by the Lorentz force. This is the<br />

‘kinematic phase’ of the dynamo process. Obviously, however, the Lorentz<br />

force increases exp<strong>on</strong>entially with growth rate 2Re p1 (c<strong>on</strong>sidering <strong>on</strong>ly the<br />

mode (5.4)) and so ultimately the back-reacti<strong>on</strong> of the Lorentz force <strong>on</strong><br />

the fluid moti<strong>on</strong> must be taken into account. This is the ‘dynamic phase’ of<br />

dynamo acti<strong>on</strong> in which the nature of the supply of energy to the system (via<br />

the dynamic equati<strong>on</strong> of moti<strong>on</strong>) must be c<strong>on</strong>sidered. While great progress<br />

has been made over the past 50 years towards a full understanding of the<br />

kinematic phase, the highly n<strong>on</strong>linear dynamic phase has proved far more<br />

intractable from an analytical point of view, and is likely to remain a focus<br />

of much research effort, both computati<strong>on</strong>al and analytical, over the next<br />

few decades.<br />

5.1 Fast and slow dynamos<br />

During the kinematic phase, the growth rate (p say) of the most unstable<br />

mode is determined in principle by the velocity field u(x) and the parameter<br />

y which appears in equati<strong>on</strong> (5.2). If u(x) is characterized by velocity scale<br />

uo and length scale lo, then <strong>on</strong> dimensi<strong>on</strong>al grounds,<br />

where, as in $2.1, R, = t’olo/y. An important distincti<strong>on</strong> between dynamos<br />

described as ‘fast’ or ‘slow’ has been introduced by Vainshtein & Zel’dovich<br />

(1972). A fast dynamo is <strong>on</strong>e for which f(Rm) = 0(1) as R, -+ x, i.e.<br />

the growth rate p scales <strong>on</strong> the dynamic time scale lo/vo. A slow dynamo,<br />

by c<strong>on</strong>trast, is <strong>on</strong>e for which f(R,) + 0 as R, -+ E; for example, if<br />

f(R,) - R,1’2 as R, -+ x, then p - (co/l~)R~’”, and diffusivity y c<strong>on</strong>tinues<br />

to influence the growth rate even in the limit y + 0. The distincti<strong>on</strong> is an<br />

important <strong>on</strong>e because, <strong>on</strong> the galactic scale, R, is extremely large, and a<br />

slow dynamo is likely to have little relevance in such c<strong>on</strong>texts. This has led to<br />

an intensive search for dynamos that can legitimately be described as fast (see<br />

Childress & Gilbert 1995); however, in the strict sense indicated above, no<br />

such dynamo has yet been found! All known dynamos are slow; diffusivity<br />

remains important no matter how small y may be. The situati<strong>on</strong> is again<br />

somewhat analogous to that governing the vorticity equati<strong>on</strong> in turbulent<br />

flow: viscous effects remain important (in providing the mechanism for


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 369<br />

dissipati<strong>on</strong> of kinetic energy) no matter how small the kinematic viscosity 1’<br />

may be.<br />

When a dynamo enters the dynamic phase (assuming that sufficient time<br />

is available for it to do so) the distincti<strong>on</strong> between fast and slow behaviour<br />

disappears; in either case, the growth rate must decrease, ultimately to zero<br />

when an equilibrium between generati<strong>on</strong> of magnetic field by the (modified)<br />

velocity field and ohmic dissipati<strong>on</strong> of magnetic field is established.<br />

’<br />

5.2 A little historical digressi<strong>on</strong><br />

The history of dynamo theory up to 1957, when Cowling’s seminal m<strong>on</strong>ograph<br />

<strong>Magnetohydrodynamics</strong> was published, was characterized by the gravest<br />

uncertainty as to whether any form of self-exciting dynamo acti<strong>on</strong> in a spherical<br />

body of fluid of uniform c<strong>on</strong>ductivity was possible at all. One of the<br />

main firm results in this regard was negative: this was Cowling’s (1934)<br />

theorem, which stated in its simplest form that steady axisymmetric dynamo<br />

acti<strong>on</strong> is impossible. This reinforced the view prevalent at the time, and<br />

originally stated in the geomagnetic c<strong>on</strong>text by Schuster (1912) that ‘the<br />

difficulties which stand in the way of basing terrestrial magnetism <strong>on</strong> electric<br />

currents inside the earth are insurmountable’. Fortunately this pessimistic<br />

c<strong>on</strong>clusi<strong>on</strong> has been eroded and now completely reversed with the passage<br />

of time. For example, Cook (1980) writes that ‘there is no theory other than<br />

a dynamo theory that shows any sign of accounting for the magnetic fields<br />

of the planets’. This view is echoed by Jacobs (1984) who writes ‘There has<br />

been much speculati<strong>on</strong> <strong>on</strong> the origin of the Earth’s magnetic field ... The<br />

<strong>on</strong>ly possible means seems to be some form of electromagnetic inducti<strong>on</strong>,<br />

electric currents flowing in the Earth’s core’.<br />

If we try to identify <strong>on</strong>e single development over the last half-century that<br />

has revoluti<strong>on</strong>ized our view of the subject, that development must surely be<br />

the ‘mean-field electrodynamics’ proposed in the seminal paper of Steenbeck,<br />

Krause & Radler (1966), and foreshadowed in earlier ‘pre-seminal’ papers<br />

of Parker (1955) and Braginski (1964). This theory, which led in the fully<br />

turbulent c<strong>on</strong>text to the discovery of the famous x-effect - the appearance<br />

of a mean electromotive force parallel to the mean magnetic field - lies<br />

at the heart of the dynamo process as currently understood. In effect, it<br />

provides a starting point for any modern approach to dynamo theory; as<br />

such, it merits the closest study. A simplified account is presented in the<br />

following sub-secti<strong>on</strong>s. A systematic treatment may be found in the research<br />

m<strong>on</strong>ographs of Moffatt (1978) and Krause & Radler (1980). Extensive<br />

treatment, with particular reference to astrophysical applicati<strong>on</strong>, may also


3 70 H. K. MofSatt<br />

be found in the books of Parker (1979) and Zeldovich, Ruzmaikin & Sokoloff<br />

(1983).<br />

5.3 Mean-field electrodynamics<br />

Let us suppose that the velocity U (X, t) appearing in equati<strong>on</strong> (2.4) is turbulent<br />

- i.e. random in both space and time (see Chapter 5). We may suppose that<br />

there is some source of energy for this turbulence (e.g. through some random<br />

stirring mechanism) so that v is statistically stati<strong>on</strong>ary in time. It is then<br />

natural to adopt the notati<strong>on</strong> ( . . . ) for a time average, over any interval<br />

l<strong>on</strong>g compared with the time scale to = lo/t'o characteristic of the energyc<strong>on</strong>taining<br />

eddies of the turbulence; here lo is the scale of these eddies and<br />

00 the r.m.s. value of v. For simplicity, we may suppose that (U) = 0.<br />

The field B(x, t) may be decomposed into mean and fluctuating parts:<br />

B(x, t) = Bo(x, t) + b(x, t), (5.6)<br />

where Bo(x, t) = (B(x, t)) and (6) 0. Here we allow Bo to depend <strong>on</strong> t; this<br />

must be interpreted as allowing for slow evoluti<strong>on</strong> <strong>on</strong> a time scale T much<br />

greater than to. Separati<strong>on</strong> of the time scales to and T (to > lo) <strong>on</strong> which the mean<br />

field develops.<br />

The mean of equati<strong>on</strong> (2.4) gives an evoluti<strong>on</strong> equati<strong>on</strong> for Bo(x, t):<br />

where 8 = (U A b), the mean electromotive force arising through interacti<strong>on</strong><br />

of the fluctuating fields. Now the problem is like the 'closure' problem of<br />

turbulence: we need to find a relati<strong>on</strong>ship between B and Bo in order to<br />

solve (5.7). But now, in c<strong>on</strong>trast to the intractable problem of turbulence, we<br />

have the separati<strong>on</strong> of scales to help us.<br />

Subtracting (5.7) from (2.4) gives an equati<strong>on</strong> for the fluctuating field b:<br />

o'b<br />

- = v A ( U ABo) t V A ( U A b - 8) + qV2b<br />

2t<br />

Without further approximati<strong>on</strong>, it is difficult, if not impossible, to solve for<br />

b in terms of Bo and v. However, we may make progress by noting that, for<br />

given U(X, tj, equati<strong>on</strong> (5.8) establishes a linear relati<strong>on</strong>ship between b and<br />

Bo; and hence, since Q is linearly related to b, between Q and Bo.<br />

Now both Q and Bo are average fields varying <strong>on</strong> the slow time scale, and


7 Rejecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 371<br />

<strong>on</strong> the large length scale L; hence this linear relati<strong>on</strong>ship between € and Bo<br />

may be represented by a series of the form<br />

where xij, Pijk,. .. are tensor (actually pseudo-tensor) coefficients which are in<br />

principle determined by the statistical properties of the turbulent field U(X, t)<br />

and the parameter q which intervenes in the soluti<strong>on</strong> of (5.8). Note that<br />

successive terms of (5.9) decrease in magnitude by a factor of order lo/L;<br />

and that any terms involving time derivatives, i.e. dBo/dt, may be eliminated<br />

in favour of space derivatives through recursive appeal to equati<strong>on</strong> (5.7). That<br />

the coefficients rij, Pijk,. . . are pseudo-tensors (rather than tensors) should<br />

be evident from the fact that € is a polar vector (like velocity) whereas Bo<br />

is an axial vector (like angular velocity).<br />

We may go further <strong>on</strong> the simplifying assumpti<strong>on</strong> that the turbulence is<br />

homogeneous and isotropic. In this case, Xij and Pijk share these properties,<br />

i.e. they are isotropic pseudo-tensors invariant under translati<strong>on</strong>, i.e. independent<br />

of x (they are already independent of t from the assumpti<strong>on</strong> that U<br />

is statistically stati<strong>on</strong>ary). Isotropy is a very str<strong>on</strong>g c<strong>on</strong>straint; it implies that<br />

(5.10)<br />

where now x is a pseudo-scalar quantity and P is a pure scalar. Hence (5.9)<br />

simplifies to<br />

€ = aBo - P(V A Bo) + . . . ; (5.11)<br />

the next term in this series involves V A (V A Bo) and so <strong>on</strong>. This is the<br />

required relati<strong>on</strong>ship between € and Bo.<br />

Substituting this back into (5.7) gives the mean-field equati<strong>on</strong> in its simplest<br />

form :<br />

dB0<br />

- = 0 A Bo) + (q + P)V2B0 + ,.,<br />

2t<br />

(5.12)<br />

where the terms indicated by + . . . involve higher derivatives of Bo, and may<br />

presumably be neglected when the scale of variati<strong>on</strong> of Bo is large. It is clear<br />

from the structure of (5.12) that P must be interpreted as an 'eddy diffusivity'<br />

associated with the turbulence (although there is no guarantee from the above<br />

treatment that P must invariably be positive!). The first term <strong>on</strong> the right of<br />

(5.12) will however always dominate the evoluti<strong>on</strong>, provided x f 0 and the<br />

scale of Bo is sufficiently large. Before going further it is therefore essential<br />

to find a means of calculating 3 explicitly and determining the c<strong>on</strong>diti<strong>on</strong>s<br />

under which this key parameter is definitely n<strong>on</strong>-zero.


372 H. K. MofSatt<br />

5.4 First-order smoothing<br />

To do this, it is legitimate to c<strong>on</strong>sider the situati<strong>on</strong> in which Bo is c<strong>on</strong>stant;<br />

the reas<strong>on</strong> for this is that tl is independent of the field Bo(x,t), and we<br />

are free to make any assumpti<strong>on</strong> about this field that simplifies calculati<strong>on</strong><br />

of a. The assumpti<strong>on</strong> that Bo is c<strong>on</strong>stant is equivalent to c<strong>on</strong>sidering the<br />

c<strong>on</strong>ceptual limit L + sc, T + sc.<br />

Under this c<strong>on</strong>diti<strong>on</strong>, the fluctuati<strong>on</strong> equati<strong>on</strong> (5.8) becomes<br />

c‘b<br />

- = (Bo * V)U + V A (v A b - 8) + yV2b.<br />

2t<br />

(5.13)<br />

Now we know that, if the magnetic Reynolds number R, = vol~/y is small,<br />

then the induced field b is weak compared with Bo (actually b = O(R,)Bo;<br />

see below). In this circumstance, the awkward term V A (U A b - €) in (5.8)<br />

is negligible compared with the other terms in the equati<strong>on</strong> and may be<br />

neglected. We then have a linear equati<strong>on</strong> with c<strong>on</strong>stant coefficients which<br />

may be solved by elementary Fourier techniques. Let us then make this<br />

assumpti<strong>on</strong> and explore the c<strong>on</strong>sequences.<br />

It is illuminating to c<strong>on</strong>sider first the situati<strong>on</strong> in which v is a circularly<br />

polarized wave of the form<br />

~(x, t) = vo(sin(kz - cot), cos(kz - cot), 0), (5.14)<br />

and Bo = (O,O, Bo) (see figure 9). Note that the vorticity associated with<br />

(5.14) is<br />

and the associated helicity density,<br />

w=VAv=kv, (5.15)<br />

is maximal (since w is parallel to U) and c<strong>on</strong>stant. Then we easily find that<br />

(5.13) is satisfied provided<br />

where cp = kz - at. We then obtain<br />

(5.18)<br />

For this case of a single ‘helicity’ wave (5.14), v A b turns out to be uniform,<br />

so the awkward term in (5.13) is identically zero! It follows from (5.18) that


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 373<br />

Z<br />

Figure 9. A circularly polarized travelling wave v(x,t) of the form (5.14) induces<br />

and interacts with the field b which is also circularly polarized (equati<strong>on</strong> (5.17)), but<br />

phase-shifted relative to v. The resulting U A b is uniform and parallel to Bo.<br />

B = rBo, where<br />

(5.19)<br />

It is obviously desirable to write the result in this form, both left- and<br />

right-hand sides of the equati<strong>on</strong> being pseudo-scalars.<br />

It is important to note the origin of the ‘r-effect’ c<strong>on</strong>tained in (5.19): it<br />

is the phase-shift between b and v caused by molecular diffusivity y which<br />

leads to a n<strong>on</strong>-zero value for r; it is a strange fact that, although diffusi<strong>on</strong> is<br />

resp<strong>on</strong>sible for the decay of magnetic field in the absence of fluid moti<strong>on</strong>, it<br />

is also resp<strong>on</strong>sible for the appearance of an &-effect which, as shown below,<br />

is a vital ingredient of the self-exciting dynamo process.<br />

The velocity field (5.14) is of course rather special; it should be clear however<br />

that if we c<strong>on</strong>sider a random superpositi<strong>on</strong> of such waves with wavevectors<br />

k, frequencies CO, and amplitudes 6(k, CO) isotropically distributed,<br />

then, provided the awkward term of (5.13) can be neglected, a result generalizing<br />

(5.19) can be obtained; all the c<strong>on</strong>tributi<strong>on</strong>s from different wave


374 H. K. Mofatt<br />

modes are additive, and the final result is<br />

k2%(k, 03) d3k do<br />

a=-:q/J co2+Y2k4 ,<br />

(5.20)<br />

where X(k,co) is the helicity spectrum functi<strong>on</strong> of the velocity field U, with<br />

the property<br />

(U - a) = /I %(k, co) d3k dco . (5.21)<br />

The factor 3 appears in (5.20) from averaging over all directi<strong>on</strong>s. The main<br />

thing to note again is the direct relati<strong>on</strong>ship between x and the helicity of<br />

the turbulent field.<br />

This theory, usually described as ‘first-order smoothing’ theory, is limited<br />

to circumstances in which, as indicated above, lbl


7 Rejecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 375<br />

turns out that the average of v w over horiz<strong>on</strong>tal planes is n<strong>on</strong>-zero and<br />

antisymmetric about the centreplane (Moffatt 1978, chap. 10).<br />

5.5 Dynamo acti<strong>on</strong> associated with the x-effect<br />

Let us return now to the mean field equati<strong>on</strong> (5.12) in the form<br />

(5.23)<br />

where now ye = y +p is the ‘effective diffusivity’, and we drop the suffix 0 <strong>on</strong><br />

Bo, to simplify notati<strong>on</strong>. The simplest way to treat (5.23) is to seek soluti<strong>on</strong>s<br />

which have the Beltrami property<br />

where K is a c<strong>on</strong>stant. For example, the field<br />

VAB = KB, (5.24)<br />

B = (c sin Kz + b cos Ky, a sin Kx + c cos Kz, b sin Ky + acosKx) (5.25)<br />

(the ‘abc’ field) has this property, as may be easily verified. We shall suppose<br />

that K is chosen to have the same sign as x, i.e. xK > 0, and that IKI is<br />

small, so that the scale L - K-’ of B is large compared with the scale lo<br />

of the underlying turbulence that gives rise to the x-effect. Of course (5.23)<br />

implies that<br />

v 2 B = -V A(VAB) = -K~B,<br />

(5.26)<br />

and so (5.23) becomes<br />

so that<br />

where<br />

(5.27)<br />

B(x, t) = B(x, 0) ePt , (5.28)<br />

P=aK-VeK2. (5.29)<br />

Obviously we have exp<strong>on</strong>ential growth of the mean field B provided<br />

aK > yeK2, i.e. provided<br />

IKI < laI/Ye. (5.30)<br />

Thus we have a dynamo effect, in which the mean field (i.e. the field <strong>on</strong> the<br />

large scale L) grows exp<strong>on</strong>entially provided L is large enough.


376 H. K. Moffatt<br />

If we adopt the low-Rm estimate for x (based <strong>on</strong> (5.20)),<br />

(5.31)<br />

assuming 2 - vi/lo, and note that in this lOw-Rm situati<strong>on</strong>, p 1: the field grows <strong>on</strong> a scale L at<br />

which the corresp<strong>on</strong>ding Rm is large.<br />

We can now summarize the implicati<strong>on</strong>s of the above discussi<strong>on</strong>. If a<br />

c<strong>on</strong>ducting fluid is in turbulent moti<strong>on</strong>, the turbulence being homogeneous<br />

and isotropic but having the crucial property of ‘lack of reflecti<strong>on</strong>al symmetry’<br />

(a property that, as observed above, can be induced in a rotating fluid<br />

through interacti<strong>on</strong> of buoyancy-induced c<strong>on</strong>vecti<strong>on</strong> and Coriolis forces),<br />

then a magnetic field will in general grow from an arbitrarily weak initial<br />

level, <strong>on</strong> a scale L large compared with the scale 10 of the turbulence. This is<br />

<strong>on</strong>e of these remarkable situati<strong>on</strong>s in which ‘order arises out of chaos’, the<br />

order being evident in the large-scale magnetic field. It may be appropriate<br />

to follow the example of Richards<strong>on</strong> (1922) by summarizing the situati<strong>on</strong> in<br />

rhyme :<br />

C<strong>on</strong>uecti<strong>on</strong> and diffusi<strong>on</strong>,<br />

In turb’lence with helicity,<br />

Yields order jkom c<strong>on</strong>fusi<strong>on</strong><br />

In cosmic electricity!<br />

This totally general principle applies no matter what the physical c<strong>on</strong>text<br />

may be, whether <strong>on</strong> the planetary, stellar, galactic or even super-galactic<br />

scale. It is this generality that makes the approach described above, which<br />

derives from that pi<strong>on</strong>eered by Steenbeck et al. (1966), so intensely appealing.<br />

5.6 The back-reacti<strong>on</strong> of Lorentz forces<br />

The exp<strong>on</strong>ential growth associated with the dynamo acti<strong>on</strong> described above<br />

must ultimately be arrested by the acti<strong>on</strong> of Lorentz forces which may


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 377<br />

have a two-fold effect: (i) the generati<strong>on</strong> of moti<strong>on</strong> <strong>on</strong> the scale L of the<br />

growing field, and (ii) the suppressi<strong>on</strong> of the turbulence (or at least severe<br />

modificati<strong>on</strong>) <strong>on</strong> the scale lo which provides the r-effect. The latter effect<br />

is the easier to analyse, since <strong>on</strong>e has as a starting point the Alfvkn-wave<br />

type of analysis described in Q 3 above: the effect of a str<strong>on</strong>g locally uniform<br />

magnetic field is to cause the c<strong>on</strong>stituent eddies of the turbulence either to<br />

propagate as Alfvkn waves, or, if resistive damping is str<strong>on</strong>g (as it is when<br />

R, is small), to decay without oscillati<strong>on</strong>.<br />

There are two aspects of the behaviour that are worth noting. First, a<br />

locally uniform magnetic field induces str<strong>on</strong>g local anisotropy in the turbulent<br />

field, modes with vorticity perpendicular to the magnetic field being most<br />

str<strong>on</strong>gly damped. The dynamic effect of the field may be estimated in terms<br />

of a (dimensi<strong>on</strong>less) magnetic interacti<strong>on</strong> parameter N defined by<br />

(5.34)<br />

and becomes str<strong>on</strong>g when the local mean field Bo grows str<strong>on</strong>g so that<br />

N >> 1. It has been shown (Moffatt 1967) that under this c<strong>on</strong>diti<strong>on</strong>, and<br />

provided NR;


378 H. K. Moffatt<br />

I t<br />

Figure 10. Generati<strong>on</strong> of toroidal fields by differential rotati<strong>on</strong>: the poloidal field is<br />

‘gripped’ by the flow, and ‘cranked’ around the axis of symmetry (see, for example,<br />

Moffatt 1978).<br />

in the range 1.3 to 15, that r(k) scales like kW, with ,u = 3.7 0.2, neatly<br />

embracing the value 11/3 = 3.66. In a branch of magnetohydrodynamics in<br />

which experimental results are sparse, this result stands out, and gives hope<br />

that more central aspects of dynamo theory may also so<strong>on</strong> be subject to<br />

experimental investigati<strong>on</strong> and verificati<strong>on</strong>.<br />

5.7 The aw-dynamo<br />

It would be misleading to leave the subject of dynamo theory without menti<strong>on</strong>ing<br />

a mechanism of field generati<strong>on</strong> that is just as important as the<br />

x-effect in the c<strong>on</strong>text of stellar and planetary magnetism, namely the generati<strong>on</strong><br />

of toroidal (or z<strong>on</strong>al) magnetic field from poloidal (or meridi<strong>on</strong>al) field<br />

by differential rotati<strong>on</strong>. This mechanism, which operates in an axisymmetric<br />

system, is easily understood with reference to figure 10: if the angular velocity<br />

w(r, Q) is n<strong>on</strong>-uniform al<strong>on</strong>g a field line of the poloidal field Bp(r, Q),<br />

i.e. if Bp . Vw # 0, then the field line will be ‘cranked’ around the axis of<br />

symmetry, and a toroidal comp<strong>on</strong>ent BT(r, Q) will be generated. This process<br />

is ultimately limited by field diffusi<strong>on</strong>; detailed analysis (see, for example,<br />

Moffatt 1978) shows that, if R,, is a magnetic Reynolds number based <strong>on</strong>


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> hfugnetohydrodynumics 379<br />

the strength and scale of the differential rotati<strong>on</strong>, then BT saturates at a<br />

level 0 (Rm) I BP I *<br />

This of course is <strong>on</strong> the assumpti<strong>on</strong> that, through some other mechanism,<br />

the field Bp is itself maintained at a steady level. There is no axisymmetric<br />

mechanism that can achieve this (Cowling 1934), and it is of course here<br />

that the a-effect is required. The 8cio-dynamo incorporates both effects in the<br />

dynamo cycle :<br />

differential rotati<strong>on</strong><br />

f \<br />

a-effect<br />

J<br />

The ct-effect here operates through the acti<strong>on</strong> of helical c<strong>on</strong>vecti<strong>on</strong> (i.e.<br />

c<strong>on</strong>vecti<strong>on</strong> influenced by Coriolis forces), which acts <strong>on</strong> the toroidal field in<br />

much the same manner as described in $5.4 above. The n<strong>on</strong>-axisymmetric<br />

character of such a c<strong>on</strong>vective process is what allows the cto-dynamo to<br />

escape the strait-jacket of Cowling’s theorem.<br />

Many variants of the xo-dynamo, involving particular distributi<strong>on</strong>s of<br />

~(r, 0) and o(r, 0), have been investigated (see particularly Krause & Radler<br />

1980), and much effort has been applied over the last 25 years to incorporate<br />

the difficult back-reacti<strong>on</strong> of Lorentz forces in these models. Many difficulties<br />

still remain to be overcome, particularly in the two most prominent spheres<br />

of applicati<strong>on</strong>, the Earth and the Sun. But it seems likely that, in the ultimate<br />

theories of the origins of both geomagnetism and heliomagnetism which may<br />

be expected to evolve over the next 50 years, both a-effect and co-effect will<br />

survive as essential ingredients.<br />

6 Relaxati<strong>on</strong> to magnetostatic equilibrium<br />

We turn now to a problem that is in some respects the opposite of the dynamo<br />

problem. As we have repeatedly observed, the Lorentz force F = j A B is<br />

in general rotati<strong>on</strong>al and so drives a fluid moti<strong>on</strong>; through this mechanism,<br />

magnetic energy can be c<strong>on</strong>verted to kinetic energy of moti<strong>on</strong>, and if the<br />

fluid is viscous this energy is dissipated into heat. Thus, even if the fluid<br />

can be regarded as perfectly c<strong>on</strong>ducting (i.e. y~ = 0 or R, = x) so that<br />

Joule dissipati<strong>on</strong> is negligible, there is an alternative indirect route by which<br />

magnetic energy may dissipate, and this dissipati<strong>on</strong> mechanism will persist<br />

for so l<strong>on</strong>g as V A F # 0.


380 H. K. MofSatt<br />

Insofar as the fluid is perfectly c<strong>on</strong>ducting however, the topology of the<br />

magnetic field is c<strong>on</strong>served during this ‘relaxati<strong>on</strong>’ process; in particular, all<br />

links and knots present in the magnetic flux tubes at the initial instant t = 0<br />

are present at least for all finite times t > 0. It is physically obvious that the<br />

magnetic energy cannot decay to zero under such circumstances; the <strong>on</strong>ly<br />

alternative possibility is that the field relaxes to an equilibrium compatible<br />

with its initial topology and for which V A F = 0. In such a state<br />

F=jAB=Vp, (6.1)<br />

i.e. the Lorentz force is balanced by a pressure gradient and the fluid is at rest:<br />

the field is in magnetostatic equilibrium (see, for example, Biskamp 1993).<br />

We shall look more closely at the details of this process in the following<br />

subsecti<strong>on</strong>s; for the moment it need merely be noted that we are faced<br />

here with an intriguing class of problems of variati<strong>on</strong>al type: to minimize<br />

a positive functi<strong>on</strong>al of the field B (here the magnetic energy) subject to<br />

c<strong>on</strong>servati<strong>on</strong> of the field topology; this topological c<strong>on</strong>straint is c<strong>on</strong>veniently<br />

captured by the requirement that <strong>on</strong>ly frozen-field distorti<strong>on</strong>s of B, i.e. those<br />

governed by the frozen-field equati<strong>on</strong><br />

2B<br />

- = V A (U A B) ,<br />

2t<br />

are to be c<strong>on</strong>sidered. We know that this equati<strong>on</strong> guarantees c<strong>on</strong>servati<strong>on</strong><br />

of magnetic helicity Zlv (62.2 above); but since (6.2) tells us that the field B<br />

in effect deforms with the flow, all its topological properties (i.e. those properties<br />

that are invariant under c<strong>on</strong>tinuous deformati<strong>on</strong>) are automatically<br />

c<strong>on</strong>served under this evoluti<strong>on</strong>.<br />

We may look at this also from a Lagrangian point of view. Let the<br />

Lagrangian particle path associated with the flow ~ (x, t) be given by<br />

x + X(X, t), 2X/& = U, X(X, 0) = x . (6.3)<br />

Then the Lagrangian statement equivalent to (6.2) is<br />

where p is the density field. The deformati<strong>on</strong> tensor ?Xi/O^xj encapsulates<br />

both the rotati<strong>on</strong> and stretching of the field element of B as it is transported<br />

from x to X in time t. If the fluid is incompressible, then of course<br />

p(X) = p(x). Note that (6.3) is a family of mappings c<strong>on</strong>tinuously dependent<br />

<strong>on</strong> the parameter t, and being the identity mapping at t = 0; in the language<br />

of topology, it is an ‘isotopy’.


7 Rejecti<strong>on</strong>s <strong>on</strong> Magnetohydvodynamics 38 1<br />

6.1 The structure of magnetostatic equilibrium states<br />

Equati<strong>on</strong> (6.1) places a str<strong>on</strong>g c<strong>on</strong>straint <strong>on</strong> the structure of possible magnetostatic<br />

equilibria; it implies that<br />

j Vp = 0 and B - Vp = 0, (6.5)<br />

so that, in any regi<strong>on</strong> where Vp # 0, both the current lines (j-lines) and field<br />

lines (B-lines) must lie <strong>on</strong> surfaces p = c<strong>on</strong>st.<br />

In any regi<strong>on</strong> where Vp 0, it follows from (6.1) that j is parallel to B,<br />

i.e.<br />

j = y(x)B (6.6)<br />

for some scalar field y(x). The field B is then described as ‘force-free’ in this<br />

regi<strong>on</strong>. There is a c<strong>on</strong>siderable literature devoted to the subject of force-free<br />

fields (see particularly Marsh 1996); here we simply note that, <strong>on</strong> taking the<br />

divergence of (6.6) and using V j = V B = 0, we obtain<br />

B*Vy = 0, (6.7)<br />

so that now the B-lines lie <strong>on</strong> surfaces y = c<strong>on</strong>st. The <strong>on</strong>ly possible escape<br />

from this (topological) c<strong>on</strong>straint is when Vy 0 and y = c<strong>on</strong>st. The field B<br />

is then a ‘Beltrami’ field. We have seen <strong>on</strong>e example in (5.25) above; a more<br />

general field satisfying V A B = yB may be c<strong>on</strong>structed as a superpositi<strong>on</strong> of<br />

circularly polarized Fourier modes in the form<br />

where I$ = k/lkl and k.$(k) = 0. With such a field, the B-lines are no l<strong>on</strong>ger<br />

c<strong>on</strong>strained to lie <strong>on</strong> surfaces; the particular field (5.25) exhibits the property<br />

of chaotic wandering of B-lines (Hen<strong>on</strong> 1966; Arnold 1966; Dombre et al.<br />

1986), and it may be c<strong>on</strong>jectured that this a generic property of force-free<br />

fields of the general form (6.8).<br />

I<br />

6.2 Magnetic relaxati<strong>on</strong><br />

Suppose now that incompressible fluid is c<strong>on</strong>tained in a domain 2 with fixed<br />

rigid boundary $2. For the reas<strong>on</strong>s already indicated, let us suppose that<br />

this fluid is viscous and perfectly c<strong>on</strong>ducting. Such a combinati<strong>on</strong> of fluid<br />

properties is physically artificial; but no matter - this is a situati<strong>on</strong> where<br />

the end justifies the means! We suppose that at t = 0 the fluid is at rest<br />

and that it supports a magnetic field Bo(x) where II. Bo = 0 <strong>on</strong> 29. Clearly<br />

this c<strong>on</strong>diti<strong>on</strong> of tangency of the field at the surface 29 persists under the


382 H. K. Mofatt<br />

subsequent evoluti<strong>on</strong>. This evoluti<strong>on</strong> is governed by equati<strong>on</strong> (6.2) together<br />

with (3.3), or equivalently<br />

(E+u+VU) =-,Vp+-jAB+vV’u.<br />

1 1<br />

P<br />

The velocity U satisfies the no-slip c<strong>on</strong>diti<strong>on</strong> U = 0 <strong>on</strong> 29.<br />

From (6.2) and (6.9) we may easily c<strong>on</strong>struct an energy equati<strong>on</strong>. Let<br />

the magnetic and kinetic energies. Then this energy equati<strong>on</strong> is<br />

(6.10)<br />

(6.11)<br />

where o = V A U . As expected, the sole mechanism of dissipati<strong>on</strong> is viscosity.<br />

Now let us suppose that the topology of the initial field Bo(x) is n<strong>on</strong>trivial,<br />

and in particular, that its helicity XA~<br />

is n<strong>on</strong>-zero. This helicity is<br />

c<strong>on</strong>served under (6.2); moreover n<strong>on</strong>-zero helicity clearly implies a positive<br />

lower bound for magnetic energy (Arnold 1974):<br />

M 2 L Z11XWl > (6.12)<br />

where Lg is a c<strong>on</strong>stant with the dimensi<strong>on</strong>s of length; this c<strong>on</strong>stant is<br />

determined by the geometry of 2 and can normally be interpreted as a<br />

representative linear scale of 9.<br />

Since M + K is m<strong>on</strong>ot<strong>on</strong>ic decreasing (by (6.11)) and bounded below (by<br />

(6.12)), it must tend to a positive c<strong>on</strong>stant; hence the right-hand side of (6.11)<br />

tends to zero as t + E. Provided no singularity of cr) develops, it follows that<br />

cr) + 0 for all x E 9 as t + E; it is then not difficult to show that, under<br />

the no-slip c<strong>on</strong>diti<strong>on</strong> <strong>on</strong> $9, U -+ 0 also for all x E 9, i.e. the limit state is<br />

indeed magnetostatic. Thus, as t + x, B + BE(x), j + jE(x), where<br />

jEIP\BE =VpE. (6.13)<br />

This evoluti<strong>on</strong>ary process may be represented pictorially by a ‘trajectory’ $<br />

in the functi<strong>on</strong> space F of solenoidal vector fields of finite energy (figure 11).<br />

This trajectory is c<strong>on</strong>fined to a subspace 9 of those fields that can be<br />

obtained from Bo(x) by isotopic (frozen-field) deformati<strong>on</strong>. Such a subspace,<br />

described as ‘isomagnetic’ is indicated as a surface in figure 11, but this is a<br />

misleading simplificati<strong>on</strong>, since it is in fact infinite-dimensi<strong>on</strong>al. Nevertheless,<br />

9 may be thought of as ‘foliated’ by such isomagnetic subspaces, the fields<br />

<strong>on</strong> any such subspace being ‘topologically accessible’ <strong>on</strong>e from another by<br />

isotopic deformati<strong>on</strong>. This c<strong>on</strong>cept of an ‘isomagnetic subspace’ is closely


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong><br />

383<br />

a<br />

Figure 11. Schematic representati<strong>on</strong> of the process of magnetic relaxati<strong>on</strong>. This<br />

is represented by the trajectory $ <strong>on</strong> the ‘isomagnetic’ folium S in the functi<strong>on</strong><br />

space 9 of solenoidal fields B(x) of finite energy; the stacked sheets represent the<br />

isomagnetic foliati<strong>on</strong> of this space (Moffatt 1985~).<br />

Figure 12. Relaxati<strong>on</strong> of linked flux tubes due to Maxwell tensi<strong>on</strong> al<strong>on</strong>g the axes of<br />

the tubes; a tangential disc<strong>on</strong>tinuity of field develops when the tubes make c<strong>on</strong>tact<br />

(Moffatt 1985~).<br />

n<br />

allied to the c<strong>on</strong>cept of ‘isovorticity’ that arises in c<strong>on</strong>siderati<strong>on</strong> of the Euler<br />

equati<strong>on</strong> of ideal fluid flow (see Chapter 11).<br />

The limit field BE(x) is itself topologically accessible from the initial field<br />

Bo(x). Note however that the mapping (6.3) may develop disc<strong>on</strong>tinuities in<br />

the limit t + a, so we cannot make the str<strong>on</strong>ger asserti<strong>on</strong> that BE(x) is<br />

‘topologically equivalent’ to Bo(x). The reas<strong>on</strong> for the possible development<br />

of disc<strong>on</strong>tinuities should be apparent from c<strong>on</strong>siderati<strong>on</strong> of the situati<strong>on</strong><br />

depicted in figure 12. Here the field Bo(x) is supposed c<strong>on</strong>fined to two linked<br />

flux tubes of volumes V1, V2 and carrying axial fluxes @I, @2 respectively. The<br />

domain 2 may be taken as the space R3. The Lorentz force is equivalent<br />

to a ‘Maxwell tensi<strong>on</strong>’ in the lines of force which causes c<strong>on</strong>tracti<strong>on</strong> of both<br />

tubes (the magnetic pressure gradient across the tubes is compensated by an<br />

adjustment of fluid pressure, the fluid being assumed incompressible). Tube<br />

c<strong>on</strong>tracti<strong>on</strong> is accompanied by increase of tube cross-secti<strong>on</strong>, both V1 and


384 H. K. Moffatt<br />

Figure 13. (a) Single flux tube with axial flux @ and uniform internal twist h; the<br />

helicity is &hQ2. (b) Multiply kinked tube, believed to be the preferred c<strong>on</strong>figurati<strong>on</strong><br />

when h >> 1; the internal ‘twist’ is c<strong>on</strong>verted to ‘writhe’ which may be estimated by<br />

the number of crossings.<br />

V2 being c<strong>on</strong>served. Thus c<strong>on</strong>tact of the two tubes is inevitable in the limit<br />

t --+ E, this c<strong>on</strong>tact being achieved by a ‘squeeze-film’ mechanism in which<br />

the (viscous) fluid is squeezed from the space between the two tubes as these<br />

c<strong>on</strong>tract towards c<strong>on</strong>tiguity.<br />

Note that the fluxes @1 and @2, and also naturally the helicity<br />

X.w = 2@1@2, are c<strong>on</strong>served during the relaxati<strong>on</strong> process.<br />

6.3 Relaxati<strong>on</strong> of a single flux tube<br />

The case of a single, possibly knotted, flux tube is of special interest. This is<br />

an aspect of the general theory whose peculiar interest was anticipated by<br />

Kelvin in his seminal paper of 1869 - although there Kelvin was c<strong>on</strong>cerned<br />

with vortex tubes rather than magnetic flux tubes; the latter are in fact more<br />

amenable to detailed investigati<strong>on</strong>.<br />

C<strong>on</strong>sider first the case of an unknotted flux tube (figure 13a) carrying<br />

axial flux @. Suppose further that the field within the tube is twisted in such<br />

a way that any pair of field lines has linking number h; such a pair may<br />

be c<strong>on</strong>sidered as the boundaries of a ‘ribb<strong>on</strong>’ which is twisted through an<br />

angle 27th before being joined end-to-end. The helicity associated with the<br />

field distributi<strong>on</strong> (Berger & Field 1984) is then given by<br />

2,~ = +_hQ2, (6.14)<br />

where the + or - is chosen according to whether the twist is right-handed<br />

or left-handed. The descripti<strong>on</strong> may be generalized in an obvious way to<br />

n<strong>on</strong>-integer values of h; if h is irrati<strong>on</strong>al, then the field lines are not closed<br />

curves, but cover a family of nested toruses.<br />

If h = 0, then there is no topological c<strong>on</strong>straint to c<strong>on</strong>tracti<strong>on</strong> of the


7 Repecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 385<br />

field lines and reducti<strong>on</strong> of magnetic energy: each field line can, in principle,<br />

c<strong>on</strong>tract to a point (as t -+ x) and it therefore seems inevitable that the<br />

unique limit state that minimizes M (at M = 0) is BE(x) = 0.<br />

If h # 0, then M is c<strong>on</strong>strained by (6.12), and, <strong>on</strong> dimensi<strong>on</strong>al grounds,<br />

the minimum value of M, ME say, is given by<br />

ME = WI(~)Q’I/-~’~, (6.15)<br />

where V is the volume of the flux tube, and m(h) is a dimensi<strong>on</strong>less functi<strong>on</strong><br />

of the dimensi<strong>on</strong>less number h. The form of this functi<strong>on</strong> may be easily<br />

estimated as follows (Chui & Moffatt 1995): the axial field BT in the tube is<br />

proporti<strong>on</strong>al to the tube length L, so that the corresp<strong>on</strong>ding c<strong>on</strong>tributi<strong>on</strong> to<br />

magnetic energy MT scales like Q2L2/V. The transverse field Bp due to twist<br />

in the tube is proporti<strong>on</strong>al to hV1/2L-3/2B~, and the additi<strong>on</strong>al c<strong>on</strong>tributi<strong>on</strong><br />

Mp therefore scales like h2Q2/L. Clearly to minimize M = MT + Mp,<br />

L adjusts itself so that the two c<strong>on</strong>tributi<strong>on</strong>s are of the same order of<br />

magnitude, i.e.<br />

- h213~113 (6.16)<br />

and then M - h4/3Q2V-1/3; thus<br />

m(h) - h4I3 (6.17)<br />

where the symbol - is used to mean ‘scales like’.<br />

Note that, if h >> 1, then from (6.16), the tube cross-secti<strong>on</strong> A = V/L -<br />

hK2I3 V2I3, and is small; this results from c<strong>on</strong>tracti<strong>on</strong> of the cross-secti<strong>on</strong>al<br />

field Bp (in the above notati<strong>on</strong>). Simple estimates suggest that the tube is<br />

subject to ‘kink instabilities’ in this limit, the limit state of minimum energy<br />

being multiply kinked as indicated in figure 13(b). The associated c<strong>on</strong>versi<strong>on</strong><br />

of ‘twist’ to ‘writhe’ (the sum being precisely the c<strong>on</strong>served h), is discussed<br />

by Moffatt & Ricca (1992).<br />

C<strong>on</strong>sider now a flux tube knotted in the form of an arbitrary knot K.<br />

Again we may suppose that the field structure within the tube is <strong>on</strong>e of<br />

‘uniform twist’, in which each pair of B-lines has the same linking number<br />

h. That this number is arbitrary should be evident from the fact that,<br />

through cutting, twisting and rec<strong>on</strong>necting the tube, it may be changed<br />

by an arbitrary amount. This freedom to specify h is what knot-theorists<br />

describe as ‘framing’, and the particular choice of twist that gives h = 0 is<br />

described as ‘zero-framing’. In any event, the helicity of the framed magnetic<br />

flux tube is still given by HAu = hQ2 (since it is just the pairwise linkages<br />

that c<strong>on</strong>tribute to XAu, as for an unknotted tube).<br />

As regards the minimum-energy state of such a tube, the argument leading


386 H. K. Mofatt<br />

Figure 14. C<strong>on</strong>jectured minimum-energy c<strong>on</strong>figurati<strong>on</strong> for the trefoil knot for the<br />

same value of h (count the crossings!).<br />

to (6.15) still applies (Moffatt 1990), but now the functi<strong>on</strong> m(h) may depend<br />

<strong>on</strong> the knot K as well as <strong>on</strong> h. We therefore write it as m~(h), a real-valued<br />

functi<strong>on</strong> which is a ‘property’ of the knot K. The argument leading to (6.17)<br />

is now valid <strong>on</strong>ly for large h (since for h = O(1) there is now a topological<br />

barrier to unlimited decrease of knot length L); thus we have<br />

mK(h) - h4I3 for h >> 1. (6.18)<br />

Note also that there may be more than <strong>on</strong>e minimum-energy state: two<br />

possibilities for the trefoil knot with h = 6 are indicated in figure 14.<br />

6.4 Ideal knots<br />

There is a c<strong>on</strong>necti<strong>on</strong>, as yet not fully understood, between the foregoing<br />

descripti<strong>on</strong> of minimum-energy knotted flux tubes and the c<strong>on</strong>cept of ‘ideal<br />

knots’ (Stasiak, Katritch & Kauffman 1998). C<strong>on</strong>sider a closed curve C in<br />

the form of a knot K; we may obviously deform C c<strong>on</strong>tinuously without<br />

changing K. Let us fix the length L of C, and c<strong>on</strong>struct a tube around C<br />

of circular cross-secti<strong>on</strong> radius r centred <strong>on</strong> C. The volume of this tube<br />

is I/ = m2L. Let us now gradually increase r until either the tube makes<br />

c<strong>on</strong>tact with itself or until r equals the minimum radius of curvature <strong>on</strong><br />

C; the volume is then I/ = V,,(C) say. We may now ask: what form does<br />

the curve C take (for given K and L) so that V,(C) is maximized? In this<br />

c<strong>on</strong>figurati<strong>on</strong>, in Stasiak et al.’s terminology, the knot is ‘ideal’.<br />

Since maximizing V for fixed L is equivalent to minimizing L for fixed<br />

V, the above c<strong>on</strong>structi<strong>on</strong> of an ideal knot is closely related to the process


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 387<br />

of magnetic relaxati<strong>on</strong> to a minimum-energy state, at least when the twist<br />

parameter h is small. The c<strong>on</strong>cept of twist, or equivalently framing, plays no<br />

part in Stasiak et al.'s c<strong>on</strong>structi<strong>on</strong>. The c<strong>on</strong>necti<strong>on</strong> between the c<strong>on</strong>structi<strong>on</strong>s,<br />

<strong>on</strong>e purely geometrical in character, the other of more physical origin,<br />

is intriguing and deserves further study. There is some evidence, presented<br />

by Stasiak et al. (1996), that knotted DNA strands have a tendency to seek<br />

the ideal c<strong>on</strong>figurati<strong>on</strong> appropriate to the knot. The reas<strong>on</strong> for this sort of<br />

behaviour may well be found within a 'minimum-energy' formulati<strong>on</strong> of the<br />

problem. Much remains to be d<strong>on</strong>e in this new and challenging field.<br />

6.5 A paradox and its resoluti<strong>on</strong><br />

The argument of 5 6.2 c<strong>on</strong>tains an altogether ast<strong>on</strong>ishing implicati<strong>on</strong>: for a<br />

given magnetic field Bo(x) of arbitrary topology, there exists a magnetostatic<br />

equilibrium field BE(x) that is topologically accessible from Bo(x) - this<br />

accessibility being realized by the magnetic relaxati<strong>on</strong> process.<br />

Now the general three-dimensi<strong>on</strong>al field Bo(x) in a finite domain 9 is<br />

likely to be chaotic, i.e. the Bo-lines are generally not closed and do not<br />

lie <strong>on</strong> surfaces. This is because the three-dimensi<strong>on</strong>al n<strong>on</strong>linear dynamical<br />

system<br />

dx/dt = Bo(x) (6.19)<br />

is in general n<strong>on</strong>-integrable. An example of the resulting chaotic behaviour<br />

may be found in Bajer & Moffatt (1990), in which the Cartesian comp<strong>on</strong>ents<br />

of Bo(x) are taken to be quadratic functi<strong>on</strong>s of the coordinates (x,y,z).<br />

A field that is chaotic cannot relax to <strong>on</strong>e whose field lines lie <strong>on</strong> surfaces;<br />

hence, in general, the relaxed field BE(x) must also be chaotic. But we have<br />

seen in $6.1 that in this case<br />

V A BE = "/,jBE (6.20)<br />

with y c<strong>on</strong>stant, i.e. at least within the regi<strong>on</strong> in which BE is chaotic, it is<br />

also a Beltrami field. As first pointed out by Arnold (1974), if we c<strong>on</strong>fine<br />

attenti<strong>on</strong> to analytic fields, the family of soluti<strong>on</strong>s of (6.20) is not nearly wide<br />

enough to include all possible topologies (and all possible degrees of chaos)<br />

in the relaxed field BE; and yet we must allow for all possible topologies,<br />

since the topology of Bo(x) was arbitrary.<br />

How then is this c<strong>on</strong>tradicti<strong>on</strong> to be resolved? Partly, no doubt, through<br />

accepting that, as indicated by the example of figure 12, tangential disc<strong>on</strong>tinuities<br />

of B (i.e. current sheets) will appear quite naturally in the course<br />

of relaxati<strong>on</strong>, i.e. there is no reas<strong>on</strong> to believe that the relaxed field BE will


388 H. K. MofSatt<br />

be analytic, even if Bo is analytic. Sec<strong>on</strong>d, and more subtly, it is known<br />

that there are always ‘islands of regularity’ within any sea of chaos; during<br />

relaxati<strong>on</strong>, the boundaries of such islands can become infinitely deformed<br />

so that the regi<strong>on</strong> of chaos within which the field BE satisfies (6.20) may<br />

exhibit an arbitrarily complex geometry; the simplicity of (6.20) is gained at<br />

the cost of geometrical complexity of the regi<strong>on</strong> within which it is valid.<br />

Magnetic relaxati<strong>on</strong> is a process that can in principle be realized numerically;<br />

however, no satisfactory way has yet been found to provide accurate<br />

simulati<strong>on</strong> in three dimensi<strong>on</strong>s of the frozen-field equati<strong>on</strong> (6.2). Field relaxati<strong>on</strong><br />

in two dimensi<strong>on</strong>s has been accomplished by Linardatos (1993), who<br />

shows a clear tendency for current sheets to form due to collapse of the<br />

separatrices passing through saddle points of the field; but the problem of<br />

chaos does not arise for two-dimensi<strong>on</strong>al fields.<br />

We note that the formati<strong>on</strong> of current sheets, which may be recognized as<br />

a natural c<strong>on</strong>comitant of magnetic relaxati<strong>on</strong> to a minimum-energy state, is<br />

of absolutely central importance in solar physics: it is a prime mechanism<br />

for the explosive activity associated with solar flares and for c<strong>on</strong>sequential<br />

heating of the solar cor<strong>on</strong>a (Parker 1994). Whereas dynamo theory is central<br />

to an understanding of the internal origins of solar magnetism, magnetic<br />

relaxati<strong>on</strong> provides the key to a proper understanding of observed external<br />

magnetic activity. The two process, both fundamental, are complementary in<br />

more senses than <strong>on</strong>e (Parker 1979; see also Priest 1982).<br />

’<br />

7 C<strong>on</strong>cluding remarks<br />

In this essay, I have attempted to c<strong>on</strong>vey the flavour, rather than the detail,<br />

of three overlapping areas of magnetohydrodynamics in which I have been<br />

successively involved over the last 40 years. In 1960, the subject was relatively<br />

young and it attracted a huge cohort of researchers, active particularly in<br />

the c<strong>on</strong>texts of astrophysics and fusi<strong>on</strong> (plasma) physics. Great progress<br />

was made, notably in dynamo theory in the 1960s, the 1970s being more<br />

a period of c<strong>on</strong>solidati<strong>on</strong>. In more recent years, as in other branches of<br />

fluid mechanics, the computer has played an increasingly important role in<br />

allowing the investigati<strong>on</strong> of n<strong>on</strong>linear effects frequently bey<strong>on</strong>d the reach<br />

of analytical study. Some fascinating new areas have emerged, most notably<br />

the area treated in $6 above in which the global topology of the magnetic<br />

field plays a central role. It may be noted that there is an exact analogy<br />

between the magnetostatic equati<strong>on</strong> (6.13) and the steady Euler equati<strong>on</strong><br />

of ideal hydrodynamics, so that the magnetic relaxati<strong>on</strong> technique provides<br />

an indirect means of determining soluti<strong>on</strong>s of the steady Euler equati<strong>on</strong>s


7 <str<strong>on</strong>g>Reflecti<strong>on</strong>s</str<strong>on</strong>g> <strong>on</strong> <strong>Magnetohydrodynamics</strong> 389<br />

,<br />

of arbitrary streamline topology (Moffatt 198.5~). There is a rich interplay<br />

between MHD and ‘ordinary’ fluid dynamics that stems from this analogy,<br />

in subtle c<strong>on</strong>juncti<strong>on</strong> with the (different) analogy between (2.4) and (2.6).<br />

<strong>Magnetohydrodynamics</strong> is a richly rewarding field of study, not <strong>on</strong>ly in its<br />

own right, but also through the illuminati<strong>on</strong> of more classical areas of fluid<br />

dynamics that this interplay provides.<br />

References<br />

ALFV~N, H. 1950 Cosmical Electrodynamics. Oxford University Press.<br />

ARNOLD, V. I. 1966 Sur la topologie des kcoulements stati<strong>on</strong>naires des fluides parfaits.<br />

C. R. Acad. Sci. Paris 261, 17-20.<br />

ARNOLD, V. I. 1974 The asymptotic Hopf invariant and its applicati<strong>on</strong>s. In Proc.<br />

Summer School in Diferential Equati<strong>on</strong>s, Erevan 1974. Armenian SSR Acad. Sci.<br />

[English transl: Sel. Math. Sou. 5 (1986), 327-3451.<br />

ARNOLD, V. I. & KHESIN, B. A. 1998 Topological Methods in Hydrodynamics. Springer.<br />

BAJER, K. & MOFFATT, H. K. 1990 On a class of steady c<strong>on</strong>fined Stokes flows with<br />

chaotic streamlines. J. Fluid Mech. 212, 337-363.<br />

BATCHELOR, G. K. 1950 On the sp<strong>on</strong>taneous magnetic field in a c<strong>on</strong>ducting liquid in<br />

turbulent moti<strong>on</strong>. Proc. R. Soc. L<strong>on</strong>d. A 201, 405-416.<br />

BERGER, M. A. & FIELD, G. B. 1984 The topological properties of magnetic helicity.<br />

J. Fluid Mech. 147, 133-148.<br />

BISKAMP, D. 1993 N<strong>on</strong>linear <strong>Magnetohydrodynamics</strong>. Cambridge University Press.<br />

BOJAREVIES, V., FREIBERGS, YA., SHILOVA, E. I. & SHCHERBININ, E. V. 1989 Electrically<br />

Induced Vortical Flows. Kluwer.<br />

BRAGINSKII, S. I. 1964 Theory of the hydromagnetic dynamo. Sou. Phys. JETP 20,<br />

1462-147 1.<br />

BRAUNBEK, W. 1932 Eine neue Methode electrodenloser Leitfahigkeitsmessung.<br />

Z. Phys. 73, 312-334.<br />

CHANDRASEKHAR, S. 1961 Hydrodynamic and Hydromagnetic Stability. Clarend<strong>on</strong>.<br />

CHILDRESS, S. & GILBERT, A. D. 1995 Stretch, Twist, Fold: the Fast Dynamo. Springer.<br />

CHUI, A. Y. K. & MOFFATT, H. K. 1995 The energy and helicity of knotted magnetic<br />

flux tubes. Proc. R. Soc. L<strong>on</strong>d. A 451, 609-629.<br />

COOK, A. H. 1980 Interiors of the Planets. Cambridge University Press.<br />

COWLING, T. G. 1934 The magnetic field of sunspots. M<strong>on</strong>. Not. R. AstrON. Soc. 94,<br />

39-48.<br />

COWLING, T. G. 1957 <strong>Magnetohydrodynamics</strong>. Interscience; 2nd Edn 1975, Adam<br />

Hilger Ltd., Bristol.<br />

DOMBRE, T., FRISCH, U,, GREENE, J. M., H~NON, M., MEHR, A. & SOWARD, A. M.<br />

1986 Chaotic streamlines in the ABC flows. J. Fluid Mech. 167, 353-391.<br />

DAVIDSON, P. A., KINNEAR, D., LINGWOOD, R. J., SHORT, D. J. & HE, X. 1999 The<br />

role of Ekman pumping and the dominance of swirl in c<strong>on</strong>fined flow driven by<br />

Lorentz forces. Eur. J. Mech. BIFluids 18, 693-711.<br />

GOLITSIN, G. 1960 Fluctuati<strong>on</strong>s of magnetic field and current density in turbulent<br />

flows of weakly c<strong>on</strong>ducting fluids. Sou. Phys. Dokl. 132, 315-318.<br />

H~NON, M. 1966 Sur la topologie des lignes de courant dans un cas particulier. C. R.<br />

Acad. Sci. Paris A 262, 312-314.<br />

JACOBS, J. A. 1984 Recersals of the Earth’s Magnetic Field. Adam Hilger Ltd., Bristol.


390 H. K. Moffatt<br />

KELVIN, LORD (then W. THOMSON) 1869 On vortex moti<strong>on</strong>. Trans. R. Soc. Edin., 25,<br />

217-260.<br />

KRAUSE, F. & RADLER, K.-H. 1980 <strong>Magnetohydrodynamics</strong> and Dynamo Theory.<br />

Pergam<strong>on</strong>.<br />

LIXARDATOS, D. 1993 Determinati<strong>on</strong> of two-dimensi<strong>on</strong>al magnetostatic equilibria<br />

and analogous Euler flows. J. Fluid Mech. 246, 569-591.<br />

MARSH, G. E. 1996 Force-free Magnetic Fields: Soluti<strong>on</strong>s, Topology and Applicati<strong>on</strong>s.<br />

World Scientific.<br />

MESTEL, A. J. 1982 Magnetic levitati<strong>on</strong> of liquid metals. J. Fluid Mech. 117, 27-43.<br />

MOFFATT, H. K. 1961 The amplificati<strong>on</strong> of a weak applied magnetic field by turbulence<br />

in fluids of moderate c<strong>on</strong>ductivity. J. Fluid Mech. 11, 625-635.<br />

MOFFATT, H. K. 1965 On fluid flow induced by a rotating magnetic field. J. Fluid<br />

Mech. 22, 521-528<br />

MOFFATT, H. K. 1967 On the suppressi<strong>on</strong> of turbulence by a uniform magnetic field.<br />

J. Fluid Mech. 28, 571-592.<br />

MOFFATT, H. K. 1969 The degree of knottedness of tangled vortex lines. J. Fluid<br />

Mech. 35, 117-129.<br />

MOFFATT, H. K. 1978 Magnetic Field Generati<strong>on</strong> in Electrically C<strong>on</strong>ducting Fluids.<br />

Cambridge University Press.<br />

MOFFATT, H. K. 1985a Magnetostatic equilibria and analogous Euler flows of arbitrarily<br />

complex topology. Part 1. Fundamentals. J. Fluid Mech. 159, 359-378.<br />

MOFFATT, H. K. 19852, High frequency excitati<strong>on</strong> of liquid metal systems. In Metallurgical<br />

Applicati<strong>on</strong>s of <strong>Magnetohydrodynamics</strong> (ed. H. K. Moffatt & M. R. E.<br />

Proctor), pp. 180-189. Metals Society, L<strong>on</strong>d<strong>on</strong>.<br />

MOFFATT, H. K. & PROCTOR, M. R. E. (Eds.) 1984 Metallurgical Applicati<strong>on</strong>s of<br />

<strong>Magnetohydrodynamics</strong>. Metals Society, L<strong>on</strong>d<strong>on</strong>.<br />

MOFFATT, H. K. 1990 The energy spectrum of knots and links. Nature 347, 367-369.<br />

MOFFATT, H. K. & RICCA, R. L. 1992 Helicity and the CglugHreanu invariant. Proc.<br />

R. Soc. L<strong>on</strong>d. A 439, 411429.<br />

MOREAU, R. 1990 <strong>Magnetohydrodynamics</strong>. Kluwer.<br />

NORTHRUP, E. E 1907 Some newly observed manifestati<strong>on</strong>s of forces in the interior<br />

of an electric c<strong>on</strong>ductor. Phys. Rev. 24, 474-497.<br />

ODIER, P., PINTOX, J.-F. & FAUVE, S. 1998 Advecti<strong>on</strong> of a magnetic field by a<br />

turbulent swirling flow. Phys. Rev. E 58, 7397-7401.<br />

PARKER, E. N. 1955 Hydromagnetic dynamo models. Astrophys. J. 122, 293-3 14.<br />

PARKER, E. N. 1979 Cosmical Magnetic Fields. Clarend<strong>on</strong>.<br />

PARKER, E. N. 1994 Sp<strong>on</strong>taneous Current Sheets in Magnetic Fields. Oxford University<br />

Press.<br />

PRIEST, E. R. 1982 Solar <strong>Magnetohydrodynamics</strong>. D. Reidel.<br />

RICHARDSOK, L. F. 1922 Weather Predicti<strong>on</strong> by Numerical Process. Cambridge University<br />

Press.<br />

SCHUSTER, A. 1912 A critical examinati<strong>on</strong> of the possible causes of terrestrial<br />

magnetism. Proc. Phys. Soc. L<strong>on</strong>d. 24, 121-137.<br />

SKEYD, A. D. & MOFFATT, H. K. 1982 Fluid-dynamical aspects of the levitati<strong>on</strong><br />

melting process. J. Fluid Mech. 117, 45-70.<br />

STASIAK, A., KATRITCH, V., BEDKAR, J., MICHOUD, D. & DUBOCHET, J. 1996 Electrophoretic<br />

mobility of DNA knots. Nature 384, 122.<br />

STASIAK, A., KATRITCH, V. & KAUFFMAN, L. H. 1998 Ideal Knots. World Scientific.<br />

STEEKBECK, M., KRAUSE, F. & RADLER, K.-H. 1966 Berechnung der mittleren


7 Rejecti<strong>on</strong>s <strong>on</strong> <strong>Magnetohydrodynamics</strong> 391<br />

Lorentz-Feldstarke fur ein elektrisch leitendes Medium in turbulenter, durch<br />

Coriolis-Krafte beeinfluster Bewegung. 2. Naturforsch. 21a, 369-376.<br />

VAINSHTEIN, S. I. & ZELDOVICH, YA. B. 1972 Origin of magnetic fields in astrophysics.<br />

SOU. Phys. USP. 15, 159-172.<br />

WOLTJER, L. 1958 A theorem <strong>on</strong> force-free magnetic fields. Proc. Nut1 Acad. Sci. 44,<br />

489491.<br />

ZELDOVICH, YA. B., RUZMAIKIN, A. A. & SOKOLOFF, D. D. 1983 Magnetic Fields in<br />

Astrophysics. Gord<strong>on</strong> and Breach.<br />

Isaac Newt<strong>on</strong> Institute for Mathematical Sciences, University of Cambridge,<br />

20 Clarks<strong>on</strong> Road, Cambridge CB3 OEH, UK

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!