11
449 ISSN 1069-3513, Izvestiya, Physics of the Solid Earth, 2006, Vol. 42, No. 6, pp. 449–459. © Pleiades Publishing, Inc., 2006. Original Russian Text © M.Yu. Reshetnyak, 2006, published in Fizika Zemli, 2006, No. 6, pp. 3–13. INTRODUCTION The origination of a large-scale magnetic field in celestial bodies is related to magnetic convection pro- cesses. It is known that the magnetic field can be gener- ated by both large-scale flows [Bullard and Gellman, 1954] and small-scale flows [Vainshtein, 1983]. In the latter case, along with small-scale fluctuating magnetic fields, large-scale fields arise [Vainshtein et al., 1980] (a synergic process), which is closely connected with the concept of the α-effect in the theory of mean fields [Krause and Rädler, 1980]. Averaging over one coordi- nate, e.g., over the longitude, and using an additional generation term with α, the 3-D dynamo problem could be reduced in many cases to the 2-D problem, while preserving the validity of the limitation imposed by the Cowling theorem, which precludes the generation of an axisymmetric magnetic field by an axisymmetric flow. The origin of the α-effect consists in the fact that the mirror symmetry is violated in a convective rotating body and, as a result, the number of, e.g., clockwise rotating eddies in the northern hemisphere is systemat- ically greater than the number of anticlockwise rotating ones [Krause and Rädler, 1980]. On the contrary, anti- clockwise rotating eddies in the opposite hemisphere dominate over clockwise rotating ones. This violation of the mirror symmetry, resulting from the averaging of the Maxwell equations over turbulent pulsations, gives rise to a component of the average magnetic field B par- allel to the average electric current J = αB (whereas the magnetic field is usually perpendicular to the current). It is this generally small parallel component of the mag- netic field that is capable of closing the self-excitation circuit of the magnetic field in the Faraday phenome- non of electromagnetic induction. The violation itself of the reflection invariance in the rotating turbulence is usually associated with the Cori- olis force action on eddies moving in a medium with a zero gradient of density [Parker, 1979]. Let the density of the body decrease with its radius. Then, an ascending eddy broadens, producing a velocity component directed along the eddy radius. Accordingly, the Corio- lis force arises due to the radial velocity and general rotation. This force makes the eddy rotate in the sense opposite to the general rotation. In a similar way, a descending eddy contracts, also giving rise to a radial velocity, and the Coriolis force makes the eddy rotate in the same sense as the general rotation. This results in a nonzero correlation, (1) i.e., the average helicity of the flow (here, v is the tur- bulent velocity and the broken brackets mean averag- ing). Correlation (1) determines the value of the α-effect (the Moffat formula): (2) where τ H is the characteristic time of correlation. Note that the helicity V · curl V is interesting in itself, without regard for the α-effect, because it is the integral of motion, like the kinetic energy [Kurganskii, 1993; Frik, 2003; Lesieur, 1997]. In mechanics, the original of χ H is the squared angular momentum [Dolzhanskii, 2001]. It is remarkable that the pseudoscalar value χ H can be estimated from symmetry considerations because a pseudoscalar can be constructed from the density gra- dient and angular rotation velocity (the so-called Krause formula, see for details [Reshetnyak and Sokolov, 2003]). This traditional interpretation of the α-effect is, however, of limited applicability. First of χ H v curl v , = α τ H χ H /3, = Hydrodynamic Helicity in Boussinesq-Type Models of the Geodynamo M. Yu. Reshetnyak Schmidt Institute of Physics of the Earth, Russian Academy of Sciences, Bol’shaya Gruzinskaya ul. 10, Moscow, 123995 Russia Received October 17, 2005 Abstract—The generation of hydrodynamic helicity is considered in models of thermal convection of a rapidly rotating incompressible liquid (small Rossby numbers). It is shown that a system of cyclonic vortices arising in geodynamo models generates a large-scale hydrodynamic helicity. The spatial distribution of the helicity is ana- lyzed as a function of the intensity of heat sources in models with various boundary conditions for the velocity field. PACS numbers: 91.25.Cw DOI: 10.1134/S1069351306060012

Hydrodynamic helicity in Boussinesq-type models of the geodynamo

Embed Size (px)

Citation preview

Page 1: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

449

ISSN 1069-3513, Izvestiya, Physics of the Solid Earth, 2006, Vol. 42, No. 6, pp. 449–459. © Pleiades Publishing, Inc., 2006.Original Russian Text © M.Yu. Reshetnyak, 2006, published in Fizika Zemli, 2006, No. 6, pp. 3–13.

INTRODUCTION

The origination of a large-scale magnetic field incelestial bodies is related to magnetic convection pro-cesses. It is known that the magnetic field can be gener-ated by both large-scale flows [Bullard and Gellman,1954] and small-scale flows [Vainshtein, 1983]. In thelatter case, along with small-scale fluctuating magneticfields, large-scale fields arise [Vainshtein et al., 1980] (asynergic process), which is closely connected with theconcept of the

α

-effect in the theory of mean fields[Krause and Rädler, 1980]. Averaging over one coordi-nate, e.g., over the longitude, and using an additionalgeneration term with

α

, the 3-D dynamo problem couldbe reduced in many cases to the 2-D problem, whilepreserving the validity of the limitation imposed by theCowling theorem, which precludes the generation of anaxisymmetric magnetic field by an axisymmetric flow.

The origin of the

α

-effect consists in the fact that themirror symmetry is violated in a convective rotatingbody and, as a result, the number of, e.g., clockwiserotating eddies in the northern hemisphere is systemat-ically greater than the number of anticlockwise rotatingones [Krause and Rädler, 1980]. On the contrary, anti-clockwise rotating eddies in the opposite hemispheredominate over clockwise rotating ones. This violationof the mirror symmetry, resulting from the averaging ofthe Maxwell equations over turbulent pulsations, givesrise to a component of the average magnetic field

B

par-allel to the average electric current

J

=

α

B

(whereas themagnetic field is usually perpendicular to the current).It is this generally small parallel component of the mag-netic field that is capable of closing the self-excitationcircuit of the magnetic field in the Faraday phenome-non of electromagnetic induction.

The violation itself of the reflection invariance in therotating turbulence is usually associated with the Cori-olis force action on eddies moving in a medium with azero gradient of density [Parker, 1979]. Let the densityof the body decrease with its radius. Then, an ascendingeddy broadens, producing a velocity componentdirected along the eddy radius. Accordingly, the Corio-lis force arises due to the radial velocity and generalrotation. This force makes the eddy rotate in the senseopposite to the general rotation. In a similar way, adescending eddy contracts, also giving rise to a radialvelocity, and the Coriolis force makes the eddy rotate inthe same sense as the general rotation. This results in anonzero correlation,

(1)

i.e., the average helicity of the flow (here,

v

is the tur-bulent velocity and the broken brackets mean averag-ing). Correlation (1) determines the value of the

α

-effect (the Moffat formula):

(2)

where

τ

H

is the characteristic time of correlation. Notethat the helicity

V

·

curl

V

is interesting in itself, withoutregard for the

α

-effect, because it is the integral ofmotion, like the kinetic energy [Kurganskii, 1993; Frik,2003; Lesieur, 1997]. In mechanics, the original of

χ

H

is the squared angular momentum [Dolzhanskii, 2001].It is remarkable that the pseudoscalar value

χ

H

can beestimated from symmetry considerations because apseudoscalar can be constructed from the density gra-dient and angular rotation velocity (the so-calledKrause formula, see for details [Reshetnyak andSokolov, 2003]). This traditional interpretation of the

α

-effect is, however, of limited applicability. First of

χH v curlv⋅⟨ ⟩ ,=

α τHχH/3,–=

Hydrodynamic Helicity in Boussinesq-TypeModels of the Geodynamo

M. Yu. Reshetnyak

Schmidt Institute of Physics of the Earth, Russian Academy of Sciences,Bol’shaya Gruzinskaya ul. 10, Moscow, 123995 Russia

Received October 17, 2005

Abstract

—The generation of hydrodynamic helicity is considered in models of thermal convection of a rapidlyrotating incompressible liquid (small Rossby numbers). It is shown that a system of cyclonic vortices arising ingeodynamo models generates a large-scale hydrodynamic helicity. The spatial distribution of the helicity is ana-lyzed as a function of the intensity of heat sources in models with various boundary conditions for the velocityfield.

PACS numbers:

91.25.Cw

DOI:

10.1134/S1069351306060012

Page 2: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

450

IZVESTIYA, PHYSICS OF THE SOLID EARTH

Vol. 42

No. 6

2006

RESHETNYAK

all, the averaged flow often cannot be separated from

the turbulent pulsations

v

in applied problems.

1

Therelationship between the helicity and

α

-effect indicatedabove has such a simple form only in the case of locallyhomogeneous and isotropic turbulence. Finally, themedium in which the magnetic field is generated isoften incompressible, so that the validity of the idea ofexpanding and compressing eddies is limited. It isworth noting that, after the

α

-effect has been estab-lished, the variability of density of the medium is, as arule, neglected in dynamo models.

The limitations noted above are particularly charac-teristic of two very important cases of planetary andlaboratory dynamos. In the laboratory experiment car-ried out by Frik [2003], liquid sodium used as a work-ing medium can be considered as completely incom-pressible. The dynamo mechanism operates in theEarth’s outer core, consisting mostly of liquid iron andhaving a density changing, according to current ideas,by 15–20% along the radius [

Geomagnetism

, 1988].The situation with compressibility is similar in planetsof the solar system. Moreover, the coefficient of volumeexpansion in the Earth’s core is too small (~

2

×

10

–5

K

–1

)to generate heating-induced helicity in terms ofParker’s model.

However, even the first 3-D models of the geody-namo in the Boussinesq approximation had regimeswith nonzero large-scale helicity [Glatzmaier and Rob-erts, 1995]. Below, we consider possible mechanismsof the helicity generation in such systems and compareour results with known models of the solar dynamo, inwhich the generation mechanisms of the

α

-effect havebeen discussed for many years [Krivodubskii, 1984;

R

ü

diger

and Kitchatinov, 1993].

EQUATIONS OF THE GEODYNAMO

Formulation of the Problem

We consider the geodynamo equations for anincompressible liquid (

·

V

= 0) in a spherical layer(

r

i

r

r

0

) rotating about the

z

axis at an angular veloc-ity

Ω

in a spherical coordinate system (

r

,

θ

,

ϕ

). Weintroduce the following units of measurement for thevelocity

V

, time

t

, pressure

P

, and magnetic field

B

:

κ

/

L

,

L

2

/

κ

,

ρκ

2

/

L

2

, and

2

Ωρκµ

0

, where

L

is the length unit,

κ

is the molecular heat conductivity coefficient,

ρ

is thedensity, and

µ

0

is the magnetic constant; then, we have

1

Many geo- and astrophysical problems are characterized by con-tinuous velocity field spectra and do not make the separation intolarge and small scales used in the dynamics of average fields.

V

∂B∂t------- curl V B×( ) q 1– ∆B+=

(3)

The dimensionless numbers of Prandtl, Ekman, Ray-

leigh, and Roberts are written as Pr =

,

E

=

,

Ra = , and

q

=

, where

ν

is the kinematic vis-

cosity coefficient,

α

is the coefficient of volume expan-sion,

g

0

is the gravity acceleration,

δ

T

is the unit pertur-bation of the temperature

T

relative to the equilibriumprofile

T

0

,

and

η

is the magnetic diffusion coefficient.The Rossby number is defined as Ro =

E

Pr

–1

.

Problem (3) becomes fully defined by introducingboundary conditions at

r

=

r

i

,

r

0

. Zero boundary condi-tions are used for the temperature perturbations

T

,

and,in conjunction with the profile

T

0

specified above, thiscorresponds to fixed values of the total temperature

T

0

+

T: (1, 0). Note that variation in spherically sym-metric boundary conditions for T (e.g., fixed tempera-ture or heat flux at boundaries or various forms of heatsource distribution in the core) affects weakly the gen-erated magnetic field beginning from the onset ofdeveloped convection [Sarson et al., 1997; Reshetnyakand Steffen, 2005] but can dramatically change theregime of geomagnetic field reversals in the presence ofinhomogeneities of the heat flux at the core–mantleboundary ranging within a few tens of percent [Glatz-maier et al., 1999]. Therefore, without loss of general-ity in what follows, we accept the condition of the first

kind with fixed temperatures: T0 = at r0 = 1.

The impermeability condition Vr = 0 is accepted forthe velocity field. The tangential velocity componentsVθ and Vϕ are subjected to the no-slip conditions Vθ = 0and Vϕ = 0 at the boundaries r = ri and r0 (the rigid corein some models can rotate about the z axis) [Glatzmaierand Roberts, 1995], which is typical of rigid bound-aries; an alternative to these boundary conditions forthe tangential velocity components is the vanishing oftangential stresses at these boundaries: τrθ, τrϕ = 0[Kuang and Bloxham, 1997; Tilgner and Busse, 1997].Boundary conditions of these types are used in modelsof convection in stars with a free boundary [Chan-drasekhar, 1981] (exact expressions for in variouscoordinate systems can be found in [Landau and Lif-shitz, 1975]). Such a type of boundary conditions,somewhat strange for the Earth’s core, is related to thefollowing. It is known that, in the first case, the gener-ated magnetic field concentrates near rigid boundaries

EPr 1– ∂V∂t------- V ∇⋅( )V+

= curlB B ∇P– 1z V RaTr1r E∆V+ +×–×

∂T∂t------ V ∇⋅( ) T T0+( )+ ∆T .=

νκ--- ν

2ΩL2-------------

αg0δTL2Ωκ

------------------- ηκ---

ri/r 1–1 ri–

-----------------

τ↔

τ↔

Page 3: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

HYDRODYNAMIC HELICITY IN BOUSSINESQ-TYPE MODELS OF THE GEODYNAMO 451

(see the comparison with a nonviscous model in[Kuang and Bloxham, 1997]) and can be largely con-trolled by processes in Ekman layers, in which intenseshear flows exist and a nonzero helicity flux is presentat the rigid boundary. Given realistic values of theEkman number E ~ 10–15, the thickness profile theEkman layer is δE = E1/2L ~ 10–1 m, i.e., too small toaccount for the existing large-scale (dipole) magneticfield of the Earth. The use of the turbulent value of theviscosity coefficient does not basically change the situ-ation: in this case, ET ~ 10–9 and the layer density is

~ 100 m, which is still at the resolution threshold ofmodern computers. Another possibility is the use ofnonviscous conditions. In this case, the magnetic fieldis generated in the main volume of the liquid core. Inboth cases, the magnetic field generated in the modelshas features at the Earth’s surface similar to thoseobserved by archaeo- and paleomagnetologists [Geo-magnetism, 1988]. Below, we apply in our calculationsthe second type of boundary conditions in order toremove boundary effects. Vacuum boundary conditionsat the outer boundary r0 are used for the magnetic field(their numerical finite-difference implementation isdescribed in [Hejda and Reshetnyak, 2000]).

To solve system (3) numerically, the method of controlvolume [Patankar, 1980; Hejda and Reshetnyak, 2003,2004] was used and equations were solved in originalphysical variables on a 100 × 100 × 120 grid. PC clustersand IBM SP 690+ supercomputers were used tosolve the problem (methods of using parallel proces-sors are described in detail in [Reshetnyak and Stef-fen, 2005]).

Modeling Results

We consider the behavior of system (3) in theabsence of a magnetic field (B = 0). An increase in theRayleigh number Ra leads to convective instability in

δET

the form of vertical rolls [Roberts, 1965; Busse, 1970;Boubnov and Golitsyn, 1995] in distinction to hexahe-dra observed in an incompressible liquid (see thereview in [Getling, 1999]). The rolls are known todecrease in diameter (by the law ~E1/3) with increasingrotation velocity of a sphere Ω. The realistic values ofthe Ekman number in the Earth being E ~ 10–15, theazimuthal wave number mcr at the convection genera-tion threshold (Ra = Racr) is estimated at ~105 [Jones,2000], which is beyond the possibilities of modernnumerical analysis and requires special models of tur-bulence [Reshetnyak and Steffen, 2004]. The situationbecomes even more complicated if real values of theRayleigh number (Ra ~ 500 Racr [Jones, 2000])2 areused.

Figure 1 presents characteristic cross sections ofvelocity components demonstrating the formation ofvertical rolls at Ra = 100 and Racr ~ 13. A furtherincrease in the Rayleigh number not only decreases thescale (Fig. 2) but also gives rise to a large-scale velocitycomponent with wave numbers n < mcr well seen in thefield spectra (Fig. 3). Note that the widely accepted ideaof an (albeit weak) attenuation of the poloidal magneticfield at the surface of the liquid core ~e–0.1n [Lowes,1974] by no means implies (a) an attenuation of thespectrum of the entire magnetic field (including the tor-oidal component) throughout the liquid core volume or(b) an attenuation of the spectrum of kinetic energy (inthe Earth, we have η/ν = 106 and the kinetic energyspectrum is much more extended than the magneticenergy spectrum). In the absence of superviscosityeffects,3 a maximum in the kinetic energy spectrum is

2 The Rayleigh number in the liquid core was estimated from thebalance of the buoyancy and Coriolis forces.

3 In the model of Glatzmaier and Roberts [1995], the coefficient ofturbulent viscosity depends on the spectral number n as νT = ν(1 +0.075n3).

Fig. 1. Results of calculations with E = 10–4, Pr = 1, and Ra = 102. The panels from left to right: the meridional cross sections ofthe nonaxisymmetric velocity components Vr (–4.0, 16.9), Vθ (–28.2, 28.4), and Vϕ (–33.6, 38.6) and the equatorial cross sectionof the nonaxisymmetric velocity component Vr (–23.9, 17.3). The numbers in parentheses bound the ranges of values of the com-ponents.

Page 4: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

452

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

RESHETNYAK

well observed in several dynamo models both withsmall Ra ~ 10 Racr [Kutzner and Cristensen, 2002; Sim-itev, 2004] and with developed convection with Ra Racr [Roberts and Glatzmaier, 2000].

The broadening of the region of convection wavenumbers is related to both linear effects (the number ofmodes increases at Ra > Racr) and the backward energytransfer along the spectrum. The latter phenomenon iswell consistent with the results of linear analysis (with-out nonlinear terms), according to which the Rayleighnumbers should be much greater than the generallyaccepted estimates Ra ~ 500 Racr in order to excite con-vection in the liquid core comparable in scale to thecore itself [Reshetnyak, 2005b].

A characteristic feature of the two regimes consid-ered above is the separation of the liquid core into tworegions (for details and literature, see [Jones, 2000]):region I above (under) the rigid core bounded by theTaylor cylinder (TC) (the TC walls correspond to Stu-artson layers) and region II outside the TC (Fig. 1).Given the excitation threshold Ra = Racr and moderateRayleigh numbers (Fig. 1), convection develops inregion II and is virtually absent in region I, but a rise inthe amplitude of heat sources (Fig. 2) shifts the convec-tion intensity maximum into region II. However, theseregions are well resolved in both cases.

As noted above, helicity (1) is generated in modelswith an incompressible liquid (e.g., see [Glatzmaierand Roberts, 1995], and references to many publica-tions can be found in [Simitev, 2004]). Distributions ofthe hydrodynamic helicity χH averaged over time andthe azimuthal coordinate ϕ are presented in Fig. 5 forthe aforementioned two excitation regimes I and II.Below, we discuss the mechanisms of helicity genera-tion and compare the inferred results with data of otherstudies.

HELICITY GENERATION

Helicity Models

Presently, several mechanisms of helicity genera-tion are known. Before discussing the most widespreadmechanisms, we should mention the possible genera-tion of helicity due to reflection of Alfvén waves fromrigid boundaries [Moffat, 1978; Anufriev, 1991]. Thismodel once served as a basis for a special choice of theform of the α-effect in the geodynamo Z model [Bra-ginsky and Roberts, 1987; Cupal and Hejda, 1992]. Interms of this model, the transition from the α-effect ofthe form α = α0cosθ that is classical in the star dynamo[Zeldovich et al., 1983] to the α-effect concentrated atthe core–mantle boundary was instrumental in passingfrom the regime of traveling waves of the solar type toa stable dipole magnetic field such that, in the main vol-ume of the liquid core, the field is directed along therotation axis z, which gave the name to the Braginskymodel.

Note also the possible generation of the α-effect byshear flows, for example, in surface Ekman layer of theatmosphere [Chkhetiani, 2001]. However, it is evidentthat the description of helicity and the α-effect in termsof other, volume models is preferable for the modelingof large-scale magnetic fields.

On the whole, the ideas of developing volume mod-els of helicity are based on the construction of a pseu-doscalar value obtained from the scalar product of theordinary (G) and axial (w) vectors. The Parker model[Parker, 1979], described above and the most popular inastrophysics, relates G to a large-scale gradient of den-sity and uses the vorticity w = curlv as the second vec-tor. However, this model cannot describe all possiblesituations. Thus, in the presence of strong magneticfields, one should take into account the back effect of alarge-scale magnetic field, tending to suppress arisingsmall-scale flows. Then, in contrast to the hydrody-

Fig. 2. Results of calculations with E = 10–4, Pr = 1, and Ra = 5 × 102. The panels from left to right: the meridional cross sectionsof the velocity components Vr (–46.5, 104.6), Vθ (–111.1, 72.1), and Vϕ (–132.4, 136.8) and the equatorial cross section of thenonaxisymmetric velocity component Vr (–129.6, 128.0). The numbers in parentheses bound the ranges of values of the compo-nents.

Page 5: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

HYDRODYNAMIC HELICITY IN BOUSSINESQ-TYPE MODELS OF THE GEODYNAMO 453

namic (in the given case, gyrotropic) helicity, therearises the magnetic helicity χM, which decreases thetotal helicity χΣ = χH + χM, so that |χΣ| |χH| + |χM|. Therelated suppression mechanism is described in [Vainsh-tein and Vainshtein, 1973; Krause and Rädler, 1980;Zeldovich et al., 1983], and the evolutionary equationfor χM is presented in [Kleeorin et al., 2000]. Estimatesof the helicity suppression in the Earth’s liquid core aregiven in [Reshetnyak and Sokolov, 2003].

Although the Parker model was initially developedspecially for the solar dynamo [Parker, 1979], subse-quent observations of helioseismology (see the historyof the problem in [Rüdiger and Kitchatinov, 1997]) pro-vided additional constraints on the distribution of thehelicity χH in the convective zone of the Sun: the helic-ity should change its sign along the radius (χH < 0 forlarge r and χH > 0 in the lower part of the convectivezone), which is obviously at variance with an ascendingand expanding eddy; this required a modification ofnumerical models [Yoshimura, 1975]. Moreover, thehypothesis on a compressible liquid in the lower part ofthe convective zone is inconsistent with modern ideasof the structure of the Sun.

Another scenario of the generation of helicity andthe α effect proposed in [Krivodubskii, 1984; Rüdigerand Kitchatinov, 1993] was successfully applied instudies of the solar dynamo. According to this model,the α effect is determined by the kinetic energy gradient∇EK rather than the density gradient. Formally, this is aconsequence of the fact that a pseudoscalar can be con-

structed from ∇EK and the angular velocity vector .This mechanism is interesting if only because spatialinhomogeneity of the kinetic energy (in both experi-ments and astrophysical bodies) is observed much moreoften than spatial inhomogeneity of density. It is worthnoting that, in the case of an extremely fast rotation (Ω→ ∞), the existence of α ≠ 0 can be substantiated not-withstanding the tendency of turbulence toward a 2-Dpattern [Rüdiger, 1978].

These ideas of the generation of large-scale mag-netic fields by small-scale turbulence in an incompress-ible liquid do not contradict the well-known theoreticalgeodynamo results [Glatzmaier and Roberts, 1995].Below, we demonstrate the formation of the α-effect ina rotating body with variable kinetic energy using, as anexample, numerical models in the Boussinesq approxi-mation.

Sources of Helicity in an Incompressible Liquid

Before analyzing more complex helicity models, weconsider possible mechanisms of the χH generation bythe Coriolis force at a qualitative level [Vainshtein

Ω

et al., 1980]. In a linear regime (small |V|), at 1Ω = 1z,

the relation ~ –Ro–11z × V yields

(4)

Historically, only the second term was initially takeninto account, which is characteristic of gaseous bodieswith high density gradients and a homogeneous distri-bution of kinetic energy. The continuity conditiondiv(Vρ) = 0 yields divV = –V · G, where G = ∇ ,and the helicity contribution of the second termamounts to –Ro–1 (G · 1z). Evidently, with increasingrotation velocity, Ro → 0 and the helicity increases.This mechanism describes well the generation of theα-effect in galaxies, where |G| > 1 [Zeldovich et al.,1983].

However, even in models of the solar dynamo[Krivodubskii, 1984; Rüdiger and Kitchatinov, 1993],the change in the helicity sign in the lower part of theconvective zone could not be accounted for withoutinvoking the helicity generation due to the spatial inho-mogeneity of velocity field pulsations. In the generalcase, the expressions for the α-effect in the Sun havethe form α ~ ∇ln(v⊥ρ), where v⊥ is the velocity pulsa-

tion component perpendicular to the rotation axis.4 It is evident that the contribution of the first term can

be fairly large in the planetary dynamo, where the mag-netic field is generated in a weakly compressiblemedium. Thus, the contribution of the first term in (4)to the helicity generation can already be comparablewith that of the second term in the Earth’s core, whosecomposition is accounted for by 95% molten iron andwhose density varies on the core scale by |G| ~ 0.2 [Bra-ginsky and Roberts, 1995].

As regards the third term in (4), a more detailedanalysis of flow patterns shows that the product vzcurlzvgives the largest contribution to the helicity and,accordingly, the third term can be neglected.

Regimes of Small Ra

Convection at the excitation threshold and compar-atively small Rayleigh numbers (Ra ~ 10 Racr) wasstudied in detail by Busse (see the review in [Simitev,2004]). The convection concentrates in vertical rolls inregion II and is weak in region I. In this case, the so-called weak dynamo is realized, when time-averagedmagnetic and kinetic energies have the same order ofmagnitude.

We consider the widely used flow type independentof z (with divv = 0) [Tilgner, 1997; Jones, 2000].

4 Since the value of turbulent pulsations in the lower convectivezone of the Sun is larger compared to the overlying region and theheat exchange is convective, rather than radiative, χH changes itssign to the opposite.

∂V∂t-------

∂∂t----- V curlV⋅( ) Ro 1– ∂Ek

∂z--------- VzdivV– curlV V×[ ]z+⎝ ⎠

⎛ ⎞ .∼

ρlog

Page 6: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

454

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

RESHETNYAK

Below, we note the stage at which the z dependence offlows becomes necessary. A roll-like flow arising in thenorthern hemisphere can be written by the formula

(5)

The helicity of such a flow is always negative:

(6)

According to properties of the equatorial symmetry offlows (Fig. 1), we have in the southern hemisphere (z < 0)curlv(–z) = curlv(z) and vz(–z) = –vz(z) (for details, see[Busse, 1970]). This leads to a sign change at the equa-tor: χH(–z) = –χH(z). This flow is geostrophic. However,the balance alone between the pressure and the Coriolisforce is insufficient for actual realization of the Busserolls; the latter requires additionally the incorporationof the next terms in the expansion in powers of theEkman number [Busse, 1970], accounting for the buoy-ancy forces and the z dependence of fields. One of theconditions following from the asymptotic analysis ofsuch flows is the existence of finite gradients along thez axis, |kx| ~ |ky| |kz| ≠ 0, where ki is a wave number(e.g., see [Reshetnyak, 2005b]), i.e., a departure fromgeostrophy. With reference to a problem in a sphere,this means the influence of boundary conditions on thevelocity field (Fig. 1). This statement can easily beunderstood, considering one of the three relations usedin the linear analysis of equations, namely, the relationcorresponding to the z component of the curl of the

v v x v y v z, ,( ) 2 πx( ) πy( ),cossin(= =

2 πx( ) πy( )sincos , 2 πx( ) πy( ) ).sinsin––

χH 2 2π πx( )2 πy( )2cossin–=

– 2 2π πx( )2 πy( )2 4 2π πx( )2 πy( )2 .sinsin–sincos

Navier–Stokes equations: Ro curlz = vz +

E∆curlzv. The system is split at kz = 0, and the z compo-nent of the vorticity attenuates (the condition of col-linearity of the vectors 1z and 1W is inessential). Thepresence of finite gradients in z on the main scale L isnecessary for the origination of rolls. The departurefrom geostrophism leads to nonvanishing gradients ofkinetic energy, which corresponds to the first term in(4). The latter is a necessary (but not sufficient) condi-tion for the generation of χH. The generation of large-scale helicity is critically dependent on the relationshipbetween the phases of vz and wz, as is evident even fromthe fact that, making the replacement vz = –vz in (5), weobtain χ → –χ. Given a layer infinite in z in the k-space,we have at the excitation threshold for the neutral mode

[Reshetnyak, 2005b]: hc = wz = |vz|2.

Then, we obtain χ = hc = , where ω =

–i v/v is the imaginary part of the growth rate. Since

the analysis of variance implies that two solutions with±k exist for a certain value of ω (waves traveling inopposite directions), both positive (kz > 0) and negative(kz < 0) helicity can be generated. The total helicity van-ishes, as is the case with the Alfvén waves in an infinitelayer [Moffat, 1978].

The generation of nonzero helicity χH can be relatedto two effects [Busse, 1975, 1976]. These are theEkman layers and the inclination of the layer boundary,changing the height of rolls. In the first case, a nonzero

v∂∂z-----

v z*ikz

E Pr 1– σ k2+( )---------------------------------

2Prωkz

E ω2 k 4Pr2+( )-------------------------------------

∂∂t-----

100

3

n

104

106

102

100

10–2

10–4

2

1

102101 100

3

n

108

106

104

102

100

2

1

102101

Fig. 3. Fourier spectra of the ϕ-components of the velocity in the azimuthal direction for two regimes (Fig. 1): (1) Vr; (2) Vθ; (3) Vϕ.

Page 7: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

HYDRODYNAMIC HELICITY IN BOUSSINESQ-TYPE MODELS OF THE GEODYNAMO 455

normal component of the velocity ~E1/2 arises at theboundary of the Ekman layer, generating helicity(~E1/2). In the northern hemisphere, the helicity is pos-itive at the outer boundary of the core and negative atthe inner core boundary. The spatial scale of the helicityin the direction normal to the boundary is comparableto the thickness of the Ekman layer δE ~ E1/2. This caseis of no interest for the geodynamo theory.

The second case is more interesting. It was shownthat, using the decomposition in the height of rollsdepending on the distance from the rotation axis, anincrease or decrease in the roll height changes the helic-ity sign. The effect is associated with Rossby wavesshifting the rolls in the radial direction, and this move-ment gives rise to their additional twisting (for moredetails, see [Busse, 2002]). This effect was usually con-sidered in the outer part (II) of the core, where the rollheight increases with decreasing distance from the saxis. We think that precisely this phenomenon isresponsible for the generation of the helicity χ inferredin [Kageyama and Sato, 1997] and from our calcula-tions at small Rayleigh numbers (Fig. 4). The effect ofboundaries leads to a negative correlation between vz

and wz. Unlike atmospheric cyclones and anticyclonescharacterized by a positive correlation, the effect ofboundaries yields an opposite helicity sign for region II.It is also worth noting that a decrease in the Prandtlnumber Pr → 0, shifting rolls into the equatorial zone(large s) [Zhang, 1993], can lead to a similar shift of χH

into the equator area. However, this effect is of no sig-nificance for the Earth’s core because, in models ofboth thermal convection (Pr ~ 0.1) and concentrationconvection (Pr ~ 10), the resulting displacement issmall, particularly, at large Reynolds numbers Re (Re =

~ 109 in the Earth). The values of the turbulent

Prandtl number PrT are ~O(1) in both cases.

Moffat Estimates

Before analyzing the regime of vigorous convec-tion, we consider some interesting consequences of theMoffat estimates with reference to the small-Ra con-vection model discussed above. Throughout this sec-tion, we assume that the distribution of the Busse rollsis homogeneous in the plane (x, y), perpendicular to therotation axis z, and the fields do not vary along thisaxis.5

We decompose the velocity field into the longitudi-nal component along the z axis and a transverse compo-nent in the plane (x, y):

(7)

5 The statistical homogeneity of a flow in the plane (x, y) corre-sponds to rather large values of Ra, so that rolls fill a spatialregion much greater in size than their diameter.

VLν

-------

V V|| V⊥, V||+ 1zVz.= =

Then, due to the homogeneity of the roll distribution inthe plane (x, y), the expression for the mean helicitytakes the form [Moffat, 1978]

(8)H V curlV⋅⟨ ⟩ V⊥ curlV⊥⋅⟨ ⟩= =

+ 2 V|| curlV⊥⋅⟨ ⟩ .

I

II II

I

Ω

z

Fig. 4. Meridional cross section of the axisymmetric com-ponent of the helicity χH for the Rayleigh numbers Ra = 102

(–6.9 × 105, 6.9 × 105) in the left panel and Ra = 5 × 102

(–3.1 × 107, 3.0 × 107) in the right panel.

Fig. 5. The sense of the meridional circulation is shown inthe upper part of the figure and the angular velocity of theliquid rotation Vϕ/s is shown in the lower part.

Page 8: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

456

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

RESHETNYAK

The second term (Vzwz ≡ V||curlV⊥) is already consid-ered above. The first term in (8) can be estimated fromthe balance of the buoyancy Coriolis forces in (3):

(9)

After applying the curl operation to (9), multiplying theresult by 1z × V, and taking into consideration (7) (fordetails, see [Moffat, 1978]), we obtain an estimate forthe first term in (8):

(10)

As before, we see that the dependence on z is necessaryfor the helicity generation. Note also that the limitingtransition from the periodic flow considered above to arandom flow and Moffat estimates (10) can be madeonly accurate to coefficients of the order of unity (Zel-dovich [1984] describes in more detail a similar case ofestimating the coefficients of turbulent diffusion forrandom and periodic flows).

Regimes of Large Ra

At large Ra, the roll mechanism of heat transfer inregion II becomes insufficient and convection movesinto region I, where flows driven by the buoyancyforces are parallel to the rotation axis and the Coriolisforce hinders to a lesser extent the motion and the heattransport from the hotter rigid core to the outer bound-ary. Overall, the mean temperature in region I is higher(T > 0) than in region II; however, along with the meanascending flow, the roll structure is also present in thisregion. Accurate estimates of the axisymmetric andasymmetric components can be found in [Glatzmaierand Roberts, 1995], but their ratio can depend stronglyon the superviscosity model in use [Zhang and Jones,1997].

Modeling at large Ra ~ 103 Racr being complicatedby the development of numerical instabilities, earlierstudies of the 3-D dynamo included artificial suppres-sion of high frequencies with the use of a superviscosity[Glatzmaier and Roberts, 1995; Kuang and Bloxham,1997]. In the case of a vanishing relative velocity at theboundary of the liquid core, a change in the helicitysign in region I is observed with increasing r: in thenorthern hemisphere, the helicity sign is positive in thelower part and negative in the upper part. This is thebasic distinction from the small-Ra case consideredabove, in which the helicity χH does not change its signthroughout the hemisphere. Maximums of |χH| are closeto rigid boundaries. It is also known that, in the modelof Glatzmaier and Roberts [1995], the axisymmetriccomponent accounts for 80% of the total velocity fieldand the angular velocity changes its sign in the TC:Vϕ(z)/s > 0 (s = rsinθ) in the lower part and <0 in theupper part (Fig. 5).

Now, we compare results of these calculations withthe vertical variation in the kinetic energy. In a plane

1z V× ∇p– RaT1z.+=

V⊥ curlV⊥⋅⟨ ⟩ Ra– T∂Vz

∂z-------- .=

layer (1z ≡ 1W), the application of the curl operation to(9) and the scalar multiplication by V yield

(11)

Assuming the axial symmetry and writing out the right-hand side of (11) in the spherical system of coordinates(s, ϕ, z), we obtain (∇sT, 0, ∇zT) · (–Vϕ, Vs, 0) = –Vϕ∇sT,where ∇sT is the so-called Archimedes wind. Estimate(11) does not contradict the numerical results reportedin [Glatzmaier and Roberts, 1995] because the temper-ature in region I rises toward the axis, ∇sT < 0, and theCoriolis force hinders to a lesser extent the motion andthe heat transfer from the hotter rigid core toward the

outer core boundary and the sign change in Ek is

determined by the inversion of the azimuthal velocityVϕ(z) (see Fig. 5; in the case of a fast rotation, the azi-muthal velocity component usually has a maximumenergy). As noted above, a kinetic energy varying alongthe z coordinate is necessary for the generation of helic-ity. The following scenario is possible: if region I is hot-ter than region II, the average of the component of Vz ispositive (in the northern hemisphere). This correspondsto the flow of the liquid into the lower part of the TC(and away from its upper part) through the Ekman layerand the TC walls. The latter implies the introduction ofa positive (negative) torque and, as a result, a positive(negative) torsion in the lower (upper) part of the TC, asis observed in Fig. 5. However, two problems arise inthis case. These are the questions of why a positivehelicity is not observed near the equator at small Ra andwhy, for the Ekman numbers considered here, the helic-ity has a scale much smaller than the observed scale.

To narrow the range of possible situations, weaddress thermal convection with large Rayleigh num-bers (Ra = 500) and nonviscous boundary conditions.As noted above, an increase in Ra shifts convection intothe TC (Figs. 1, 2). Although the use of nonviscousboundary conditions resulted in a sign-constant angularvelocity of the TC liquid, which was observed in[Glatzmaier and Roberts, 1995] but was not observed in[Kuang and Bloxham, 1997], the helicity distributionremained qualitatively the same (Fig. 4): the TC valueof χH changes its sign in the vertical direction. Thenumerical experiments discussed above demonstratethat the elimination of the axisymmetric component ofthe velocity V changes insignificantly (by ~15%) the ϕ-averaged helicity χH. The termwise analysis of theexpression for χH

(12)

∂∂z-----Ek Ra∇T 1z V⋅( ).⋅=

∂∂z-----

χH Vr

r θsin------------- ∂

∂θ------ Vϕ θsin( )

∂Vθ

∂ϕ---------–⎝ ⎠

⎛ ⎞=

+Vθ

r------ 1

θsin-----------

∂Vr

∂ϕ-------- ∂

∂r----- rVϕ( )–⎝ ⎠

⎛ ⎞ Vϕ

r------ ∂

∂r----- rVθ( )

∂Vr

∂θ--------–⎝ ⎠

⎛ ⎞+

Page 9: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

HYDRODYNAMIC HELICITY IN BOUSSINESQ-TYPE MODELS OF THE GEODYNAMO 457

in this model shows that the first of the three terms in(12) makes the greatest contribution to the formation oftwo χH structures of opposite signs in the TC; i.e., as inthe cases of small Ra considered above, we deal with anonzero correlation of small-scale cyclonic flowsresulting in a large-scale effect but with a smaller scalealong the z axis: χH changes its sign. As seen from Fig. 2,eddies in the lower and upper regions of the northernhemisphere rotate in opposite senses, but the directionof the vertical velocity component remains the same allalong one roll.

Since the calculations were performed for nonvis-cous boundary conditions, we think that the positivehelicity arises at the boundary of the rigid core due tothe variation in the roll height associated with theboundary curvature [Busse, 1975]: with an increase inthe distance from the z axis caused by the Rossbywaves, the roll height increases, which corresponds tothe positive sign of the helicity near the rigid core. Thespatial scale of the distributions of positive and negativehelicities is ~(r0 – ri)/2 (Fig. 4).

DISCUSSION

Apparently, direct numerical modeling fails to pro-vide regimes with E ~ 10–15 and Ra ~ 500E–1/3 = 5 × 107

and one can only hope that the model results alreadybelong to the asymptotic regime E 0, Ra/Racr =const. However, the fact that the effect of the helicityformation described above depends only on the tilt ofthe boundary and is independent of E, makes the givenmechanism attractive for modeling the helicity at real-istic values of parameters.

The helicity generation mechanisms consideredabove can exist in the Earth’s liquid core along with thetraditional mechanism proposed by Parker [1979] (seealso [Shimizu and Loper, 2000]). The concurrent oper-ation of both mechanisms leads to an increase in thehelicity near the outer boundary and they weaken eachother near the rigid core; i.e., they operate in anantiphase mode. Since the helicity in both mechanismsis antisymmetric with respect to the equatorial plane,the total helicity also possesses this property.

It is also worth noting that the form of the α-effect,concentrated near the core–mantle boundary in the Bra-ginsky Z model and observed in our study in the TC,can be interpreted in terms of not only the reflection ofAlfvén waves from a rigid boundary but also purelyhydrodynamic processes near rigid boundaries. How-ever, in this case, the model should be complementedby helicity of an opposite sign near the boundary of therigid core (note that initially the Z model was devoid ofa rigid core). The ideas of sources of α- and Ω-effectsconcentrated in different spatial regions were alsoapplied in studies of the solar dynamo [Stix, 1971].

As regards the formal similarity between theinferred dependences and solar dynamo models, weshould note that the existence of the kinetic energy gra-

dient alone is insufficient for the realization of large-scale helicity. This condition is necessary for the forma-tion of the roll flow structure, but the correlation of vor-ticity with the vertical velocity component requiresadditional conditions, in particular, an inclination of thevolume boundaries.

Above, we did not touch on the problem of calculat-ing the α-effect (2). Practical use of the above helicityestimates in large-scale dynamo models (such as thedynamo of mean fields [Krause and Rädler, 1980])requires information on the spectra of the fields V and B.Using the simplest estimate for the mixing length, wehave τH = l/v, i.e., α ~ –v/3. Assuming the spectrum ofthe velocity field to be a white noise, vn = const = Vwd,we obtain that the magnetic Reynolds number rm =v/(nη) on a scale of 1/n is ~1 even for a scale of 10–3L,so that the turbulence-driven generation of the mag-netic field is ineffective. The latter imposes a constrainton the maximum number of rolls: nmax ~ 103 [Jones,2000]. In the case of turbulence involving a cascade ofrolls of different diameters, this constraint agrees withthe estimate of the energy-carrying scale for the termcurl(αB) in the induction equation. The resulting ampli-tude of the α-effect (~Vwd/3) and the related value of the

dynamo number Rα = used in the αω-dynamo

models are comparable to estimates derived in [Anu-friev et al., 1997]. The use of models with a decreasingspectrum (vn ~ n–γ), i.e., the Kolmogorov turbulencewith γ = 1/3, increases the energy-carrying scale(decreases nmax).

ACKNOWLEDGMENTS

I am grateful to F.H. Busse and D.D. Sokolov forhelp in my work. This work was supported by the Rus-sian Foundation for Basic Research, project no. N03-05-64074, and INTAS, grant no. N03-51-5807.

REFERENCES

1. A. P. Anufriev, “An α-Effect on the Core Mantle Bound-ary,” Geophys. Astrophys. Fluid Dynamics 57, 135–146(1991).

2. A. P. Anufriev, M. Yu. Reshetnyak, and D. D. Sokolov,“Estimation of the Dynamo Number in the Turbulent α-Effect for the Earth’s Liquid Core,” Geomagn. Aeron.37, 141–146 (1997).

3. J. M. Aurnou, D. Brito, and P. L. Olson, “Mechanics ofInner Core Super-Rotation,” Phys. Earth Planet. Inter.23, 3401–3407 (1996).

4. B. M. Boubnov and G. S. Golitsyn, Convection in Rotat-ing Fluids (Kluwer, London, 1995).

5. S. I. Braginsky and P. H. Roberts, “Model Z Geody-namo,” Geophys. Astrophys. Fluid Dynamics 38, 327–349 (1987).

αLη

-------

Page 10: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

458

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

RESHETNYAK

6. S. I. Braginsky and P. H. Roberts, “Equations GoverningConvection in Earth’s Core and the Geodynamo,” Geo-phys. Astrophys. Fluid Dynamics 79, 1–97 (1995).

7. E. C. Bullard and H. Gellman, “Homogeneous Dynamosand Terrestrial Magnetism,” Phil. Trans. R. Soc. LondonA247, 213–278 (1954).

8. F. H. Busse, “Thermal Instabilities in Rapidly RotatingSystems,” J. Fluid Mech. 44, 441–460 (1970).

9. F. H. Busse, “A Model of the Geodynamo,” Geophys. J.R. Astron. Soc. 42, 437–459 (1975).

10. F. H. Busse, “A Model of the Geodynamo,” Phys. EarthPlanet. Inter. 12 (4), 350–358 (1976).

11. F. H. Busse, “Convective Flows in Rapidly RotatingSpheres and Their Dynamo Action,” Phys. Fluids 14 (4),1301–1314 (2002).

12. S. Chandrasekhar, Hydrodynamics and HydromagneticStability (Dover, New York, 1981).

13. O. G. Chkhetiani, “On the Helical Structure of theEkman Boundary Layer,” Izv. Akad. Nauk, Fiz. Atmosf.Okeana 37 (5), 614–620 (2001).

14. I. Cupal and P. Hejda, “Magnetic Field and α-Effect inModel Z,” Geophys. Astrophys. Fluid Dynamics 67, 87–97 (1992).

15. F. V. Dolzhanskii, “On Mechanical Originals of Funda-mental Hydrodynamic Invariants,” Izv. Akad. Nauk, Fiz.Atmosf. Okeana 37 (4), 446–458 (2001).

16. P. G. Frik, Turbulence: Approaches and Models (Inst.Komp’yuternykh Issledovanii, Izhevsk, 2003) [in Rus-sian].

17. Geomagnetism, Ed. by J. A. Jacobs (Academic, NewYork, 1988).

18. A. V. Getling, Rayleigh–Benard Convection: Structuresand Dynamics (Editorial URSS, Moscow, 1999) [in Rus-sian].

19. G. A. Glatzmaier and P. H. Roberts, “A Three-Dimen-sional Convective Dynamo Solution with Rotating andFinitely Conducting Inner Core and Mantle,” Phys.Earth Planet. Inter. 91, 63–75 (1995).

20. G. A. Glatzmaier, R. S. Coe, L. Hongre, and P. Roberts,“The Role of the Earth’s Mantle in Controlling the Fre-quency of Geomagnetic Reversals,” Nature 401, 885–890 (1999).

21. P. Hejda and M. Reshetnyak, “The Grid-SpectralApproach to 3-D Geodynamo Modeling,” Computers &Geosciences 26, 167–175 (2000).

22. P. Hejda and M. Reshetnyak, “Control Volume Methodfor the Dynamo Problem in the Sphere with the FreeRotating Inner Core,” Studia Geophys. Geodet. 47, 147–159 (2003).

23. P. Hejda and M. Reshetnyak, “Control Volume Methodfor the Thermal Convection Problem in a RotatingSpherical Shell: Test on the Benchmark Solution,” Stu-dia Geophys. Geodet. 48, 741–746 (2004).

24. C. A. Jones, “Convection-Driven Geodynamo Models,”Phil. Trans. R. Soc. London A358, 873–897 (2000).

25. A. Kageyama and T. Sato, “Velocity and Magnetic FieldStructures in a Magnetohydrodynamic Dynamo,” Phys.Plasmas 4 (5), 1569–1575 (1997).

26. N. Kleeorin, D. Moss, I. Rogachevskii, and D. Sokoloff,“Helicity Balance and Steady-State Strength of the

Dynamo Generated Galactic Magnetic Field,” Astron.Astrophys. 361, L5–L8 (2000).

27. F. Krause and K.-H. Rädler, Mean Field Magnetohydro-dynamics and Dynamo Theory (Akademie-Verlag, Ber-lin, 1980).

28. V. I. Krivodubskii, “Source Intensity of Magnetic Fieldsof the Solar αω-Dynamo,” Astron. Zh. 61 (3), 540–548(1984).

29. W. Kuang and J. Bloxham, “An Earth-Like NumericalDynamo Model,” Nature 389, 371–374 (1997).

30. W. Kuang and J. Bloxham, “Numerical Modeling ofMagnetohydrodynamic Convection in a Rapidly Rotat-ing Spherical Shell: Weak and Strong Field DynamoAction,” J. Comput. Phys. 153, 51–81 (1999).

31. M. V. Kurganskii, Introduction to Large Scale Dynamicsof the Atmosphere: Adiabatic Invariants and TheirApplications (Gidrometeoizdat, St. Petersburg, 1993)[in Russian].

32. C. Kutzner and U. R. Cristensen, “From Stable Dipolartowards Reversing Numerical Dynamos,” Phys. EarthPlanet. Inter. 131, 29–45 (2002).

33. L. D. Landau and E. M. Lifshitz, Course of TheoreticalPhysics, Vol. 2: The Classical Theory of Fields (Nauka,Moscow, 1988; Pergamon, Oxford, 1975).

34. M. Lesieur, Turbulence in Fluids (Kluwer, London,1997).

35. F. J. Lowes, “Spatial Power Spectrum of the Main Geo-magnetic Field,” Geophys. J. R. Astron. Soc. 36, 717–725 (1974).

36. H. K. Moffat, Magnetic Field Generation in ElectricallyConducting Fluids (Cambridge Univ., Cambridge,1978).

37. E. N. Parker, Cosmical Magnetic Fields: Their Originand Their Activity (Clarendon Press, Oxford, 1979; Mir,Moscow, 1982).

38. S. V. Patankar, Numerical Heat Transfer a Fluid Flow(Taylor & Francis, New York, 1980).

39. M. Yu. Reshetnyak, “Estimation of the Turbulent Viscos-ity in the Liquid Core of the Earth,” Dokl. Akad. Nauk400 (1), 105–109 (2005a).

40. M. Yu. Reshetnyak, “Dynamo Catastrophe, or Why theGeomagnetic Field is So Long-Lived,” Geomagn.Aeron. 45 (4), 571–575 (2005b).

41. M. Yu. Reshetnyak and D. D. Sokolov, “GeomagneticField Intensity and Suppression of Helicity in the Geo-dynamo,” Fiz. Zemli, No. 9, 82–86 (2003) [Izvestiya,Phys. Solid Earth 39, 774–777 (2003).

42. M. Reshetnyak and B. Steffen, “The Subgrid Problem ofThermal Convection in the Earth’s Liquid Core,”Numerical Methods and Programming 5, 41–45 (2004).http://www.srcc.msu.su/num-meth/english/index.html.

43. M. Reshetnyak and B. Steffen, “Dynamo Model in theSpherical Shell,” Numerical Methods and Programming6, 27–32 (2005). http://www.srcc.msu.su/num-meth/english/index.html.

44. P. H. Roberts, “On the Thermal Instability of a HighlyRotating Fluid Sphere,” Astrophys. J. 141, 240–250(1965).

45. P. H. Roberts and G. A. Glatzmaier, “A Test of the Fro-zen-Flux Approximation Using a New Geodynamo

Page 11: Hydrodynamic helicity in Boussinesq-type models of the geodynamo

IZVESTIYA, PHYSICS OF THE SOLID EARTH Vol. 42 No. 6 2006

HYDRODYNAMIC HELICITY IN BOUSSINESQ-TYPE MODELS OF THE GEODYNAMO 459

Model,” Phil. Trans. R. Soc. London 358, 1109–1121(2000).

46. R. Rüdiger, “On the α-Effect for Slow and Fast Rota-tion,” Astron. Nachr. 299 (4), 217–222 (1978).

47. R. Rüdiger and L. L. Kitchatinov, “α-Effect and α-Quenching,” Astron. Astrophys. 269, 581–588 (1993).

48. R. Rüdiger and L. L. Kitchatinov, “The Slender SolarTachocline: A Magnetic Model,” Astron. Nachr. 318,273–279 (1997).

49. G. R. Sarson, C. A. Jones, and A. W. Longbottom, “TheInfluence of the Boundary Region on the Geodynamo,”Phys. Earth Planet. Inter. 101, 13–32 (1997).

50. H. Shimizu and D. E. Loper, “Small-Scale Helicity andα-Effect in the Earth’s Core,” Phys. Earth Planet. Inter.121, 139–155 (2000).

51. R. Simitev, Convection and Magnetic Field Generationin Rotating Spherical Fluid Shells, Ph. D., Bayreuth:Univ. Bayreuth, 2004, p. 193. http://www.phy.uni-bayreuth.de/theo/tp4/members/simitev.html.

52. M. Stix, “A Nonaxisymmetric α-Effect Dynamo,”Astron. Astrophys. 13 203–208 (1971).

53. A. Tilgner, “A Kinematic Dynamo with a Small ScaleVelocity Field,” Phys. Lett. A226, 75–79 (1997).

54. A. Tilgner and F. H. Busse, “Finite Amplitude in Rotat-ing Spherical Fluid Shells,” J. Fluid Mech. 332, 359–376(1997).

55. S. I. Vainshtein, Magnetic Fields in Cosmos (Nauka,Moscow, 1983) [in Russian].

56. L. L. Vainshtein and S. I. Vainshtein, “On the Possibilityof a Linear Turbulent Dynamo,” Geomagn. Aeron. 13(1), 149–153 (1973).

57. S. I. Vainshtein, Ya. B. Zeldovich, and A. A. Ruzmaikin,Turbulent Dynamo in Astrophysics (Nauka, Moscow,1980) [in Russian].

58. H. Yoshimura, “A Model of the Solar Cycle Driven bythe Dynamo Action of the Global Convection in theSolar Convection Zone,” Astrophys. J. Suppl. 29, 467–494 (1975).

59. Ya. B. Zeldovich, Selected Works: Chemical Physics andFluid Dynamics (Nauka, Moscow, 1984) [in Russian].

60. Ya. B. Zeldovich, A. A. Ruzmaikin, and D. D. Sokoloff,Magnetic Fields in Astrophysics (Gordon and Breach,New York, 1983).

61. K. Zhang, “On Equatorially Trapped Boundary InertialWaves,” J. Fluid Mech. 248, 203–217 (1993).

62. K. Zhang and C. A. Jones, “The Effect of Hyperviscosityon Geodynamo Models,” Geophys. Res. Lett. 24, 2869–2872 (1997).