134 Physics of the Earth and Planetary Interiors, 20 (1979) 134-151 © ElsevierScientificPublishingCompany, Amsterdam - Printed in The Netherlands CONVECTION DRIVEN DYNAMOS A.M. SOWARD School o f Mathematics, The University, Newcastle upon Tyne, NE1 7R U (Great Britain) (Accepted for publication in revised form April 17, 1979) Soward, A.M., 1979. Convection driven dynamos. Phys. Earth Planet. Inter., 20: 134-151. Various models of thermal convection in rapidly rotating fluids permeated by strong magnetic fields are discussed. Particular attention is paid to the possibility that the magnetic field can be maintained by dynamo action rather than by externally applied electric currents. Two dynamo models are givenparticular attention. They are the plane layer model of Childressand Soward (1972) and the annulus model of Busse (1975). Though these models do not totally resolvethe geodynamo problem, they do highlight important features of hydromagnetic dynamos. As a result some speculations are made about the true character of the geodynamo. 1. Introduction The fluid motions responsible for driving the geodynamo are generally believed to be the result of convection, thermal or otherwise. The only serious alternative is precession. Though a large amount of energy can be transmitted into the fluid by forces acting at the core surface, the recent investigations of Loper (1975) and Rochester et al. (1975) suggest that most of the available precessional energy is dissipated in boundary layers leaving an inadequate supply to drive the dynamo. On the other hand, convection maintained directly by buoyancy forces leads to motions in the main body of the core, which are likely to be sufficiently complex to regenerate magnetic field. Furthermore the wastage of energy in boundary layers is likely to be less severe. Out of all possible convective processes thermal convection is generally adopted in theoretical dynamo models (e.g. Busse, 1973, 1975; Childress and Soward, 1972) owing to its relative simplicity. It is clear that progress towards understanding the geophysical problem can only be made through investigations of highly idealised and simplified models. Here such a model is proposed which, it is hoped, contains all the key ingredients necessary to describe qualitative features of the convection and dynamo process accurately. We shall begin by considering a self-gravitating, electrically conducting Boussinesq fluid, which is confined within a spherical container radius L and suppose that the entire system rotates rigidly with angular velocity t2. Of course, the model would be more geophysically realistic if it included a central rigid core, but, since this refinement does not appear to be of fundamental importance in the ensuing dynamics, it will not be considered. We also suppose that heat sources are distributed uniformly throughout the core and as a result the temperature and corresponding density distribution are spherically symmetric. Despite the fact that the fluid is top heavy an equilibrium state of no relative motion is possible. Nevertheless once the adverse density gradient/3 attains a critical value ~3e(say), the system becomes unstable to small perturbations. For yet larger values of/3, finite amplitude convection ensues. The hope is that, for sufficiently large values of/3, the convection can support a magnetic field by magnetic induction and so operate as a hydromagnetic dynamo. Relative to coordinate axes rotating with angular velocity f~, the fluid velocity u is governed by the equation of motion Du --+ 2f~ X u = -Vp - tv0g+ o-~j X B + vV2u Dt (V "u = O) (1.1a,b) where p and p are the fluid density and pressure, B is the magnetic field, j =/a-~ V X B (1.1c) is the electric current, gt is the magnetic permeability, a is the coefficient of expansion, f I1" dx + 0 is the temperature, x and t denote position and time, D/Dt is the material derivative and v is the viscosity. The magnetic field B and perturbation temperature 0 are governed by 0B - V X (u X 13)+ XV2B 0t (V. B = 0) (1.1d,e) and DO = - u . P + r V 2 0 Dt (1.1f) respectively, where Xand r are the magnetic and thermal diffusivities, respectively. In addition the gravitational acceleration g and the temperature gradient II are g = -gox/L, II = - ( 3 o x l L (lAg,h) where go and/3o are their respective magnitudes at the outer boundary. The mathematical problem is completed upon specification of suitable boundary conditions. Here it is important that the condition on the magnetic field should rule out the possibility that it is maintained externally by an applied electric current. The spherical hydromagnetic dynamo problem described above is extremely complex and only facets of the problem have been tackled so far. In this paper we will outline the two main approaches to the problem. The first approach is simply to consider convection in the presence of a prescribed magnetic field. Much of the early work on this magnetohydrodynamic convection is described by Chandrasekhar (1961). Nevertheless in Section 2 we follow Braginsky's (1964a) development as it highlights both the parameters and the physical processes which are central to the understanding of the geophysical problem. The second approach is to suppose that the convection is 135 given and to see if there is any possibility of magnetic field regeneration. This, of course, is the kinematic dynamo problem which is now well understood (e.g. see Moffatt, 1978). When the two approaches are considered simultaneously they yield the hydromagnetic dynamo problem. In Section 3 we shall describe the plane layer hydromagnetic dynamo developed by Childress and Seward (1972). This model has had limited success and, even when the magnetic field is very weak, it suffers from the phenomenon of dynamo instability, which occurs whenever the Lorentz force aids rather than hinders convection. In Section 4 the nature of convection in a sphere is discussed both with and without an applied magnetic field. The annulus configuration developed by Busse (1970) provides a very simple but effective model for reproducing many of the important features of convection in a sphere. Busse (1975) used it as the basis of his geodynamo model and later Busse (1976) used it again with the inclusion of a zonal magnetic field to investigate the nature of convection in a sphere with magnetic field. Since the present understanding of the hydromagnetic dynamo problem is strongly influenced by the results of these two papers, the annulus model is described in detail in Section 5 together with a comprehensive discussion of the various thermal instabilities that are possible. New results are presented in Section 5.1 concerning those instabilities which can occur on intermediate length scales (see eq. 5.5a), while in Section 5.2 results which are qualitatively equivalent to some earlier results of Feam (1979a) are outlined. Since the analysis of Section 5 contains more detail than the remaining sections, it is perhaps advisable for the nonspecialist reader to bypass Sections 5.1, 5.2 initially and look instead at Figs. 2 and 3 which summarise the key results. Nevertheless we feel that the analysis in this section, none of which has appeared elsewhere, provides a useful direct extension of Busse's (1976) results. In particular they help us to assess in Section 6 the viability of his proposed weak field dynamo model together with the scaling law (see eqs. 5.9a and 6.2) for the upper limit of the strength of planetary magnetic fields. Though this model has never been worked out in complete detail, the results of Section 5 suggest that it is subject both to dynamic and dynamo instabilities. This being so the model is unlikely to operate in the manner proposed. It is therefore suggested that the geodynamo is a modified form of the 136 strong field model developed by Braginsky (1964b,c) with magnetic field strength of order ~ (see eq. 2.11 below) which is intermediate between Busse and Braginsky. 2. Wave m o t i o n s and convective instabilities in an unbounded fluid Many o f the important characteristics of magnetohydrodynamic convection in a rotating system can be isolated by considering an unbounded fluid, in which all conditions are uniform and II is parallel to g. We, therefore, consider small perturbations u=u', B=B0+b ', 0=0 ', p=po+p ' (2.1) of the static state and linearise eq. (1.1) accordingly. Since the resulting equations have constant coefficients, we may seek solutions proportional to exp i(k .x - cot) (2.2) As a result the governing equations reduce to the algebraic system (-i6o + va2)u ' + 2 f / X u' = = -tkP' - o~0'g + i(p/z)-l(k • Bo)b' (2.3a) (-/to + ~2)b' = i(k" Bo)u' (2.3b) (-i6o + ra2)0 '= - I1"u' (2.3c) k - u' = k . b' = 0 (and Ikl = a) which have non.trivial solutions provided (2.3d,e) and, when in addition dissipative processes are ignored, the resulting quadratic equation for 6o has two roots. When the Alfven angular velocity ~"~M = Bo/L V ~ is small compared to the rotation speed I2 (2.6a) ~2M< < ~ (2.6b) as it is in the geophysically interesting case, the magnitudes of the two frequencies differ considerably. One frequency is of order I2 and corresponds to inertial waves modified by the presence of the magnetic field. The other frequency is of order I2MC,where ~"2MC(=~"~I/~'2)< < ~Q,M (2.6c) Unlike the fast inertial wave, the primary force balance for this slow wave is between the Lorentz (magnetic) and Coriolis forces, while the fluid inertia pau/at is negligible. Once the effects of dissipation are taken into account all waves decay. With the inclusion of an adverse density gradient, the size of a3g at which instability sets in becomes our main concern. Since positive growth rates are always possible for sufficiently long vertical length scales, whenever o¢3gis positive, eq. (2.4) must be interpreted cautiously. For this reason, in the estimates below it is always assumed that the components of k in the g, Bo and f~ directions are each o f order L -1 . Though this assumption is not always justified, it is for the cases considered in this section. In the absence of dissipation, buoyancy (Archimedean) forces reduce the frequency 6o of the slow waves, eq. (2.6c), and yield 4 (k a2 -i6o + Ka2 o3g (2.4a) where - i s = -i6o + va2 + (ptO-l(k • Bo)2/(-i6o + )~a2) (2.4b) and k± is the component of the real wave vector k lying in the plane perpendicular to 15and g. The eqs. (2.4) were first obtained by Braginsky (1964a) and following his development we now isolate some of the key magnetohydrodynamic processes. In the absence of buoyancy forces eq. (2.4a) reduces to s = -+2(k " a ) / a (2.5) These are Braginsky's (1967) MAC waves which become unstable when 6o2 is negative and this can occur for ot3g= O(I2~t) (2.7b) On the other hand, the fast inertial waves require much larger buoyancy forces to make them unstable. Indeed inspection of eq. (2.4) shows that the criterion corresponding to (2.7b) is ot3g= 0 ( ~ 2) (2.8) Since ~2/I2Mis large, it is reasonable to suppose that 137 MAC waves are more readily excited and for this reason Braginsky (1967) envisaged that they play a key role in the dynamo process. When dissipative effects are reinstated, instability may set in through a state of either steady or oscillatory convection. The analysis of Roberts and Stewartson (1974) for a plane layer indicates clearly that the preferred mode depends mainly on the ratio q = K[X (2.9) of the thermal to magnetic diffusivities. For values of q less than unity steady convection is always preferred. Therefore, as for MAC waves, viscosity is neglected and having set co = 0 in eq. (2.4) we obtain k_~g _ 2ou ~ ( k . n ) (k. no)2 + (2.1o) 2a3(k .n)K (k- no)2 2pu ~ ( k -n) This suggests that a suitable measure of the magnetic field strength and the adverse density gradient are A = ~McL2/X(=B~/pUX~) (2.11 a) and R = o,,~L2/Ka (2.1 lb) The former is the ratio of the time scales for magnetic diffusion and slow magnetic waves. The latter is a modified Rayleigh number. Evidently R is minimised and takes an order one value, when A is of order unity. In this case, o~g = O(E[22) (2.12a) where E = K/L212 (2.12b) is a modified Ekman number based on the thermal diffusivity g. The main deficiency in the simple calculations described above lies in the assumption that the length scale of convection is L in all directions of interest. In this respect the preferred mode is not always obliging! Often very short length scales are invoked as in the model discussed in the following section. Nevertheless the analysis does focus our attention on the three key parameters E, q, and A of the magnetohydrodynamic convection problem. The realised values for the geodynamo are E = O(10-14), q = O(10 -6) and A = 0(3 X 10-2Bo)2 (2.13 a,b,c) where B0 is measured in gauss, and for this reason it will be assumed throughout this paper that E < < 1, q < 1, e = 13(1) (2.14a,b,c) where e(=v/K) is the Prandtl number. Whether A is large or small depends on the size of Bo. If B0 is weak and takes a value of about 5 gauss appropriate to the core surface, then A is small, of order 10-2. On the other hand, if there is a large toroidal field, say of 300 gauss, then A may be of order 102. Order one values of A also enjoy special significance for it is with feld strengths of this size that convection can occur most readily. Indeed, according to eqs. (2.10)-(2.12) the minimum Rayleigh number R is of order unity. To compare this with the MAC wave result, eq. (2.7b) is written in the form R = O(A/q) (2.15) Evidently, for small q, MAC waves can only be driven by density gradients with highly supercritical Rayleigh numbers. 3. Plane layer dynamo The problem of magnetohydromagnetic convection in a rapidly rotating plane layer has been investigated by a number of authors, including Chandrasekhar (1961), Eltayeb (1972, 1975) and Roberts and Stewart. son (1974). Here the particular case of a horizontal layer of fluid permeated by a horizontal magnetic field and rotating about a vertical axis will be discussed. Relative to rectangular cartesian coordinates (x, y, z), the lower boundary is z = 0 and II = ~2~., g = -g'/., Bo = Bl (3.1 a,b,c) where the superscript "denotes unit vectors. For simplicity in the subsequent discussion both boundaries are assumed to be stress free and perfect conductors of both heat and electricity. When the magnetic field is absent and the Prandtl number is greater than unity (o > 1) instability always sets in as steady convection. The ensuing margir~al convection is characterised by the temperature perturba. tions 0 = Oo sin(rrz/L) etkj-"x + cc (3.2) where cc denotes the complex conjugate of the expression preceeding it and Oo is arbitrary on linear theory. 138 According to eq. (2.4) R and k± are related by (k.t/a)ZR = (oE)-l(27r/aL) 2 + oE(aL) 4 (3.3a) and a 2 = k~. + r¢2[L 2 (3.3b) Minim±sat±on of R yields the critical Rayleigh number Ro and the corresponding wave number k±o. They are given approximately by Ro = 3(4rr4/oE)1/3, Lk±o = (X/2~r/oE) 1/3 (3.4a,b) The large values of Ro and k±o result because of the severe constraints imposed by the rapid rotation. Indeed to lowest order the flow is geostrophic and u satisfies 21"1X u = - V p (3.5a) The solution has the form u = 7 X ~b~+ wi (3.5b) where ~ and w vary rapidly on the short horizontal length scale L(aE) 1/~, as indicated by eq. (3.4b). At the next order of approximation the two dimensionality of the flow is broken by the competing effects of buoyancy and viscosity. Thus ~ and w vary on the relatively long vertical length scale L and are given by = ~o cos(fez/L) eik±°'x + cc, (3.6a) W = W 0 sin(yz/L) e ik-l-O'x + cc (3.6b) where k m ~ o = (x/2/Tr)wo, 0o = ([3/~k~.o)Wo (3.6c,d) The existence of two distinct length scales prompted Childress and Soward (1972) to use these results as the basis of a hydromagnetic dynamo model. As well as the mathematical advantages of two scales, the flow (3.5b) itself is well endowed with the helicity necessary for the dynamo process. In fact, if we define the horizontal average • 1 tt <...>= f f ... d dy -1 -/ (3.7a) then the mean value of the helicity for the convection roU described by eq. (3.6) is H = >l) and So = rl/Ek(>>l) (5.5b,c) The ensuing approximations that can be made indicate the existence of two large and one small root of the cubic (5.4). The former is loosely related to Rossby waves while the latter corresponds to a magnetic wave. Both categories were mentioned in Section 4 above and will be discussed in more detail in Sections 5.1.1 and 5.1.2 below. 5.1.1. Rossby modes The two large roots may be isolated by neglecting all terms in eq. (5.4) related to dissipative processes. Consequently, if we let P = -is (5.6a) where s is dimensionless and measures the frequency (5.3b) in units of K/L 2, (5.4) is given approximately by - E s + 2~7/k - (R - Ak2/q)/s = 0 (5.6b) and has the two roots s± =s0(1 -+ [ ( R o - R + k 2 A / q ) / R o ] 1/2} (5.6c) Since instability is associated with complex values of s, it is clear that So is the frequency of the modified Rossby waves which occur on the non-dissipative stability boundary R = Ro + k2A/q (5.7) Above this boundary waves grow very fast on the time scale soI . Whether or not the neutral waves (5.6), which exist below the boundary (5.7), grow or decay on the slow diffusion time k -2 depends on the small terms neglected in (5.6). To determine this small growth rate, we set P = -is + p (5.8a) and approximate (5.4) on the basis that Ip/sl<< 1 (5.8b) At leading order this gives Ro(R o - R + k:A/q) = I (Ro - R + k2A/q) I (1-o)k 2 . "2 . .)z (R - lak A[q)] (5.8c) where # = (q-1 _ o)/(1 - o) (5.8d) The curves of constant growth rate are parabolas all of which pass through the origin and touch the stability boundary (5.7) at its intersection with the line R = #k~A/q (see Fig. 2). We may also note in passing that the maximum value of A on the neutral parabola p = 0 is A1 = (Ro/k2)q2](1 + o)(1 - q) and here R and s take the values (5.9a) R1 = Ro(o](1 + o) 2 + 1/(1 + oX1 - q)) (5.9b) sl = So/(1 + o) (5.9c) As illustrated in Fig. 2 the unstable waves are located 144 R I '~ IIIII T Z~"/%f ."/ 1/ / ,,,,11~ E A, A Fig. 2. The non-dissipative and dissipative stability bounda- ries, for fixed k, are indicated by the straight line I and curve II, respectively. The former is R =Re +k2A/q, while the latter is a composite ofeq. (5.8) with p = 0 and eq. (5.12) with r = re. The straight line III def'med by R = k~Afq divides the region lying between the two stability boundaries into two parts. Above III the unstable waves are fast with frequency s_ (see eq. 5.5c), while below they are slow with frequency s(r) (see eq. 5.12), where r > rc. All curves of constant growth rate, typified by II and IV, pass through the point P, which lies at the intersection of I with V defined by R = #k2A/q. Note, however, that the point P lies in the positive A, R quadrant only i f , > 0. The point Q(A 1, RI) (see eq. 5.9) locates the maximum value of A, at which neutral fast waves exist. to the right of the neutral parabola and within the strip k2A]q 0) and so they always propagate eastward. Now, according to eq. (5.8), the growth rate is infinite on the non-dissipative stability boundary (5.7) as well as on the line R = k2A/q. Consequently the approximation tp/sl < < 1, upon which eq. (5.8) is based, breaks down in the vicinity of both lines. In the case of the stability boundary (5.6), modes corresponding to eq. (5.8) with large growth rates on the resistive time scale make a transition to modes corresponding to (5.6) with small growth rates on the Rossby wave time scale. In absolute terms, of course, the growth rate increases monotonically with R. On the other hand, the nature of the instability in the vicinity of R =k2A[q will become apparent from the analysis below. For the moment it is sufficient to notice that here the frequency of the waves corresponding to s_ is very small and s_ = 0 onR =k2A/q (5.11) 5.1.2. Magnetic modes So far attention has been restricted to the two large roots of eq. (5.4). To complete the picture, the nature of the remaining small root must be considered. In this limit it is reasonable to neglect the inertia term au[at as well as the viscous term in comparison with the Lorentz force. By this device we isolate the magnetic modes mentioned in Section 4. Therefore, upon setting E = 0 and P = -is +p, as before, the real and imaginary parts of the resulting quadratic lead to two equations which determine s and p in terms of A and R. More convenient expressions are obtained, however, if we introduce a new parameter r. They are p(r) = k 2 {r(1 - q)/2q - (1 + q)/2q} (5.12a) s(r) = S(r) = So((r + 1)R - (r - 1)k2A/q}/4rRo and - s ( r ) s ( - r ) = k4 ((1 - q)/2q}2(r 2 - 1) (5.12b) (5.12c) Here we regard r as an independent variable in the definition of S(r) but, in view ofeq. (5.12c), r, A and R are related in the case o f s(~'). Specifically given a value of r, substitution of (5.12b) into (5.12c) leads to a quadratic expression relating R to A. It defines a hyperbola with asymptotes S(r) = 0, S(-r) = 0 and has a branch in the region A >t 0, R I> 0 of interest only when Irl i> 1 (5.12d) From (5.12b and c) it is clear that the two hyperbolas defined by r = -+rl, where rl is a given positive constant, are coincident. Consequently, everywhere on the hyperbola Irl = ra, the growth rate takes one of the two values p(-+rl) given by eq. (5.12a). The corresponding value of the'two frequencies s(rl) and s(-rl) are determined by eq. (5.12b) and take positive and negative values, respectively, for all A, R on the hyperbola r = re. When r = re, where rl -- (1 + q)/(1 - q) (5.13a) the growth rate is zero while the corresponding frequency is s = ~so(R - k2A)/Ro(1 + q) (>9) (5.13b) The region of instability is defined by r > rc and here, since s(r) > 0, the unstable waves propagate eastward. When R is of order Re, eq. (5.12b) would suggest at first sight that that frequency s is large of order So. This being the case, the approximations upon which eqs. (5.12) are based would be violated. A closer inspection of (5.12b and c) reveals that for order one values of z(>0, say) the point (A, R) lying on the hyperbola r = constant is close to one or other of the asymptotes S(+-r) = 0 (5.14a) where the + or - sign is taken depending on whether R is less than or greater than k2A/q, respectively. In each case the realised value of the frequency s(+-r)at (A, R) is very small of order k4[so while the corresponding value of s(T-r) is large and given approxi- mately by s(~r) = ~so(R - k2A/ q)/Ro (5.14b) In general only the low.frequency solution is admissi- ble and this supplements the two solutions (5.6c) found earlier. On the other hand, when R and k2A/q are small compared with Re, both the solutions s(+r) of (5.i 2) are acceptable. Indeed, when R > k2A/q, (5.14b) now approximates not just s(-r) but also s_ in (5.6c). When Ik2A/qR - 11 < < 1 (R = O(Ro)) (5.15) the approximations leading to eq. (5.14) fail and the complete shape of the hyperbolas must be considered. In this region, eq. (5.12) indicates that in addition to k2r both s(O and s ( - r ) are of order ks~/2 and though large they are still small compared with So. Therefore, both solutions are acceptable and provide the key to our understanding of the switch over from magnetic to Rossby modes that occurs across the line R = k2A/q. In particular we may note that, when R = k2A/q, the frequency and the growth rate of the growing mode are both given by s =p = ~ ( 1 - q R '~'/2 ks~/2 (5.16) The general picture, which emerges and is summarised in Fig. 2, is the following. Consider a tLxed value of A with k2A[q of order Re. All modes are stable 145 below the asymptote S(rc) = 0, namely the line R = k2A. Above this line the slow magnetic modes defined by eq. (5.12) with r = re are unstable. Their frequen- cies and growth rates are o f order k4[so and k 2, respectively. Across the line R = k2A[q, the growing mag- netic modes make a rapid transition to the growing Rossby modes defined by s_ in eq. (5.6c). The frequencies are now of order So but the growth rates remain the same, of order k 2. Other than the rapid frequency change the most striking feature of the transition is the relatively large growth rates of order ks~/2 which are achieved (see eq. 5.16). Once above the non-dissipative stability boundary both s and p are of order So. 5.2. The onset of instability In Section 5.1 above, attention was restricted to fixed values of the wave number k. When the critical Rayleigh number is obtained without any special restrictions on k, a rather different picture emerges. First, in the absence of dissipation, minimisation of R, as given by eq. (5.7), R with respect to k yields the critical Rayleigh number Re = 2n(A/Eq)X/2 (5.17) Second, when the effects of dissipation are included, the true critical Rayleigh number is obtained by minimising R, as given by eq. (5.4), with respect to k, keeping s(=/P) real. Feam (1979a) has undertaken a similar analysis within the broader context of convection in the sphere. We will outline below all his main conclusions, which relate to our simpler problem, but will derive the results for the annulus model using a technique which we believe is new in the context of stability theory. We begin by introducing the physically relevant parameters X =k2[s and Y= (1 + o)-l(So[S) (5.18a,b) whose magnitudes measure the ratio of the oscillation time scale to the diffusion and Rossby wave time scales, respectively. In terms of these new variables the real and imaginary parts of eq. (5.4) determine the following parametric representations of A and R. They are A=2[E(1 +o)r/211/3A, R =2[ff*/E(1 + o)]1/3/~ (5.19a,b) 146 where = (1 - q)-l(y_ ~.)(q2 + X2)(X2y)-~a, (5.19c) / } = ( 1 _ q ) - l ( y _ ~l-+ e \~-)( 1 + x 2 x x I , a ) -2'a (S.19d) and e = (1 - q)o[(1 + o) (5.19e) By varying X and Y('~) a region in the/~, k plane is generated. Its lower edge defines the stability bound. ary and the value of R here for a particular choice of determines the critical Rayleigh number. On this boundary and on some other curves, which are of no particular interest, the Jacobian ~(.~, R)/O(X, Y) vanishes. Application of this condition yields F(Y, e) =G(X2, q2) (5.20a) where F(Y, e) = (2Y - 1)(Y - 2 + 2e) (2Y - 1 + e)(Y+ 1) (5.20b) (X 2 + q2X2X2 - 1) c(x = (x + - (5.20c) Q / A r 0 E , b 8 C Fig. 3. With o = 0.1, R is plotted vs..~, for various values of q according to eqs. (5.19) and (5.20). The stability boundary corre- sponds to the continuous curve AB in (a) qtop > q = 0.25; the segmented curves AG, GF, FB in Co)q = qtov = 0 . 5 3 1 3 5 . . . ; the segmented curves AF, FB in (c) qtov < q 0.6. (The insert in (b) which is not drawn to scale provides a blow up of the curves near G and illustrates how reeonnection takes place at H.) The values of X2 and Y appropriate to the asymptotic behaviour at B, E and C are described in the caption to Fig. 5 while at B, k = 1/.~, and at E and C, R ~ ~1/2 (see also eqs. 5.21-5.23). 147 The various solutions of eqs. (5.19) and (5.20) are discussed in detail in the Appendix. It transpires that three separate cases can be distinguished and they are illustrated in Fig. 3 for o = 0.1. The results indicate that the stability boundary is smooth for q ~ 1, X and Y are of order q1,,2 and q-l ,respectively. Again referring to eq. (5.18) this shows that the time scale of oscillation is long compared with the Rossby time scale s-I and therefore convection corresponds to the slow magnetic mode discussed in Section 5.1.2. It should be emphasised that the sudden transition from Rossby to magnetic modes that occurs at A = 1 is peculiar to the case q < < 1. For the cases illustrated in Fig. 3, the transition is continuous. Finally we consider briefly the case q > 4~ (5.26) Now the curve DC intersects AB at the two points G and F of Fig. 3(b). On the segment GF, X is of order qS and Yis approximately 1/2. The corresponding convection is clearly described as a Rossby mode. As stated earlier the character of the three different types of solution described above was determined earlier by Fearn (1979a). He was primarily concerned, however, with the behaviour of the critical Rayleigh number for the full sphere. In that case there is an added degree of flexibility owing to the arbitrariness of the radius ~0 of the circular cylinder close to which convection takes place. The annulus model considered above represents the special case in which that radius is held fixed and for this reason must be interpreted cautiously. Nevertheless, though some features of the convection in a sphere cannot be illustrated without allowing for variations in ~o, there are some features which are duplicated. These include the modal discontinuity at F (see Fig. 3c) which is found to be very common, as well as the smooth dependence on A (see Fig. 3a) which also occurs, but less often. On the other hand, Fearn (1979a) found no examples with the modal discontinuity associated with the point G in Fig. 3(b) except when the variations of ~o are restricted as they would be if the presence of a central rigid core is allowed for. 6. Discussion Busse's (1975) original geodynamo model based on the annulus configuration of Section 5 was developed through perturbations about the non.magnetic system. As in the case of the plane layer dynamo discussed in Section 3 the Rayleigh number is assumed to be slightly in excess of the critical value appropriate to the non-magnetic problem. Provided the excess is suf- 148 ficiently large, finite amplitude convection ensues capable of sustaining large-scale magnetic field. The magnitude of this field is exactly that required to prevent any further intensification of the convection. Like the plane layer model it is dynamically stable but unlike the plane layer model the dynamo process itself is also stable. The success of the annulus model pivots on the fact that, near A = 0, the critical Rayleigh number Rc is an increasing function of A. This is only the case, however, when A ~t O(1). In addition, the system should also be dynamically stable because the Rayleigh number is close to R e. This picture, which again pivots on a quasi-linear theory, is far too simplistic and cannot work. To highlight the difficulty, we note that for dynamo action the magnetic Reynolds number Rx = UL/X (6.3a) based on a typical fluid velocity Umust be at least of order unity and so the Peclet number Rx = UL/r (6.3b) must be at least of order q-l. It follows that convection must be highly non-linear and to obtain flows of the required intensity we may anticipate that the Rayleigh number must be well in excess of its critical value. Since the intense non-linear effects are largely concentrated in promoting heat transport, the mean temperature profile must be maintained at a value very close indeed to the adiabat and this makes the realised value of the Rayleigh number a very uncertain quantity. By contrast, this is not the case for the proposed order one value of A, which is based on an anticipated balance of the Coriolis and Lorentz forces. Here since the diffusivity ratio q is so small, the relative sizes of the two forces are unlikely to be significantly altered by nonlinear processes. Little is known about the non-linear convection problem. Roberts and Stewartson (1974, 1975),however, have initiated the study of non-linear convection in the presence of a uniform horizontal magnetic field Bo = Bo$' in the simple plane layer geometry of Section 3. The analysis hinges on the fact that, when viscosity is neglected, there is a non-convective geostrophic mode with velocity U(x)~7which suffers no damping. This mode though not excited by buoyancy forces is readily driven by the Lorentz force, when the uniform magnetic field is modified by the finite amplitude convection. When A >vr3, two modes of convection are possible, which consist of rolls with axes oblique but making equal angles with the magnetic field. Roberts and Stewartson find that there exist finite amplitude equilibria in which only one family of rolls is present. They also find that, for certain values of A, the equilibria are unstable to perturbations composed of the second set of rolls together with the shear flow U(x). In this case no stable finite equilibria were found. A more recent extension by Soward (1979b), which is restricted to q < < l, isolated stable equilibria, when the damping of the shear flow U(x) by Ekrnan suction is taken into account. It was shown, however, that as the Rayleigh number is increased above its critical value, the geostrophic flow U(x) quickly becomes large compared with the convective velocities. The result suggests that, for Rayleigh numbers well above their critical values, fluid motions within a sphere will be dominated by large zonal flows with relatively small asymmetric convective motions, while the magnetic field itself is largely zonal. The magnetic field and flow configuration just described is, of course, very similar to that proposed by Braginsky (1964b,c, 1967). The main difference between the two pictures is that whereas Braginsky envisages convection in the form of non-dissipative MAC waves, we envisage that diffusion plays a central part in the convective process. The reason is simply that, when A = O(1), the MAC waves and the magnetic diffusion time scales are comparable (see eq. 2.1 la). The ensuing dynamo is still of the ~,~-type but the a-effect is not that appropriate to the high-conductivity limit. Instead it more closely resembles the lowconductivity limit of the annulus and plane layer models. In conclusion we suggest that the geodynamo operates with A = O(1) or slightly larger and R = O(q-l), roughly at the location of Braginsky's MAC waves. Here motion should be sufficiently intense to regenerate magnetic field and perhaps conditions are also favourable for dynamo stability. Acknowledgement The author would like to thank Dr. D. Fearn for many useful discussions relating particularly to Section 5. Appendix In order to distinguish the various types of solutions of eqs. (5.19)and (5.20)that can occur it is necessary to appreciate the character of the functions F and G. When Y = ½,F is zero and as Y increases F reaches a maximum Fmax(eX 2q 2, we have G > I > Fmax(e ) implying that eq. (5.20a) has no solutions. The character of G in this region is therefore of no interest. The case 1/V~" < q < 1 is the same as the previous case except that there is no local maximum. Instead F decreases throughout the interval 0 ~ 1/9, or equivalently q <(8o- 1)/9o (<1) (A2) Fmax is less than ~ and so X is determined by eq. (5.20) as a single-valued function of Y. Consecjuently eq. (5.19c,d) determines a single curve in the A, R plane, similar to AB in Fig. 3(a), which marks the stability boundary. When e < 1/9, or equivalently (80- 1)/9o