physics of mantle convection contents - usc geodynamics

113
PHYSICS OF MANTLE CONVECTION Yanick Ricard Laboratoire de Science de la Terre Universit´ e de Lyon; Universit´ e de Lyon 1; CNRS; Ecole Normale Sup´ erieure de Lyon, 2 rue Rapha¨ el Dubois, 69622, Villeurbanne Cedex, France. abstract This chapter presents the fundamental physics necessary to under- stand the complex fluid dynamics of mantle convection. The first section derives the equations of conservation for mass, momentum and energy, and the boundary and interface conditions for the various physical quantities. The thermodynamic and rheological properties of solids are discussed in the second section. The mech- anism of thermal convection and the classical Boussinesq and anelastic approx- imations are presented in the third section. As the subject is not often included in geophysical text books, an introduction to the physics of multicomponent and multiphase flows is given in a fourth section. At last the specific applications to mantle convection are reviewed in the fifth section. Contents 1 Introduction 4 2 Conservation equations 5 2.1 General expression of conservation equations ........... 5 2.2 Mass conservation .......................... 7 2.3 Momentum conservation ...................... 8 2.3.1 General momentum conservation ............. 8 2.3.2 Inertia and non-Galilean forces ............... 10 2.3.3 Angular momentum conservation ............. 11 2.4 Energy conservation ........................ 12 2.4.1 First law and internal energy ................ 12 2.4.2 State variables ....................... 13 2.4.3 Temperature ........................ 15 2.4.4 Second law and entropy .................. 16 1

Upload: others

Post on 12-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

PHYSICS OF MANTLE CONVECTION

Yanick RicardLaboratoire de Science de la Terre

Universite de Lyon; Universite de Lyon 1; CNRS; Ecole Normale Superieure de Lyon,2 rue Raphael Dubois, 69622, Villeurbanne Cedex, France.

abstract This chapter presents the fundamental physics necessary tounder-stand the complex fluid dynamics of mantle convection. The first section derivesthe equations of conservation for mass, momentum and energy, and the boundaryand interface conditions for the various physical quantities. The thermodynamicand rheological properties of solids are discussed in the second section. The mech-anism of thermal convection and the classical Boussinesq and anelastic approx-imations are presented in the third section. As the subject is not often includedin geophysical text books, an introduction to the physics ofmulticomponent andmultiphase flows is given in a fourth section. At last the specific applications tomantle convection are reviewed in the fifth section.

Contents

1 Introduction 4

2 Conservation equations 52.1 General expression of conservation equations . . . . . . . .. . . 52.2 Mass conservation . . . . . . . . . . . . . . . . . . . . . . . . . . 72.3 Momentum conservation . . . . . . . . . . . . . . . . . . . . . . 8

2.3.1 General momentum conservation . . . . . . . . . . . . . 82.3.2 Inertia and non-Galilean forces . . . . . . . . . . . . . . . 102.3.3 Angular momentum conservation . . . . . . . . . . . . . 11

2.4 Energy conservation . . . . . . . . . . . . . . . . . . . . . . . . 122.4.1 First law and internal energy . . . . . . . . . . . . . . . . 122.4.2 State variables . . . . . . . . . . . . . . . . . . . . . . . 132.4.3 Temperature . . . . . . . . . . . . . . . . . . . . . . . . 152.4.4 Second law and entropy . . . . . . . . . . . . . . . . . . 16

1

Page 2: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

2.5 Gravitational forces . . . . . . . . . . . . . . . . . . . . . . . . . 172.5.1 Poisson equation . . . . . . . . . . . . . . . . . . . . . . 172.5.2 Self-gravitation . . . . . . . . . . . . . . . . . . . . . . . 182.5.3 Conservative forms of momentum and energy equations .19

2.6 Boundary and interface conditions . . . . . . . . . . . . . . . . . 212.6.1 General method . . . . . . . . . . . . . . . . . . . . . . . 212.6.2 Interface conditions in the 1D case and for bounded vari-

ables . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.6.3 Phase change interfaces . . . . . . . . . . . . . . . . . . 232.6.4 Weakly deformable surface of a convective cell . . . . . .24

3 Thermodynamic and rheological properties 263.1 Equation of State and solid properties . . . . . . . . . . . . . . .263.2 Rheology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.2.1 Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . 293.2.2 Viscous Newtonian rheology . . . . . . . . . . . . . . . . 303.2.3 Maxwellian Visco-elasticity . . . . . . . . . . . . . . . . 313.2.4 Non-linear rheologies . . . . . . . . . . . . . . . . . . . . 33

4 Physics of convection 354.1 Basic balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364.2 Two simple solutions . . . . . . . . . . . . . . . . . . . . . . . . 37

4.2.1 The diffusive solution . . . . . . . . . . . . . . . . . . . 374.2.2 The adiabatic solution . . . . . . . . . . . . . . . . . . . 384.2.3 Stability of the adiabatic gradient . . . . . . . . . . . . . 39

4.3 Approximate equations . . . . . . . . . . . . . . . . . . . . . . . 404.3.1 Depth dependent reference profiles . . . . . . . . . . . . 404.3.2 Nondimensionalization . . . . . . . . . . . . . . . . . . . 404.3.3 Anelastic approximation . . . . . . . . . . . . . . . . . . 424.3.4 Dimensionless numbers . . . . . . . . . . . . . . . . . . 434.3.5 Boussinesq approximation . . . . . . . . . . . . . . . . . 444.3.6 Internal heating . . . . . . . . . . . . . . . . . . . . . . . 454.3.7 Alternative forms . . . . . . . . . . . . . . . . . . . . . . 474.3.8 Change of nondimensionalization . . . . . . . . . . . . . 47

4.4 Linear stability analysis for basally heated convection . . . . . . . 484.5 Road to chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5 Introduction to physics of multicomponent and multiphaseflows 515.1 Fluid dynamics of multicomponent flows in solution . . . . .. . 53

5.1.1 Mass conservation in a multicomponent solution . . . . .535.1.2 Momentum and energy in a multicomponent solution . . . 55

2

Page 3: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

5.1.3 Entropy conservation in a multicomponent solution . .. . 565.1.4 Advection-diffusion equation and reaction rates . . .. . . 575.1.5 Conservation properties of the advection-diffusionequation 615.1.6 Laminar and turbulent stirring . . . . . . . . . . . . . . . 635.1.7 Diffusion in Lagrangian coordinates . . . . . . . . . . . . 64

5.2 Fluid dynamics of two phase flows . . . . . . . . . . . . . . . . . 665.2.1 Mass conservation for matrix and fluid . . . . . . . . . . 685.2.2 Momentum conservation of matrix and fluid . . . . . . . . 685.2.3 Energy conservation for two-phase flows . . . . . . . . . 705.2.4 Entropy production and phenomenological laws . . . . . .725.2.5 Summary equations . . . . . . . . . . . . . . . . . . . . . 74

6 Specifics of Earth’s mantle convection 766.1 A mantle without inertia . . . . . . . . . . . . . . . . . . . . . . 76

6.1.1 Dynamic models . . . . . . . . . . . . . . . . . . . . . . 776.1.2 Mantle flow and post-glacial models . . . . . . . . . . . . 786.1.3 Time dependent models . . . . . . . . . . . . . . . . . . 80

6.2 A mantle with internal heating . . . . . . . . . . . . . . . . . . . 806.3 A complex rheology . . . . . . . . . . . . . . . . . . . . . . . . 81

6.3.1 Temperature dependence of viscosity . . . . . . . . . . . 846.3.2 Depth-dependence of viscosity . . . . . . . . . . . . . . . 856.3.3 Stress-dependence of viscosity . . . . . . . . . . . . . . . 866.3.4 Grain size dependence of viscosity . . . . . . . . . . . . . 86

6.4 Importance of sphericity . . . . . . . . . . . . . . . . . . . . . . 876.5 Other depth-dependent parameters . . . . . . . . . . . . . . . . . 87

6.5.1 Thermal expansivity variations . . . . . . . . . . . . . . . 876.5.2 Increase in average density with depth . . . . . . . . . . . 876.5.3 Thermal conductivity variations . . . . . . . . . . . . . . 88

6.6 Thermo-chemical convection . . . . . . . . . . . . . . . . . . . . 896.7 A complex lithosphere: plates and continents . . . . . . . . .. . 91

3

Page 4: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

1 Introduction

In many text books of fluid dynamics, and for most students, the word “fluid”refers to one of the states of matter, either liquid or gaseous, in contrast to the solidstate. This definition is much too restrictive. In fact, the definition of a fluid rests inits tendency to deform irrecoverably. Basically any material that appears as elasticor non-deformable, with a crystalline structure (i.e., belonging to the solid state)or with a disordered structure (e.g., a glass, which from a thermodynamic point ofview belongs to the liquid state) can be deformed when subjected to stresses for along enough time.

The characteristic time constant of the geological processes related to mantleconvection, typically 10 Myrs (3×1014 s), is so long that the mantle, althoughstronger than steel and able to transmit seismic shear waves, can be treated asa fluid. Similarly, ice, which is the solid form of water, is able to flow frommountain tops to valleys in the form of glaciers. A formalismthat was developedfor ordinary liquids or gases can therefore be used in order to study the insideof planets. It is not the equations themselves, but their parameters (viscosity,conductivity, spatial dimensions...) that characterize their applicability to mantledynamics.

Most materials can therefore behave like elastic solids on very short time con-stants and like liquids at long times. The characteristic time that controls theappropriate rheological behavior is the ratio between viscosity, η, and elasticity(shear modulus),µr, called the Maxwell timeτM (Maxwell, 1831-1879),

τM =η

µr

. (1)

The rheological transition in some materials like silicon putty occurs in only afew minutes; a silicon ball can bounce on the floor, but it turns into a puddle whenleft on a table for tens of minutes. The transition time is of the order of a fewhundred to a few thousand years for the mantle (see 3.2). Phenomena of a shorterduration than this time will experience the mantle as an elastic solid while mosttectonic processes will experience the mantle as an irreversibly deformable fluid.Surface loading of the Earth by glaciation and deglaciationinvolves times of a fewthousands years for which elastic aspects cannot be totallyneglected with respectto viscous aspects.

Although the word “convection” is often reserved for flows driven by internalbuoyancy anomalies of thermal origin, in this chapter, we will more generally use“convection” for any motion of a fluid driven by internal or external forcing. Con-vection can be kinematically forced by boundary conditionsor induced by densityvariations. The former is “forced” and the latter is “free convection” which canbe of compositional or thermal origin. We will however, mostly focus on the

4

Page 5: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

aspects of thermal convection (or Rayleigh-Benard convection, (Rayleigh, 1842-1919; Benard, 1874-1939)) when the fluid motion is driven bythermal anomaliesand discuss several common approximations that are made in this case. We know,however, that many aspects of mantle convection can be more complex and in-volve compositional and petrological density anomalies ormulti-phase physics.We will therefore review some of these complexities.

The physics of fluid behavior, like the physics of elastic media is based on thegeneral continuum hypothesis. This hypothesis requires that quantities like den-sity, temperature, or velocity are defined everywhere, continuously and at “points”or infinitesimal volumes that have a statistically meaningful number of moleculesso that the quantity represents an average independent of microscopic molecularfluctuations. This hypothesis seems natural for ordinary fluids at the laboratoryscale. We will adopt the same hypothesis for the mantle although we know that itis heterogeneous at various scales and made of compositionally distinct grains.

2 Conservation equations

The basic equations of this section can be found in more detail in many classicaltext books (Batchelor, 1967;Landau and Lifchitz, 1980). We will only emphasizethe aspects that are pertinent for Earth’s and terrestrial mantles.

2.1 General expression of conservation equations

Let us consider a fluid transported by the velocity fieldv, a function of positionX and timet. There are two classical approaches to describe the physicsin thisdeformable medium. Any variableA in a flow, can be considered as a simplefunction of position and time,A(X, t), in a way very similar to the specificationof an electromagnetic field. This is the Eulerian point of view (Euler, 1707-1783).The second point of view is traditionally attributed to Lagrange (Lagrange, 1736-1813). It considers the trajectory that a material element of the flow initially atX0 would followX(X0, t). An observer following this trajectory would naturallychoose the variableA(X(X0, t), t). The same variableA seen by an Eulerian ora Lagrangian observer would have very different time-derivatives. According toEuler the time derivative would simply be the rate of change seen by an observerat a fixed position, i.e., the partial derivative∂/∂t. According to Lagrange, thetime derivative, noted withD, would be the rate of change seen by an observerriding on a material particle

DA

Dt=

(

dA(X(X0, t), t)

dt

)

X0

=∑

i=1,3

∂A

∂Xi

∂Xi

∂t+∂A

∂t, (2)

5

Page 6: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

where theXi are the coordinates ofX. SinceX is the position of a material ele-ment of the flow, its partial time derivative is simply the flowvelocityv. The La-grangian derivative is also sometimes material called the derivative, total deriva-tive, or substantial derivative.

The previous relation was written for a scalar fieldA but it could easily beapplied to a vector fieldA. The Eulerian and Lagrangian time derivatives are thusrelated by the symbolic relation

D

Dt=

∂t+ (v · ∇) (3)

The operator(v · ∇) is the symbolic vector(v1∂/∂x1, v2∂/∂x2, v3∂/∂x3) and aconvenient mnemonic is to interpret it as the scalar productof the velocity field bythe gradient operator(∂/∂x1, ∂/∂x2, ∂/∂x3). The operator(v ·∇) can be appliedto a scalar or a vector. Notice that(v · ∇)A is a vector that is neither parallel toA nor tov.

In a purely homogeneous fluid, the flow lines are not visible and the mechan-ical properties are independent of the original position offluid particles. In thiscase the Eulerian perspective seems natural. On the other hand, a physicist de-scribing elastic media can easily draw marks on the surface of deformable ob-jects and flow lines become perceptible for him. After an elastic deformation, thestresses are also dependent of the initial equilibrium state. The Lagrangian per-spective is therefore more appropriate. We will mostly adopt the Eulerian perspec-tive for the description of the mantle.. However when we discuss deformation ofheterogeneities embedded in and stirred by the convective mantle the Lagrangianpoint of view will be more meaningful (see section 5.1.7).

A starting point for describing the physics of a continuum are the conservationequations. Consider a scalar or a vector extensive variable(i.e., mass, momentum,energy, entropy, number of moles...) with a density per unitvolumeA, and avirtual but fixed volumeΩ enclosed by the surfaceΣ. This virtual volume isfreely crossed by the flow. The temporal change of the net quantity of A insideΩis

d

dt

ΩA dV =

Ω

∂A

∂tdV . (4)

SinceΩ is fixed, the derivative of the integral is the integral of thepartial timederivative.

The total quantity of the extensive variableA in a volumeΩ can be related toa local production,HA (with units ofA per unit volume and unit time), and tothe transport (influx or efflux) ofA across the interface. This transport can eitherbe a macroscopic advective transport by the flow or a more indirect transport, forexample at a microscopic diffusive level. Let us callJA the total flux ofA per unit

6

Page 7: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

surface area. The conservation ofA can be expressed in integral form as

Ω

∂A

∂tdV = −

ΣJA · dS +

ΩHA dV , (5)

where the infinitesimal surface element vectordS is oriented with the outwardunit normal; hence the minus sign associates outward flux with a sink of quantityA. Equation (5) is the general form of any conservation equation. When thevolumeΩ, surfaceΣ and fluxJA are regular enough (in mathematical terms whenthe volume is compact, the surface piecewise smooth and the flux continuouslydifferentiable), we can make use of the divergence theorem

ΣJA · dS =

Ω∇ · JA dV , (6)

to transform the surface integral into a volume integral. The divergence operatortransforms the vectorJ with Cartesian coordinates(J1, J1, J2) (Descartes, 1596-1650) into the scalar∇ · J = ∂J1/∂x1 + ∂J2/∂x2 + ∂J3/∂x3, scalar productof the symbolic operator∇ by the real vectorJ (notice the difference betweenthe scalar∇ · J and the operatorJ · ∇). Since the integral equation (5) is validfor any virtual volumeΩ, we can deduce that the general differential form of theconservation equation is,

∂A

∂t+ ∇ · JA = HA. (7)

A similar expression can be used for a vector quantityA with a tensor fluxJ and avector source termHA. In this case, the divergence operator converts the second-order tensor with componentsJij into a vector whose Cartesian components are∑

j=1,3 ∂Jij/∂xj .We now apply this formalism to various physical quantities.Three quantities

are strictly conserved: the mass, the momentum and the energy. This means thatthey can only change in a volumeΩ but influx or efflux across the surfaceΣ.We must identify the corresponding fluxes but no source termsshould be present(in fact, in classical mechanics the radioactivity appearsas a source of energy)(see also section 2.5.3). One very important physical quantity is not conserved -the entropy - but the second law of thermodynamics insures the positivity of theassociated sources.

2.2 Mass conservation

The net rate at which mass is flowing is

Jρ = ρv. (8)

7

Page 8: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Using either the Eulerian or the Lagrangian time derivatives, mass conservationbecomes

∂ρ

∂t+ ∇ · (ρv) = 0, (9)

orDρ

Dt+ ρ∇ · v = 0. (10)

In an incompressible fluid, particles have constant density, and so in the particleframe of reference, the Lagrangian observer does not see anydensity variation andDρ/Dt = 0. In this case, mass conservation takes the simple form∇·v = 0. Thisequation is commonly called the continuity equation although this terminology isa little bit vague.

Using mass conservation, a few identities can be derived that are very usefulfor transforming an equation of conservation for a quantityper unit mass to aquantity per unit volume. For example for any scalar fieldA,

∂(ρA)

∂t+ ∇ · (ρAv) = ρ

DA

Dt, (11)

and for any vector fieldA,

∂(ρA)

∂t+ ∇ · (ρA ⊗ v) = ρ

DA

Dt, (12)

whereA⊗ v is a dyadic tensor of componentsAivj.

2.3 Momentum conservation

2.3.1 General momentum conservation

The changes of momentum can be easily deduced by balancing the changes ofmomentum with the body forces acting in the volumeΩ and the surface forceacting on its surfaceΣ, i.e. Newton’s 2nd law (Newton, 1642-1727). The totalmomentum is ∫

Ωρv dV , (13)

and its variations are due to

• advective transport of momentum across the surfaceΣ,

• forces acting on this surface,

• internal body forces.

8

Page 9: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

−1 −0.5 0 0.5 1 1.5 2−1

−0.5

0

0.5

1

1.5

2

σyydxdz

σzydxdz

σxydxdz

x

y

z

o

Figure 1: The force per unit area applied on a surface directed by the normalvectorni is by definitionσ ·ni. The component of this force along the unit vectorej thereforeej · σ · ni.

The momentum conservation of an open, fixed volume, can therefore be expressedin integral form as

Ω

∂(ρv)

∂tdV = −

Σρv(v · dS) +

Σσ · dS +

ΩF dV . (14)

The tensorσ corresponds to the total stresses applied on the surfaceΣ, (see Figure1). Our convention is thatσij is thei-component of the force per unit area across aplane normal to thej-direction. The termF represents the sum of all body forces,and in particular the gravitational forcesρg (we will not consider electromagneticforces).

Using the divergence theorem (for the first term of the right side, ρv(v · dS)can also be writtenρ(v ⊗ v) · dS) and the equality (12), the differential form ofmomentum conservation becomes

ρDv

Dt= ∇ · σ + F. (15)

It is common to divide the total stress tensor into a thermodynamic pressure−P I

whereI is the identity stress tensor, and a velocity-dependent stressτ . The re-lationship between the tensorτ and the velocity field will be discussed later insection 3.2. Without motion, the total stress tensor is thusisotropic and equal

9

Page 10: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

to the usual pressure. In most geophysical literature, it has been assumed thatthe velocity dependent tensor has no isotropic component, i.e., it is tracelesstr(τ ) = 0. In this case the thermodynamic pressureP is the average isotropicstress, tr(σ) = −3P , which is not the hydrostatic pressure (see section 3.2 formore details). The velocity-dependent stress tensorτ is thus also the deviatoricstress tensor.

As ∇ · (P I) = ∇P , momentum conservation (14), in terms of pressure anddeviatoric stresses is

ρDv

Dt= −∇P + ∇ · τ + F. (16)

This equation is called the Navier-Stokes equation (Navier, 1785-1836; Stokes,1819-1903) when the stress tensor is linearly related to thestrain rate tensor andthe fluid incompressible (see section 3.2).

2.3.2 Inertia and non-Galilean forces

In almost all studies of mantle dynamics the fact that the Earth is rotating is simplyneglected. It is however worth discussing this point. Let usdefine a referenceframe of vectorsei attached to the solid Earth. These vectors rotate with the Earthand with respect to a Galilean frame such that

dei

dt= ω × ei, (17)

whereω is the angular velocity of Earth’s rotation. A point of the Earth, X =∑

iXiei, has a velocity in the Galilean frame(

dX

dt

)

Gal.

=∑

i=1,3

[(

dXi

dt

)

ei +Xi

(

dei

dt

)]

= vEarth + ω ×X (18)

wherevEarth is the velocity in the Earth’s frame, and by repeating the derivation,an accelerationγGal. in a Galilean frame (Galileo, 1564-1642)

γGal = γEarth + 2ω × v + ω × (ω × X) +dω

dt× X. (19)

In this well known expression, one recognizes on the right side, the accelerationin the non-Galilean Earth reference frame, the Coriolis, (Coriolis, 1792-1843),centrifugal and Poincare accelerations (Poincare, 1854-1912).

To quantify the importance of the three first acceleration terms (neglecting thePoincare term), let us consider a characteristic length scale (the Earth’s radius,a = 6371 km), and mantle velocity (the maximum plate tectonic speed,U = 10cm yr−1) and let us compare the various acceleration terms. One immediately gets

inertia

Coriolis=

U

2ωa=

1

2.9 × 1011, (20)

10

Page 11: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Coriolis

gravitational force=

2ωU

g=

1

2.1 × 1013, (21)

centrifugal

gravitational force=ω2a

g=

1

291. (22)

Thus, the inertial term is much smaller than the Coriolis term (this ratio is alsoknown as the Rossby number, (Rossby, 1898-1957)), which is itself negligible rel-ative to gravitational force. Even if we argue that a more meaningful comparisonwould be between the whole Coriolis force2ρωU and the lateral variations of thegravitational forceδρg, (this ratio would be the inverse of the Eckman number,(Eckman, 1874-1954)), inertia and Coriolis accelerationsstill play a negligiblerole in mantle dynamics. Neglecting inertia means that forces are instantaneouslyin balance and that changes in kinetic energy are negligiblesince inertia is thetime-derivative of the kinetic energy. We can perform a simple numerical esti-mate of the mantle kinetic energy. The kinetic energy of a lithospheric plate (asquare of size 2000 km, thickness 100 km, velocity 5 cm yr−1 and density 3000kg m−3) is 1.67 kJ, which is comparable to that of a middle size car (2000 kg)driven at only 4.65 km h−1!

The centrifugal term is also quite small but not so small (1/291 of gravitationalforce). It controls two effects. The first is the Earth’s flattening with an equatorialbulge of 21 km (1/300 of Earth’s radius) which is a static phenomenon that hasno interactions with convective dynamics. The second effect is the possibilitythat the whole planet rotates along an equatorial axis in order to keep its maininertial axis coincident with its rotational axis (Spada et al., 1992a;Ricard et al.,1993b). This rotational equilibrium of the Earth will not bediscussed here (seee.g.,Chandrasekhar(1969) for the static equilibrium shape of a rotating planetand e.g.,Munk and MacDonald(1960) for the dynamics of a deformable rotatingbody).

We neglect all the acceleration terms in the following but weshould rememberthat in addition to the convective motion of a non-rotating planet, a rotation of theplanet with respect to an equatorial axis, is possible. Thismotion documented bypaleomagnetism, is called True Polar Wander (Besse and Courtillot, 1991).

2.3.3 Angular momentum conservation

The angular momentum per unit massJ = X × v obeys a law of conservation.This law can be obtained in two different ways. First, as we did for mass andmomentum conservation, we can express the balance of angular momentum inintegral form. In the absence of intrinsic angular momentumsources, its variationsare due to

• advective transport of angular momentum across the surfaceΣ,

11

Page 12: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

• torque of forces acting on this surface,

• torque of internal body forces.

The resulting balance is therefore,∫

Ω

∂(ρJ)

∂tdV = −

ΣρJ(v · dS) +

ΣX × (σ · dS) +

ΩX × F dV . (23)

The only difficulty to transform this integral form into a local equation is with theintegral involving the stress tensor. After some algebra, equation (23) becomes

ρDJ

Dt= X × ∇ · σ + X × F + T , (24)

where the torqueT is the vector(τzy −τyz , τxz −τzx, τyx −τxy). A second expres-sion can be obtained by the vectorial multiplication of the momentum equation(15) byX. Since

X × Dv

Dt=DJ

Dt− DX

Dt× v =

DJ

Dt− v × v =

DJ

Dt(25)

we get,

ρDJ

Dt= X × ∇ · σ + X × F, (26)

which differs from (24) by the absence of the torqueT . This proves that in theabsence of sources of angular momentum, the stress (eitherσ or τ ) must be rep-resented by a symmetrical tensor,

σ = σt, τ = τ

t, (27)

where[ ]t denotes tensor transposition.

2.4 Energy conservation

2.4.1 First law and internal energy

The total energy per unit mass of a fluid is the sum of its internal energy,U , andits kinetic energy (this approach implies that the work of the various forces isseparately taken into account; another approach that we usein 2.5.3, adds to thetotal energy the various possible potential energies and ignores forces). In thefixed volumeΩ, the total energy is thus

Ωρ(U +

v2

2) dV . (28)

A change of this energy content can be caused by

12

Page 13: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

• advection of energy across the boundaryΣ by macroscopic flow,

• transfer of energy through the same surface without mass transport, by saydiffusion or conduction,

• work of body forces,

• work of surface forces,

• volumetrically distributed radioactive heat production

Using the divergence theorem, the balance of energy can therefore be written

∂t

(

ρ(U +v2

2)

)

= −∇ ·(

ρ(u+v2

2)v + q + Pv − τ · v

)

+F ·v+ρH, (29)

whereq is the diffusive flux,H the rate of energy production per unit mass andwhere the stresses are divided into thermodynamic pressureand velocity depen-dent stresses.

This expression can can be developed and simplified by using (11) and theequations of mass and momentum conservation, (9) and (16) toreach the form

ρDUDt

= −∇ · q − P∇ · v + τ : ∇v + ρH. (30)

The viscous dissipation termτ : ∇v is the contraction of the two tensorsτ and∇v (of components∂vi/∂xj). Its expression is

ij τij∂vi/∂xj .

2.4.2 State variables

The internal energy can be expressed in terms of the more usual thermodynamicstate variables, namely, temperature, pressure and volume. We use volume to fol-low the classical thermodynamics approach, but since we apply thermodynamicsto points in a continuous medium, the volumeV is in fact the volume per unitmass or1/ρ. We use the first law of thermodynamics which states that duringan infinitesimal process the variation of internal energy isthe sum of the heatδQ and reversible workδW exchanged. Although irreversible processes occur inthe fluid, we assume that we can adopt the hypothesis of a localthermodynamicequilibrium.

The increments in heat and work are not exact differentials:the entire preciseprocess of energy exchange has to be known to compute these increments, notonly the initial and final stages. Using either aT − V or T − P formulation, wecan write

δQ = CV dT + ldV = CPdT + hdP, (31)

13

Page 14: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

whereCP anCV are the heat capacities at constant pressure and volume, respec-tively, andh andl are two other calorimetric coefficients necessary to account forheat exchange at constant temperature. For fluids the reversible exchange of workis only due to the work of pressure forces

δW = −PdV. (32)

This implies that only the pressure term corresponds to an energy capable to bestored and returned without loss when the volume change is reversed. On the con-trary the stresses related to the velocity will ultimately appear in the dissipative,irrecoverable term of viscous dissipation. This point willbe further considered inthe section 3.2 about rheology.

Thermodynamics states that the total variations of energy,dU = δQ + δW ,enthalpy,dH = dU + d(PV ) or entropy,dS = δQ/T , are exact differentials andU , H andS are potentials. This means that the net change in energy (enthalpy,entropy) between an initial and a final state depends only on the initial and finalstates themselves, and not on the intermediate stages. Thisimplies mathematicallythat the second partial derivatives of these potentials with respect to any pair ofvariables are independent of the order of differentiation.Using these rules a largenumber of relations can be derived among the thermodynamic coefficients andtheir derivatives. These are called the Maxwell relations and are discussed inmost thermodynamics textbooks (e.g.,Poirier, 1991). We can in particular derivethe values ofl (starting fromdU anddS in T − V formulation) andh (startingfrom dH anddS in T − P formulation),

l = αTKT , and h = −αTρ. (33)

In these expressions forl andh, we introduced the thermal expansivityα and theisothermal incompressibilityKT ,

α =1

V

(

∂V

∂T

)

P

= −1

ρ

(

∂ρ

∂T

)

P

,

KT = − V

(

∂P

∂V

)

T

= ρ

(

∂P

∂ρ

)

T

.

(34)

The thermodynamic laws and differentials apply to a closed deformable vol-umeΩ(t). This corresponds to the perspective of Lagrange. We can thereforeinterpret the differential symbols ”d” of the thermodynamic definitions (31) of(32) as Lagrangian derivatives ”D”.

Therefore, in total, when the expressions forl andh are taken into account,(33), and when the differential symbols are interpreted as Lagrangian derivatives,

14

Page 15: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

the change of internal energy,dU = δQ+ δW can be recast as

DUDt

= CVDT

Dt+ (αTKT − P )

∇ · vρ

, (35)

orDUDt

= CPDT

Dt− αT

ρ

DP

Dt− P

∇ · vρ

. (36)

In these equations we also have replaced the volume variation using mass conser-vation (9)

DV

Dt=D(1/ρ)

Dt= − 1

ρ2

Dt=

∇ · vρ

. (37)

2.4.3 Temperature

We can now employ either thermodynamic relation (35) or (36), in our conserva-tion equation deduced from fluid mechanics, (30), to expressthe conservation ofenergy in terms of temperature variations

ρCPDT

Dt= − ∇ · q + αT

DP

Dt+ τ : ∇v + ρH,

ρCVDT

Dt= − ∇ · q − αTKT∇ · v + τ : ∇v + ρH.

(38)

Apart from diffusion, three sources of temperature variations appear on the rightside of these equations. The last termρH is the source of radioactive heat produc-tion. This term is of prime importance for the mantle, mostlyheated by the decayof radioactive elements like235U, 238U, 236Th and40K. All together these nuclidesgenerate about20 × 1012 W (McDonough and Sun, 1995). Although this numbermay seem large, it is in fact very small. Since the Earth now has about6 × 109

inhabitants, the total natural radioactivity of the Earth is only∼3 kW/person, notenough to run the appliances of a standard kitchen in a developed country. It isamazing that this ridiculously small energy source drives plate tectonics, raisesmountains and produces a magnetic field. In addition to the present-day radioac-tivity, extinct radionucleides, like that of36Al (with a half life of 0.73 Myr), haveplayed an important role in the initial stage of planet formation (Lee et al., 1976).

The viscous dissipation termτ : ∇v converts mechanical energy into a tem-perature increase. This term explains the classical Joule experiment (equivalencebetween work and heat, (Joule, 1818-1889)) in which the potential energy of aload (measured in joules), drives a propeller in a fluid, and dissipates the mechan-ical energy as thermal energy (measured in calories).

The remaining source term, containing the thermodynamic coefficients (α orαKT in (38)), cancels when the fluid is incompressible (e.g., whenα = 0 or when∇ ·v = 0). This term is related to adiabatic compression and will be discussed insection 4.2.2.

15

Page 16: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

2.4.4 Second law and entropy

We now consider the second law of thermodynamics and entropyconservation.Assuming local thermodynamic equilibrium, we havedU = TdS − PdV . Usingthe equation of conservation for the internal energyU , (30), and expressing thevolume change in terms of velocity divergence, (37), we obtain

ρTDSDt

= −∇ · q + τ : ∇v + ρH. (39)

To identify the entropy sources, we can express this equation in the form of aconservation equation (see (7)),

∂(ρS)

∂t= −∇ ·

(

ρSv +q

T

)

− 1

T 2q · ∇T +

1

Tτ : ∇v +

1

TρH. (40)

The physical meaning of this equation is therefore that the change of entropy isrelated to a flux of advected and diffused entropy,ρSv andq/T , and to threeentropy production terms, including from radiogenic heating.

A brief introduction to the general principles of non-equilibrium thermody-namics will be given in section 5.1.4. Here, we simply state that the second lawrequires that in all situations, the total entropy production is positive. When dif-ferent entropy production terms involve factors of different tensor orders (tensors,vectors or scalars), they must separately be positive. Thisis called the Curie prin-ciple (Curie, 1859-1906) (see e.g.,de Groot and Mazur(1984);Woods(1975)). Itimplies that

−q · ∇T ≥ 0, and τ : ∇v ≥ 0. (41)

The usual Fourier law (Fourier, 1768-1830) with a positive thermal conductivityk > 0,

q = −k∇T, (42)

satisfies the second law.When the conductivityk is uniform, the thermal diffusion term of the energy

equation−∇ ·q becomesk∇2T where∇2 = ∇ ·∇ is the scalar Laplacian oper-ator (Laplace, 1749-1827). Instead of a thermal conductivity, a thermal diffusivityκ can be introduced

κ =k

ρCP, (43)

(in principle isobaric and isochoric thermal diffusivities should be defined). Insituations with uniform conductivity, without motion and radioactivity sources,the energy equation (38) becomes the standard diffusion equation

∂T

∂t= κ∇2T. (44)

16

Page 17: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

The relation between stress and velocity will be discussed in detail in section3.2. We will show that the relationship

τ = 2η(

ǫ− 1

3∇ · v

)

(45)

is appropriate for the mantle, where the strain rate tensorǫ is defined by

ǫ =1

2

(

[∇v] + [∇v]t)

. (46)

Using this relation and assumingη uniform, the divergence of the stress tensorthat appears in the momentum conservation equation has the simple form

∇ · τ = η∇2v +η

3∇(∇ · v), (47)

where the vectorial Laplacien∇2v is ∇(∇ ·v)−∇× (∇×v). This relationshipsuggests a meaningful interpretation of the viscosity. Themomentum equation(16), can be written

∂v

∂t=η

ρ∇

2v + other terms.... (48)

forgetting the other terms, a comparison with the thermal diffusion equation (44)shows that the kinematic viscosityν, defined by

ν =η

ρ, (49)

should rather be called the momentum diffusivity; it plays the same role withrespect to the velocity as thermal diffusivity does with respect to temperature.

2.5 Gravitational forces

2.5.1 Poisson equation

In this chapter, the only force is the gravitational body force. The gravity is thesum of this gravitational body force and the centrifugal force already discussed(see 2.3.2). The gravitational force per unit mass is the gradient of the gravitationalpotentialψ, a solution of Poisson’s equation (Poisson, 1781-1840), i.e.,

g = −∇ψ, and ∇2ψ = 4πGρ, (50)

whereG is the gravitational constant. In the force term that appears in the mo-mentum equation (16),F = ρg, the gravitational force per unit mass should be inagreement with the distribution of masses: the Earth shouldbe self-gravitating.

17

Page 18: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

2.5.2 Self-gravitation

When dealing with fluid dynamics at the laboratory scale, thegravitational forcecan be considered as constant and uniform. The gravitational force is related tothe entire distribution of mass in the Earth (and the Universe) and is practically in-dependent of the local changes in density in the experimental environment. There-fore, at the laboratory scale, it is reasonable to ignore Poisson’s equation and toassume thatg is a uniform and constant reference gravitational field.

Inside a planet, the density can be divided into an average depth-dependentdensity,ρ0(r), the source of the reference depth-dependent gravitational field,g0(r), and a density perturbationδρ, the source a gravitational perturbationδg.The force term,F is therefore to first orderρ0g0 + δρg0 + ρ0δg; it is temptingto assume that each term in this expression is much larger than the next one andhopefully that only the first two terms are of importance (neglecting the secondterm would suppress any feed back between density perturbations and flow). Prac-tically, this assumption would imply consideration of the total density anomaliesbut only the depth dependent gravitational field. Solving Poisson’s equation tocompute the perturbed gravitational field would thus be avoided. We can test theabove idea and show that unfortunately, the third term,+ρ0δg, may be of thesame order as the second one (Ricard et al., 1984;Richards and Hager, 1984;Panasyuk et al., 1996). To perform this exercise we have to introduce the spher-ical harmonic functionsYlm(θ, φ). These functions of latitudeθ and longitudeφ oscillate on a sphere just like 2D sinusoidal functions on a plane. Each har-monic function changes signl −m times from North to South pole, andm timesover the same angle (180 degrees) around the equator. The degreel can thus beinterpreted as corresponding to a wavelength of order2πa/l wherea is radius.Spherical harmonics constitute a basis for functions defined on the sphere and arealso eigenfunctions of the angular part of Laplace’s equation which facilitates thesolution of Poisson’s equation.

Let us consider a density anomalyδρ = σδ(r − a)Ylm(θ, φ) at the surface ofa sphere of radiusa and uniform densityρ0 (δ(r − a) is the Dirac delta function(Dirac, 1902-1984),σ has unit of kg m−2). This mass distribution generates insidethe planet the radial gravitational perturbation field of

δg = 4πGσl

2l + 1

(

r

a

)l−1

Ylm(θ, φ), (51)

We can compare the termsρ0〈δg〉 and g0〈δρ〉, both averaged over the planetradius. For a uniform planet, the surface gravitational force per unit mass isg0 = 4/3πGρ0a. Since〈δρ〉 = σYlm(θ, φ)/a, we get

ρ0〈δg〉g0〈δρ〉

=3

2l + 1. (52)

18

Page 19: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

This estimate is certainly crude and a precise computation taking into account adistributed density distribution could be done. However this rule of thumb wouldremain valid. At low degree the effect of self-gravitationρ0δg is about 50% ofthe direct effectδρg0 and reaches 10% of it only nearl ∼ 15. Self-gravitation hasbeen taken into account in various models intended to explain the Earth’s gravityfield from mantle density anomalies (see also Ribe, this volume and Forte, Vol3). Some spherical convection codes seem to neglect this effect although it isimportant at the longest wavelengths.

2.5.3 Conservative forms of momentum and energy equations

In the general remarks on conservation laws in section 2.1, we wrote that con-served quantities like mass, momentum and energy can only betransported butdo not have production terms (contrary to entropy). However, in the momentumconservation (16) and in the energy conservation (29) two terms,ρg andρg · v,appear as sources (we also said that the radioactive termρH, appears because theclassical physics does not identify mass as energy and vice versa. A negligibleterm−ρH/c2, wherec is the speed of light, should, moreover, be present in themass conservation).

It is interesting to check that our equations can be recast into an exact con-servative form. An advantage of writing equations in conservative form is that itis appropriate to treat with global balances, interfaces and boundaries (see sec-tion 2.6). We can obtain conservative equations by using Poisson’s relation andperforming some algebra (usingg · ∇g = ∇g · g, sinceg = −∇ψ)

ρg = − 1

4πG∇ ·

(

g ⊗ g − 1

2g2I

)

(53)

ρg · v = −ρDψDt

− 1

4πG∇ ·

(

g∂ψ

∂t

)

− 1

8πG

∂g2

∂t(54)

If we substitute these two expressions in the momentum and the energy conserva-tion equations, (16) and (29), we obtain the conservative forms

∂(ρv)

∂t= −∇ ·

(

ρv ⊗ v + P I − τ +1

4πGg ⊗ g − 1

8πGg2 I

)

. (55)

∂t

(

ρ(U + ψ +v2

2) +

g2

8πG

)

=

−∇ ·(

ρv(U + ψ +v2

2) +q + Pv − τ · v +

g

4πG

∂ψ

∂t

)

+ ρH,

(56)

When the gravitational force is time independent, a potential ψ can simply beadded to the kinetic and internal energies to replace the work of gravitational

19

Page 20: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

forces. When gravitational force and its potential are time-dependent (due to massredistribution during convection, segregation of elements, etc), two new termsmust be added; a gravitational energy proportional tog2 and a gravitational fluxproportional tog∂ψ/∂t (this is equivalent to the magnetic energy proportional toB2 (whereB is the magnetic induction, in Tesla (Tesla, 1856-1943)), and to thePoynting vector of magneto-hydrodynamics (Poynting, 1852-1914)).

In a permanent or in a statistically steady regime, the time-dependent termsof energy equation (56) can be neglected and the equation canthen be integratedover the volume of the Earth. The natural assumption is that the Earth’s surfacethe velocities are perpendicular to the Earth’s surface normal vector and that thesurface is either stress-free or with no horizontal velocity (we exclude the caseof convection forced by imposing a non-zero surface velocity). Using the diver-gence theorem to transform the volume integral of the divergence back to a surfaceintegral of flux, most terms cancel and all that remains is

Σq · dS =

ΩρH dV . (57)

The surface flux in a statistically steady regime is simply the total radiogenic heatproduction.

It is surprising at first, that viscous dissipation does not appear in this balance.To understand this point, we can directly integrate the energy equation written interms of temperature (38),

ΩρCP

DT

DtdV =

−∫

Σq · dS +

Ω

(

αTDP

Dt+ τ : ∇v

)

dV +∫

ΩρH dV .

(58)

On the right side, the first and last terms cancel each other out by (57). On the leftside, we can use (11) to replaceρDT/Dt by ∂(ρT )/∂t+ ∇ · (ρvT ). The modestassumptions thatCP is a constant and that the temperature is statistically constantlead to

Ω

(

αTDP

Dt+ τ : ∇v

)

dV = 0. (59)

The total heat production due to dissipation is balanced by the work due to due tocompression and expansion over the convective cycle (Hewitt et al., 1975). Thisbalance is global not local. Dissipation occurs mostly nearthe boundary layersof the convection and compressional work is done along the downwellings andupwellings of the flow.

20

Page 21: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

2.6 Boundary and interface conditions

2.6.1 General method

A boundary condition is a special case of an interface condition when certainproperties are taken as known on one side of the interface. Sometimes the prop-erties are explicitly known (e.g. the three velocity components are zero on a noslip surface), but often an interface condition simply expresses the continuity ofa conserved quantity. To obtain the continuity conditions for a quantityA, thegeneral method is to start from the conservation equation ofA in its integral form(see (5)). We choose a cylindrical volumeΩ (a pill-box) of infinitely small radiusR where the top and bottom surfaces are located at a distance±ǫ from a disconti-nuity surface (see Figure 2). We choose two Cartesian axisOx andOy, we callnthe upward unit vector normal to the interface andt, a radial unit vector, normalto the cylindrical side of the pill-box, andθ is the angle betweent andOx. Ifwe now make the volumeΩ(ǫ) shrink to zero by decreasingǫ at constantR, thevolume integrals of the time dependent and the source terms will also go to zero(unless the source term contains explicit surface terms like in the case of surfacetension, but this is irrelevant at mantle scales). Since thesurface of the pill-boxΣ(ǫ) remains finite we must have

limǫ→0

Σ(ǫ)JA · dS = 0, (60)

(the demonstration is here written for a vector flux, but is easily extended to tensorflux). This condition can also be written

πR2[JA] · n +R∫ +ǫ

−ǫ

∫ 2π

0J · t dzdθ (61)

where[X] is the jump ofX across the interface, sometimes notedX+ −X−. Inmost cases, the second term goes to zero withǫ because the components ofJ arebounded, or is exactly zero when the flux is not a function ofθ (since the doubleintegral becomes the product of an integral inz times

∫ 2π0 tdθ = 0). In these cases,

the boundary condition forA becomes

[JA] · n = 0. (62)

At an interface, the normal flux ofA must therefore be continuous. Howeverin some cases, e.g., whenJ varies withx andy but contains az-derivative, thesecond term may not cancel and this happens in the case of boundaries associatedwith phase changes.

21

Page 22: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

0 0.5 10

0.5

1

n

x

R

y

Figure 2: The pillbox volume used to derive the interface andboundary condi-tions.

2.6.2 Interface conditions in the 1D case and for bounded variables

Using the mass, momentum, energy and entropy conservationsin their conser-vative forms in (9), (55), (56) and (40) and assuming for now that no variablebecomes infinite at an interface, the interface conditions in the reference framewhere the interface is motionless are

[ρv] · n = 0,

[τ ] · n− [P ]n = 0,

[ρvU + q − τ · v + Pv] · n = 0,

[ρvS +q

T] · n = 0,

(63)

(the gravitational force per unit mass and its potential arecontinuous). In theseequations, we neglected the inertia and the kinetic energy terms in the secondand third equations of (63) as appropriate for the mantle. When these terms areaccounted for (adding[−ρv ⊗ v] · n to the second equation and[ρvv2/2] · n tothe third), these equations are known as Hugoniot-Rankine conditions (Hugoniot,1851-1887; Rankine, 1820-1872).

At the surface of a fluid, and on any impermeable interfaces wherev · n = 0,the general jump conditions (63) without inertia, imply that the heat flux,[q]·n, the

22

Page 23: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

entropy flux[q/T ]·n (and therefore the temperatureT ) and the stress components[τ ] · n − [P ]n are continuous. In 3-D, four boundary conditions are necessary ona surface to solve for the three components of velocity and for the temperature.The temperature (or the heat flux) can be imposed and, for the velocity and stress,either free slip boundary conditions (v · n = 0, which is the first condition of(63)and(τ · n) × n = 0), or no slip boundary conditions (v = 0), are generally used.

2.6.3 Phase change interfaces

Mantle minerals undergo several phase transitions at depthand at least two ofthem, the olivinewadsleyite and the ringwooditeperovskite+magnesiowustitetransitions around 410 and 660 km depth, respectively, are sharp enough to bemodeled by discontinuities. Conditions (63) suggest that[ρv] ·n = 0 and[τ ·n−Pn] = 0 and these seem to be the conditions used in many convection models.However as pointed by (Corrieu et al., 1995), the first condition is correct, not thesecond one. The problem arises from the term in∂vz/∂z present in the rheologicallaw (100) that becomes infinite when the material is forced tochange its densitydiscontinuously. To enforce the change in shape that occurslocally, the normalhorizontal stresses have to become infinite and therefore their contributions to theforce equilibrium of a pill-box do not vanish when the pill-box height is decreased.

To derive the appropriate interface condition we have to consider again (61)whereJA is substituted byσ. The only terms may be unbounded on the interfaceareσxx, σyy andσzz. Omitting the other stress components, that would make nocontribution to the interface condition whenǫ goes to zero, the stress continuitybecomes

πR2[σ] ·n+exR∫ +ǫ

−ǫ

∫ 2π

0σxx cos θ dzdθ+Rey

∫ +ǫ

−ǫ

∫ 2π

0σyy sin θ dzdθ = 0. (64)

SinceR is small we can replace the stresses on the cylindrical side of the pill-box by their first order expansions, e.g.,σxx = σxx(o) + (∂σxx/∂x)R cos θ +(∂σxx/∂y)R sin θ and perform the integration inθ. After simplification byπR2,one gets

[σ] · n + ex∂

∂x

∫ +ǫ

−ǫσxx dz + ey

∂y

∫ +ǫ

−ǫσyy dz = 0. (65)

This expression already demonstrates the continuity ofσzz. Usingσxx = σzz +2η∂vx/∂x − 2η∂vz/∂z, and assuming that the viscosity remains uniform, we seethat

limǫ→0

∫ ǫ

−ǫσxx dz = −2η lim

ǫ→0

∫ ǫ

−ǫ

∂vz

∂zdz = −2η[vz]. (66)

The same result holds for theσyy term. Sincevz is discontinuous, forcing a suddenchange in volume implies a discontinuity of the tangential stresses. The boundary

23

Page 24: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

conditions are thus

[τxz]−2η∂

∂x[vz] = [τyz ]−2η

∂y[vz ] = [τzz−P ] = [vx] = [vy] = [ρvz] = 0. (67)

When the kinetic energy is neglected, and the viscous stresses are much smallerthan the pressure term, which are two approximations valid for the mantle, the lasttwo boundary conditions are, assuming continuity of temperature,

[ρv(U +P

ρ)] · n + [q] · n =0,

[ρvS] · n +1

T[q] · n =0.

(68)

The diffusive fluxq can be eliminated from these two equations. Sinceρv iscontinuous and remembering thatU+P/ρ is the enthalpyH, we simply recognizethe Clapeyron condition, which is latent heat release,

∆H = T∆S, (69)

where the enthalpy and entropy jumps,[H] and[S], were replaced by their moretraditional notations,∆H and∆S. The heat flux is discontinuous across an inter-face,

∆Hρv · n + [q] · n = 0, (70)

and the discontinuity amounts to the enthalpy released by the mass flux that hasundergone a chemical reaction or a phase change.

2.6.4 Weakly deformable surface of a convective cell

When a no slip condition is imposed at the surface, both normal and shear stressesare present at the boundary. These stresses, according to the second interfacecondition (63), must balance the force−τ · n + Pn exerted by the fluid. Thisis reasonable for a laboratory experiment with a fluid totally enclosed in a tankwhose walls are rigid enough to resist fluid traction. However in the case of freeslip boundary conditions, it may seem strange that by imposing a zero verticalvelocity, a finite normal stress results at the free surface.It is therefore worthdiscussing this point in more detail.

The natural boundary conditions should be that both the normal and tangentialstresses applied on the free deformable surface,z = h(x, y, t), of a convectivefluid are zero

(τ · n− Pn)on z=h = 0, (71)

(neglecting atmospheric pressure). In this expression thetopographyh is un-known and the normaln, computed at the surface of the planet isn = (ez −∇Hh)/

1 + |∇h|2 whereez is the unit vector alongz, opposite to gravity.

24

Page 25: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

The variation of topography is related to the convective flowand satisfies

∂h

∂t+ v0

H · ∇Hh− v0z = 0. (72)

This equation expresses the fact that a material particle onthe surface remainsalways on it. In this expressionv0

H andv0z are the horizontal and vertical velocity

components at the surface of the planet. We will see in section 4 that lateral pres-sure and stress variations are always very small compared tothe average pressure(this is because in most fluids, and in the mantle, the lateraldensity variationsremain negligible compared to the average density). This implies that the surfacetopography is not much affected by the internal dynamics andremains close tohorizontal, |∇Hh| << 1. Boundary condition (71) and topography advection(72) can therefore be expanded to first order to give

(τ · ez − Pez)on z=0 ≃ −ρ0g0hez , (73)

dh

dt= v0

z , (74)

where we again make use of the fact that the total stress remains close to hydro-static, i.e.,n ·τ ·n << P (ρ0 andg0 are the surface values of density and gravity).To first order, the stress boundary condition on a weakly deformable top surface istherefore zero shear stress but with a time dependent normalstress related to thesurface topography and vertical velocity.

The convection equations with these boundary conditions could be solved butthis is not always useful. Since the boundary conditions involve both displacementh and velocityv0

z , the solution is akin to an eigenvalue problem. It can be shownthat for an internal density structure of wavelengthλ, v0

z goes to zero in a time oforder η/ρ0g0λ whereη is the typical viscosity of the underlying liquid over thedepthλ (Richards and Hager, 1984). For the Earth, this time is the characteristictime of postglacial rebound and is typically a few thousand years for wavelengthsof a few thousand kilometers (e.g.,Spada et al., 1992b).

For convection, where the characteristic times are much longer, it is thus ap-propriate to assume that the induced topography is in mechanical equilibrium withthe internal density structure. A zero normal velocity can therefore be imposedand the resulting normal stress can be used to estimate the topography generatedby the convective flow. Internal compositional interfaces can be treated in a sim-ilar manner if they are only weakly deformable (i.e., when their intrinsic densityjumps are much larger than the thermal density variations).This is the case forthe core-mantle boundary (CMB).

For short wavelength structures and for rapid events (e.g.,for a localized ther-mal anomaly impinging the Earth’s surface), the time for topographic equilibra-tion becomes comparable to the time scale of internal convective processes. In

25

Page 26: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

this case the precise computation of a history-dependent topography is necessaryand the finite elasticity of the lithosphere, the coldest part of the mantle, plays animportant role (Zhong et al., 1996).

3 Thermodynamic and rheological properties

Section 2 on conservation equations is valid for all fluids (although the interfaceconditions are mostly discussed when inertia and kinetic energy are negligible).The differences between mantle convection and core, oceanic or atmospheric con-vection come from the thermodynamic and transport properties of solids that arevery different from those of usual fluids. We review some basic general propertiesof solids in this section 3 and will be more specific in the lastsection 6.

3.1 Equation of State and solid properties

The equation of state of any material (EoS) relates its pressure, density and tem-perature. The equation of state of a perfect gas,PV/T =constant, is well known,but irrelevant for solids. Unfortunately there is no equation for solids based ona simple and efficient theoretical model. In the Earth mineralogical community,the third order finite strain Birch-Murnaghan EoS seems highly favored (Birch,1952). This equation is cumbersome and is essentially empirical. More physicalapproaches have been used inVinet et al.(1987),Poirier and Tarantola(1998)andStacey and Davis(2004) but it seems that for each solid, the EoS has to beobtained experimentally.

In the simplest cases, the density varies aroundρ0 measured at temperatureT0

and pressureP0 as

ρ = ρ0

(

1 − α(T − T0) +P − P0

KT

)

, (75)

where the thermal expansivityα and incompressibilityKT have been defined in(34). This expression is a first order expansion of any EoS. Equation (75) can how-ever be misleading if one forgets that the parametersα andKT cannot be constantbut must be related through Maxwell relations (for example,their definitions (34)imply that∂(αρ)/∂P = −∂(ρ/KT )/∂T ).

Equation (75) can be used for a very simple numerical estimate that illus-trates an important characteristics of solid Earth geophysics. Typically for sili-catesα ∼ 10−5 K−1, KT ∼ 1011 Pa, while temperature variations in the mantle,∆T , are of a few 1000 K with a pressure increase between the surface and thecore,∆P , of order of1011 Pa. This indicates that the overall density variationsdue to temperature differences are negligible compared to those due to pressure

26

Page 27: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

differences (α∆T << 1 but ∆P/KT ∼ 1). In planets, to first order, the radialdensity is only a function of pressure, not of temperature. This is opposite to mostliquid or solid laboratory experiments, where the properties are usually controlledby temperature.

A very important quantity in the thermodynamics of solids isthe Gruneisenparameter (Gruneisen, 1877-1949)

Γ =αKT

ρCV=

1

ρCV

(

∂P

∂T

)

V

. (76)

The Gruneisen parameter is dimensionless, does not vary much through the mantle(around 1) and can reasonably be considered as independent of the temperature.An empirical law (Anderson, 1979) relatesΓ with the density,

Γ = Γ0

(

ρ0

ρ

)q

, (77)

whereq is around 1. The Gruneisen parameter can also be related to the mi-croscopic vibrational properties of crystals (Stacey, 1977). At high temperature,above the Debye temperature (Debye, 1884-1966), all solidshave more or lessthe same heat capacity at constant volume. This is called theDulong and Petitrule (Dulong, 1785-1838; Petit, 1791-1820). At highT , each atom vibrates andthe thermal vibrational energy is equipartioned between the three dimensions (de-grees of freedom) which leads toCV m = 3R per mole of atoms, independent ofthe nature of the solid (R is the gas constant). Assuming that the mantle is madeof pure forsterite Mg2SiO4 that contains 7 atoms for a molar mass of 140 g, itsheat capacity at constant volume is therefore close toCV m = 21R = 174.56 JK−1mol−1 or CV = 1247 J K−1kg−1. The approximate constancy ofCV and thefact thatΓ is only a function ofρ allow us to integrate (76)

P = F (ρ) + α0K0T (T − T0)

(

ρ

ρ0

)1−q

, (78)

whereα0 andK0T are the thermal expansivity and incompressibility at standard

conditions and whereF (ρ) is a density dependent integration constant. A rathersimple but acceptable choice for the functionF (ρ), at least for mantle dynami-cists, is the Murnaghan EoS (Murnaghan, 1951) at constantT that allows us towrite an EoS for solids of the form

P =K0

T

n

[(

ρ

ρ0

)n

− 1

]

+ α0K0T (T − T0)

(

ρ

ρ0

)1−q

, (79)

with an exponentn of order of 3. This equation could easily be used to deriveany thermodynamic property likeα(P, T ) or KT (P, T ). This equation has been

27

Page 28: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

used implicitly in various models of mantle convection (e.g., Glatzmaier, 1988;Bercovici et al., 1989a, 1992). An important consequence of this EoS assumingq ∼ 1 is thatαKT is more or less constant and that

KT ∼ K0T

(

ρ

ρ0

)n

, α ∼ α0

(

ρ

ρ0

)−n

. (80)

In the mantle, the incompressibility increases and the thermal expansion decreasessignificantly with depth. The geophysical consequences arefurther discussed insection 6.5.1.

Two other thermodynamic equalities can also be straightforwardly deduced bychain rules of derivatives and will be used in the following.A relation betweenthe two heat capacitiesCP andCV of the energy equations (38) can be derivedfrom the two expressions for heat increments, (31) and the definition of h, (33),

Cp − Cv =αT

ρ

(

∂P

∂T

)

V

. (81)

The same expressions for heat increments,(31) and thel andh definitions ,(33),imply that for an adiabatic transformation (whenδQ anddS are zero),

(

∂P

∂T

)

S

=ρCp

αT, and,

(

∂V

∂T

)

S

= − Cv

αKTT. (82)

The equations (81)-(82), take simpler forms when the Gruneisen parameter (76)and the adiabatic compressibility defined by

KS = ρ

(

∂P

∂ρ

)

S

, (83)

are used; they are,CP

CV=KS

KT= 1 + ΓαT. (84)

SinceΓ ∼ 1 and sinceαT << 1, the two heat capacities are basically equal. Itseems safer to assume thatCV is constant (the Dulong and Petit rule) and inferCP from it. The incompressibilityKS is defined similarly toKT but at constantentropy. The theory of elastic waves introduces this parameter that can be obtainedfrom seismic observations

KS = ρ(

v2p −

4

3v2

s

)

, (85)

wherevp andvs are thep andswave velocities. This important parameter providesa connection between geodynamics and seismology.

28

Page 29: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

3.2 Rheology

In section 2.3 on momentum conservation, no assumption is made on the rheologyof the fluid, i.e., on the relation between the stress tensor and the flow itself. Incontrast, the discussion of energy conservation, (section2.4) relies on the assump-tion that the pressure-related work is entirely recoverable, (32); as a consequence,the work of the deviatoric stresses ends up entirely as a dissipative term, hence asource of entropy. In a real fluid, this may be wrong for two reasons: part of thedeviatoric stresses may be recoverable and part of the isotropic work may not berecoverable. In the first case, elasticity may be present, inthe second case, bulkviscosity.

3.2.1 Elasticity

On a very short time scale, the mantle is an elastic solid in which compressionaland shear waves propagate (e.g.,Kennett(2001)). In an elastic solid, the linearstrain tensor,

ǫe =

1

2(∇u + [∇u]t), (86)

whereu is the displacement vector (this is valid for small deformations, (see e.g.,Malvern, 1969;Landau and Lifchitz, 2000) for the large deformation case) is lin-early related to the stress tensor,

σeij = Ae

ijklǫekl. (87)

whereAe is the fourth rank stiffness tensor. Since both the stress and the straintensors are symmetric and because of the Maxwell thermodynamic relations forinternal energy (including the elastic energy),∂2U/∂ǫij∂ǫkl = ∂2U/∂ǫkl∂ǫij , theelastic tensor is invariant to permutations ofi and j, k and l, ij andkl. Thisleaves in the most general case of anisotropy, 21 independent stiffness coefficients(Malvern, 1969). In crystals, this number decreases with the number of symme-tries of the unit cell. For isotropic elastic solids, only two parameters are needed,the incompressibilityK and the rigidityµR and the elastic behavior satisfies

σe = Ktr(ǫe)I + 2µR

(

ǫe − 1

3tr(ǫe)I

)

, (88)

wheretr(ǫe) = ∇ · u.Two remarks can be made on this rapid presentation of elasticity which are

more deeply developed in textbooks of mechanics (e.g.,Malvern, 1969;Landauand Lifchitz, 2000) or of seismology (e.g.,Dahlen and Tromp, 1998). First, theexpression (88) assumes that the displacement vector is computed from an ini-tial situation where the solid is perfectly stress-free, i.e., σe = 0 whenǫ

e = 0.

29

Page 30: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

In practical problems, only incremental displacements with respect to an ini-tial pre-stressed state are known andσ

e has to be understood as a variation ofthe stress tensor. Second, temperature variations are associated with changes inelastic stresses and the incompressibilityK takes these variations into account.The incompressibility should beK = KS for rapid adiabatic seismic waves andK = KT for isothermal variations. The other elastic parameters that are of-ten introduced, Poisson’s ratio, Young’s modulus (Young, 1773-1829), Lame’sparameters (Lame, 1795-1870), are simple functions of incompressibility andrigidity. Since the term proportional toµR is traceless, equation (88) leads to,tr(σe) = 3Ktr(ǫe), the rheology law can also be written in terms of compliance(i.e., gettingǫe as a function ofσe),

ǫe =

1

9Ktr(σe)I +

1

2µR

(

σe − 1

3tr(σe)I

)

. (89)

In these equations, the trace of the stress tensor can also bereplaced by the pres-sure definition

tr(σe) = −3P. (90)

The momentum equation (15) remains valid in a purely elasticsolid (exceptthat the advective transport is generally neglected,D/Dt ∼ ∂/∂t), but the dis-cussion of energy conservation and thermodynamics is different for elastic andviscous bodies. The work the elastic stress is entirely recoverable: a deformedelastic body returns to its undeformed shape when the external forces are released.The internal energy change due to the storage of elastic stress isδW = V σe : dǫe

instead ofδW = −PdV and, this is provided by the deformation work termτ : ∇v, which is therefore non dissipative. Thus, for an elastic body, the tem-perature equations (38) and the entropy equation (39) hold but with theτ : ∇v

source term removed.

3.2.2 Viscous Newtonian rheology

On a very long time scale, it is reasonable to assume that the internal deviatoricstresses become eventually relaxed and dissipated as heat.This is the assump-tion that we have implicitly made and that is usual in fluid mechanics. Since thedissipative term isτ : ∇v = τ : ǫv and must be positive according to the sec-ond law, this suggests a relationship between velocity-related stresses and velocityderivatives such that the total stress tensor has the form

σvij = −Pδij + Av

ijklǫvkl, (91)

δij being the Kronecker symbol (Kronecker, 1823-1891). Exceptfor the timederivative, the only formal difference between this expression and (87), is that

30

Page 31: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

pressure exists in a motionless fluid but is always associated with deformation inan elastic solid.

Using the same arguments as for the elastic case, the viscousrheology in theisotropic case can therefore be written in term of stiffness

σv = (−P + ζtr(ǫv)) I + 2η

(

ǫv − 1

3tr(ǫv)I

)

, (92)

wheretr(ǫv) = ∇ · v. Usingtr(σv) = 3(−P + ζtr(ǫv)), the rheology can alsobe expressed in term of compliance

ǫv =

1

9ζ(3P + tr(σv))I +

1

(

σv − 1

3tr(σv)I

)

. (93)

The two parametersη and ζ are positive according to the second law and arecalled the shear and bulk viscosities. When they are intrinsic material properties(i.e., independent of the flow itself), the fluid is called linear or Newtonian. Thehypothesis of isotropy of the rheology is probably wrong fora mantle composed ofhighly anisotropic materials (seeKarato, 1998) but only a few papers have tried totackle the problem of anisotropic viscosity (Christensen, 1997a;Muhlhaus et al.,2004).

Since tr(σv)/3 = −P + ζ∇ · v, the isotropic average of the total stress is notthe pressure term, unlessζ∇ · v = 0. Therefore, part of the stress work,τ : ǫ

v,during isotropic compaction could be dissipated in the formof the heat sourceζ(∇ · v)2. A density-independent bulk viscosity allows an infinite compressionunder a finite isotropic stress. The bulk viscosity parameter ζ is generally onlyintroduced to be immediately omitted and we will do the same.However, using(92) withζ = 0 but keeping∇ · v 6= 0 does not seem valid since it would removeall resistance to isotropic compression. We will see that consideringζ = 0 in (92)is formally correct although the real physical explanationis more complex: elasticstresses must be present to provide a resistance to isotropic viscous compression.The bulk viscosity, or some equivalent concept, is however necessary to handletwo phase compression problems (McKenzie, 1984;Bercovici et al., 2001a)(seesection 5.2).

3.2.3 Maxwellian Visco-elasticity

To account for the fact that the Earth behaves elastically onshort time constantsand viscously at long times, it is often assumed that under the same stress, thedeformation has both elastic and viscous components. By summing the viscouscompliance equation (93) with the time derivative of the elastic compliance equa-tion, (89) and in the case of an infinite bulk viscosityζ ,we get

ǫ =1

9Ktr(σ)I +

1

(

σ − 1

3tr(σ)I

)

+1

2µR

(

σ − 1

3tr(σ)I

)

, (94)

31

Page 32: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

whereσ = σv = σ

e andǫ = ǫv + ǫ

e. This time-dependent rheological law is theconstitutive law of a linear Maxwell solid.

A few simple illustrations of the behavior of a Maxwellian body will illustratethe physical meaning of equation (94) (see also Ribe, this volume). First, we canconsider the case where stress and strain are simple time-dependent sinusoidalfunctions with frequencyω (i.e., σ = σ0 exp(iωt) andǫ = ǫ0 exp(iωt)). Thesolution to this problem can then be used to solve other time-dependent problemsby Fourier or Laplace transforms. The equation (94) becomes

ǫ0 =1

9Ktr(σ0)I +

1

2µR(1 − i

ωτ)(

σ0 −1

3tr(σ0)I

)

, (95)

whereτ = η/µR is the Maxwell time, (1). This equation can be compared to(89), and shows that the solution of a visco-elastic problemis formally equivalentto that of an elastic problem with a complex elastic rigidity. This is called thecorrespondence principle.

We can also solve the problem of a purely 1D Maxwellian body (only σzz

and ǫzz are non zero), submitted to a sudden loadσzz = σ0H(t) (whereH isthe Heaviside distribution, (Heaviside, 1850-1925)), or to a sudden strainǫzz =ǫ0H(t). The solutions are, fort ≥ 0,

ǫ0 =1

3kµR

σ0 +1

3ησ0t, (96)

andσ0 = 3kµR exp(−k t

τ)ǫ0, (97)

respectively, withk = 3K/(3K+µR). In both cases, the instantaneous elastic re-sponse is followed the viscous flow. In the first case, the finite elastic deformationis followed by a steady flow. In the second case, the initial elastic stresses are thendissipated by viscous relaxation over a time constant,τ/k. This time constant isdifferent from the Maxwell time constant as both deviatoricand non-deviatoricstresses are present. For mantle material the timeτ/k would however be of thesame order as the Maxwell time constantτ (in the mid-mantle,K ∼ 2µR ∼ 200GPa).

From equation (94), we can now understand what rheology mustbe used for acompressible viscous mantle. For phenomena that occur on time constants muchlarger than the Maxwell time, the deviatoric stresses can only be supported by theviscosity. As a typical viscosity for the deep mantle is in the range1019−1022 Pa s(see sections 4 and 6), the appropriate Maxwell times are in the range 30 yr-30 kyr,much shorter than those of convection. By constrast, the isotropic stress remainsonly supported by elasticity in the approximation where thebulk viscosityζ is

32

Page 33: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

infinitely large. The appropriate rheology for mantle convection is therefore givenby

ǫ =1

9Ktr(σ)I +

1

(

σ − 1

3tr(σ)I

)

. (98)

This equation is simultaneously a rheology equation for thedeviatoric stress andan EoS for the isotropic stress. UsingP = −tr(σ)/3, the stress tensor verifies

σ = −P I + 2η

(

ǫ − P

3KI

)

. (99)

This equation is intrinsically a visco-elastic equation, that can be replaced by apurely viscous equation plus an EoS

σ = −P I + 2η(

ǫ − 1

3tr(ǫ)I

)

, (100)

tr(ǫ) =P

K. (101)

The equation (100) is therefore the appropriate limit of theequation (92) for slowdeformation, whenζ = +∞ and when isotropic compaction is resisted by theelastic stresses.

The use of a Maxwell visco-elastic body to represent the mantle rheology onshort time scale remains however rather arbitrary. Insteadof summing the elasticand viscous deformations for the same stress tensor, another linear viscoelasticbody could be obtained by partitioning the total stress intoelastic and viscouscomponents for the same strain rate. Basically instead of having the elasticityand the viscosity added like a spring and a dashpot in series (Maxwell rheology),this Kelvin-Voigt rheology would connect in parallel a viscous dashpot with anelastic spring (Kelvin, 1824-1907; Voigt, 1850-1919). Of course further degreesof complexity could be reached by summing Maxwell and Voigt bodies, in seriesor in parallel. Such models have sometimes be used for the Earth but the data thatcould support or dismiss them is scarce (Yuen et al., 1986).

3.2.4 Non-linear rheologies

Even without elasticity and bulk viscosity, the assumptionof a linear Newtonianrheology for the mantle is problematic. The shear viscositycannot be a directfunction of velocity since this would contradict the necessary Galilean invarianceof material properties. However the shear viscosity could be any function of theinvariants of the strain rate tensor. There are three invariants of the strain ratetensor; its trace (but tr(τ ) = 0, in the absence of bulk viscosity), its determinant

33

Page 34: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

and the second invariantI2 =√

ǫ : ǫ (where as in (30), the double dots denotetensor contraction). The viscosity could therefore be a function of det(ǫ) andI2.

The main mechanisms of solid state deformation pertinent for mantle condi-tions (excluding the brittle and plastic deformations) areeither diffusion creep ordislocation creep (seePoirier, 1991). In the first case, finite deformation is ob-tained by summing the migrations of individual atoms exchanging their positionswith crystalline lattice vacancies. In crystals, the average number of lattice va-canciesC varies with pressure,P and temperature,T , according to Boltzmannstatistics (Boltzmann, 1844-1906),

C ∝ exp(−PVRT

), (102)

(V is the atomic volume,R the gas constant). A mineral is composed of grains ofsized with an average concentration of lattice vacanciesC0. Submitted to a devi-atoric stressτ , a gradient of vacancies of order|∇C| ∝ (C0/d)(τV/RT ) appearsdue to the difference in stress regime between the faces in compression and thefaces in extension (τV << RT ) (see Poirier, 1991;Ranalli, 1995;Turcotte andSchubert, 1982;Schubert et al., 2001).. This induces the flux of atoms (numberof atoms per unit surface and unit time)

J ∝ DC0

d

τ

RT, (103)

whereD is a diffusion coefficient. This flux of atoms goes from the grain facesin compression to the grain faces in extension. Along the direction of maximumcompression, each crystal grain shortens by a quantityδd which corresponds to atotal transport ofd2δd/V atoms. These atoms can be transported in a timeδt bythe fluxJV across the grain of sectiond2 (with volume diffusionDV ). They canalso be transported by grain boundary fluxJh (with grain boundary diffusionDh)along the grains interfaces through a surfacehd (h being the thickness of the grainboundary), according to

d2δd

V∼ JDd

2δt, ord2δd

V∼ Jhdhδt. (104)

As ǫ = (δd/δt)/d, the previous equations lead to the stress-strain rate relationship

ǫ ∝ V

d2RT

(

Dv +Dhh

d

)

τ . (105)

This diffusion mechanisms lead to a Newtonian rheology but with a grain-sizedependence of the viscosity;η ∝ d2 for Nabarro-Herring creep with diffusioninside the grain (Nabarro, 1916-2006; Herring, 1914) andη ∝ d3 for Coble

34

Page 35: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

grain-boundary creep (Coble, 1928-1992). The viscosity isalso very stronglyT -dependent not so much because of the explicit factorT in (105), but becausediffusion is a thermally activated process,D ∝ exp(Edif/RT ), whereEdif is anactivation enthalpy of diffusion.

In the case of dislocation creep, lines or planar imperfections are present in thecrystalline lattice and macroscopic deformation occurs byslip motion along theseimperfections, called dislocations. Instead of the grain size d for diffusion creep,the mean spacingdd between dislocations provides the length-scale. This distanceis often found to vary as1/I2. Therefore, instead of a diffusion creep with a vis-cosity indn the resulting rheology is rather inI−n

2 and is also thermally activatedwith an activation energyEdis. Dislocation creep leads to a non-linear regimewhere the equivalent viscosity varies with the second invariant with a power−n,wheren is typically of order 2,

ǫ ∼ In2 exp(Edis/RT )τ , (106)

(this relationship is often written, in short,ǫ ∼ τm with a stress exponentm of

order 3 butτ m really meansIm−12 τ ).

In general, for a given stress and a given temperature, the mechanism withthe smallest viscosity (largest strain rate) prevails. Whether linear (grain size de-pendent), or non-linear (stress dependent), viscosities are also strongly dependentupon temperature, pressure, melt content, water content, mineralogical phase andoxygen fugacity (e.g.,Hirth and Kolhstedt(1996)). In section 6.3 we will furtherdiscuss the rheological mechanisms appropriate for the Earth.

4 Physics of convection

The complex and very general system of equations that we havediscussed insections 2 and 3, can be used to model an infinite number of mantle flow situations.Mantle flow can sometimes be simply modeled as driven by the motion of plates(some examples are discussed in Zhong et al., Parmentier andKing, this volume).It can also be induced by compositional density anomalies (some examples arediscussed in Tackley this volume and Forte, Vol 3). However,a fundamental causeof motion is due to the interplay between density and temperature and this is calledthermal convection.

The phenomenon of thermal convection is common to all fluids (gas, liquidand creeping solids) and it can be illustrated by simple experiments (see Davaille,this volume). The simplest can be done using water and an experimental setupcalled the shadowgraph method. Parallel light enters a transparent fluid put ina glass tank and is deflected where there are refractive indexgradients due totemperature variations in the fluid. A pattern of bright regions and dark shadows

35

Page 36: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

is formed on a screen put on the other side of the tank. From this shadowgraph thestructure of the temperature pattern can be qualitatively assessed (see examples ofshadowgraphs inTritton (1988)).

4.1 Basic balance

From a simple thought experiment on thermal convection, we can derive the basicdynamic balance of convection. Let us consider a volume of fluid, Ω, of char-acteristic sizea, in which there is a temperature excess∆T with respect to thesurrounding fluid. The fluid is subject to a gravityg, it has an average densityρand thermal expansivityα. The volumeΩ, because of its anomalous temperature,experiences an Archimedian force, or buoyancy (Archimede around 287-212 BC)given by

F = −c1a3ρα∆Tg, (107)

(c1 is a constant taking into account the shape ofΩ, e.g.,c1 = 4π/3 for a sphere).If the volumeΩ is in a fluid of viscosityη, it will sink or rise with a velocity givenby Stokes law (Stokes, 1819-1903)

vs = −c1c2a2ρα∆Tg

η, (108)

(c2 is a drag coefficient accounting for the shape ofΩ, i.e.,c2 = π/6 for a sphere).During its motion, the volumeΩ exchanges heat by diffusion with the rest of

the fluid and the diffusion equation (44) tells us that a time of order

td = c3ρCpa

2

k. (109)

is needed before temperature equilibration with its surroundings. During this time,the fluid parcel travels the distancel = vstd.

A natural indication of the possibility that the parcel of fluid moves can beobtained by comparing the distancel to the characteristic sizea. Whenl >> a,i.e., when the fluid volume can be displaced by several times its size, motion willbe possible. On the contrary whenl << a, thermal equilibration will be so rapidthat no motion will occur.

The conditionl >> a, whenvs andtd are replaced by the above expressions,depends on only one quantity, the Rayleigh numberRa

Ra =ρ2α∆Tga3CP

ηk=α∆Tga3

κν, (110)

in terms of which motion occurs whenRa >> 1 (assuming thatc1c2c3 ∼ 1).The Rayleigh number compares the driving mechanism (e.g., the Archimedian

36

Page 37: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

buoyancy) to the two resistive mechanisms, the diffusion ofheat, represented byκ (see (43)), and the diffusion of momentum, represented byν (see (49)).

This simple balance suggests that a large nondimensional numberRa favorsfluid motion. How largeRa needs to be, is a question that we cannot address atthis moment but it will be discussed in section 4.4. Convection lifts hot fluid andcauses cold fluid to sink (assumingα > 0, which is true for most fluids and for themantle). A convective system will rapidly reach an equilibrium where all thermalheterogeneities are swept up or down (ifRa is large) or thermally equilibrated (ifRa is small), unless a forcing mechanism continuously injectsnew cold parcelsat the top and new hot parcels at the bottom. This can be done bycooling thetop surface or heating the bottom one. When a fluid is heated from the side, alateral temperature anomaly is constantly imposed and the liquid lateral thermalequilibration is prevented. The fluid remains in motion regardless of the amplitudeof the imposed temperature anomaly.

4.2 Two simple solutions

4.2.1 The diffusive solution

Trying to directly and exactly solve the mass, momentum, energy and Poisson’sequations and accounting for a realistic EoS would certainly be a formidable task.This complex system of equations has however two rather obvious but oppositesolutions.

A steady and motionless solution is indeed possible. The assumption∂/∂t =0 andv = 0 satisfies the mass equation (9), the momentum equation (16) whenthe pressure is hydrostatic,

0 = −∇P + ρg, (111)

and the energy equation (38) when the temperature is diffusive (using the Fourierlaw (42)),

∇ · (k∇T ) + ρH = 0. (112)

Solving analytically for the hydrostatic pressure and the diffusive temperature istrivial whenH, k andρ are uniform. For example, choosing a depthz positivedownward, we get

P = ρgz, T = T0 + ∆Tz

h+

1

2ρHz(h− z), (113)

across a conductive solution withT = T0, P = 0 at z = 0 andT = T0 + ∆T atz = h. Computing analytically the conductive solution remains feasible, but couldbe quite cumbersome if one introduces a realistic EoS and computes gravity inagreement with the density distribution using the Poisson equation (50). In section

37

Page 38: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

4.4 we will understand why the fluid does not necessarily choose the diffusivesolution.

4.2.2 The adiabatic solution

The previous diffusive solution was obtained for a steady motionless situation.However, the opposite situation where the velocities are very large, also corre-sponds to a rather simple situation. The energy equation (38) can also be written

ρCV T

(

D lnT

Dt− Γ

D ln ρ

Dt

)

= −∇ · q + τ : ∇v + ρH, (114)

or

ρCPT

(

D lnT

Dt− α

ρCP

DP

Dt

)

= −∇ · q + τ : ∇v + ρH. (115)

The right sides of these equations were previously shown to be equal toρTDS/Dtin (39). If we decrease the viscosity in a fluid, the convective velocity increases.The advection terms,v ·∇T , v ·∇ ln ρ andv ·∇P , become then much larger thanthe time dependent, diffusion and radioactive production terms. With a low vis-cosity, the fluid becomes also unable to sustain large stresses. As a consequence,when convection is vigorous enough, the fluid should evolve toward a situationwhere

(∇ lnT )S− Γ (∇ ln ρ)

S= 0,

(∇ lnT )S− α

ρCP(∇P )

S= 0.

(116)

Since such equations imply that the entropy is exactly conserved,DS/Dt = 0,this equilibrium is called the adiabatic equilibrium. We added a subscript[ ]S todenote the isentropic state.

Notice that, since the Gruneisen parameter is only density-dependent, see (77),density and temperature are simply related along the adiabat. For exemple, ifthe Gruneisen parameter is a constant,Γ0 (usingq = 0 in (77)), the first of theequations (116) implies

T = T0

(

ρ

ρ0

)Γ0

, (117)

whereT0 andρ0 are two reference values. This equation implies that the adiabatictemperature increases by a factor 1.72 (e.g., from 1300 K to 2230 K) from theasthenosphere (ρ0 ∼ 3200 kg m−3) to the CMB (ρ ∼ 5500 kg m−3), if we assumeΓ0 = 1.

38

Page 39: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

4.2.3 Stability of the adiabatic gradient

When a fluid is compressed, it heats up and it cools down when decompressed.This is the same physics that explains why atmospheric temperature decreaseswith altitude. Of course this adiabatic effect is vanishingly small in laboratoryexperiments, but not always in nature.

If a parcel of fluid is rapidly moved up or down alongz by a distancea,it changes its temperature adiabatically, by the quantitya(dT/dz)S. Howeverthe surrounding fluid will be at the temperaturea(dT/dz) wheredT/dz is justthe temperature gradient, not necessarily adiabatic, of the fluid at rest. We candefine∆Tna as the non-adiabatic temperature∆Tna = a(dT/dz − (dT/dz)S).The parcel being warmer or colder than the surroundings willrise or sink with aStokes velocity that, rather than (108) will be of order

vs = −c1c2a2ρα∆Tnag

η. (118)

Sincez is depth,dz is positive alongg, the adiabatic gradient is positive and thefluid parcel locally unstable when the gradient in the surrounding fluid is larger(superadiabatic) than the adiabatic gradient. On the contrary a sub-adiabatic gra-dient is stable with respect to convection. It is therefore not the total temperaturedifference between the top and bottom of the fluid that drivesmotion, but only itsnon-adiabatic part.

To compare the Stokes velocity with the thermal equilibration time, we needto introduce a modified Rayleigh number

Ra =α∆Tnag0a

3

κν. (119)

This number is based on the non-adiabatic temperature difference in excess of theadiabatic variation imposed over the heighta.

We have shown that inside a convective cell, the thermal gradient should besuperadiabatic. Superadiabaticity is the source of convective instability, but vig-orous convective stirring involves largely rapid adiabatic vertical motion; so muchof convecting fluid is indeed adiabatic while most of superadiabaticity is boundup in the thermal boundary layers where the vertical motion goes to zero. Thismechanism suggests that an adiabatic reference backgroundshould not be such abad assumption for a convective fluid.

This adiabaticity hypothesis should, however, not be takentoo literally (Jean-loz and Morris, 1987). In most numerical simulations, the resulting averagedgeotherm can be far (a few hundred K) from adiabatic (Bunge et al., 2001). First,radioactive heating, dissipation and diffusion are never totally negligible, second,even if each fluid parcel follows its own adiabatic geotherm,the average geothermmay not correspond to any particular adiabat.

39

Page 40: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

4.3 Approximate equations

4.3.1 Depth dependent reference profiles

We assume that the thermodynamic state is not far from an hydrostatic adiabat,thus we choose this state as a reference and rewrite the equations of fluid dy-namics in term of perturbations to this state (see alsoJarvis and Mckenzie, 1980;Glatzmaier, 1988;Bercovici et al., 1992). We denote all the reference variableswith an overbar. We choose a reference hydrostatic pressuregiven by

∇P = ρg, (120)

and adiabatic temperature and densities obeying (116)

∇T =αg

CpT

∇ρ =αg

CpΓρ,

(121)

where all the parameters are computed along the reference geotherm and whereghas been solved using Poisson’s equation (50) with the reference densityρ.

The reference parameters are depth-dependent and usually,even for a simpleEoS cannot be analytically obtained. They can however be computed numericallyfrom (116) assuming that the Gruneisen parameter is onlyρ-dependent. Using theEoS (79), all the thermodynamic quantities become functions of depth only, sothat the reference profile can be obtained by quadratures.

4.3.2 Nondimensionalization

As a principle, the validity of equations cannot depend on the units in which thequantities are expressed. The laws of physics can only relate dimensionless com-binations of parameters (e.g.,Barenblatt, 1996). This is fundamental in fluid dy-namics where a large number of quantities appears in the equations (see also Ribe,this volume). A necessary starting point is therefore to rephrase any fluid dynam-ics problem involvingN dimensional parameters in term ofM dimensionlessquantities (N −M ≤ 0 is the number of independent physical dimensions of theproblem).

One cannot perform the nondimensionalization of the convection equationsusing the variable reference profiles. We must therefore introduce constant param-eters, with indices[ ]0, corresponding to some typical or mantle-averaged valuesof the depth dependent reference values. We introduce for exampleα0, ρ0, g0 orC0

p .The adiabatic equations (121) impose the natural scale for temperature varia-

tion, Cp/(α||g||). This scale varies with depth but should not be too differentfrom

40

Page 41: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

the value computed with the constant parameters (with indices 0). We thereforeintroduce a nondimensional number, the dissipation numberD0,

D0 =α0g0a

C0p

, (122)

that compares the natural scale of temperature variations with the layer thickness,a. The dissipation numberD0 is around 0.5 for the Earth’s mantle. For anylaboratory experiment this number would be infinitely small, only geophysicalor astrophysical problems have large dissipation numbers.The hydrostatic andadiabatic reference profiles satisfy approximately

dT

dz∼ D0

T

a, and

dz∼ D0

Γ0

ρ

a. (123)

A zero dissipation number leads to uniform reference temperature and pressure,as the adiabatic compression effects are not large enough toaffect these quantities.

From top to bottom, the reference temperature increases adiabatically by∆TSwhile a total temperature jump∆TS + ∆Tna is imposed. Only the excess nonadiabatic temperature∆Tna is really useful to drive convection. The term drivingconvective instability is thus the dimensionless quantityǫ,

ǫ = α0∆Tna, (124)

which is always a small number (all the necessary numerical values are listed inTable 1).

All these preliminaries have been somewhat lengthy but we are now readyto nondimensionalize the various equations. Then, we will get the approximateequations, (Jarvis and Mckenzie, 1980), by simply taking into account thatǫ << 1(anelastic equations) or more crudely that bothǫ << 1 andD0 << 1 (the so-called Boussinesq equations, Boussinesq, 1842-1929). Themathematical formu-lation is heavy because of the number of symbols with or without overbars, tildesor indices, but is straightforward. The reader may jump directly to the results ofsection 4.3.3.

To nondimensionalize the equations we can use the quantities a and∆Tna;we also need a characteristic velocity and pressure. Following our discussion ofthe basic force balance in section 4.1, we use a typical Stokes velocity,V0 =a2g0ρ0α0∆Tna/η0, time a/V0, and pressureP0 = η0V0/a. Using the definitions

41

Page 42: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

of D0, (122),Γ0, (76) andǫ, (124), we perform the following change of variables

v =ǫaK0

T

η0

D0

Γ0

C0P

C0V

v,

P =P + ǫ K0T

D0

Γ0

C0P

C0V

P ,

α0T =α0T + ǫ T ,

∇ =1

a∇,

∂t=ǫ

K0T

η0

D0

Γ0

C0P

C0V

∂t

(125)

The EoS (75) can then be expanded at first order for a thermodynamic state closeto the hydrostatic adiabatic reference

ρ = ρ+ρ

KT

(P − P ) − αρ(T − T ). (126)

After nondimensionalization, the EoS becomes

ρ = ρ

(

1 + ǫ (K0

T

KT

D0

Γ0

C0P

C0V

P − α

α0

T )

)

. (127)

As C0P ∼ C0

V , KT ∼ K0T , α ∼ α0, P ∼ T ∼ 1 and ǫ << 1 it shows that

the density remains close to the reference profile within terms of orderǫ, ρ =ρ(1 +O(ǫ)).

4.3.3 Anelastic approximation

To approximate the equations of convection we perform the change of variables,(125), and use the fact thatǫ is small (see alsoSchubert et al., 2001). The choiceof the reference state leads to the cancellation of the termsO(1) of the mass,momentum and energy equations, (9), (16) and (38).

At orderO(ǫ), the first terms that remain in each equation are

∇ ·(

ρ

ρ0

v

)

= 0.

Ra

Pr

Dv

Dt= −∇P + ∇ · τ +

ρ

ρ0

g

g0

K0T

KT

D0

Γ0

C0P

C0V

P − ρ

ρ0

g

g0

α

α0

T .

ρ

ρ0

CP

C0P

DT

Dt=

1

Ra∇ ·

[

k

k0∇

(

T

∆Tna+ T

)]

α0

ρ

ρ0

g

g0D0T vg

ρ0

1

Ra

ρ0Ha2

k0∆T+D0τ : ∇v

(128)

42

Page 43: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

In the last equation,vg is the component of the velocity field is along the radialreference gravityg ·v = gvg. This new set of equations constitutes the equation offluid dynamics in the anelastic approximation. The term “anelastic” comes fromthe fact that the propagation of sound waves is impossible since the term in∂ρ/∂tis neglected.

In equations (128), we introduced the Rayleigh,Ra, and Prandtl,Pr, numbers

Ra =α0∆Tnaρ

20g0a

3C0P

η0k0=α0∆Tnag0a

3

ν0κ0, (129)

Pr =η0C

0P

k0=ν0

κ0. (130)

The physical meaning of the Rayleigh number as a measure of convective vigorhas already been discussed. The Prandtl number (Prandtl, 1975-1953) comparesthe two diffusive processes: namely the diffusion of momentum and heat.

In the momentum equation,τ = τa/(η0V0) is the nondimensionalized stresstensor which in the Newtonian case (without bulk viscosity), becomes

τ =η

η0(∇v + [∇v]t) − 2

3

η

η0∇ · v. (131)

The formalism is already so heavy that we have not included the self-gravitationalterm. This term is not negligible at long wavelengths (see section 2.5.2). To ac-count for this term we should have added on the right side of second equationof (128) a term−ρ/ρ0∇ψ, where the perturbed gravitational potential due to thedeparture of the density from the reference profile, satisfies Poisson’s equation

∇2ψ = 3

ρ

< ρ >

(

K0T

KT

D0

Γ0

C0P

C0V

P − α

α0T

)

, (132)

where< ρ > is the average density of the Earth.

4.3.4 Dimensionless numbers

The ratioRa/Pr is also called the Grashof numberGr = αg∆Tnaa3/ν2 (Grashof,

1826-1893). This number can also be written asV0a/ν and could be called theReynolds number,Re, of the flow (Reynolds, 1842-1912). Using one or the othernames depends on the quantities that are best known. For example, if the ve-locity V is a parameter imposed by a boundary condition, using it to performthe nondimensionalization and speaking in terms of Reynolds number would bemore natural than using the Grashof number. If a thermal structure is imposedby a velocity boundary condition (for example by the thickening of the oceanic

43

Page 44: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

lithosphere with age), it would seem natural to introduce a Peclet numberV a/κ(Peclet, 1973-1857), which is nothing more than the Rayleigh number of the flowif the velocity is imposed by the internal dynamics.

The Rayleigh and Prandtl numbers can be estimated in different ways. In factthe only difficult parameter to know is the viscosity. In mosttextbooks the valueof 1021 Pa s, first proposed byHaskell (1937), is given with a unanimity thathides very large uncertainties and most probably a large geographical variability.The mantle viscosity and its depth dependence can be constrained by post-glacialrebound, geoid, true polar wander, change of flattening of the Earth, and plateforce balance models or extrapolated from laboratory measurements. An increaseof viscosity with depth, between one or two orders of magnitudes is likely, with anasthenosphere significantly less viscous (1019 Pa s) at least under oceanic platesand a lower mantle probably around1022 Pa s (see details in section 6.1). It isimpossible to give justice to all the papers on this subject but some geodynamicestimates of mantle viscosity can be found in e.g.,Peltier (1989);Sabadini andYuen(1989);Lambeck and Johnston(1998) orRicard et al.(1993a). Whateverthe value of the real viscosity, the ratioRa/Pr is so small that inertia plays norole in the mantle and the left side of the momentum equation in the anelasticapproximation (128) can safely be set to zero (see numericalvalues in Table 1).

We can also introduce the Rayleigh and Prandtl number in the EoS (127) anddefine a new number,M ,

ρ = ρ

(

1 +K0

T

KTRaPrM2P − α

α0T ǫ

)

. (133)

This number is the Mach numberM = vD/vM , ratio of the velocity of thermal

diffusion, κ0/a, to the velocityvM =√

KT/ρ (Mach, 1838-1916). This last

velocity is very close to the bulk velocityvφ =√

KS/ρ (see (85)). The anelas-tic approximation is sometimes called small Mach number approximation. TheEarth’s Mach number is indeed of order 10−15 since thermal diffusion is muchslower than the sound speed. However the anelastic approximation really requiresa smallRaPrM2 and this quantity is justǫD0C

0P/Γ0C

0V , i.e., of orderǫ = 10−2

(Bercovici et al., 1992). A planet could have a very low Mach number but so largea Prandtl number that the anelastic approximation would notbe valid. Similarlya planet can have a low Reynolds number (creeping convection) with a very largeRayleigh number (chaotic convection).

4.3.5 Boussinesq approximation

As was expected from section 2.5.3, where we had shown that the dissipation andthe adiabatic terms balance each other in a statistical steady-state regime, these

44

Page 45: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

terms are proportional to the same dissipation numberD0. AlthoughD0 is not sosmall, most of the physics of mantle convection, except for the additional adiabatictemperature gradient, is captured with models whereD0 is arbitrarily set to zero.The fact that the Boussinesq approximation captures most ofthe mantle dynamicsis due to the fact that the mantle viscosity is so high that a huge∆Tna is main-tained across the top and bottom boundary layers. The other sources and sinks ofenergy and density anomaly (viscous dissipation and adiabatic heating) are there-fore always small (Jarvis and Mckenzie, 1980;Glatzmaier, 1988;Bercovici et al.,1992;Tackley, 1996).

In the Boussinesq approximation,D0 = 0, the reference density and tem-perature become constants according to (123) (when the reference temperature isuniform we can also chooseCP = CV = CP = CV ). The non adiabatic tem-perature increase∆Tna becomes simply the total temperature increase∆T . Thisapproximation is of course excellent for laboratory scale experiments where ef-fectivelyD0 << 1. The EoS (127) indicates that the density is only a function oftemperature and the fluid dynamics equations become

∇ · v =0,

Ra

Pr

Dv

Dt= − ∇P + ∇ ·

(

η

η0(∇v + [∇v]t)

)

− g

g0

α

α0T ,

DT

Dt=

1

Ra∇ ·

[

k

k0∇T

]

+1

Ra

ρ0Ha2

k0∆T.

(134)

Here again as in the mantle,Gr = Ra/Pr << 1, the inertia in the momentumequation (134) can be neglected. The self-gravitational term −ρ/ρ0∇ψ shouldbe added to the momentum equation (second equation of (134))for large scalesimulations, the gravitational potential being solution of (132) where only thethermal part of the density variation needs to be taken into account.

The physical behavior of a large Rayleigh number is obvious in (134). WhenRa → ∞, the temperature becomes a purely advected and conserved quantity,DT/Dt = 0.

4.3.6 Internal heating

In the nondimensionalization, we assumed that the non-adiabatic temperature∆Tna and the radioactive sources are two independent quantities. Of course, inthe case where the mantle is only heated from within, the excess temperature isnot anymore a free parameter but must result from the properties of the flow itself.In the nondimensionalization, we can replace∆Tna by ρ0Ha

2/k0 in such a waythat the radioactive heat source of the anelastic or Boussinesq energy equations,

45

Page 46: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Table 1: Typical parameter values for numerical models of mantle convection.To emphasize the drastic differences between the highly viscous mantle and areal liquid (in which shear waves do not propagate), we addedestimates for thecore assuming that core convection is so efficient that only 1K of non-adiabatictemperature difference can be maintained across it. Noticethat with only 1 K oftemperature difference, the Rayleigh number of the fluid core would already reach1027!.

Mantle Coresize a 3 106 3 106 mdyn. viscosity η0 1021 10−3 Pa sheat capacity C0

P ouC0V 1000 700 J K−1 kg−1

density ρ0 4000 11000 kg m−3

heat cond. k0 3 50 W m−1 K−1

expansivity α0 2 10−5 10−5 K−1

temperature excess ∆Tna 1500 1? Kradiactivity prod. H 7 10−11 0? W kg−1

gravity g0 9.8 5 m s−2

incompressibility K0T 1011 1012 Pa

kin. viscosity ν = η0/ρ0 2.5 1017 9.1 10−8 m2 s−1

thermal diff. κ = k0/(ρ0C0P ) 7.5 10−7 6.5 10−6 m2 S−1

ǫ 3.0 10−2 1.0−5

dissip. number D0 0.59 0.21Gruneisen par. Γ 0.50 1.3Rayleigh Ra 4.2 107 2.2 1027?Intern. Rayleigh RaH 2.4 1010 0?Prandtl Pr 3.3 1023 1.4 10−2

Reynolds Re = Ra/Pr 1.3 10−15 1.6 1029

Mach M 5.0 10−15 2.3 10−16

RaPrM2 3.5 10−2 1.6 10−6

46

Page 47: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

(128) or (134), are simply1/Ra. This choice requires the introduction of a some-what different Rayleigh number, the internally heated Rayleigh number, (Roberts,1967),

RaH =α0Hρ

30g0a

5C0P

η0k20

. (135)

4.3.7 Alternative forms

Using the formulation of internal energy in terms ofCV , we would have reachedthe equivalent anelastic energy equation

ρ

ρ0

CV

C0P

DT

Dt=

1

Ra∇ ·

[

k

k0∇

(

T

∆T+ T

)]

+

α0

ρ

ρ0

g

g0

KT

KSD0T vz+

ρ

ρ0

1

Ra

ρ0Ha2

k0∆T+D0τ : ∇v,

(136)

where the reference incompressibilityKS is the incompressibility measured alongthe reference adiabatic profile

KS = ρ

(

∂P

∂ρ

)

S

= ρ||∇P ||||∇ρ|| =

ρ2g

||∇ρ|| . (137)

This relationship is given here as a definition ofKS, and this incompressibilityis built from a theoretical hydrostatic and adiabatic model. However if the realEarth is indeed hydrostatic and adiabatic, then this relationship (137) connectsa seismological observationKS(r)/ρ(r) to the density gradient of the real Earthρ(r)g(r)/||∇ρ(r)||. This is the important Bullen hypothesis (Bullen, 1940) usedto build the reference density of the Earth (e.g.,Dziewonski and Anderson, 1981).

4.3.8 Change of nondimensionalization

We use a Stokes velocity to nondimensionalize the equations. We have thereforeintroduced a velocityV0 of order of 300 m yr−1 and a timea/V0=10000 yr (seeTable 1). This is certainly very fast and short compared withgeological scales.Most physical and geophysical textbooks (e.g.,Schubert et al., 2001) use insteada diffusive timetD = ρCPa

2/k0 and velocitya/tD. This is perfectly valid butTable 1, shows that the diffusive time and velocity amount totD = 400 byr andVD = 7 10−6 m yr−1. These values would, on the contrary be very slow and longcompared with geological scales. A nondimensionalizationusing a diffusive time

47

Page 48: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

scale, leads to the anelastic equations

∇ · ( ρρ0

v) =0.

1

Pr

Dv

Dt= − ∇P + ∇ · τ +

ρ

ρ0

g

g0

K0T

KT

D0

Γ0

C0P

C0V

P − ρ

ρ0

g

g0

α

α0

RaT

ρ

ρ0

CP

C0P

DT

Dt=∇ ·

[

k

k0∇

(

T

∆T+ T

)]

+

α

α0

ρ

ρ0

g

g0D0T vg +

ρ

ρ0

ρ0Ha2

k0∆T+D0

Raτ : ∇v

(138)

and the Boussinesq equations,

∇ · v =0,

1

Pr

Dv

Dt= − ∇P + ∇ ·

(

η

η0

(∇v + [∇v]t)

)

− g

g0

α

α0

RaT ,

DT

Dt=∇ ·

[

k

k0∇T

]

+1

Ra

ρ0Ha2

k0∆T.

(139)

Notice that theRa number appears in different places than in (128) or (134). Ofcourse after their appropriate changes of variables, the dimensional solutions arethe same.

4.4 Linear stability analysis for basally heated convection

To understand why the diffusive solution is not necessarilythe solution chosenby the fluid, the standard way to test the stability of a solution. This is whatphysicists call a study of marginal stability (see also Ribe, this volume). It consistsof substituing into the basic equations a known solution plus an infinitely smallperturbation and checking whether or not this perturbationamplifies, decreases orpropagates. Its is only if the perturbation decreases in amplitude, that the testedsolution is stable.

We use the Boussinesq approximation, with constant viscosity and conduc-tivity, neglecting inertia and without internal heating. The nondimensionalizedequations (134) (the tilde sign has been omitted for simplicity) write

∇ · v =0,

−∇P + ∇2v − Tez =0,

∂T

∂t+ v · ∇T =

1

Ra∇

2T,

(140)

48

Page 49: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

(ez is the normal vector directed alongg). The steady diffusive nondimensionaltemperature solution isT = z, and we test a solution of the formT = z+δT . Thetemperature boundary condition,T = 0 on top andT = 1 at the bottom requiresthatδT vanishes forz = 0 andz = 1. As in the diffusive case the velocity is zero,the velocity induced byδT will be infinitely smallδv. In the nonlinear term, wecan approximatev ·∇T = δv ·∇(z+ δT ) by δvz = vz. With this approximation,the equations are linear and we can find a solution in the form of a plane wave.

For a fluid confined betweenz = 0 and z = 1 and unbounded in thex-direction, a solution,δT = θ(t) sin(πz) sin(kx) is appropriate and satisfies theboundary conditions. This solution is 2D, has a single mode in thez-direction,and is periodic inx with wavelengthλ = 2π/k. More complex patterns could betried but the mode we have chosen would destabilize first (seeRibe, this volume).It is then straightforward to deduce that for such a thermal anomaly, the energyequation imposes a vertical velocity

vz = −(

θ +(k2 + π2)

Raθ

)

sin(πz) sin(kx). (141)

From mass conservation thex-component of the velocity must be

vx = −πk

(

θ +(k2 + π2)

Raθ

)

cos(πz) cos(kx). (142)

This flow has a vertical component that vanishes on the top andbottom surfaceswhere the horizontal component is maximum. The choice of thetemperaturestructure corresponds to free-slip velocity conditions. When the velocity and thetemperature are introduced in the momentum equation, the time evolution of thetemperature perturbation is found

θ = θ

(

k2

(π2 + k2)2− (π2 + k2)

Ra

)

. (143)

For any wavenumberk, a small enough Rayleigh number corresponds to astable solution,θ/θ < 0. When the Rayleigh number is increased, the temperaturecomponent of wavenumberk becomes unstable at the threshold Rayleigh number

Ra =(π2 + k2)3

k2. (144)

ThisRa(k) curve is plotted in Fig 3. This curve has a minimum when

k =π√2, Rac =

27

4π4 ∼ 657. (145)

49

Page 50: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

10−2

10−1

100

101

102

103

Size of the Convection Cell π/k

101

103

105

107

Ray

leig

h N

umbe

r

UNSTABLE

STABLE657

STABLE

critical

Rayleigh

critical aspect ratio

Figure 3: Critical Rayleigh number as a function of the half wavelengthπ/k (thesize of the convection cells). Above this curve, convectionoccurs with a wholerange of unstable wavelengths. Below this curve, the conductive temperature isstable since temperature perturbations of any wavelength,decrease. When theRayleigh number is increased, the first unstable wavelengthcorresponds to a con-vection cell of aspect ratio

√2 and a critical Rayleigh number of 657.

50

Page 51: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

What can be interpreted as the size of one convective cell isπ/k since one wave-length corresponds to two contrarotating cells. The critical cell has an aspect ratio,width over height, of

√2.

A Rayleigh number of 657 is the critical Rayleigh number for convectionheated from below with free-slip boundary conditions. As soon asRa > Rac

there is a wavenumber interval over which convection begins. Of course, whenconvection grows in amplitude, the marginal stability solution becomes less andless pertinent as the assumption thatδv · ∇δT << δv · ∇z becomes invalid.

4.5 Road to chaos

In Cartesian geometry, when the Rayleigh number reaches itscritical value, con-vection starts, and forms rolls. When the Rayleigh number isfurther increased,complex series of convection patterns can be obtained, firststationary, then peri-odic, anf finally, chaotic (see Davaille, this volume). Using the values of Table 1,the critical Rayleigh number of the mantle would be attainedfor a non-adiabatictemperature difference between the surface and the CMB of only 0.025 K! Themantle Rayleigh number is several orders of magnitude higher than critical andthe mantle is in a chaotic state of convection.

Figure 4 shows a stationary convection pattern atRa = 105 and three snap-shots of numerical simulation of convection at higher Rayleigh number. The colorscale has been chosen differently in each panel to emphasizethe thermal structuresthat decrease in length scale withRa. This view is somewhat misleading since allthe thermal anomalies become confined in a top cold boundary layer and in a hotbottom one at large Rayleigh numbers. Most of the interior ofthe cell becomesjust isothermal (or adiabatic when anelastic equations areused). The various tran-sitions of convection as the Rayleigh number increases willbe discussed in otherchapters of this Treatise (see e.g., Davaille, Ribe and Zhang et al., this volume).

5 Introduction to physics of multicomponent and mul-tiphase flows

The mantle is not a simple homogeneous material. It is made ofgrains of variablebulk composition and mineralogy and contains fluids, magma and gases. Dis-cussion of multicomponent and multiphase flows could deal with solids, liquidsor gases, include compressibility or not, and consider elastic, viscous or morecomplex rheology. For each combination of these characteristics a geophysicalapplication is possible. Here we will restrict the presentation to viscous creepmodels (i.e., without inertia), where the various components are treated with con-

51

Page 52: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 4: Convection patterns of a fluid heated from below at Rayleigh number105, 106, 107, 108. The temperature color bars range from 0 (top boundary) to 1(bottom boundary). The Boussinesq approximation was used (numerical simula-tions by F. Dubuffet). The increase in Rayleigh number corresponds to a decreaseof the boundary layer thicknesses and the width of plumes. Only in the case of thelowest Rayleigh number (top left) is the convection stationary with cells of aspectratio ∼

√2 as predicted by marginal stability. For higher Rayleigh number, the

patterns are highly time-dependent.

52

Page 53: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

tinuous variables (i.e., each component is implicitly present everywhere). We donot consider approaches where the various components are separated by movingand deformable interfaces. Our presentation excludes cases where the problemis to match properties at macroscopic interfaces between regions of different buthomogeneous compositions.

We will focus on two cases. First, when all the components areperfectly mixedin variable proportions. This corresponds to the classicalchemical approach ofmultiple components in a solution. This will provide some tools to understandmantle phase transitions and the physics of chemical diffusion and mixing. Wewill be rather formal and refer the applications and illustrations to other chaptersof this Treatise (e.g., Parmentier and Tackley, this volume). Our goal is to explainwhy and when the advection diffusion equation can be used in mantle dynamics.The irreversible thermodynamics of multicomponent flows isdiscussed in variousclassical books (e.g.,Haase, 1990;de Groot and Mazur, 1984). However as usualwith geophysical flows, the mantle has many simplifications and a few complexi-ties that are not necessarily well documented in these classical textbooks.

The second case will be for two phase flows in which the two phases are sepa-rated by physical interfaces which are highly convolved andwith spatial character-istics much smaller than the typical size of geodynamic models. This is typicallythe case where magma can percolate through a compacting matrix (see also Ding-well, Vol 2). This approach was used to model melt extractionand core-mantleinteraction (McKenzie, 1984;Scott and Stevenson, 1984). Magma migration hasalso been treated in a large number of publications where solid and magma areconsidered as separated in studies of dike propagation through hydraulic fractur-ing, (e.g.,Lister and Kerr, 1991), or where fusion is parameterized in some way(e.g.,Ito et al., 1999;Choblet and Parmentier, 2001). We do not discuss theselatter approaches.

5.1 Fluid dynamics of multicomponent flows in solution

5.1.1 Mass conservation in a multicomponent solution

If we want to study the evolution of major or trace element concentration in theconvecting mantle, we can consider the mantle, instead of a homogeneous fluid, asa solution of various componentsi in volumetric proportionsφi (with

i φi = 1)having the densitiesρi and velocitiesvi (and later, thermal expansivitiesαi, heatcapacitiesCi

p...).Using a mass balance very similar to what we had discussed fora homoge-

neous fluid, we obtain a mass conservation equation of the form

∂(φiρi)

∂t+ ∇ · (φiρivi) = Γi, (146)

53

Page 54: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

whereΓi is the rate of mass production of componenti. This rate of mass produc-tion is zero if no reactions produce the componenti.

In the fluid, the average density is

ρ =∑

φiρi, (147)

and various average velocities can be defined (weighted by the mass, the volume,the number of moles... of each componenti). In this section, we introduce thebarycentric velocity,vb (velocity of the center of mass), defined by

vb =

φiρivi

ρ. (148)

The average mass conservation can be obtained by summing theequations ofcomponent conservation (146),

∂ρ

∂t+ ∇ · (ρvb) = 0, (149)

since the sum of the rates of mass production is zero∑

i

Γi = 0. (150)

In equation (146), instead of the various component velocitiesvi, we can intro-duce the barycentric velocityvb and the diffusive flux of the componenti withrespect to this average flow,

∂(φiρi)

∂t+ ∇ · (φiρivb) = −∇ · Ji + Γi, (151)

where we define the diffusive flux,Ji, by

Ji = φiρi(vi − vb). (152)

By definition of the barycentric velocity (148), the sum of the diffusive flows justcancels out,

i

Ji = 0. (153)

A diffusive transport is nothing else than an advective transfer with respect tothe average barycentric velocity. We will show later in simple cases, that thediffusive transports are driven by concentration gradients (Woods, 1975;de Grootand Mazur, 1984;Haase, 1990).

If we introduce the mass fractionCi = φiρi/ρ (in kg of i per kg of mixture),we can easily show from (149) and (151) that

ρDCi

Dt= −∇ · Ji + Γi, (154)

where the Lagrangian derivative is defined with the barycentric velocity,

D

Dt=

∂t+ vb · ∇ (155)

54

Page 55: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

5.1.2 Momentum and energy in a multicomponent solution

In our multicomponent solution, all constituents are present at each point and theyare all locally submitted to the same pressure and stresses.We assume that theviscous stress is simply related tovb and we neglect inertia as appropriate for themantle. Newton’s second law (here, simply the balance of forces) can be appliedto the barycenter and implies

∇ · τ − ∇P + ρg = 0, (156)

where the only force is due to the (constant) gravity. The momentum equationthus remains identical to that of a fluid with uniform composition and withoutinertia (16).

Since there is only one momentum equation fori components thei − 1 othervelocity equations will be found by using the constraints ofthe laws of thermody-namics and in particular the positivity of the entropy source. To derive the energyconservation, we perform the standard balance to account for all the energy ex-changes in a volumeΩ and across its surfaceΣ. Instead of the one componentequation (30), we have to sum up various contributions and weget

i

∂(φiρiUi)

∂t= − ∇ ·

(

i

φiρiUivi + P∑

i

φivi + q − τ · vb

)

+ g ·∑

i

φiρivi + ρH.(157)

In this expression, we recognize the temporal changes in energy (Ui is the com-ponent internal energy per unit mass, the kinetic energies are neglected), the bulkenergy flux , the pressure work, the thermal diffusion, the viscous stress work,the gravity work and the radioactivity production (ρH =

φiρiHi). The variousφi come from the assumption that each componenti is present in proportionφi

in both the volumeΩ and its surfaceΣ. We assume that thermal diffusion actsequally for each component and that the surface work of the stress tensor is onlyrelated to the barycentric velocity.

Using the definition of the barycentric velocity (148), of the diffusive fluxes,(152), of the momentum conservation, (156), and using

i φi = 1, the energyexpression can be simplified to

i

φiρiDHi

Dt= −∇ ·q−

i

Ji ·∇Hi +DP

Dt−∑

i

ΓiHi +τ : ∇vb + ρH, (158)

whereHi are the component enthalpies

Hi = Ui +P

ρi

. (159)

55

Page 56: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

The enthalpy variation for each componenti can be expressed as a function ofthe state variablesP andT . FromdHi = δQi + VidP and the expression of theexchanged heat (31), we can write

ρiDHi

Dt= ρiC

ip

DT

Dt+ (1 − αiT )

DP

Dt, (160)

Finally, the expression for the temperature evolution is

ρCpDT

Dt= −∇ ·q−

i

Ji ·∇Hi + αTDP

Dt−∑

i

ΓiHi +τ : ∇vb + ρH, (161)

where the average heat capacity and thermal expansivity areCp =∑

i φiρiCip/ρ

andα =∑

i φiαi.Compared to the homogeneous case (38), two new heat source terms are

present, the enthalpy exchange through chemical reactions,∑

i ΓiHi, and the en-thalpy redistribution by component diffusion,

i Ji · ∇Hi.

5.1.3 Entropy conservation in a multicomponent solution

Entropy conservation is essential for deriving the expressions of the diffusivefluxes. The general expression of entropy conservation (7) is

i

∂(φiρiSi)

∂t= −∇ · JS +HS , (162)

whereJS andHS are the yet unknown entropy flux and source. The entropy ofthe various components take into account their specific entropies as well as theirconfigurational entropies or mixing entropies due to the dispersion of the com-ponenti in the solution. Introducing the barycentric velocities and the diffusivefluxes, this equation can be recast as

i

φiρiDSi

Dt= ∇·

(

i

φiρiSivb +∑

i

SiJi − JS

)

+HS−SiΓi−Ji·∇Si. (163)

However a second expression of the entropy conservation canbe obtained fromthe enthalpy conservation, (158): usingdHi = TdSi + VidP , which, in our casecan be expressed as

ρiDHi

Dt= ρiT

DSi

Dt+DP

Dt, (164)

we derive

i

φiρiTDSi

Dt= −∇ · q −

i

Ji · ∇Hi −∑

i

ΓiHi + τ : ∇vb + ρH. (165)

56

Page 57: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

A comparison of the two expressions for the entropy conservation, (163) and (165)allows us to identify the total entropy flux

JS =q

T+ vb

i

φiρiSi +∑

i

SiJi, (166)

and the entropy sources

THS = −(q

T+∑

i

SiJi) ·∇T −∑

i

Ji ·∇µi −∑

i

Γiµi +τ : ∇vb + ρH. (167)

where we introduced the chemical potentialsµi = Hi − TSi. The total entropyflux, (166), is related to thermal diffusion and to advectionand chemical diffusionof component entropies.

In (167), the gradients of chemical potential and temperature are not indepen-dent as the chemical potential gradients implicitly include the temperature gradi-ent, so that alternative expressions can be found. For example, using

∇µi = T∇µi

T+ µi

∇T

T, (168)

andµi = Hi − TSi, the entropy source (167) can be written as

THS = −(q+∑

i

HiJi)∇T

T−T

i

Ji ·∇µi

T−∑

i

Γiµi +τ : ∇vb + ρH. (169)

We can also introduce the gradient ofµ at constant temperature∇Tµ as

∇Tµi = ∇µi + Si∇T , (170)

which leads to

THS = − 1

Tq · ∇T −

i

Ji · ∇Tµi −∑

i

Γiµi + τ : ∇vb + ρH. (171)

This last equation has the advantage of separating the temperature contribution,∇T , from the compositional contribution,∇Tµi (∇Tµi varies mostly with com-position as composition can change over very short distances, however this termis also related to pressure variations) (seede Groot and Mazur, 1984).

5.1.4 Advection-diffusion equation and reaction rates

Among the entropy sources, only terms involving similar tensorial ranks can becoupled in an isotropic medium, according to Curie’s principle. The positivity of

57

Page 58: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

the entropy production imposes three conditions, couplingtensors, vectors, andscalars.

τ : ∇vb ≥ 0, − q · ∇T

T−∑

i

Ji · ∇Tµi ≥ 0 −∑

i

Γiµi ≥ 0. (172)

The first term relates tensors and we have already discussed its implications forthe rheology in section 3.2.

The second term relates vectors and we assume, in agreement with the generalprinciple of non-equilibrium thermodynamics (de Groot and Mazur, 1984), thata matrix a phenomenological matrixM relates the thermodynamic fluxesJ =J1...Ji...q to the thermodynamic forcesX = −∇Tµ1...−∇Tµi...− ∇T/T

J1

J2

...q

= −

m11 m12 ... m1q

m21 m22 ... m2q

... ... ... ...mq1 mq2 ... mqq

∇Tµ1

∇Tµ2

...∇T/T

(173)

This linear relationship implies that the term of vectorialrank (with superscriptv), in the entropy source,TH(v)

S appears as

TH(v)S

= XtMX = XtM +M t

2X. (174)

According to the second law of thermodynamics, the symmetric part of the matrixM , (M + M t)/2 must be positive definite, i.e., the right hand side of equation(174) must be positive for any vectorsX.

At microscopic scale, a process and its reverse occur at the same rate. A con-sequence, known as the Onsager reciprocal relations, is theexistence of symmetryor antisymmetry betweenmij andmji (Onsager, 1903-1976). A general discus-sion can be found in e.g.,de Groot and Mazur(1984) orWoods(1975). When theforces are even functions of the velocities and in the absence of magnetic field,the matrixM must be symmetric. As∇T/T, and∇Tµi are even functions, asindependent of the velocities,mij = mji.

In the general case, the transport of heat by concentration gradients (the Du-four effect, (Dufour, 1832-1892)) or the transport of concentration by temperaturegradients (the Soret effect, (Soret, 1827-1890)) are possible. In many situationsthese cross-effects are small and we will assume that the matrix M does not couplethermal and compositional effects (the last row and column of M are zero exceptfor mqq/T = k, the thermal conductivity). In some case, however, when thechemical potential changes very rapidly with temperature,it becomes impossibleto neglect the coupling between chemical diffusion and temperature variations. Inthis case it may be safer to consider a formalism where the thermodynamic force

58

Page 59: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

that drives the chemical diffusion of the componenti is T∇(µi/T ) rather than∇Tµi (see alsoRichard et al., 2006).

Even without coupling between thermal and compositional effects, chemicaldiffusion in a multicomponent system remains difficult to discuss in the most gen-eral case (the positive definiteness of a symmetrici by imatrix is not a very strongconstraint). We therefore restrict our study to a simple two-component systemwhere

(

J1

J2

)

= −(

m11 m12

m21 m22

)(

∇Tµ1

∇Tµ2

)

(175)

For such a simple case, the sum of the fluxes must cancel (see (153)), and since theOnsager relations impose the symmetry of the matrix, the coefficientsmij mustverify

m11 +m21 = m12 +m22 = m12 −m21 = 0. (176)

Only one coefficient, for examplem11 can be freely chosen, and the fluxes can bewritten

J1 = m11∇T (µ2 − µ1), (177)

J2 = m11∇T (µ1 − µ2), (178)

and the second law requiresm11 > 0. If the component 1 is in small quantity (thesolute) and the component 2 is in large quantity (the solvent, with ∇(µ2) = 0), wecan easily track the evolution of solute concentrationC1. Its chemical diffusionflux is J1 = −m11∇Tµ1 and according to (154), its concentration satisfies

ρ

(

∂C

∂t+ vb · ∇C

)

= ∇ · (m∇Tµ) + Γ, (179)

where the subscripts1 have been omitted.For a solute the chemical potential is a standard chemical potentialµ0 plus a

mixing term expressing the entropy gain (configurational entropy associated withthe increased disorder) made by dispersing the solute into the solvent, of the formRT log a(C) (for crystalline solids, the activitya(C) of the mixing term can becomplex since it depends on the number and multiplicity of crystallographic sites(Spear, 1993), but we just need to know that it is related toC). In a domain wherethe average density remains uniform, the advection-diffusion equation is obtained

∂C

∂t+ vb · ∇C = ∇ · (D∇C) + Γ (180)

with a diffusion coefficientD = m/ρ(∂µ/∂C), most likelyT -dependent. Thenegative linear relationship between chemical diffusion and concentration gradi-ent is called the first Fick’s law (Fick, 1829-1901).

59

Page 60: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

When a component is present in two domains separated by a compositional in-terface, its standard chemical potentialµ0 is generally discontinuous. In this casethe gradient of the chemical potential at constantT , ∇Tµ is a mathematical dis-tribution that contains a term∇Tµ0, infinite on the compositional interface. Thisdiscontinuity drives an infinitely fast diffusion of the solute component across theinterface until the equilibrium[µ] = [µ0 + RT log a(C)] = 0. The concentrationratio ofC (or partition coefficient ofC), must therefore verify

a(C)+

a(C)−= exp (−µ

+0 − µ−

0

RT), (181)

where[ ]+ and[ ]− denote the values on the two sides of the discontinuity. Thisequation corresponds to the general rule of chemical equilibrium.

The last entropy source in (172) relates two scalars (production rates andchemical potentials). In a mixture ofi components involvingk stable atomicspecies, the conservation of these atomic species implies that onlyr = i − k lin-early independent reactions exist. Letnj

i be the stoichiometric coefficient of thecomponenti in thejth = (1...r) chemical reaction with reaction rate,Γj . We canexpressΓi as

Γi =∑

j=1...r

njiΓj , (182)

and the second law imposes

−∑

j=1...r

Γj

i

njiµi ≥ 0, (183)

The positivity of the entropy source is satisfied if the kinetic rates of thejth =1...r chemical reaction are proportional to the their chemical affinities, ∆Gj =∑

i njiµi, with positive reaction rate factorsRj

Γj = −Rj∆Gj . (184)

Chemical reaction rates are very rarely simply proportional to the affinities and theRj are likely some complex, but positive, functions ofP , T and concentrationsCi. In the case of exact thermodynamic equilibrium,

i njiµi = 0, the second law

is of course satisfied.In the same way as we defined the affinity∆Gj of the reactionj, we can

define its enthalpy∆Hj =∑

i njiHi (seede Groot and Mazur, 1984). The en-

thalpy exchange term of the energy equation (161),∑

i ΓiHi can also be written∑

j Γj∆Hj , which represents the products of the reaction rates and thereactionenthalpies. Various phase changes take place in the mantle,most notably at 410and 660 km depth. Their effects on mantle convection have been studied by vari-ous authors and will be discussed in section 6.6.

60

Page 61: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

5.1.5 Conservation properties of the advection-diffusionequation

We now make the hypothesis that the evolution of concentration of a solute inthe convective fluid is controlled by the advection-diffusion equation (180), andthat this solute is not involved in any chemical reaction,Γ = 0. For simplicity,we assume that the barycentric flow is incompressible (C can therefore be a con-centration per unit volume or per unit mass) and the diffusion coefficientD is aconstant. The fluid and the solute cannot escape the domainΩ; the normal veloc-ity and normal diffusive flux are thus zero on the boundaries of the domain, i.e.,v · n = 0 and∇C · n = 0 on the surfaceΣ with normal vectorn.

First, it is obvious that when integrated over the total domain Ω and with thedivergence theorem, the advection diffusion equation (180) implies

d

dt

ΩC dV = −

Σ(Cv −D∇C) · dS = 0. (185)

The initial heterogeneity does not disappear, it is just redistributed through time.To understand how the heterogeneity is redistributed we canexpress the evo-

lution of the concentration variance. Multiplying (180) by2C, we get after somealgebra

DC2

Dt= D∇2C2 − 2D|∇C|2. (186)

This expression when integrated over the closed volumeΩ implies that

d

dt

ΩC2 dV = −2D

Ω|∇C|2 dV . (187)

Note that the actual variance isC2− C2 whereC is the average concentration, butTheC2 term makes no contribution to (187). Since the right side is always nega-tive, the variance must continuously decrease until|∇C| = 0 which correspondsto a state of complete homogenization.

The concept of mixing is associated with the idea of homogenization wherethe concentration variance decreases with time. Since the average mixing rateis proportional to the diffusionD, (187), we note however, that a non-diffusiveflow does not homogenize at all. A diffusive flow just stirs theheterogeneities.In other terms, if the initial concentration is eitherC = 1 or C = 0, a perfecthomogenization is achieved after a timet if the concentration is everywhereC =C, the average concentration. When there is no diffusion, theinitial heterogeneityis stirred and stretched, but the local concentrations remain, for all time, eitherC = 1 orC = 0, but never an intermediate value (see Figure 5)

In the case of the Earth’s mantle, the solid state diffusion coefficients are allvery low (D = 10−19 m2 s−1 for uranium,D = 10−13 m2 s−1 for helium (seeTable 2)) and many studies have totally neglected chemical diffusion. We see, at

61

Page 62: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 5: An initial heterogeneity (top) is introduced att = 0 into a time-dependent convection cell. Without diffusion,D = 0, (bottom left), the het-erogeneity is stirred by convection and then stretched on the form of thin ribbons.However the variance of the heterogeneity concentration remains constant. It isonly with diffusion,D 6= 0, (bottom right), that a real homogenization occurs witha decrease of the heterogeneity variance.

62

Page 63: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

face value, that these models are not really homogenizing, only stirring the hetero-geneities. Without diffusion a chemical heterogeneity (e.g., a piece of subductedoceanic crust) will forever remain the same petrological heterogeneity, only itsshape will change.

Since the mixing rate is related to the compositional gradient (187) we shoulddiscuss the evolution of this gradient. We multiply (180) bythe operator2∇C ·∇to obtain

D|∇C|2Dt

= −2∇C · ǫ · ∇C + 2D[

∇ ·(

∇2C∇C)

− (∇2C)2]

, (188)

which can be integrated as

d

dt

Ω|∇C|2 dV = −2

Ω∇C · ǫ · ∇C dV − 2D

Ω(∇2C)2 dV . (189)

The rate of gradient production is related to the flow properties through the strainrate tensorǫ and to the diffusion. The diffusion term is negative and decreases thesharpness of compositional gradients.

The term related to the flow properties through the strain tensor (first term ofthe right side of (189)), could in principle be either positive or negative. Howeveras time evolves, this term must become positive. The strain rate tensor has locallythree principal axes and three principal strain rates, the sum of them being zerosince the flow is incompressible. The stretched heterogeneities becomes elongatedalong the direction of the maximum principal strain rate andthe concentrationgradients reorient themselves along the minimum, and negative, principal strainrate. The term under the first integral on the right side of (189) is thus of orderof +|ǫmin||∇C|2 (ǫmin is the local, negative eigenvalue of the strain rate tensor).Stirring is thus the source of production of concentration gradient.

We can now understand the interplay between advection and diffusion. Evenwhen the diffusion coefficientD is vanishingly small in (189), the stirring of theflow by convection will enhance the concentration gradientsuntil the average dif-fusion term, proportional to the concentration gradients,will become large enough(see (187)), for a rapid decrease of the concentration variance. We illustrate thisbehavior in the next two paragraphs by choosing a simple expression for the strainrate and computing the evolution of concentration through time.

5.1.6 Laminar and turbulent stirring

The efficiency of mixing, mostly controlled by stirring, is therefore related to theability of the flow to rapidly reduce the thickness of heterogeneities (Olson et al.,1984). In this section we set aside diffusion and discuss thestirring propertiesof a flow (see also Tackley, this volume). Let us consider a vertical piece of

63

Page 64: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

heterogeneity of width2d0, height2L (L << d0) in a simple shear flowvx =ǫz. Its top and bottom ends are at(0, L0), (0,−L0) and they will be advected to(ǫtL0, L0), (−ǫtL0,−L0) after a timet. As the heterogeneity length increases as2L0(1 + ǫ2t2), mass conservation implies that its half widthd(t) decreases as

d(t) =d0√

1 + ǫ2t2. (190)

Such flows, in which heterogeneities are stretched at rate∼ 1/t, are called flowswith laminar stirring. They are not very efficient in enhancing the diffusion be-cause they do not increase the concentration gradients, typically of order1/d(t),fast enough.

On the contrary, in a pure shear flow,vz = dz/dt = zǫ, the length of theheterogeneity would increase asL = L0 exp(ǫt) and its width would shrink as

d(t) = d0 exp(−ǫt). (191)

Such a flow is said to induce turbulent stirring. This is unfortunate terminologybecause turbulent stirring can occur in a creeping flow withRe = 0. Mantleconvection is not turbulent but it generates turbulent stirring.

Chaotic mixing flows have globally turbulent stirring properties and the qual-itative idea that highly time-dependent convection with high Rayleigh numbermixes more efficiently than low Rayleigh number convection is often true (Schmalzlet al., 1996). However steady 3D flows can also induce turbulent mixing. Thissurprising phenomenon called Lagrangian chaos is well illustrated for some the-oretical flows (Dombre et al., 1986) and for various simple flows (Ottino, 1989;Toussaint et al., 2000). For example, in a steady flow under an oceanic ridge offsetby a transform fault, the mixing is turbulent (Ferrachat and Ricard, 1998).

5.1.7 Diffusion in Lagrangian coordinates

In section 5.1.5 we discussed the mixing properties from an Eulerian viewpoint.We can also understand the interplay between diffusion and stretching (stirring) byadopting a Lagrangian viewpoint (Kellogg and Turcotte, 1987;Ricard and Coltice,2004), i.e., by solving the advection-diffusion equationsin a coordinate frame thatfollows the deformation.

Let us consider a strip of thickness2l0 with an initial concentrationC0 em-bedded in a infinite matrix of concentrationC∞. In the absence of motion, thesolution of the advection-diffusion equation (180) can be expressed using the er-ror function and the time dependent concentrationC(x, t) is given by

C(x, t) − C∞

C0 − C∞

=1

2

[

erf

(

l0 − x

2√Dt

)

+ erf

(

l0 + x

2√Dt

)

,

]

(192)

64

Page 65: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

wherex is a coordinate perpendicular to the strip and is zero at its center.The concentration at the center of the strip(x = 0) is

C(0, t) − C∞

C0 − C∞

= erf

(

l0

2√Dt

)

, (193)

and the concentration decreases by a factor of about 2 in the diffusive time

t0 ∼l20D. (194)

(erf(1/2) is not far from1/2). The time needed to homogenize a 7 km thick pieceof oceanic crust introduced into a motionless mantle is extremely long (see Table2); even the relatively mobile helium would be frozen in place since the Earthformed as it would only have migrated around 50 cm.

However, this idea of a√t diffusion is faulty since the flow stirs the hetero-

geneity and increases compositional gradients (see (189))which in turn acceler-ates the mixing process (e.g., (187)). Assuming that the problem remains two-dimensional enough so that diffusion only occurs perpendicular to the deformingheterogeneity, letl(t) be the thickness of the strip containing a chemical hetero-geneity. The velocity perpendicular to the strip would locally be at first order

vx =x

l(t)

d

dtl(t), (195)

(each side of the stripe, atx = ±l(t) moves at±dl(t)/dt).We can choose as a new space variablex = xl0/l(t), in such a way that

the Lagrangian coordinatex will vary between the fixed values -l0 and l0. Thediffusion equation becomes

(

∂C

∂t

)

x

= D

(

l0l(t)

)2∂2C

∂x2, (196)

where the partial time derivative is now computed at constant x. We see that theadvection diffusion equation has been turned into a pure diffusive equation wherethe diffusivityD has been replaced byD (l0/l(t))

2. This equivalent diffusivity islarger than D and increases with time asl(t) decreases.

To solve analytically equation (196) it is appropriate to rescale the time vari-able by definingt = F (t) with

F (t) =∫ t

0(l0/l(u))

2 du, (197)

and the resulting advection-diffusion equation in Lagrangian coordinates becomesthe simple diffusion equation with constant diffusivity. Its solution is given by

65

Page 66: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

(192) wheret andx are replaced byt andx. For example, the concentration at thecenter of the deformable strip varies like

C(0, t) − C∞

C0 − C∞

= erf

l0

2√

DF (t)

, (198)

and the concentration diminishes in amplitude by a factor of2 after a timet thatsatisfies

F (t) ∼ l20D. (199)

To perform a numerical application let us consider that the flow is either a simpleshear (190), or a pure shear deformation, (191). ComputingF (t) from equation(197) is straightforward and, assumingǫt >> 1, we get from (199), the homoge-nization times

tL ∼ 31/3l2/30

ǫ2/3D1/3, (200)

and

tT ∼ 1

2ǫlog

2l20ǫ

D, (201)

respectively. For the same oceanic crust of initial thickness 7 km, we get homoge-nization times of about 1.49 byr for He and 1.92 byr for U if we use the pure shearmechanism and assume rather arbitrarily thatǫ = 5× 10−16 s−1 (this correspondsto a typical plate velocity of 7 cm yr−1 over a plate length of 5000 km). AlthoughHe and U have diffusion coefficients 6 orders of magnitude apart, their residencetimes in a piece of subducted oceanic crust may be comparable.

The use of tracers to simulate the evolution of chemical properties in the man-tle, is our best method since solid state diffusion is too slow to be efficiently ac-counted for in a numerical simulations (e.g.,van Keken et al., 2002;Tackley andXie, 2002). However, by using tracers, we do not necessarily take into accountthat some of them may represent points that have been stretched so much that theirinitial concentration anomalies have completely diffusedinto the background. Inother words, even if diffusion seems negligible, diffusionwill erase all hetero-geneities after a finite time that is mostly controlled by thestirring properties ofthe flow.

5.2 Fluid dynamics of two phase flows

Up to now, in all of section 5.1, all components were mixed in asingle phase.However, another important geophysical application occurs when the multicom-ponents belong to different phases. This case can be illustrated with the dynamicsof partial melt in a deformable compacting matrix. Partial melts are obviously

66

Page 67: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Table 2: Homogenization times for helium and uranium assuming an heterogene-ity of initial thickness2l0 = 7 km and a strain rate of5 × 10−16 s−1.

Uranium HeliumD m2 s−1 10−19 10−13

t0 ma 3.88 × 1013 3.88 × 107

tL ma 3.60 × 105 3.60 × 103

tT ma 1920 1490

present under ridges and hotspots, but they may also be present in the middle anddeep mantle (Williams and Garnero, 1996;Bercovici and Karato, 2003) and theywere certainly more frequent when the Earth was younger. We discuss the situa-tion where two phases, fluid and matrix, can interact. In contrast to section 5.1,where the proportion and velocity of each component in solution was defined ev-erywhere at a microscopic level, in a partial melt aggregate, the local velocity ata microscopic level is either the velocity of a matrix grain,vm, or the interstitialvelocity of the melt,vf .

We assume that the two phases are individually homogeneous,incompressibleand with densitiesρf andρm. They have Newtonian rheologies with viscositiesηf andηm. They are isotropically mixed and connected. Their volume fractionsareφ (the porosity) and1 − φ. The rate of magma melting or freezing is∆Γ (inkg m−3 s−1). Although the two phases have very different physical properties wewill require the equations to be material invariant until weneed to use numericalvalues. This means that swappingf andm, φ and1−φ, ∆Γ and−∆Γ, must leavethe equations unchanged. This rule is both a physical requirement and a strongguidance in establishing the general equations (Bercovici et al., 2001a).

We make the hypothesis that there is a mesoscopic size of volumeδV whichincludes enough grains and interstitial fluid that averagedand continuous quanti-ties can be defined. Classical fluid dynamics also has its implicit averaging vol-umeδV that must contain enough atoms that quantum effects are negligible, butwhat is needed here is a much larger volume. This averaging approach remainsmeaningful because the geophysical macroscopic phenomenon that we want tounderstand (say, melting under ridges) has characteristicsizes large compared tothose of the averaging volume (say, a few cm3) (Bear, 1988).

To do the averaging, we define at microscopic level a functionθ that takes thevalue 1 in the interstitial fluid and the value 0 in the matrix grain. Mathematically,this function is rather a distribution and it has a very convolved topology. Fromit, we can define first, the porosity (volume fraction of fluid)φ, then, the fluid and

67

Page 68: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

matrix averaged velocities,vf andvm (Bercovici et al., 2001a) by

φ =1

δV

δVθ dV , (202)

φvf =1

δV

δVθvf dV , (1 − φ)vm =

1

δV

δV(1 − θ)vm dV . (203)

5.2.1 Mass conservation for matrix and fluid

Having defined the average quantities, the derivation of thetwo mass conservationequations is fairly standard (McKenzie, 1984;Bercovici et al., 2001a). They are

∂φ

∂t+ ∇ · [φvf ] =

∆Γ

ρf, (204)

−∂φ∂t

+ ∇ · [(1 − φ)vm] = −∆Γ

ρm

. (205)

We get the same equations as in (146) except that we refer toφ, 1 − φ, ∆Γ in-stead ofφ1, φ2 andΓ1. When averaged, the mass conservation equations of twoseparated phase takes the same form as the mass conservationequations of twocomponents in a solution.

We define an average and a difference quantity for any generalvariableq, by

q = φqf + (1 − φ)qm, ∆q = qm − qf . (206)

The velocityv is volume averaged and is different from the barycentric velocity(148),vb = (φρfvf + (1− φ)ρmvm)/ρ. By combining the fluid and matrix massconservation equations we get the total mass conservation equation

∂ρ

∂t+ ∇ · (ρvb) = 0, (207)

(as before, (149)), and the time rate of change in volume during melting

∇ · v = ∆Γ∆ρ

ρfρm. (208)

5.2.2 Momentum conservation of matrix and fluid

Total momentum conservation, i.e., the balance of the forces applied to the mix-ture is

∇ · τ − ∇P + ρg = 0. (209)

We have considered that the only force is due to gravity, although surface tensionbetween the two phases could also be introduced (Bercovici et al., 2001a). In this

68

Page 69: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

equation,P , τ andρ are the average pressure, stress and density. The equation isnot surprising and looks identical to its counterpart for a multicomponent solution(156). However the average pressure and stresses,P = φPf + (1 − φ)Pm andτ = φτ f + (1 − φ)τm are now the sum of two separate contributions, fromtwo separate phases having most likely very different rheologies and differentpressures. Hypothesizing that the two phases may feel the same pressure doesnot rest on any physical justification and certainly cannot hold if surface tensionis present. We will show later that their pressure difference controls the rate ofporosity change.

We split the total momentum equation into two equations one for the fluid andone for the matrix

−∇[φPf ] + φρfg + ∇ · [φτ f ] + hf = 0, (210)

−∇[(1 − φ)Pm] + (1 − φ)ρmg + ∇ · [(1 − φ)τm] + hm = 0. (211)

wherehf andhm satisfyhf + hm = 0 and represent the interaction forces actingon the fluid and on the matrix, across the interfaces separating the two phases.Because of the complexities of the interfaces, these two interaction forces must beparametrized in some way.

The simplest contribution to the interfacial forces that preserves Galilean in-variance is a Darcy-like termc∆v = c(vm −vf ) (Drew and Segel, 1971;McKen-zie, 1984) (Darcy, 1803-1858). The interaction coefficientc is related to perme-ability which is itself a function of porosity (Bear, 1988). A symmetrical formcompatible with the usual Darcy term is (seeBercovici et al., 2001a)

c =ηfηm

k0[ηf (1 − φ)n−2 + ηmφn−2], (212)

where the permeability of the formk0φn was used (usuallyn ∼ 2–3). Assuming

n = 2 andηm ≫ ηf , the interaction coefficient becomes a constant,c = ηf/k0.In the absence of gravity and when the pressures are uniform and equal, no

motion should occur even in the presence of non-uniform porosity. In this situa-tion where∆v = τ f = τm = 0 and whereP is uniform,P = Pf = Pm, theforce balances are−P∇φ+ hf = −P∇(1− φ) + hm = 0. Therefore, the inter-face forceshf andhm, must also include−P∇φ and−P∇(1− φ) when the twopressures are equal. This ledBercovici and Ricard(2003) to write the interactionterms

hf =c∆v + Pf∇φ+ ω∆P∇φ,

hm = − c∆v + Pm∇(1 − φ) + (1 − ω)∆P∇φ,(213)

with 0 ≤ ω ≤ 1. These expressions verifyhf + hm = 0, are Galilean andmaterial invariant and allow equilibrium of a mixture with non uniform porositybut uniform and equal pressures (see also (McKenzie, 1984)).

69

Page 70: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

At microscopic level, the matrix-melt interfaces are not sharp discontinuitiesbut correspond to layers (called “selvedge” layers) of disorganized atom distribu-tions. The coefficient0 < ω < 1 controls the partitioning of the pressure jump(and potentially of the surface tension) between the two phases (Bercovici andRicard, 2003) and represents the fraction of the volume-averaged surface forceexerted on the fluid phase. The exact value ofω is related to the microscopicbehavior of the two phases (molecular bond strengths and thickness of the inter-facial selvage layers) and measures the extent to which the microscopic interfacelayer is embedded in one phase more than the other. The only general physicalconstraints that we have are thatω must be zero when the fluid phase disappears(whenφ = 0) and when the fluid phase becomes unable to sustain stresses (whenηf = 0). A symmetrical form like

ω =φηf

φηf + (1 − φ)ηm

(214)

satisfies these conditions.To summarize, general expressions for the equations of fluidand matrix mo-

mentum conservation are (Bercovici and Ricard, 2003)

−φ[∇Pf − ρfg] + ∇ · [φτ f ] + c∆v + ω∆P∇φ = 0, (215)

−(1− φ)[∇Pm − ρmg] + ∇ · [(1− φ)τm]− c∆v + (1−ω)∆P∇φ = 0. (216)

The relationship between stress and velocities does not include an explicit bulkviscosity term (Bercovici et al., 2001a), and for each phasej the deviatoric stressis simply

τ j = ηj

(

∇vj + [∇vj ]t − 2

3∇ · vj I

)

, (217)

wherej stands forf orm. There is no difference in constitutive relations for theisolated component and the component in the mixture.

5.2.3 Energy conservation for two-phase flows

In the case where surface energy and entropy exist on interfaces the conservationof energy deserves more care (Sramek et al., 2007). Otherwise the global conser-vation is straightforward and can be expressed by the following equation wherethe left side represents the temporal change of internal energy content in a fixedcontrol volume and the right side represents the different contributions to thischange, namely internal heat sources, loss of energy due to diffusion, advection

70

Page 71: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

of energy, and rate of work of both surface and body forces,

∂t[φρfUf + (1 − φ)ρmUm]

= ρH − ∇ · q − ∇ · [φρfUfvf + (1 − φ)ρmUmvm]

+ ∇ · [−φPfvf − (1 − φ)Pmvm + φvf · τ f + (1 − φ)vm · τ m]

+ φvf · ρfg + (1 − φ)vm · ρmg.

(218)

The last equation is manipulated in the standard way using the mass and momen-tum equations. Because the two phases are incompressible, their internal energiesare simplydUf = CfdT anddUm = CmdT . After some algebra we get

φρfCfDfT

Dt+ (1 − φ)ρmCm

DmT

Dt

= −∇ · q − ∆PDωφ

Dt+ ∆H∆Γ + Ψ + ρH,

(219)

whereΨ is the rate of deformational work

Ψ = φ∇vf : τ f + (1 − φ)∇vm : τm + c(∆v)2. (220)

It contains the dissipation terms of each phase plus a term related to the frictionbetween the two phases. The fundamental derivatives are defined by

Dj

Dt=

∂t+ vj · ∇ , (221)

wherevj is to be substituted with the appropriate velocityvf , vm or vω with

vω = ωvf + (1 − ω)vm, (222)

In contrast to section 5.1 it would not make much sense to try to keep the equa-tions in terms of an average velocity likevb plus some diffusion terms. Here thetwo components may have very different velocities and we have to define variousmaterial derivatives.

Sinceω represents a partitioning of pressure jump, it is not surprising to findthe velocityvω (included inDω/Dt) in the work term related to this pressurejump. Associated with this partitioning factor, we can alsointroduce interfacevalues,qω, that we will use later. Any quantity∆q = qm − qf can also be written(qm−qω)−(qf −qω). When the property jump is embedded entirely in the matrix(ω = 0), there should be no jump within the fluid and we must haveqω = qf .Reciprocally, whenω = 1, we should haveqω = qm. This prompts us to defineinterface values by

qω = (1 − ω)qf + ωqm. (223)

71

Page 72: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Notice in the expressions of the interface velocityvω, (222), and interface value,qω (223), thatω and1 − ω are interchanged.

The right side of (219) contains two new expressions in addition to the usualterms (heat production, diffusion and deformational work). The first term includesthe changes in porosityDωφ/Dt times the difference in pressures between phases,∆P . The other term contains the difference in the specific enthalpies ∆H =Hm−Hf where the enthalpy of phasej is defined byHj = Uj +Pj/ρj. A similarterm was found for components reacting in a solution (158).

5.2.4 Entropy production and phenomenological laws

Entropy conservation is needed to constrain the the pressure jump between phasesand the melting rate. Starting from entropy conservation and following Srameket al. (2007)

∂t[φρfSf + (1 − φ)ρmSm] = −∇ · JS +HS , (224)

whereJS is the total entropy flux andHS is the internal entropy production, wecompare the energy and the entropy equations (219) and (224)taking into ac-count that, for each incompressible phase,dsj = CjdT/T = duj/T . After somealgebra, one gets

JS = φρfSfvf + (1 − φ)ρmSmvm +q

T, (225)

THS = − 1

Tq · ∇T − ∆P

Dωφ

Dt+ ∆µ∆Γ + Ψ + ρH, (226)

where we have introduced the difference in chemical potentials between the twophases

∆µ = ∆H− T∆S (227)

and∆S = Sm − Sf is the change in specific entropies.Following the standard procedure of non-equilibrium thermodynamics, we

chooseq = −k∇T and we assume that there is a linear relationship betweenthe two thermodynamic fluxes,Dωφ/Dt and∆Γ, and the two thermodynamicforces−∆P and∆µ since they have the same tensorial rank. We write

(

Dωφ/Dt∆Γ

)

=

(

m11 m12

m21 m22

)(

−∆P∆µ

)

(228)

The matrix of phenomenological coefficientsmij is positive definite and symmet-rical by Onsager’s theorem,m12 = m21 (seeSramek et al., 2007). For a 2x2

72

Page 73: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

matrix, it is rather simple to show that the positivity implies,m22 > 0, m11 > 0andm11m22 −m2

12 > 0 (positivity of the determinant).The form of the phenomenological coefficientsmij can be constrained through

thought experiments. First using mass conservations (204)and (205) and the def-initions ofvω andρω, (222) and (223), we can combine equations (228) to get

∆P = − m22

m11m22 −m212

[

(1 − ω)(1 − φ)∇ · vm − ωφ∇ · vf +

(

ρω

ρfρm− m12

m22

)

∆Γ

]

.(229)

In the limiting case where the two phases have the same density ρf = ρm = ρω,melting can occur with no motion,vm = vf = 0, (229) should therefore predictthe equality of pressure between phases,∆P = 0. In this case we must choosem12/m22 = 1/ρf = 1/ρm. Let us consider now a situation of homogeneousisotropic melting where the melt has such a low viscosity that it cannot sustainviscous stresses and cannot interact with the solid by Darcyterms. For such aninviscid melt,ω = 0, ρω = ρf andvω = vm. In this case, since the melt canescape instantaneously, the matrix should not dilate,∇ · vm = 0, and thus thetwo pressures should also be the same,∆P = 0. In this situation all the terms inequation (229) are 0 except for the term proportional to∆Γ. Thus, in the generalcase,

m12

m22=

ρω

ρfρm. (230)

Using this condition and introducing two positive coefficients,ζ = m22/(m11m22−m2

12), andR = m22, we can recast equations (228) as

∆P = −ζ[

Dωφ

Dt− ρω

ρfρm∆Γ

]

, (231)

∆Γ = R[

∆µ− ρω

ρfρm

∆P

]

(232)

The first equation establishes a general relation controlling the pressure dropbetween phases. The coefficientζ that links the pressure jump between the twophases to the porosity changes in excess of the melting rate,is in fact equiva-lent to a bulk viscosity as introduced in section 3.2 (see also the summary section5.2.5). The physical requirement that the two phase mixtureshould have the in-compressible properties of either the matrix or the fluid when φ = 0 or φ = 1imposes a porosity dependence toζ with limφ→0 ζ(φ) = limφ→1 ζ(φ) = +∞.Simple micromechanical models (e.g.,Nye, 1953;Bercovici et al., 2001a) allow

73

Page 74: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

us to estimate the bulk viscosity as

ζ = K0(ηf + ηm)

φ(1 − φ). (233)

The dimensionless constantK0 accounts for grain/pore geometry and is ofO(1).A more general, but more hypothetical, interpretation of the entropy positivity

could argue that some deformational work,Ψ, might affect the pressure drop of(231). This hypothesis led to a damage theory developed in (Bercovici et al.,2001a,b;Ricard and Bercovici, 2003;Bercovici and Ricard, 2003, 2005). Herewe assume that the system remains close enough to mechanicalequilibrium thatdamage does not occur.

The second equation (232) controls the kinetics of the melting/freezing and byconsequence, defines the equilibrium condition. In the caseof mechanical equi-librium, when there is no pressure drop between the two phases, the melting ratecancels when there is equality of the chemical potentials ofthe two single phases.In case of mechanical disequilibrium, (∆P 6= 0), the chemical equilibrium doesnot occur when the two chemical potentials are equal. We define a new effectivechemical potential,

µ∗

i = Ui +P ω

ρi− TSi, (234)

wherei stands forf orm and write the kinetic equation (232),

∆Γ = R∆µ∗. (235)

Chemical equilibrium imposes the equality of the effectivepotentials on the inter-face, at the pressureP ω at which the phase change effectively occurs.

Using (231) and (232), we can show that the entropy production is indeedpositive and given by

THS = k1

T|∇T |2 +

∆P 2

ζ+

(∆µ∗)2

R + Ψ + ρH. (236)

Chemical relaxation and bulk compression are associated with dissipative terms.

5.2.5 Summary equations

For convenience we summarize the governing equations when the matrix is muchmore viscous than the fluid phase (ηf ≪ ηm) as typical for melting scenarios,which implies thatτ f = 0, ω = 0, ρω = ρf , P ω = Pf , andvω = vm.

The mass conservations equations are (204) and (205). The equation of con-servation of momentum for the fluid phase is

−φ∇[Pf + ρfgz] + c∆v = 0, (237)

74

Page 75: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

(assumingz positive upward). This is Darcy’s law withc = ηfφ2/k(φ), wherek is

the permeability (often varying ask0φn with n = 2 or 3). The second momentum

equation could be the matrix momentum (211) or a combined force-difference (oraction-reaction) equation

−∇[(1 − φ)∆P ] + (1 − φ)∆ρg + ∇ · [(1 − φ)τm] − c∆v

φ= 0, (238)

where the deviatoric stress in the matrix is given by

τm = ηm

(

∇vm + [∇vm]T − 2

3∇ · vmI

)

. (239)

and the pressure jump between phases, (231), becomes

∆P = −ζ(1 − φ)∇ · vm, (240)

if an equivalent bulk viscosity is used, or

∆P = −K0ηm∇ · vm

φ, (241)

from the micromechanical model, (233), ofBercovici et al.(2001a).The action-reaction equation (238) can also be written in a different way for

example by elimination of∆v taken from the Darcy equilibrium (237),

−∇Pf + ∇ · [(1 − φ)τ ∗

m] + ρg = 0, (242)

whereτ∗

m includes the∆P term and is defined by

τ∗

m = ηm

(

∇vm + [∇vm]T − 2

3∇ · vmI

)

+ ζ(1 − φ)∇ · vmI. (243)

This shows that if the pressure is defined everywhere as the fluid pressure, thenit is equivalent to use for the matrix a rheology, (see (93)),with a bulk viscosity(1 − φ)ζ ∼ ζ (McKenzie, 1984). This analogy only holds without surface tensionbetween phases (Bercovici et al., 2001a;Ricard et al., 2001).

The rate of melting is controlled by

∆Γ = R(

∆µ+ Pf

(

1

ρm− 1

ρf

)

− T∆s

)

, (244)

and the energy equation is

ρfφCfDfT

Dt+ ρm(1 − φ)Cm

DmT

Dt− ∆H∆Γ =

ρH − ∇ · q +∆Γ2

R +K0ηm1 − φ

φ(∇ · vm)2 + Ψ,

(245)

75

Page 76: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

where we have assumed the relation (241).These equations have been used by many authors with various levels of ap-

proximation. The most benign have been to replace1 − φ by 1. Most authorshave also considered the bulk viscosityζ as a porosity independent parameter,(e.g.,McKenzie, 1984;Scott and Stevenson, 1984;Richter and Mckenzie, 1984;Ribe, 1985a,b;Scott and Stevenson, 1986, 1989;Stevenson, 1989;Kelemen et al.,1997;Choblet and Parmentier, 2001;Spiegelman et al., 2001;Spiegelman andKelemen, 2003;Katz et al., 2004). This overestimates the possibilities of matrixcompaction at low porosity. Porosity dependent parametershave been explicitlyaccounted for in other papers (e.g.,Fowler, 1985;Connolly and Podladchikov,1998;Schmeling, 2000;Bercovici et al., 2001a;Ricard et al., 2001;Rabinowiczet al., 2002). The melting rates are sometimes imposed instead of solving theenergy equation (e.g.,Turcotte and Morgan, 1992) or solved assuming univarianttransition (Fowler, 1989;Sramek et al., 2007). Surface tension is added inRileyet al. (1990) and (Hier-Majumder et al., 2006). Similar equations have also beenused to describe the interaction between iron and silicatesnear the CMB (Buffettet al., 2000), the formation of dendrites (Poirier, 1991) or the compaction of lavaflows (Massol et al., 2001).

6 Specifics of Earth’s mantle convection

In this last section we discuss various aspects of physics unique to large-scalemantle convection. We leave the problems of partial meltingto Parmentier and Itoand Van Keken, this volume. We are aware of the impossibilityto be exhaustivebut most of the important points are more deeply developed inother chapters ofthe Treatise (see also the books bySchubert et al.(2001) andDavies(1999)).

6.1 A mantle without inertia

The most striking difference between mantle convection andmost other forms ofconvection is that inertia is totally negligible. This is because the Prandtl numberis much larger than the (already very large) Rayleigh number. This implies thanthe mantle velocity flow obeys

∇ · (ρv) = 0,

−∇P + ∇ · τ + δρg + ρδg = 0,

∇2ψ = 4πGδρ,

δg = −∇ψ,

(246)

in agreement with (128). In this set of equation we kept the self-gravitation termas appropriate at long wavelengths. If the internal loadsρ are knowns, the flow can

76

Page 77: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

be computed independently of the temperature equation. This time-independentsystem has been used by many authors to try to infer the mantleflow properties.

6.1.1 Dynamic models

The system of equation (246) can been solved analytically for a depth-dependentviscosity, when variables are expressed on the basis of spherical harmonics (seeHager and Clayton(1989) and also Ribe, this volume and Forte, Vol. 3). Variouspossible surface observables (geoid height or gravity freeair anomalies, velocitydivergence, amplitude of deviatoric stress at the surface,surface dynamic topog-raphy, CMB topography...) can be expressed on the basis of spherical harmon-ics with componentsOlm. Through (246), they are related to the spherical har-monics components of the internal density variationsδρlm(r) by various degree-dependent Green’s functions (Green, 1793-1841)

Olm =∫

GOl (r)δρlm(r) dr (247)

(see Ribe, this volume for analytical details). The Green’sfunctionsGOl (r) can be

computed from the averaged density and viscosity profiles.Before seismic imaging gave us a proxy of the 3D density structure of the man-

tle, various theoretical attempts have tried to connect models of mantle convectionto plate velocities (Hager and Oconnell, 1979;Hager and O’Connell, 1981), tothe Earth gravity field (or to the geoid, proportional toψ), to the lithospheric stressregime or to the topography (Runcorn, 1964;Parsons and Daly, 1983;Lago andRabinowicz, 1984;Richards and Hager, 1984;Ricard et al., 1984).

An internal load of negative buoyancy induces a downwellingflow that de-flects the Earth’s surface, the CMB and any other internal compositional bound-aries, if they exist. The amount of deflection corresponds tothe usual isostaticrule for a load close to an interface: the weight of the induced topography equalsat first order the mass of the internal load. The total gravityanomaly resultingfrom a given internal load is affected by the mass anomalies associated with theflow-induced boundary deflections as well as by the load itself. Due to the de-flection of the Earth’s surface, the geoid perturbation induced by a dense sinkinganomaly is generally negative (e.g., free air gravity has a minimum above a denseload). However, when the mantle viscosity increases significantly with depth, by1-2 orders of magnitude, a mass anomaly close to the viscosity increase, inducesa larger CMB deformation and a lower surface deformation. The resulting gravityanomaly corresponds to a geoid high. The fact that cold subduction zones corre-spond to a relative geoid high suggests a factor≥ 30 viscosity increase around theupper-lower mantle interface (Lago and Rabinowicz, 1984;Hager et al., 1985).Shallow anomalies and anomalies near the CMB, being locallycompensated, do

77

Page 78: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

not contribute to the long-wavelength gravity field. The lithospheric stress field,like the geoid, is affected by mid-mantle density heterogeneities. The surface de-flection induced by a deep seated density anomaly decreases with the depth ofthis anomaly but even lower mantle loads should significantly affect the surfacetopography.

6.1.2 Mantle flow and post-glacial models

As soon as seismic tomography started to image the mantle structures, these seis-mic velocity anomalies have been used to further constrain the mantle viscosity.The fact that the geoid and seismic velocity anomalies are positively correlatedaround the transition zone but negatively in the deep mantleheterogeneities sug-gests a viscosity larger than 10 but not too large (less than 100) otherwise the man-tle would be everywhere positively correlated with gravity(Hager et al., 1985;Hager and Clayton, 1989;King and Hager, 1994). The same modeling approach,assuming a proportionality between seismic velocity anomalies and density vari-ations, was also used to match the observed plate divergence(Forte and Peltier,1987), the plate velocities (Ricard and Vigny, 1989;Ricard et al., 1991) and thelithospheric stresses (Bai et al., 1992;Lithgow-Bertelloni and Guynn, 2004). Theinitial Boussinesq models were extended to account for compressibility (Forte andPeltier, 1991).

Joint inversions of gravity with postglacial rebound were also performed tofurther constrain the mantle viscosity profile. The viscosity increase required bysubduction was initially thought to be too large to reconcile with post glacialrebound (Peltier, 1996). The various approaches (time-dependent for the post-glacial models and time-independent for the geoid models) seem to have con-verged to a standard viscosity profile with a significant increase with depth (Mitro-vica and Forte, 1997). Whether this viscosity increase occurs across a disconti-nuity (at the upper-lower mantle interface, or deeper) or asa gradual increase isprobably beyond the resolution of these approaches.

Although these dynamic models explain the observed geoid, they require sur-face topography of order 1km, induced by mantle convection and called dynamictopography (in contrast to the isostatic topography related to crustal and litho-spheric density variations). Its direct observation, fromthe Earth’s topographycorrected for isostatic crustal contributions, is difficult and remains controversial(e.g.,Colin and Fleitout, 1990;Kido and Seno, 1994;Lestunff and Ricard, 1995;Lithgow-Bertelloni and Silver, 1998).

78

Page 79: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

-2900

-2610

-2320

-2030

-1740

-1450

-1160

-870

-580

-290

0

Dep

th z

5 10 15 20

degree l

-2900

-2610

-2320

-2030

-1740

-1450

-1160

-870

-580

-290

0

Dep

th z

5 10 15 20

degree l

-1.0 -0.5 -0.3-0.2-0.10.00.10.20.3 0.5 1.0

Correlation Tomography-Geoid

-5e-05 -2,5e-05 0 2,5e-05 5e-05

Greoid Green function

0,5

0,6

0,7

0,8

0,9

1

Nor

mal

ized

rad

ius

µlower

/µupper

=1

µlower

/µupper

=30

µlower

/µupper

=100

-1e-05 -5e-06 0 5e-06Geoid Green function

0,5

0,6

0,7

0,8

0,9

1

Nor

mal

ized

rad

ius

µlower

µupper

=1

µlower

/µupper

=30

µlower

/µupper

=100l=2 l=10

Figure 6: Correlations between gravity and the synthetic tomographic Smeanmodel (Becker and Boschi, 2002) as a function of degreel and normalized ra-dius (top). The seismic velocities, proxy of the density variations, are positivelycorrelated with gravity around the upper-lower mantle interface (warm colors) butnegatively correlated, near the surface and in the deep lower mantle (cold colors).Geoid Green functions for degree 2 (bottom left) and degree 10 (bottom right)and three possible viscosity increases between upper and lower mantle. The geoidGreen function for a uniform viscosity (dashed line) is everywhere negative andall the anomalies around the upper-lower mantle would induce a gravity signal op-posite to that observed. A too large viscosity increase (a factor 100 for the dottedlines) cannot explain the rather good negative correlationbetween lower mantleanomalies and the geoid at long wavelength. A moderate increase (a factor 30 forthe solid line) leads to the best fit as the sign of the Green functions is everywherethat of the observed density-gravity correlations. The different Green functionsare computed for an incompressible mantle, with a lithosphere, 10 times moreviscous than the upper mantle.

79

Page 80: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

6.1.3 Time dependent models

The thermal diffusion in the mantle is so slow that even over 100-200 myrs it canbe neglected in some long wavelength global models. The equations (246) canthus be solved by imposing the known paleo-plate velocitiesat the surface and ad-vect the mass anomalies with the flow without solving explicitly the energy equa-tion. This forced-convection approach has shown that the deep mantle structureis mostly inherited from the Cenozoic and Mesozoic plate motion (Richards andEngebretson, 1992;Lithgow-Bertelloni and Richards, 1998). From plate paleo-slab reconstructions only, a density models can be obtainedthat gives a strikingfit to the observed geoid and is in relative agreement with long-wavelength to-mography (Ricard et al., 1989). This approach was also used to study the hotspotfixity (Richards, 1991;Steinberger and O’Connell, 1998), the sea level changes(Lithgow-Bertelloni and Gurnis, 1997) or the polar wander of the Earth (Spadaet al., 1992a;Richards et al., 1997).

6.2 A mantle with internal heating

When the top and bottom boundary conditions are the same (i.e., both free-slip orboth no-slip), purely basally heated convection in a Cartesian box leads to a per-fectly symmetric system. We could simultaneously reverse the vertical axis andthe color scale of Figure 4 and get temperature patterns thatare also convectivesolutions. The convective fluid has a near adiabatic core andthe temperature vari-ations are confined into two boundary layers, a hot bottom layer and a cold toplayer. The thicknesses of these two boundary layers and the temperature dropsacross them, are the same. The mid-depth temperature is simply the average ofthe top and bottom temperatures. Instabilities develop from the bottom layer (hotrising plumes) and the cold layer (cold downwelling plumes). They have a tem-perature hotter or colder than the depth-dependent averagetemperature. Theyare active structures driven by their intrinsic positive ornegative buoyancy. TheEarth’s mantle has however a large number of characteristics that break the sym-metry between upwellings and downwellings.

What is probably the major difference between mantle convection and purelybasally heated convection is that the Earth is largely powered by radiogenic heat-ing from the decay of uranium, thorium and potassium. Convection purely heatedfrom within is depicted in Figure 7. In the extreme case wherethe fluid is entirelyheated from within, the fluid has no hot bottom boundary layer. There are onlyconcentrated downwelling currents sinking from the top cold boundary layer. Thedownwellings are active as they are moved by their own negative buoyancy. Tocompensate for the resulting downwelling flow, the background is rising passively,i.e., without being pushed up by a positive buoyancy (Bercovici et al., 2000). In

80

Page 81: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

the case of basal heating, any plume leaving the top or bottomboundary layer trav-els adiabatically (neglecting diffusion and shear heating). However, in the case ofinternal heating, while the rapid downwellings remain close to adiabatic, the ra-dioactive decay can accumulate heat during the slow upwellings. This heatingis opposite to the adiabatic cooling and the average temperature in an internallyheated system remains more homogeneous and with a significant subadiabaticgradient (Parmentier et al., 1994).

The Earth’s mantle is however not in such an extreme situation. Some heatflow is extracted across the CMB from the molten iron outer core. This basal heatflux drives active upwellings (hotspots). The ratio of the internal radioactive heatto the total heat extracted at the Earth’s surface is called the Urey number (Urey,1951). Geochemical models of mantle composition (McDonough and Sun, 1995;Rudnick and Fountain, 1995) imply that about 50% of the surface heat flux is dueto mantle and core cooling and only 50% or even less (Lyubetskaya and Kore-naga, 2007), to radioactive decay. Generally geophysicists have difficulties withthese numbers as they seem to imply a too large mantle temperature in the past(Davies, 1980;Schubert et al., 1980). From convection modeling of the Earth’ssecular cooling, they often favor ratios of order of 80% radioactive and 20% cool-ing, although the complex properties of the lithosphere mayallow to reconcilethe thermal history of the Earth with a low radiogenic content (Korenaga, 2003;Grigne et al., 2005) (see Jaupart et al., this volume). The basal heat flux at theCMB represents the core cooling component, part of the totalcooling rate of theEarth. The secular cooling and the presence of internal sources tends to decreasethe thickness of the hot bottom layer compared to that of the cold top layer, in-crease the active role of downwellings (the subducting slabs), and decrease thenumber or the strength of the active upwellings (the hotspots).

6.3 A complex rheology

We have shown that the rheological laws of crystalline solids may be linear or non-linear, depending on temperature, grain size and stress level. Various deformationmechanisms (grain diffusion, grain boundary diffusion, dislocation creep...) actsimultaneously. The equivalent viscosity of each individual mechanism can bewritten in the form

η = AIn2 d

−m expE∗ + PV ∗

RT, (248)

whereE∗ andV ∗ are the activation energy and volume,P andT the pressureand temperature,R the perfect gas constant,d the grain size,m the grain sizeexponent,I2 the second stress invariant andn a stress exponent (Weertman andWeertman, 1975;Ranalli, 1995). The multiplicative factorA varies with water

81

Page 82: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 7: Convection patterns of a fluid entirely heated frominside at Rayleighnumber 106, 107, 108, 109 (simulations ran by Fabien Dubuffet). Cold finger-likeinstabilities are sinking from the top boundary layer, and spread on the bottomboundary layer. No active upwellings are present (compare with convection pat-terns for a fluid heated from below, Figure 4).

content, melt content and mineralogy. In general the composite rheology is dom-inated by the mechanism leading to the lowest viscosity.

In Figure 8, we plot as a function of temperature, and for various possiblegrain sizes (0.1 mm, 1 mm, 1 cm) the stress rate at which the strain rate predictedfor the dislocation and diffusios mechanisms are the same (see (106) and (105)).The data corresponds to dry upper mantle (Karato and Wu, 1993). Low stressand temperature favor diffusion creep while high stress andhigh temperature fa-vor dislocation creep. Below the lithosphere, in the upper mantle or at least inits shallowest part, non-linear creep is likely to occur. Asdepth increase, the de-crease in the average deviatoric stress favors a diffusive regime with a Newtonianviscosity. The observation of rheological parameters at lower mantle conditionsare more difficult but the lower mantle should mostly be in diffusive linear regimeexcept the zones of intense shear around subductions (McNamara et al., 2001).

The lateral variations of viscosity due to each separate parameter, stress ex-ponent, temperature, water content or grain size can potentially be very large.Surprisingly, attempts to deduce these variations directly from geodynamic obser-vations have not be very successful. Attempts to explain theEarth’s gravity frominternal loads do not seem to require lateral viscosity variations in the deep mantle(Zhang and Christensen, 1993). Near the surface, viscosity variations are present,

82

Page 83: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

1000 1200 1400 1600 1800 2000Temperature K

103

104

105

106

107

108

Dev

iato

ric S

tres

s P

a s

Dislocation

1 cm

1 mm

0.1 mm

Diffusion

Depth increase

Figure 8: Creep regime map for dry olivine. High deviatoric stress or temperaturefavor a dislocation mechanism. A decrease in the grain size favors diffusion. Inthe upper mantle the stress and temperature conditions tendto bring the creepregime from dislocation to diffusion at depth (n = 2.5, m = 0 andE∗ = 540 kJmol−1 for dislocation creep,n = 0,m = 2.5 andE∗ = 300 kJ mol−1 for diffusioncreep).

83

Page 84: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

at least between continental and oceanic lithosphere (Ricard et al., 1991;Cadekand Fleitout, 2003). The gain in fitting the Earth’s gravity or post glacial rebounddata with 3D viscosity models remains rather moderate compared to the complex-ities added to the modeling (Gasperini et al., 2004) and most models of mantleviscosity are restricted to radial profiles (Mitrovica and Forte, 1997). Even themodeling of slabs, their curvatures and their stress patterns do not really requirethat they are much stiffer than the surrounding mantle (Tao and Oconnell, 1993).

6.3.1 Temperature dependence of viscosity

Mantle viscosity is a strong function of temperature and thecold lithosphere seemsto have a viscosity of order 1025 Pa s (Beaumont, 1978), 4 to 6 orders of magni-tude stiffer than the asthenosphere. The activation energyE∗ is typically from300 to 600 kJ mol−1 (Drury and FitzGerald, 1998) the lowest values being fordiffusion creep. This implies a factor∼10 in viscosity decrease for a 100 K tem-perature increase (usingT ∼ 1600 K). The effect of temperature dependenceof viscosity on the planform of convection was recognized experimentally usingoils or syrups (Booker, 1976;Richter, 1978;Nataf and Richter, 1982;Weinsteinand Christensen, 1991). In the case of a strongly temperature dependent viscos-ity, the definition of the Rayleigh number is rather arbitrary as the maximum,the minimum or some average viscosity can be chosen in its definition. Anothernondimensional number (e.g., the ratio of viscosity variations,ηmax/ηmin) mustbe known to characterize the convection.

Not surprisingly, two extreme regimes are found. For a viscosity ratio lowerthan about 100, the convection pattern remains quite similar to convection withuniform viscosity. On the other hand, if the viscosity of thecold boundary layer(the lithosphere) is more than 3000 times that of the underlying asthenosphere,the surface becomes stagnant (Solomatov, 1995). Below this immobile lid, theflow resembles convection below a rigid top surface (Davaille and Jaupart, 1993).In between, when the viscosity ratios are in the range 100-3000, the lithospheredeforms slowly and in this sluggish regime, the convection cells have large aspectratios.

Convection with temperature-dependent viscosity has beeninvestigated byvarious authors (Parmentier et al., 1976;Christensen, 1984b;Tackley et al., 1993;Trompert and Hansen, 1998b;Kameyama and Ogawa, 2000). Since the topboundary layer is stiffer than the bottom boundary layer, the top boundary layeris also thicker than the bottom one. This impedes surface heat removal, eases theheat flux across the bottom boundary layer, and raises the average mantle temper-ature. Convection patterns computed withT -dependent viscosity remain howeverquite far from Earth-like convection. The major differenceis that when theT -dependence is too strong, the surface freezes and becomes immobile while on the

84

Page 85: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

real Earth, the lithosphere is highly viscous but broken into tectonic plates sepa-rated by weak boundaries. Without mechanisms other than a simpleT -dependenceof viscosity, the Earth would be in a stagnant-lid regime.

Various modelers have thus tried to useT -dependent rheologies but have im-posed a plate-like surface velocity. This has been very useful to understand theinitiation of subduction (Toth and Gurnis, 1998), the interaction of slabs with thephase changes in the transition zone (Christensen, 1996, 1997b) and the relation-ship between subduction and gravity (King and Hager, 1994). These numericalexperiments, mostly intended to model slabs, compare satisfactorily with labora-tory experiments (Kincaid and Olson, 1987;Guilloufrottier et al., 1995).

To conclude this brief section on temperature dependence ofviscosity we dis-cuss the general concept of self-regulation of planetary interiorsTozer(1972). If aplanet were convecting too vigorously it would lose more heat than radioactivelyproduced. It would therefore cool down until the viscosity is large enough to re-duce the heat transfer. On the contrary, a planet convectingtoo slowly would notextract its radioactive energy, and would heat up until the viscosity is reduced suf-ficiently (see also Jaupart et al, this volume). The internaltemperature of planetsis mostly controled to the activation energy (or rather enthalpy) of the viscosity(assuming that planets have similar amount of heat sources). To first order, largeand small terrestrial rocky planets probably have the same internal temperatures.

6.3.2 Depth-dependence of viscosity

The activation volume of the viscosity is typically around 10−5 m3 mol−1. Ex-trapolating to CMB conditions, this suggests a large viscosity increase throughoutthe mantle. However measurements of viscosity at both highT andP conditionsare very difficult (see also Weidner, Vol 2). The viscosity increase by a factor 30to 100 suggested by geodynamics (see paragraph 6.1) is probably a constrain asrobust as what could be deduced from mineralogic experiments.

The effect of a depth dependent viscosity on the planform of convection hasbeen studied by e.g.,Gurnis and Davies(1986);Cserepes(1993) orDubuffet et al.(2000). At least two important geodynamic observations canbe explained by anincrease of viscosity with depth. One is the relative stability of hotspots. A slug-gish lower mantle where convection is decreased in intensity by a larger viscosity(and also by a smaller expansivity and a potentially larger thermal conductivityas discussed in section 6.5.3) favors the relative hotspot fixity (Richards, 1991;Steinberger and O’Connell, 1998). A second consequence is a depth dependenceof the wavelengths of the thermal heterogeneities. A viscosity increase (togetherwith the existence of plates and continents that impose their own wavelengths,(see section 6.7), induces the existence of large-scale thermal anomalies at depth(Bunge and Richards, 1996). A slab crossing a factor 30-100 viscosity increase

85

Page 86: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

should thicken by a factor of order 3-5Gurnis and Hager(1988). This thickeningis observed in tomographic models (van der Hilst et al., 1997) and can be inferredfrom geoid modeling (Ricard et al., 1993a).

6.3.3 Stress-dependence of viscosity

Starting fromParmentier et al.(1976) the effect of a stress-dependent viscosityhas been studied byChristensen(1984a);Malevsky and Yuen(1992);van Kekenet al.(1992);Larsen et al.(1993), assuming either entirely non-linear or compos-ite rheologies (where deformation is accommodated by both linear and non-linearmechanisms). At moderate Rayleigh number, the effect of a non-linear rheologyis not very significant. In fact, the non-linearity in the rheology is somewhat op-posed to the temperature-dependence of the rheology. As shown by Christensen(1984a), aT -dependent, non-linear rheology with an exponentn ∼ 3 leads toconvection cells rather similar to what would be obtained with a linear rheologyand an activation energy divided by∼ n. Convection with both non-linear andT -dependent rheology looks more isoviscous than convectionwith only stress-dependent or onlyT -dependent, rheologies.

At large Rayleigh number, however, non-linear convection becomes more un-stable (Malevsky and Yuen, 1992) and the combination of non-linear rheology,T -dependent rheology and viscous dissipation can accelerate the rising velocityof hot plumes by more than an order of magnitude (Larsen and Yuen, 1997).

6.3.4 Grain size dependence of viscosity

The factorA of the viscosity law (248), can also be strongly variable andit isfor example a function of the grain sized with A ∝ dm andm of order3 in thediffusion regime (Karato et al., 1986) (when the rheology is linear in terms ofstress, it becomes non-linear in terms of grain size). Thereis a clear potential feedback interaction between deformation, grain size reduction by dynamic recrystal-lization, viscosity reduction and further localization (Jaroslow et al., 1996). Grainsize reduction is ofset by grain grow (e.g., the fact that surface energy drives massdiffusion from small grains to larger grains) which provides an effective healingmechanism (Hillert , 1965). A grain size-dependent viscosity has been introducedinto geodynamic models (e.g.,Braun et al., 1999;Kameyama et al., 1997). Theeffect is potentially important in the mantle and even more important in the litho-sphere.

86

Page 87: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

6.4 Importance of sphericity

An obvious difference between the convection planform in a planet and in an ex-perimental tank is due to the sphericity of the former. In thecase of purely basallyheated convection, the same heat flux (in a statistical sense) has to be transportedthrough the bottom boundary layer and the top boundary layer. However as theCMB surface is about 4 times smaller than the top surface, this implies a 4 timeslarger thermal gradient through the bottom boundary layer than across the litho-sphere. A bottom boundary layer thicker than the top boundary layer reinforcesthe upwelling hot instabilities with respect to the downgoing cold instabilities.Sphericity affects the average temperature and the top and bottom boundary layerthicknesses in a way totally opposite to the effects of internal sources (see section6.2) orT -dependent viscosity (see section 6.3). Although numerically more dif-ficult to handle, spherical convection models are more and more common (Glatz-maier, 1988;Bercovici et al., 1989a,b, 1992;Tackley et al., 1993;Bunge et al.,1997;Zhong et al., 2000).

6.5 Other depth-dependent parameters

6.5.1 Thermal expansivity variations

The thermal expansivity varies with depth, as predicted by the EoS (79), fromwhich we can easily deduce that

α =α0

(ρ/ρ0)n−1+q + α0(T − T0). (249)

It decreases with both temperature and density, and thus with depth. The expan-sivity varies from∼ 4 × 10−5 K−1 near the surface to∼ 8 × 10−6 K−1 nearthe CMB (Chopelas and Boehler, 1992). This diminishes the buoyancy forcesand slows down the deep mantle convection (Hansen et al., 1993). Like the in-crease of viscosity with depth, a depth-dependent thermal expansivity broadensthe thermal structures of the lower mantle, and suppresses some hot instabilitiesat the CMB. On the other hand, hot instabilities gain buoyancy as they rise in themantle, which favors their relative lateral stationarity.In addition to its averagedepth dependence, the temperature dependence of the expansivity also affects thebuoyancy of slabs (Schmeling et al., 2003).

6.5.2 Increase in average density with depth

To take into account compressibility and the depth dependence of density, theBoussinesq approximation has been replaced by the anelastic approximation inseveral studies. Such investigations have been carried outby Glatzmaier(1988);

87

Page 88: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Bercovici et al.(1992) and since extended to higher Rayleigh numbers (e.g.,Bal-achandar et al., 1992, 1993;Zhang and Yuen, 1996).

One of the difficulties with compressible fluids is that the local criterion forinstability, (see section 4.2.3), is related to the adiabatic gradient. Depending onassumptions about the curvature of the reference geotherm with depth (the slopeof the adiabatic gradient), part of the fluid can be unstable while the other partis stable. Assuming a uniform adiabatic gradient does not favor the preferentialdestabilization of either the upper or the lower mantle. On the other hand, assum-ing that the reference temperature increases exponentially with depth (i.e., takingthe order of magnitude equations (123) as real equalities) would lead to an easierdestabilization of the top of the mantle than of its bottom asa much larger heat fluxwould be carried along the lower mantle adiabat. In the real Earth, the adiabaticgradient, (in K km−1), should decrease with depth (due to the decrease in ex-pansivityα with depth insufficiently balanced by the density increase,see (121)).Since less heat can be carried out along the deep mantle adiabat, compressibilityshould favor the destabilization of the deep mantle.

Compressible convection models generally predict downgoing sheets and cylin-drical upwellings reminiscent of slabs and hotspots (Zhang and Yuen, 1996). Vis-cous dissipation is positive (as an entropy-related source) but maximum just be-low the cold boundary layer and just above the hot boundary layer, where risingor sinking instabilitities interact with the layered structures of the boundary lay-ers. On the contrary the adiabatic source heats the downwellings and cools theupwellings. On average, it reaches a maximum absolute valuein the mid-mantle.Locally, viscous dissipation and adiabatic heatings can belarger than radiogenicheat production although integrated other the whole mantleand averaged overtime, the adiabatic and dissipative sources cancel out (see(59)).

6.5.3 Thermal conductivity variations

The thermal conductivity of a solid is due to two different effects. First, a hotmaterial produces blackbody radiation that can be absorbedby neighboring atoms.This radiative transport of heat is probably a minor component since the meanfree path of photons in mantle materials is very small. Second, phonons, whichare collective vibrations of atoms, are excited and can dissipate their energiesby interacting with other phonons, with defects and with grain boundaries. Thefree paths of phonons being larger, they are the main contributors to the thermalconductivity.

According toHofmeister(1999), thermal conductivity should increase withdepth by a factor∼2-3. The recent observations of phase transitions in the bottomof the lower mantle should also be associated with another conductivity increase(Badro et al., 2004). This is one more effect (with the viscosity increaseand

88

Page 89: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

the thermal expansivity decrease) that should decrease thedeep mantle convec-tive vigor. It also broadens the thermal anomalies, increases the average mantletemperature and thins the bottom boundary layer (Dubuffet et al., 1999).

6.6 Thermo-chemical convection

Except in section 5, a simple negative relationship was assumed between den-sity variations and temperature variations, through the thermal expansivity,∆ρ =−αρ∆T . However, in the mantle several sources of density anomalies are present(see also, Stixrude, Vol 1). The density in the mantle varieswith the temperatureTfor a given mineralogical composition, or phase content, symbolized by the sym-bol φ (e.g., for a given proportion of oxides and perovskite in thelower mantle).The mineralogy for a given bulk elemental compositionχ (e.g., the proportion ofMg, Fe, O... atoms), evolves with pressure and temperature to maintain the Gibbsenergy minimum. The variations of density in the mantle at a given pressure, havepotentially three contributions that can be summarized as

∆ρ =

(

∂ρ

∂T

)

φ

∆T +

(

∂ρ

∂φ

)

T

(

∂φ

∂T

)

χ

∆T +

(

∂ρ

∂φ

)

T

(

∂φ

∂χ

)

T

∆χ (250)

The first term on the right side is the intrinsic thermal effect computed assuminga fixed mineralogy; we have already discussed this term. The second term is athermochemical effect. The density is a function of the mineralogical composi-tion contoled at uniform pressure and elemental composition, by the temperaturevariations. This effect is responsible for a rise in the 410 km deep interface and itdeepens the 660 km interface in the presence of cold downwellings (Irifune andRingwood, 1987). The last term is the intrinsic chemical effect (related to varia-tions of the mineralogy due to changes in the elemental composition at constanttemperature). The three contributions have very similar amplitudes and none ofthem is negligible (Ricard et al., 2005).

The effect of the second term has been rather well studied (Schubert et al.,1975;Christensen and Yuen, 1984;Machetel and Weber, 1991;Peltier and Sol-heim, 1992;Tackley et al., 1993;Tackley, 1995;Christensen, 1996). Phase changesin cold downgoing slabs occur at shallower depth in the case of exothermic phasechanges and at greater depth for endothermic phase changes (the ringwoodite tooxides plus perovskite phase change at 660 km depth is endothermic, all the im-portant other phase changes of the transition zone are exothermic). These sourcesof anomalies and their signs are related to the Clapeyron slope of the phase tran-sitions. The existence of latent heat release during phase change (see (161)) is asecondary and minor effect. The recent discovery of a phase transformation in thedeep lower mantle (Murakami et al., 2004) (the post-perovskite phase) suggests

89

Page 90: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

that part of the complexities of the D” layer are related to the interaction betweena phase change and the hot boundary layer of the mantle (Nakagawa and Tackley,2006) (see also Irifune, Vol 2).

The fact that below the normal 660 km depth interface there isa region whereslabs remain in a low density upper-mantle phase instead of being transformedinto the dense lower mantle phase is potentially a strong impediment to slab pen-etration. The idea that this effect induces a layering of convection at 660 kmor a situation where layered convection is punctuated by large ”avalanche” eventsdates back toRingwood and Irifune(1988) and was supported by numerical simu-lations in the 1990’s (e.g.,Machetel and Weber, 1991;Honda et al., 1993;Tackley,1995) . It seems however that the importance of this potential effect has been re-duced in recent simulations with more realistic Clapeyron slopes, phase diagrams(taking into account both the pyroxene and garnet phases), thermodynamic refer-ence values (the phase change effect has to be compared with thermal effects andthus an accurate choice for the thermal expansivity is necessary), and viscosityprofiles.

The last contribution to the density anomalies are related to variations in chem-ical composition (see Tackley, this volume). There are indications of large-scaledepth and lateral variations of Fe or Si contents in the mantle (Bolton and Masters,2001;Saltzer et al., 2004). A large well-documented elemental differentiation isbetween the oceanic crust (poor in Mg, rich in Al and Si) and the mantle. Theoceanic crust in its high pressure eclogitic facies, is∼ 5% denser that the averagemantle density in most of the mantle except in the shallowest100 kilometers of thelower mantle where it is lighter (Irifune and Ringwood, 1993). In the deepest man-tle it is not yet totally clear whether the eclogite remains denser, neutrally buoy-ant or even slightly lighter than the average mantle (e.g.,Ricolleau et al., 2004).Thermochemical simulations starting with the pioneering paper ofChristensenand Hofmann(1994) show the possibility of a partial segregation of oceanic crustduring subduction, forming pyramidal piles on the CMB. These results have beenconfirmed by e.g.,Tackley(2000b) andDavies(2002). These compositional pyra-mids may anchor the hotspots (Jellinek and Manga, 2002;Davaille et al., 2002).The presence of a petrologically dense component of the source of hotspots alsoseems necessary to explain their excess temperature (Farnetani, 1997).

Not only present-day subductions can generate compositional anomalies inthe mantle. Geochemists have often argued for a deep layer ofprimitive mate-rial. This layer should be intrinsically denser to resist entrainment by convection.The stability of such a layer has been discussed by various authors (Davaille,1999;Kellogg et al., 1999;Tackley and Xie, 2002;LeBars and Davaille, 2002;Samuel and Farnetani, 2003). Numerical simulations of thermo-chemical con-vection are certainly going to replace the thermal convection models in the nextyears. They will help to bridge the gap between geochemical observations and

90

Page 91: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

convection modeling (Coltice and Ricard, 1999;van Keken et al., 2002).

6.7 A complex lithosphere: plates and continents

The lithosphere is part of the convection cell and plate tectonics and mantle con-vection cannot be separated. The fact that the cold lithosphere is much moreviscous and concentrates most of the mass heterogeneities of the mantle, makesit behaving to some extent like a membrane on top of a less viscous fluid. Thissuggests some analogy between mantle convection and what iscalled Marangoniconvection (Marangoni, 1840-1925). Marangoni convection(Nield, 1964) is con-trolled by temperature dependent surface tension on top of thin layers of fluids.

The Earth’s mantle is certainly not controlled by surface tension and Marangoniconvection, strictly speaking has nothing to do with mantleconvection. Howeverthe equations of thermal convection with cooling from the top and with a highlyviscous lithosphere can be shown to be mathematically related (through a changeof variables) to those of Marangoni convection (Lemery et al., 2000). There arelarge differences between mantle convection and surface driven convection butthis analogy has sometimes been advocated as a ”top-down” view of the mantledynamics (Anderson, 2001). More classically, the interpretation of plate cool-ing in terms of ridge-push force (Turcotte and Schubert, 1982), or the analysis oftectonic stresses using thin sheet approximations (England and Mckenzie, 1982)belong to the same approach.

Due to the complexities of the lithosphere properties, the boundary conditionat the surface of the Earth is far from being a uniform free-slip condition. Bothcontinents and tectonic plates impose their own wavelengths and specific bound-ary conditions on the underlying convecting asthenosphere. Of course the positionof the continents and the number and shape of the plates are themselves, conse-quences of mantle convection. The plates obviously organize the large scale flowin the mantle (Hager and Oconnell, 1979;Ricard and Vigny, 1989). They imposea complex boundary condition where the angular velocity is piece-wise constant.The continents with their reduced heat flow (Jaupart and Mareschal, 1999) alsoimpose a laterally variable heat flux boundary condition.

Convection models with continents have been studied numerically (Gurnisand Hager, 1988;Grigne and Labrosse, 2001) and experimentally (Guillou andJaupart, 1995). Continents with their thick lithosphere tend to increase the thick-ness of the top boundary layer and the temperature below them(see Figure 9).Hot rising currents are predicted under continents and downwellings are localizedalong continental edges. The existence of a thick and stablecontinental root mustbe due to a chemically lighter and more viscous subcontinental lithosphere (Doinet al., 1997). The ratio of the heat flux extracted across continents compared tothat extracted across oceans increases with the Rayleigh number. This suggests

91

Page 92: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

that the continental geotherms were not much different in the past when the ra-diogenic sources were larger; it is mostly the oceanic heat flux that was larger(Lenardic, 1998). Simulating organized plates self-consistently coupled with aconvective mantle has been a very difficult quest. The attempts to generate platesusingT -dependent or simple non-linear rheologies have failed. Although in 2Dsome successes can be obtained in localizing deformation inplate-like domains,(Schmeling and Jacoby, 1981;Weinstein and Olson, 1992;Weinstein, 1996), theyare obtained with stress exponents (e.g.,n ≥ 7) that are larger than what can beexpected from laboratory experiments (n ∼ 2). The problems are however worstin 3D. Generally these early models do not predict the important shear motionsbetween plates that we observe (Christensen and Harder, 1991;Ogawa et al.,1991).

Some authors have tried to mimic the presence of plates by imposing plate-like surface boundary conditions. These studies have been performed in 2D and3D (Ricard and Vigny, 1989;Gable et al., 1991;King et al., 1992;Monnereau andQuere, 2001). Although they have confirmed the profound effect of plates on thewavelengths of convection, on its time-dependence and on the surface heat flux,these approaches cannot predict the evolution of surface plate geometry. Figure10 illustrates the organizing effect of plates in spherical, internally heated com-pressible convection with depth dependent viscosity (Bunge and Richards, 1996).To obtain a self-consistent generation of surface plates, more complex rheologiesthat include brittle failure, strain-softening and damagemechanisms must be in-troduced (e.g.,Bercovici, 1993, 1995;Moresi and Solomatov, 1998;Auth et al.,2003). The existence of plates seems also to require the existence of a weaksub-lithospheric asthenosphere (Richards et al., 2001). In the last years, the firstsuccesses in computing 3D models that spontaneously organize their top bound-ary layer into plates have been reached (Tackley, 1998, 2000c,d,e;Trompert andHansen, 1998a;Stein et al., 2004). Although the topological characteristics ofthe predicted plates and their time evolution may be still far from the observedcharacteristics of plate tectonics, and often too episodic(stagnant-lid convectionpunctuated by plate-like events), a very important breakthrough has been made bymodelers (see Figure 11).

The Earth’s plate boundaries keep the memory of their weakness over geologi-cal times (Gurnis et al., 2000). This implies that the rheological properties cannotbe a simple time-independent function of stress or temperature but has a long termmemory. The rheologies that have been used to predict platesin convective mod-els remain empirical and their interpretation in terms of microscopic behavior anddamage theory remains largely to be done (Bercovici and Ricard, 2005). Reviewson the rapid progress and the limitations of self-coherent convection models canbe found inBercovici et al.(2000);Tackley(2000a);Bercovici(2003).

92

Page 93: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 9: Convection patterns in the presence of 4 continents. The total aspectratio is 7, the continents are defined by a viscosity increaseby a factor 103 overthe depth 1/10. The viscosity is otherwise constant. The Rayleigh number basedon the total temperature drop (bottom panels) or on the internal radioactivity (toppanels) is 107. The downwellings are localized near the continent margins. Alarge difference in heat flux is predicted between oceans andcontinents. In thecase of bottom heating, hotspots tend to be preferentially anchored below conti-nents where they bring an excess heat. This tends to reduce the surface heat fluxvariations.

93

Page 94: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 10: This figure depicts spherical compressible internally heated convectionmodels where the viscosity increases with depth (simulations by Peter Bunge). Inthe first row, a uniform free-slip condition on top has been used. In the second row,the present day observed plate motion is imposed at the surface. The left columnshows the temperature field in the middle of the upper mantle,the right column inthe middle of the lower mantle. This figure summarizes various points discussedin the text: the presence of linear cold downwellings, the absence of active up-wellings in the absence of basal heating, the enlargement ofthermal structure inthe more viscous lower mantle (top row). Although the modeling is not self con-sistent (i.e., the presence of plates and the constancy of plate velocities are totallyarbitrary) it is clear that the presence of plates can changeradically the convectionpatterns (compare top and bottom rows).

94

Page 95: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Figure 11: Convection models with self-coherent plate generation (Stein et al.,2004). Snapshot of the temperature field for a model calculation in a box of aspectratio 4. The viscosity is temperature, pressure and stress dependent. The flowpattern reveals cylindrical upwellings and sheet-like downflow. Three plates haveformed sketched in the small plot.

95

Page 96: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Acknowledgements:This chapter benefited from a very careful check of the equations by FredericChambat, and from detailed and constructive reviews by NeilRibe, Gerry Schu-bert. and Dave Bercovici. I also thank Ondrej Sramek, William Landhuy, NicolasColtice, Fabien Dubuffet, David Bercovici and probably others for their commentsand suggestions.

References

Anderson, D. L. (2001), Geophysics - Top-down tectonics?,Science, 293, 2016–2018.

Anderson, O. L. (1979), Evidence supporting the approximation γρ = const for thegruneisen parameter of the earth lower mantle,J. Geophys. Res., 84, 3537–3542.

Auth, C., D. Bercovici, and U. R. Christensen (2003), Two-dimensional convection witha self-lubricating, simple-damage rheology,Geophys. J. Int., 154, 783–800.

Badro, J., J. P. Rueff, G. Vanko, G. Monaco, G. Fiquet, and F. Guyot (2004), Electronictransitions in perovskite: Possible nonconvecting layersin the lower mantle,Science,305, 383–386.

Bai, W. M., C. Vigny, Y. Ricard, and C. Froidevaux (1992), On the origin of deviatoricstresses in the lithosphere,J. Geophys. Res., 97, 11,729–11,737.

Balachandar, S., D. A. Yuen, and D. Reuteler (1992), Time-dependent 3-dimensionalcompressible convection with depth-dependent properties, Geophys. Res. Lett., 19,2247–2250.

Balachandar, S., D. A. Yuen, and D. Reuteler (1993), Viscousand adiabatic heating effectsin 3-dimensional compressible convection at infinite prandtl number,Phys. Fluids A-fluid Dynamics, 5, 2938–2945.

Barenblatt, G. I. (1996),Scaling, Cambridge University Press.

Batchelor (1967),An Introduction to Fluid Dynamics, Cambridge University Press.

Bear, J. (1988),Dynamics of fluids in porous media, Dover publishers.

Beaumont, C. (1978), Evolution of sedimentary basins on a viscoelastic lithosphere -theory and examples,Geophys. J. Royal Astron. Soc., 55, 471–497.

Becker, T. W., and L. Boschi (2002), A comparison of tomographic and geodynamicmantle models,Geochem. Geophys. Geosyst., 3, 2001GC000,168.

Bercovici, D. (1993), A simple-model of plate generation from mantle flow,Geophys. J.Int., 114, 635–650.

96

Page 97: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Bercovici, D. (1995), A source-sink model of the generationof plate-tectonics from non-newtonian mantle flow,J. Geophysical Research-solid Earth, 100, 2013–2030.

Bercovici, D. (2003), The generation of plate tectonics from mantle convection,EarthPlanet. Sci. Lett., 205, 107–121.

Bercovici, D., and S. Karato (2003), Whole-mantle convection and the transition-zonewater filter,Nature, 425, 39–44.

Bercovici, D., and Y. Ricard (2003), Energetics of a two-phase model of lithosphericdamage, shear localization and plate-boundary formation,Geophys. J. Int., 152, 581–596.

Bercovici, D., and Y. Ricard (2005), Tectonic plate generation and two-phase damage:Void growth versus grain size reduction,J. Geophys. Res., 110.

Bercovici, D., G. Schubert, and G. A. Glatzmaier (1989a), 3-dimensional spherical-models of convection in the earths mantle,Science, 244, 950–955.

Bercovici, D., G. Schubert, G. A. Glatzmaier, and A. Zebib (1989b), 3-dimensionalthermal-convection in a spherical-shell,J. Fluid Mechanics, 206, 75–104.

Bercovici, D., G. Schubert, and G. A. Glatzmaier (1992), 3-dimensional convection of aninfinite-prandtl-number compressible fluid in a basally heated spherical-shell,J. FluidDyn., 239, 683–719.

Bercovici, D., Y. Ricard, and M. Richards (2000),The relation between mantle dynamicsand plate tectonics: A primer, in: History and Dynamics of Global Plate Motions,Geophys. Monogr. Ser., vol. 121, 5–46 pp., M.A. Richards, R. Gordon, and R. van derHilst, AGU, Washington D.C.

Bercovici, D., Y. Ricard, and G. Schubert (2001a), A two-phase model of compaction anddamage, 1. general theory,J. Geophys. Res., 106, 8887–8906.

Bercovici, D., Y. Ricard, and G. Schubert (2001b), A two-phase model for compaction anddamage 3. Applications to shear localization and plate boundary formation,J. Geophys.Res., 106, 8925–8939.

Besse, J., and V. Courtillot (1991), Revised and synthetic apparent polar wander paths ofthe african, eurasian, north-america and indian plates, and true polar wander since 200ma,J. Geophys. Res., 96, 4029–4050.

Birch, F. (1952), Elasticity and constitution of the earth’s interior,J. Geophys. Res., 57,227–286.

Bolton, H., and G. Masters (2001), Travel times of P and S fromthe global digital seismicnetworks: Implications for the relative variation of P and Svelocity in the mantle,J.Geophys. Res., 106, 13,527–13,540.

97

Page 98: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Booker, J. R. (1976), Thermal-convection with strongly temperature-dependent viscosity,J. Fluid Dyn., 76, 741–754.

Braun, J., J. Chery, A. Poliakov, D. Mainprice, A. Vauchez, A. Tomassi, and M. Daig-nieres (1999), A simple parameterization of strain localization in the ductile regimedue to grain size reduction: A case study for olivine,J. Geophys. Res., 104, 25,167–25,181.

Buffett, B. A., E. J. Garnero, and R. Jeanloz (2000), Sediments at the top of Earth’s core,Science, 290, 1338–1342.

Bullen, K. E. (1940), The problem of the earth’s density variation,Seismol. Soc. Am. Bull.,30, 235–250.

Bunge, H. P., and M. A. Richards (1996), The origin of large scale structure in mantleconvection: Effects of plate motions and viscosity stratification,Geophys. Res. Lett.,23, 2987–2990.

Bunge, H. P., M. A. Richards, and J. R. Baumgardner (1997), A sensitivity study of three-dimensional spherical mantle convection at 10(8) Rayleighnumber: Effects of depth-dependent viscosity, heating mode, and an endothermic phase change,J. Geophys. Res.,102, 11,991–12,007.

Bunge, H. P., Y. Ricard, and J. Matas (2001), Non-adiabaticity in mantle convection,Geophys. Res. Lett., 28, 879–882.

Cadek, O., and L. Fleitout (2003), Effect of lateral viscosity variations in the top 300 kmon the geoid and dynamic topography,Geophys. J. Int., 152, 566–580.

Chandrasekhar, S. (1969),Ellipsoidal figures of equilibrium, Yale Press, New Haven.

Choblet, G., and E. M. Parmentier (2001), Mantle upwelling and melting beneath slowspreading centers: effects of variable rheology and melt productivity,Earth Planet. Sci.Lett., 184, 589–604.

Chopelas, A., and R. Boehler (1992), Thermal expansivity inthe lower mantle,Geophys.Res. Lett., 19, 1983–1986.

Christensen, U. (1984a), Convection with pressure-dependent and temperature-dependentnon-newtonian rheology,Geophys. J. Royal Astron. Soc., 77, 343–384.

Christensen, U. R. (1984b), Heat-transport by variable viscosity convection and implica-tions for the earths thermal evolution,Phys. Earth Planet. Inter., 35, 264–282.

Christensen, U. R. (1996), The influence of trench migrationon slab penetration into thelower mantle,Earth Planetary Science Lett., 140, 27–39.

98

Page 99: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Christensen, U. R. (1997a), Some geodynamical effects of anisotropic viscosity,Geophys.J. Royal Astron. Soc., 91, 711–736.

Christensen, U. R. (1997b), Influence of chemical buoyancy on the dynamics of slabs inthe transition zone,J. Geophys. Res., 102, 22,435–22,443.

Christensen, U. R., and H. Harder (1991), 3-d convection with variable viscosity,Geo-phys. J. Int., 104, 213–226.

Christensen, U. R., and A. Hofmann (1994), Segregation of subducted oceanic crust inthe convecting mantle,J. Geophys. Res., 99, 19,867–19,884.

Christensen, U. R., and D. A. Yuen (1984), The interaction ofa subducting lithosphericslab with a chemical or phase-boundary,J. Geophys. Res., 89, 4389–4402.

Colin, P., and L. Fleitout (1990), Topography of the ocean-floor - thermal evolution ofthe lithosphere and interaction of deep mantle heterogeneities with the lithosphere,Geophys. Res. Lett., 17, 1961–1964.

Coltice, N., and Y. Ricard (1999), Geochemical observations and one layer mantle con-vection,Earth Planetary Science Lett., 174, 125–137.

Connolly, J. A. D., and Y. Y. Podladchikov (1998), Compaction-driven fluid flow in vis-coelastic rock,Geodinamica Acta, 11, 55–84.

Corrieu, V., C. Thoraval, and Y. Ricard (1995), Mantle dynamics and geoid green-functions,Geophys. J. Int., 120, 516–523.

Cserepes, L. (1993), Effect of depth-dependent viscosity on the pattern of mantle convec-tion, Geophys. Res. Lett., 20, 2091–2094.

Dahlen, F. A., and J. Tromp (1998),Theoretical Global Seismology, Princeton UniversityPress.

Davaille, A. (1999), Two-layer thermal convection in viscous fluids,J. Fluid Dyn., 379,223–253.

Davaille, A., and C. Jaupart (1993), Transient high-rayleigh-number thermal-convectionwith large viscosity variations,J. Fluid Dyn., 253, 141–166.

Davaille, A., F. Girard, and M. LeBars (2002), How to anchor hotspots in a convectingmantle?,Earth Planet. Sci. Lett., 203, 621–634.

Davies, G. F. (1980), Thermal histories of convective earthmodels and constraints onradiogenic heat-production in the earth,J. Geophys. Res., 85, 2517–2530.

Davies, G. F. (1999),Davies, G. F., 1999, Dynamic Earth Plates, Plumes and MantleConvection, Cambridge University Press, 458 p.

99

Page 100: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Davies, G. F. (2002), Stirring geochemistry in mantle convection models with stiff platesand slabs,Geochimica Et Cosmochimica Acta, 66, 3125–3142.

de Groot, S. R., and P. Mazur (1984),Non-Equilibrium Thermodynamics, Dover Publica-tions, New York.

Doin, M. P., L. Fleitout, and U. Christensen (1997), Mantle convection and stability ofdepleted and undepleted continental lithosphere,J. Geophys. Res., 102, 2771–2787.

Dombre, T., U. Frisch, J. M. Greene, M. Henon, A. Mehr, and A. M. Soward (1986),Chaotic streamlines in the abc flows,J. Fluid Dyn., 167, 353–391.

Drew, D. A., and L. A. Segel (1971), Averaged equations for 2-phase flows,Studies InAppl. Mathematics, 50, 205–220.

Drury, M. R., and J. D. FitzGerald (1998),The Earth’s mantle: composition, structure,and evolution, chap. Mantle rheology: insights from laboratory studies of deformationand phase transition, Cambridge University Press, Cambridge, England.

Dubuffet, F., D. A. Yuen, and M. Rabinowicz (1999), Effects of a realistic mantle thermalconductivity on the patterns of 3-D convection,Earth Planet. Sci. Lett., 171, 401–409.

Dubuffet, F., M. Rabinowicz, and M. Monnereau (2000), Multiple scales in mantle con-vection,Earth Planet. Sci. Lett., 178, 351–366.

Dziewonski, A. M., and D. Anderson (1981), Preliminary reference earth model,Phys.Earth Planet. Inter., 25, 297–356.

England, P., and D. Mckenzie (1982), A thin viscous sheet model for continental defor-mation,Geophys. J. Royal Astron. Soc., 70, 295–321.

Farnetani, C. G. (1997), Excess temperature of mantle plumes: The role of chemicalstratification across d”,Geophys. Res. Lett., 24, 1583–1586.

Ferrachat, S., and Y. Ricard (1998), Regular vs, chaotic mantle mixing,Earth Planet. Sci.Lett., 155, 75–86.

Forte, A. M., and W. R. Peltier (1987), Plate tectonics and aspherical earth structure – theimportance of poloidal-toroidal coupling,J. Geophys. Res., 92, 3645–3679.

Forte, A. M., and W. R. Peltier (1991), Viscous flow models of global geophysical ob-servables .1. forward problems,J. Geophys. Res., 96, 20,131–20,159.

Fowler, A. C. (1985), A mathematical model of magma transport in the asthenosphere,Geophys. Astrophys. Fluid Dyn, 33, 63–96.

Fowler, A. C. (1989), Generation and creep of magma in the earth, Siam J. On Appl.Mathematics, 49, 231–245.

100

Page 101: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Gable, C. W., R. J. Oconnell, and B. J. Travis (1991), Convection in 3 dimensions withsurface plates - generation of toroidal flow,J. Geophys. Res., 96, 8391–8405.

Gasperini, P., G. D. Forno, and E. Boschi (2004), Linear or non-linear rheology in theEarth’s mantle: the prevalence of power-law creep in the postglacial isostatic readjust-ment of laurentia,Geophys. J. Int., 157, 1297–1302.

Glatzmaier, G. A. (1988), Numerical simulation of mantle convection: Time-dependent,three dimensional, compressible,spherical shell,Geophys. Astrophys. Fluid Dyn., 43,223–264.

Grigne, C., and S. Labrosse (2001), Effects of continents on Earth cooling: thermal blan-keting and depletion in radioactive elements,Geophys. Res. Lett., 28, 2707–2710.

Grigne, C., S. Labrosse, and P. J. Tackley (2005), Convective heat transfer as a functionof wavelength: Implications for the cooling of the earth,J. Geophysical Research-solidEarth, 110.

Guillou, L., and C. Jaupart (1995), On the effect of continents on mantle convection,J.Geophys. Res., 100, 24,217–24,238.

Guilloufrottier, L., J. Buttles, and P. Olson (1995), Laboratory experiments on the struc-ture of subducted lithosphere,Earth Planet. Sci. Lett., 133, 19–34.

Gurnis, M., and G. F. Davies (1986), Numerical study of high rayleigh number convectionin a medium with depth-dependent viscosity,Geophysical J. Royal Astronomical Soc.,85, 523–541.

Gurnis, M., and B. H. Hager (1988), Controls of the structureof subducted slabs,Nature,335, 317–321.

Gurnis, M., L. Moresi, and R. D. Mueller (2000),The History and Dynamics of GlobalPlate Motions, Richards, M. A., Gordon, R., and van der Hilst, R. editors, chap. Modelsof mantle convection incorporating plate tectonics: The Australian region since theCretaceous, pp. 211–238, American Geophysical Union, Washongton D.C.

Haase, R. (1990),Thermodynamics of irreversible processes, Dover Publications, NewYork.

Hager, B. H., and R. W. Clayton (1989),Constraints on the structure of mantle convectionusing seismic observation, flow models and the geoid, in: Mantle Convection, PlateTectonics and Global Dynamics, W.R. Peltier, Gordon and Breach Science Publishers,New York.

Hager, B. H., and R. J. Oconnell (1979), Kinematic models of large-scale flow in theearths mantle,J. Geophys. Res., 84, 1031–1048.

101

Page 102: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Hager, B. H., and R. J. O’Connell (1981), A simple global model of plate tectonics andmantle convection,J. Geophys. Res., 86, 4843–4867.

Hager, B. H., R. W. Clayton, M. A. Richards, R. P. Comer, and A.M. Dziewonski (1985),Lower mantle heterogeneity, dynamic topography and the geoid, Nature, 313, 541–546.

Hansen, U., D. A. Yuen, S. E. Kroening, and T. B. Larsen (1993), Dynamic consequencesof depth-dependent thermal expansivity and viscosity on mantle circulations and ther-mal structure,Phys. Earth Planet. Inter., 77, 205–223.

Haskell, N. A. (1937), The viscosity of the asthenosphere,Am. J. Sci., 33, 22–28.

Hewitt, J. M., D. P. McKenzie, and N. O. Weiss (1975), Dissipative heating in convectiveflows,J. Fluid Dyn., 68, 721–738.

Hier-Majumder, S., Y. Ricard, and D. Bercovici (2006), Roleof grain boundaries inmagma migration and storage,Earth Planet. Sci. Lett., 248, 735–749.

Hillert, M. (1965), On theory of normal and abnormal grain growth, Acta Metallurgica,13, 227–235.

Hirth, G., and D. L. Kolhstedt (1996), Water in the oceanic upper mantle: implications forrheology, melt extraction and the evolution of the lithosphere,Earth Planet. Sci. Lett.,144, 93–108.

Hofmeister, A. M. (1999), Mantle values of thermal conductivity and the geotherm fromphonon lifetimes,Science, 283, 1699–1706.

Honda, S., D. A. Yuen, S. Balachandar, and D. Reuteler (1993), 3-dimensional instabili-ties of mantle convection with multiple phase-transitions, Science, 259, 1308–1311.

Irifune, T., and A. Ringwood (1993), Phase transformationsin subducted oceanic crustand buoyancy relationships at depths of 600-800 km in the mantle, Earth Planet. Sci.Lett., 117, 101–110.

Irifune, T., and A. E. Ringwood (1987), Phase transformations in a harzburgite compo-sition to 26GPa: implications for dynamical behaviour of the subducting slab,EarthPlanet. Sci. Lett., 86, 365–376.

Ito, G., Y. Shen, G. Hirth, and C. J. Wolfe (1999), Mantle flow,melting, and dehydrationof the Iceland mantle plume,Earth Planetary Science Lett., 165, 81–96.

Jaroslow, G. E., G. Hirth, and H. J. B. Dick (1996), Abyssal peridotite mylonites: Impli-cations for grain-size sensitive flow and strain localization in the oceanic lithosphere,Tectonophysics, 256, 17–37.

Jarvis, G. T., and D. P. Mckenzie (1980), Convection in a compressible fluid with infiniteprandtl number,J. Fluid Dyn., 96, 515–583.

102

Page 103: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Jaupart, C., and J. C. Mareschal (1999), The thermal structure and thickness of continentalroots,Lithos, 48, 93–114.

Jeanloz, R., and S. Morris (1987), Is the mantle geotherm subadiabatic,Geophys. Res.Lett., 14, 335–338.

Jellinek, A. M., and M. Manga (2002), The influence of a chemical boundary layer on thefixity, spacing and lifetime of mantle plumes,Nature, 418, 760–763.

Kameyama, M., and M. Ogawa (2000), Transitions in thermal convection with stronglytemperature-dependent viscosity in a wide box,Earth Planet. Sci. Lett., 180, 355–367.

Kameyama, M., D. A. Yuen, and H. Fujimoto (1997), The interaction of viscous heatingwith grain-size dependent rheology in the formation of localized slip zones,Geophys.Res. Lett., 24, 2523–2526.

Karato, S. I. (1998), Seismic anisotropy in the deep mantle,boundary layers and thegeometry of mantle convection,Pure Appl. Geophys., 151, 565–587.

Karato, S. I., and P. Wu (1993), Rheology of the upper mantle:A synthesis,Science, 260,771–778.

Karato, S. I., M. S. Paterson, and J. D. F. gerald (1986), Rheology of synthetic olivineaggregates - influence of grain-size and water,J. Geophys. Res., 91, 8151–8176.

Katz, R. F., M. Spiegelman, and S. M. Carbotte (2004), Ridge migration, asthenosphericflow and the origin of magmatic segmentation in the global mid-ocean ridge system,Geophys. Res. Lett., 31.

Kelemen, P. B., G. Hirth, N. Shimizu, M. Spiegelman, and H. J.B. Dick (1997), Areview of melt migration processes in the adiabatically upwelling mantle beneathoceanic spreading ridges,Philosophical Transactions Royal Soc. London Series A-mathematical Phys. Engineering Sciences, 355, 283–318.

Kellogg, L. H., and D. L. Turcotte (1987), homogeneization of the mantle by convectivemixing and diffusion,Earth Planet. Sci. Lett., 81, 371–378.

Kellogg, L. H., B. H. Hager, and R. D. van der Hilst (1999), Compositional stratificationin the deep mantle,Science, 283, 1881–1884.

Kennett, B. L. N. (2001),The seismic wavefield, Cambridge University Press.

Kido, M., and T. Seno (1994), Dynamic topography compared with residual depth anoma-lies in oceans and implications for age-depth curves,Geophys. Res. Lett., 21, 717–720.

Kincaid, C., and P. Olson (1987), An experimental-study of subduction and slab migra-tion, J. Geophys. Res., 92, 13,832–13,840.

103

Page 104: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

King, S. D., and B. H. Hager (1994), Subducted slabs and the geoid .1. numerical experi-ments with temperature-dependent viscosity,J. Geophys. Res., 99, 19,843–19,852.

King, S. D., C. W. Gable, and S. A. Weinstein (1992), Models ofconvection-driven tec-tonic plates - a comparison of methods and results,Geophys. J. Int., 109, 481–487.

Korenaga, J. (2003), Energetics of mantle convection and the fate of fossil heat,Geophys-ical Research Lett., 30.

Lago, B., and M. Rabinowicz (1984), Admittance for convection in a layered sphericalshell,Geophys. J. Royal Astron. Soc..

Lambeck, K., and P. Johnston (1998),The Earth’s mantle: composition, structure, andevolution, chap. The viscosity of the mantle, pp. 461–502, Cambridge University Press,Cambridge, England.

Landau, L., and E. Lifchitz (1980),An Introduction to Fluid Dynamics, Pergamon Press,Oxford.

Landau, L., and E. Lifchitz (2000),Theory of elasticity, Butterworth- Heinemann.

Larsen, T. B., and D. A. Yuen (1997), Fast plumeheads: Temperature-dependent versusnon-Newtonian rheology,Geophys. Res. Lett., 24, 1995–1998.

Larsen, T. B., A. V. Malevsky, D. A. Yuen, and J. L. Smedsmo (1993), Temperature-dependent newtonian and non-newtonian convection - implications for lithosphericprocesses,Geophys. Res. Lett., 20, 2595–2598.

LeBars, M., and A. Davaille (2002), Stability of thermal convection in two superimposedmiscible viscous fluids,J. Fluid Dyn., 471, 339–363.

Lee, T., D. A. Papanastassiou, and G. L. Wasserburg (1976), Demonstration of26mgexcess in allende and evidence for al26, Geophys. Res. Lett., pp. 41–44.

Lemery, C., Y. Ricard, and J. Sommeria (2000), A model for theemergence of thermalplumes in Rayleigh-Benard convection at infinite Prandtl number,J. Fluid Mechanics,414, 225–250.

Lenardic, A. (1998), On the partitioning of mantle heat lossbelow oceans and continentsover time and its relationship to the Archaean paradox,Geophys. J. Int., 134, 706–720.

Lestunff, Y., and Y. Ricard (1995), Topography and geoid dueto lithospheric mass anoma-lies,Geophys. J. Int., 122, 982–990.

Lister, J. R., and R. C. Kerr (1991), Fluid-mechanical models of crack-propagation andtheir application to magma transport in dykes,J. Geophysical Research-solid EarthPlanets, 96, 10,049–10,077.

104

Page 105: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Lithgow-Bertelloni, C., and M. Gurnis (1997), Cenozoic subsidence and uplift of conti-nents from time-varying dynamic topography,Geology, 25, 735–738.

Lithgow-Bertelloni, C., and J. H. Guynn (2004), Origin of the lithospheric stress field,J.Geophys. Res., 109, doi: 10.1029/2003JB002,467.

Lithgow-Bertelloni, C., and M. A. Richards (1998), Dynamics of cenozoic and mesozoicplate motions,Reviews of Geophys., 36, 27–78.

Lithgow-Bertelloni, C., and P. G. Silver (1998), Dynamic topography, plate driving forcesand the African superswell,Nature, 395, 269–272.

Lyubetskaya, T., and J. Korenaga (2007), Chemical composition of earth’s primitive man-tle and its variance, 1, method and results,J. Geophys. Res..

Machetel, P., and P. Weber (1991), Intermittent layered convection in a model mantle withan endothermic phase change at 670 km,Nature, 350, 55–57.

Malevsky, A. V., and D. A. Yuen (1992), Strongly chaotic non-newtonian mantle convec-tion, Geophysical Astrophysical Fluid Dynamics, 65, 149–171.

Malvern, L. (1969),Introduction to the mechanics of a continuum medium, Prentice-HallSeries in Engineering of the Physical Sciences.

Massol, H., C. Jaupart, and D. W. Pepper (2001), Ascent and decompression of viscousvesicular magma in a volcanic conduit,J. Geophys. Res., 106, 16,223–16,240.

McDonough, W. F., and S. S. Sun (1995), The composition of theearth,Chem. Geology,120, 223–253.

McKenzie, D. (1984), The generation and compaction of partially molten rock,J. Petrol.,25, 713–765.

McNamara, A. K., S. I. Karato, and P. E. van Keken (2001), Localisation of dislocationcreep in the lower mantle: Implications for the origin of seismic anisotropy,EarthPlanet. Sci. Lett., 191, 85–99.

Mitrovica, J. X., and A. M. Forte (1997), Radial profile of mantle viscosity: Results fromthe joint inversion of convection and postglacial rebound observables,J. Geophys. Res.,102, 2751–2769.

Monnereau, M., and S. Quere (2001), Spherical shell models of mantle convection withtectonic plates,Earth Planet. Sci. Lett., 184, 575 – 588).

Moresi, L., and V. Solomatov (1998), Mantle convection witha brittle lithosphere:thoughts on the global tectonic styles of the Earth and venus, Geophys. J. Int., 133,669–682.

105

Page 106: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Muhlhaus, H. B., L. Moresi, and M. Cada (2004), Emergent anisotropy and flow align-ment in viscous rocks,Pure Appl. Geoph., 161, 2451–2463.

Munk, W. H., and G. J. F. MacDonald (1960),The rotation of the Earth: A geophysicaldiscussion, Cambridge.

Murakami, T., K. Hirose, K. Kawamura, N. Sata, and Y. Ohishi (2004), Post-perovskitephase transition inMgSiO3, Science, 304, 855–858.

Murnaghan, F. D. (1951),Finite deformation of an elastic solid, John Wiley and Sons,New York.

Nakagawa, T., and P. J. Tackley (2006), Three-dimensional structures and dynamics inthe deep mantle: Effects of post-perovskite phase change and deep mantle layering,Geophys. Res. Lett., 33(L12S11, doi:1029/2006GL025719).

Nataf, H. C., and F. M. Richter (1982), Convection experiments in fluids with highlytemperature-dependent viscosity and the thermal evolution of the planets,Phys. EarthPlanet. Inter., 29, 320–329.

Nield, D. A. (1964), Surface tension and buoyancy effects incellular convection,J. FluidDyn., 19, 341–352.

Nye, J. F. (1953), The flow law of ice from measurements in glacier tunnels, laboratoryexperiments and the jungfraufirn borehole experiment,Proc. Royal Soc. London SeriesA-mathematical Phys. Sciences, 219, 477–489.

Ogawa, M., G. Schubert, and A. Zebib (1991), Numerical simulations of 3-dimensionalthermal-convection in a fluid with strongly temperature-dependent viscosity,J. FluidDyn., 233, 299–328.

Olson, P., D. A. Yuen, and D. Balsiger (1984), Mixing of passive heterogeneties by mantleconvection,J. Geophys. Res., 89, 425–436.

Ottino, J. M. (1989),The kinematics of mixing: stretching, chaos and transport, Cam-bridge University Press, New-York.

Panasyuk, S., B. Hager, and A. Forte (1996), Understanding the effects of mantle com-pressibility on geoid kernels,Geophys. J. Int., 124, 121–133.

Parmentier, E. M., D. L. Turcotte, and K. E. Torrance (1976),Studies of finite-amplitudenon-newtonian thermal convection with application to convection in earths mantle,J.Geophys. Res., 81, 1839–1846.

Parmentier, E. M., C. Sotin, and B. J. Travis (1994), Turbulent 3-d thermal-convectionin an infinite prandtl number, volumetrically heated fluid - implications for mantle dy-namics,Geophys. J. Int., 116, 241–251.

106

Page 107: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Parsons, B., and S. Daly (1983), The relationship between surface topography, gravityanomaly and the temperature structure of convection,J. Geophys. Res., 88, 1129–1144.

Peltier, W. R. (1989),Mantle convection: plate tectonics and Global geodynamics, chap.Mantle viscosity, pp. 389–478, Gordon and Breach Science Publishers, New York.

Peltier, W. R. (1996), Mantle viscosity and ice-age ice sheet topography,Science, 273,1359–1364.

Peltier, W. R., and L. P. Solheim (1992), Mantle phase transition and layered chaoticconvection,Geophys. Res. Lett., 19, 321–324.

Poirier, J. P. (1991),Introduction to the physics of the Earth’s interior, Cambridge Uni-versity Press.

Poirier, J. P., and A. Tarantola (1998), A logarithmic equation of state,Phys. Earth Planet.Inter., 109, 1–8.

Rabinowicz, M., Y. Ricard, and M. Gregoire (2002), Compaction in a mantle with a verysmall melt concentration: Implications for the generationof carbonatitic and carbonate-bearing high alkaline mafic melt impregnations,Earth Planet. Sci. Lett., 203, 205–220.

Ranalli, G. (1995),Rheology of the Earth, Kluwer academic publishers.

Ribe, N. M. (1985a), The deformation and compaction of partial molten zones,Geophys.J. Royal Astron. Soc., 83, 487–501.

Ribe, N. M. (1985b), The generation and composition of partial melts in the earths mantle,Earth Planet. Sci. Lett., 73, 361–376.

Ricard, Y., and D. Bercovici (2003), Two-phase damage theory and crustal rock failure:the theoretical ’void’ limit, and the prediction of experimental data,Geophys. J. Int.,155, 1057–1064.

Ricard, Y., and N. Coltice (2004),Geophysical and geochemical models of mantle convec-tion: Successes and future challenges, The state of the planet: Frontiers and challengesin geophysics, vol. 150, 59-68 pp., Geophys. Monogr. Ser.

Ricard, Y., and C. Vigny (1989), Mantle dynamics with induced plate-tectonics,J. Geo-phys. Res., 94, 17,543–17,559.

Ricard, Y., L. Fleitout, and C. Froidevaux (1984), Geoid heights and lithospheric stressesfor a dynamic Earth,Ann. Geophysica, 2, 267–286.

Ricard, Y., C. Vigny, and C. Froidevaux (1989), Mantle heterogeneities, geoid and platemotion: a monte carlo inversion,J. Geophys. Res., 94, 13,739–13,754.

107

Page 108: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Ricard, Y., C. Doglioni, and R. Sabadini (1991), Differential rotation between lithosphereand mantle - a consequence of lateral mantle viscosity variations,J. Geophys. Res., 96,8407–8415.

Ricard, Y., M. Richards, C. Lithgow-Bertelloni, and Y. L. Stunff (1993a), A geodynamicmodel of mantle density heterogeneity,J. Geophys. Res., 98, 21,895–21,909.

Ricard, Y., G. Spada, and R. Sabadini (1993b), Polar wander of a dynamic earth,Geophys.J. Int., 113, 284–298.

Ricard, Y., D. Bercovici, and G. Schubert (2001), A two-phase model for compaction anddamage 2. Applications to compaction, deformation, and therole of interfacial surfacetension,J. Geophys. Res., 106, 8907–8924.

Ricard, Y., E. Mattern, and J. Matas (2005),Earth’s Deep Mantle: Structure, Composi-tion, and Evolution, chap. Synthetic tomographic images of slabs from mineral physics,pp. 285–302, AGU monograph, van der Hilst, Bass, Matas and Trampert eds.

Richard, G., D. Bercovici, and S. I. Karato (2006), Slab dehydration in the Earth’s mantletransition zone,Earth Planetary Science Lett., 251, 156–167.

Richards, M. (1991),Hotspots and the case for a high viscosity lower mantle, in GlacialIsostasy, Sea-Level and Mantle Rheology, R. Sabadini et al., eds, 571-587 pp., KluwerAcademic Publishers, Dordrecht.

Richards, M. A., and D. C. Engebretson (1992), Large-scale mantle convection and thehistory of subduction,Nature, 355, 437–440.

Richards, M. A., and B. H. Hager (1984), Geoid anomalies in a dynamic earth,J. Geophys.Res., 89, 5987–6002.

Richards, M. A., R. Y, C. Lithgow-Bertelloni, G. Spada, and R. Sabadini (1997), Anexplanation for earth’s long-term rotational stability,Science, 275, 372–375.

Richards, M. A., W. S. Yang, J. R. Baumgardner, and H. P. Bunge(2001), Role of a low-viscosity zone in stabilizing plate tectonics: Implications for comparative terrestrialplanetology,Geochem. Geophys. Geosys., 2.

Richter, F. M. (1978), Experiments on the stability of convection rolls in fluids whoseviscosity depends on temperature,J. Fluid Dyn., 89, 553–560.

Richter, F. M., and D. Mckenzie (1984), Dynamical models formelt segregation from adeformable matrix,J. Geology, 92, 729–740.

Ricolleau, A., G. Fiquet, J. Perillat, I. Daniel, N. Menguy,H. Cardon, A. Addad, C. Vanni,and N. Guignot (2004), The fate of subducted basaltic crust in the earth’s lower mantle:an experimental petrological study, inEos Trans. AGU, Fall Meet. Suppl, vol. 85(47),abstract U33B-02.

108

Page 109: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Riley, G. N., D. L. Kohlstedt, and F. M. Richter (1990), Melt migration in a silicateliquid-olivine system - an experimental test of compactiontheory,Geophys. Res. Lett.,17, 2101–2104.

Ringwood, A. E., and T. Irifune (1988), Nature of the 650-km seismic discontinuity -implications for mantle dynamics and differentiation,Nature, 331, 131–136.

Roberts, P. H. (1967), Convection in horizontal layers withinternal heat generation . the-ory, J. Fluid Mechanics, 30, 33–49.

Rudnick, R. L., and D. M. Fountain (1995), Nature and composition of the continental-crust - a lower crustal perspective,Rev. Geophys., 33, 267–309.

Runcorn, S. K. (1964), Satellite gravity measurements and the laminar viscous flow modelof the earth’s mantle,J. Geophys. Res..

Sabadini, R., and D. A. Yuen (1989), Mantle stratification and long-term polar wander,Nature, 339, 373–375.

Saltzer, R. L., E. Stutzmann, and R. D. van der Hilst (2004), Poisson’s ratio in the lowermantle beneath Alaska: Evidence for compositional heterogeneity, J. Geophys. Res.,109.

Samuel, H., and C. G. Farnetani (2003), Thermochemical convection and helium concen-trations in mantle plumes,Earth Planet. Sci. Lett., 207, 39–56.

Schmalzl, J., G. A. Houseman, and U. Hansen (1996), Mixing invigorous, time-dependent three-dimensional convection and application to Earth’s mantle,J. Geophys.Res., 101, 21,847–21,858.

Schmeling, H. (2000),Physics and chemistry of partially molten rocks, chap. Partial Melt-ing and melt migration in a convecting mantle, Kluwer Academic Publisher, Dordrecht.

Schmeling, H., and W. R. Jacoby (1981), On modeling the lithosphere in mantle convec-tion with non-linear rheology,J. Geophysics-zeitschrift Fur Geophysik, 50, 89–100.

Schmeling, H., G. Marquart, and T. Ruedas (2003), Pressure-and temperature-dependentthermal expansivity and the effect on mantle convection andsurface observables,Geo-phys. J. Int., 154, 224–229.

Schubert, G., D. A. Yuen, and D. L. Turcotte (1975), Role of phase-transitions in a dy-namic mantle,Geophys. J. Royal Astron. Soc., 42, 705–735.

Schubert, G., D. Stevenson, and P. Cassen (1980), Whole planet cooling and the radio-genic heat-source contents of the earth and moon,J. Geophys. Res., 85, 2531–2538.

Schubert, G., D. Turcotte, and P. Olson (2001),Mantle convection in the Earth and plan-ets, Cambridge University Press.

109

Page 110: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Scott, D. R., and D. J. Stevenson (1984), Magma solitons,Geophys. Res. Lett., 11, 9283–9296.

Scott, D. R., and D. J. Stevenson (1986), Magma ascent by porous flow,J. Geophys. Res.,91, 9283–9296.

Scott, D. R., and D. J. Stevenson (1989), A self-consistent model of melting, magmamigration and buoyancy-driven circulation beneath mid-ocean ridges,J. Geophys. Res.,94, 2973–2988.

Solomatov, V. S. (1995), Scaling of temperature-dependentand stress-dependent viscosityconvection,Phys. Fluids, 7, 266–274.

Spada, G., Y. Ricard, and R. Sabadini (1992a), Excitation oftrue polar wander by sub-duction,Nature, 360, 452–454.

Spada, G., R. Sabadini, D. A. Yuen, and Y. Ricard (1992b), Effects on postglacial reboundfrom the hard rheology in the transition zone,Geophysical J. Int., 109, 683–700.

Spear, F. S. (1993),Metamorphic Phase Equilibria and Pressure-Temperature-TimePaths, 799p, Mineralogical Society of America, Washington, D. C.

Spiegelman, M., and P. B. Kelemen (2003), Extreme chemical variability as a conse-quence of channelized melt transport,Geochem. Geophys. Geosys., 4.

Spiegelman, M., P. B. Kelemen, and E. Aharonov (2001), Causes and consequences offlow organization during melt transport: The reaction infiltration instability in com-pactible media,J. Geophys. Res., 106, 2061–2077.

Sramek, O., Y. Ricard, and D. Bercovici (2007), Simultaneous melting and compaction indeformable two phase media,Geophys. J. Int..

Stacey, F. D. (1977), Application of thermodynamics to fundamental earth physics,Geo-phys. Surveys, 3, 175–204.

Stacey, F. D., and P. M. Davis (2004), High pressure equations of state with applicationsto the lower mantle and core,Phys. Earth Planet. Inter., 142, 137–184.

Stein, C., J. Schmalzl, and U. Hansen (2004), The effect of rheological parameters onplate behaviour in a self-consistent model of mantle convection, Phys. Earth Planet.Inter., 142, 225–255.

Steinberger, B., and R. J. O’Connell (1998), Advection of plumes in mantle flow: impli-cations for hotspot motion, mantle viscosity and plume distribution, Geophys. J. Int.,132, 412–434.

Stevenson, D. J. (1989), Spontaneous small-scale melt segregation in partial melts under-going deformation,Geophys. Res. Lett., 16, 1067–1070.

110

Page 111: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Tackley, P. (1995), On the penetration of an endothermic phase transition by upwellingsand downwellings,J. Geophys. Res., 100, 15,477–15,488.

Tackley, P. (2000a),The quest for self-consistent incorporation of plate tectonics in man-tle convection, in: History and Dynamics of Global Plate Motions, vol. 121, AGUGeophys. Monograph Series, M.A. Richards, R. Gordon, and R.van der Hilst, AGU,Washington D.C.

Tackley, P. J. (1996), Effects of strongly variable viscosity on three-dimensional com-pressible convection in planetary mantles,J. Geophysical Research-solid Earth, 101,3311–3332.

Tackley, P. J. (1998), Self-consistent generation of tectonic plates in three-dimensionalmantle convection,Earth Planetary Science Lett., 157, 9–22.

Tackley, P. J. (2000b), Mantle convection and plate tectonics: Toward an integrated phys-ical and chemical theory,Science, 288, 2002–2007.

Tackley, P. J. (2000c), Mantle convection and plate tectonics: Toward an integrated phys-ical and chemical theory,Science, 288, 2002–2007.

Tackley, P. J. (2000d), Self consistent generation of tectonic plates in time-dependent,three dimensional mantle convection simulations, part 1: Pseudoplastic yielding,G3,1.

Tackley, P. J. (2000e), Self consistent generation of tectonic plates in time-dependent,three dimensional mantle convection simulations, part 2: Strain weakening and as-thenosphere,G3, 1.

Tackley, P. J., and S. X. Xie (2002), The thermochemical structure and evolution of Earth’smantle: constraints and numerical models,Philosophical Transactions Royal Soc. Lon-don Series A-mathematical Phys. Engineering Sciences, 360, 2593–2609.

Tackley, P. J., D. J. Stevenson, G. A. Glatzmaier, and G. Schubert (1993), Effects of anendothermic phase-transition at 670 km depth in a sphericalmodel of convection in theearths mantle,Nature, 361, 699–704.

Tao, W. C., and R. J. Oconnell (1993), Deformation of a weak subducted slab and variationof seismicity with depth,Nature, 361, 626–628.

Toth, J., and M. Gurnis (1998), Dynamics of subduction initiation at preexisting faultzones,J. Geophys. Res., 103, 18,053–18,067.

Toussaint, V., P. Carriere, J. Scott, and J. N. Gence (2000),Spectral decay of a passivescalar in chaotic mixing,Phys. Fluids, 12, 2834–2844.

111

Page 112: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Tozer, D. C. (1972), The present thermal state of the terrestrial planets,Phys. Earth Planet.Inter., 6, 182–197.

Tritton, D. J. (1988),Physical fluid dynamics, Oxford University Press.

Trompert, R., and U. Hansen (1998a), Mantle convection simulations with rheologies thatgenerate plate-like behaviour,Nature, 395, 686–689.

Trompert, R. A., and U. Hansen (1998b), On the rayleigh number dependence of convec-tion with a strongly temperature-dependent viscosity,Phys. Fluids, 10, 351–360.

Turcotte, D. L., and J. P. Morgan (1992),Mantle flow and melt generation at mid-oceanridges, chap. The physics of magma migration and mantle flow beneatha mid-ocearidge, pp. 155–182, eds J. Phipps Morgan, D. K. Blackmann andJ. M. Simpson, AGUNew York.

Turcotte, D. L., and G. Schubert (1982),Geodynamics: applications of continuum physicsto geological problems, 1-450 pp., Wiley, New York.

Urey, H. C. (1951), The origin and development of the earth and other terrestrial planets,Geochim. Cosmochim. Acta, 1, 209–277.

van der Hilst, R. D., S. Widiyantoro, and E. R. Engdahl (1997), Evidence for deep mantlecirculation from global tomography,Nature, 386, 578–584.

van Keken, P., D. A. Yuen, and A. Vandenberg (1992), Pulsating diapiric flows - conse-quences of vertical variations in mantle creep laws,Earth Planetary Science Lett., 112,179–194.

van Keken, P. E., E. H. Hauri, and C. J. Ballentine (2002), Mantle mixing: The generation,preservation, and destruction of chemical heterogeneity,Ann. Rev. Earth PlanetarySciences, 30, 493–525.

Vinet, P., J. Ferrante, J. H. Rose, and J. R. Smith (1987), Compressibility of solids,J.Geophys. Res., 92, 9319–9325.

Weertman, J., and J. R. Weertman (1975), High-temperature creep of rock and mantleviscosity,Ann. Rev. Earth Planet. Sci., 3, 293–315.

Weinstein, S. A. (1996), Thermal convection in a cylindrical annulus with a non-Newtonian outer surface,Pure Appl. Geophys., 146, 551–572.

Weinstein, S. A., and U. Christensen (1991), Convection planforms in a fluid with atemperature-dependent viscosity beneath a stress-free upper boundary,GeophysicalResearch Lett., 18, 2035–2038.

Weinstein, S. A., and P. L. Olson (1992), Thermal-convection with non-newtonian plates,Geophys. J. Int., 111, 515–530.

112

Page 113: PHYSICS OF MANTLE CONVECTION Contents - USC Geodynamics

Williams, Q., and E. J. Garnero (1996), Seismic evidence forpartial melt at the base ofEarth’s mantle,Science, 273, 1528–1530.

Woods, L. C. (1975),The thermodynamics of fluid systems, Clarendon Press, Oxford.

Yuen, D., R. C. A. Sabadini, and P. Gasperini (1986), On transient rheology and glacialisostasy,J. Geophys. Res., 91, 1420–1438.

Zhang, S. X., and U. Christensen (1993), Some effects of lateral viscosity variations ongeoid and surface velocities induced by density anomalies in the mantle,Geophys. J.Int., 114, 531–547.

Zhang, S. X., and D. A. Yuen (1996), Various influences on plumes and dynamics intime-dependent, compressible mantle convection in 3-D spherical shell,Phys. EarthPlanetary Interiors, 94, 241–267.

Zhong, S. J., M. Gurnis, and L. Moresi (1996), Free surface formulation of mantle con-vection, part 1: basic theory and implication to plumes,Geophys. J. Int., 127, 708–718.

Zhong, S. J., M. T. Zuber, L. Moresi, and M. Gurnis (2000), Role of temperature-dependent viscosity and surface plates in spherical shell models of mantle convection,J. Geophys. Res., 105, 11,063–11,082.

113