You are on page 1of 20

Geophysical Journal International

Geophys. J. Int. (2011) 184, 699–718 doi: 10.1111/j.1365-246X.2010.04870.x

Disequilibrium melting of a two phase multicomponent mantle

John F. Rudge,1,2, ∗ David Bercovici1 and Marc Spiegelman3


1 Department of Geology and Geophysics, Yale University, New Haven, CT 06511, USA. E-mail: rudge@esc.cam.ac.uk
2 Institute
of Theoretical Geophysics, Bullard Laboratories, University of Cambridge, Madingley Road, Cambridge, CB3 0EZ, UK
3 Lamont-Doherty Earth Observatory, Columbia University, Palisades, NY 10964, USA

Accepted 2010 October 22. Received 2010 October 18; in original form 2010 June 11

GJI Mineral physics, rheology, heat flow and volcanology


Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020
SUMMARY
Melt generation and segregation in Earth’s mantle is typically modelled using the mixture
theory of two phase flows, which combine a set of conservation laws for mass, momentum and
energy with phenomenological laws for fluxes of mass and heat. Most current two phase flow
models assume local thermodynamic equilibrium between melt and matrix, but geochemical
observations suggest disequilibrium transport may play an important role. Here we generalise
the existing two phase flow theories to encompass multiple thermodynamic components and
disequilibrium. Our main focus is on the phenomenological laws describing phase change and
we present general disequilibrium melting laws, which reduce to the familiar fractional and
equilibrium melting laws in appropriate limits. To demonstrate the behaviour of our melting
laws, we address two simple model problems for a binary system: melting at constant pressure
and melting in a 1-D upwelling column at steady state. The framework presented here will
prove useful in future for modelling reaction infiltration instabilities in a thermodynamically
consistent manner. This framework will be useful not only for magma dynamics but for a wide
range of reactive two phase flow problems.
Key words: Mid-ocean ridge processes; Mechanics, theory, and modelling; Magma genesis
and partial melting; Magma migration and fragmentation.

face tension and damage (Bercovici et al. 2001; Ricard et al. 2001;
1 I N T RO D U C T I O N
Bercovici & Ricard 2003) and have been made more rigorous by
Beneath the Earth’s mid-ocean ridges the mantle melts and that melt the formal homogenisation of microscale models (Simpson et al.
rises to the surface to form new crust. Why the mantle melts is well 2010a,b).
understood: it is a natural consequence of the thermodynamics of The present work seeks to extend the current two phase flow
decompression melting (e.g. Stolper & Asimow 2007). But how models of magma dynamics in two ways: (1) to allow chemical dis-
the melt rises is still poorly understood, despite many decades of equilibrium between the two phases and (2) to encompass multiple
work on the problem. Part of the reason for this poor understanding thermodynamic components. In most two phase flow models, local
is the complex coupling that exists between melt segregation and thermodynamic equilibrium is assumed to hold everywhere. This
melt generation. To have a full description of the system we must is a very useful assumption, as it means that the thermodynamic
consider not only the fluid dynamics of melt segregation but also variables are constrained to lie on phase diagrams, which has been
the thermodynamics and perhaps even kinetics, of melt generation. much exploited in recent geodynamic models (Katz 2008; Tirone
The main geodynamical modelling approach for this magma dy- et al. 2009). However, there is compelling geochemical evidence
namics problem has been to use the mixture theory of two phase which suggests that melts are not always in chemical equilibrium
flows (Drew 1983). A particularly useful and highly simplified two with the matrix through which they pass (Kelemen et al. 1997) and
phase flow theory appropriate to magma dynamics was written down indeed mantle melting is thought to be closer to a fractional process
by McKenzie (1984) and others (Scott & Stevenson 1984; Fowler than an equilibrium process. Chemical disequilibrium has been in-
1985) and has since been applied to a wide variety of problems. voked in models of the reaction infiltration instability (Aharonov
In these models, melt percolates through the matrix according to et al. 1995; Spiegelman et al. 2001), a mechanism that may explain
Darcy’s law and the matrix resists compaction through an effective the focusing of melt into channels, promoting rapid transport of the
bulk viscosity. More recently, these two phase flow theories have melt.
been generalised to account for additional phenomena such as sur- Recent models of reactive transport have used somewhat ad hoc
linear kinetic laws to describe mass transfer between phases, but the
aim here is to provide a more rigorous treatment. For single com-
∗ Now at: Institute of Theoretical Geophysics, Bullard Laboratories, UK. ponent melting, two phase flow equations encompassing chemical


C 2010 The Authors 699
Geophysical Journal International 
C 2010 RAS
700 J. F. Rudge, D. Bercovici and M. Spiegelman

disequilibrium were derived by Sramek et al. (2007) (although no 1 Ds φ Ds (1/ρs ) 


∇ · vs = + ρs − . (7)
disequilibrium calculations were performed). Sramek et al. (2007) 1 − φ Dt Dt (1 − φ)ρs
invoked the theory of non-equilibrium thermodynamics (e.g. de
Similarly, the mean velocity v ≡ φv f +(1− φ)vs = φ(v f − vs ) + vs
Groot & Mazur 1984) to provide linear phenomenological laws for
satisfies
disequilibrium melting and we follow their approach in our gener-
alisation to multiple components. D f (1/ρ f ) Ds (1/ρs )
∇ · v = φρ f + (1 − φ)ρs − (1/ρ), (8)
The manuscript is organised as follows: First, we write down the Dt Dt
equations governing conservation of mass, momentum and energy. where (1/ρ) = 1/ρs − 1/ρ f . It is often more convenient to work
The set of governing equations are completed by equations of state with mass conservation in the form of (7) and (8) rather than (1)
and phenomenological laws for transfer of heat and mass between and (2).
and within the phases. The main focus here is on the phenomeno-
logical laws governing transfer of mass between the phases (i.e.
melting, crystallisation, or dissolution). Non-equilibrium thermo- 2.2 Components
dynamics provides a theoretical basis for linear phenomenological
laws, but it seems that non-linear laws are required if fractional The two phases are made up of n thermodynamic components

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


melting is to be described accurately. Finally, the theory is applied e.g. if the phases were pure olivine, component 1 could be Mg2 SiO4
to two well-studied model problems for a binary system: melting (forsterite) and component 2 could be Fe2 SiO4 (fayalite). Conser-
due to increasing temperature at constant pressure and melting by vation of components is
decompression in a upwelling 1-D column. Df cf
j
j j
φρ f + c f = −∇ · J f +  j , (9)
Dt
2 MASS
Ds csj
In what follows, subscripts refer to the individual phases (either the (1 − φ) ρs − csj = −∇ · Jsj −  j , (10)
Dt
fluid melt f or solid matrix s) and superscripts to the components
(1, 2, . . . , n). where the left hand sides of the above equations have been written
j
in Lagrangian form using (3) and (4). cf and csj are the concen-
trations by mass of component j in the two phases.  j represents
2.1 Phases the rate of mass exchange of component j from solid to fluid (in-
j
terphase exchange), whereas Jf and Jjs represent the diffusive mass
Conservation of mass for the two phases is
  fluxes of component j within the fluid and solid phases respectively
∂ φρ f   (intraphase fluxes). For consistency with the conservation of mass
+ ∇ · φρ f v f = , (1) eqs (1) and (2), the following constraints hold
∂t
 j 
∂ ((1 − φ)ρs ) cf = csj = 1, (11)
+ ∇ · ((1 − φ)ρs vs ) = −, (2) j j
∂t
where φ is the porosity (the volume fraction of melt), ρ f and ρs  j

are densities and vf and vs are velocities of the fluid and solid Jf = Jsj = 0, (12)
j j
respectively.  represents the total rate of mass exchange from solid
phase to fluid phase (the melting rate). It follows that given any 
scalar quantities af and as per unit mass we have = j. (13)
  j
∂ φρ f a f   Df af
+ ∇ · φρ f a f v f = φρ f + a f , (3) j
The first of these constraints simply states that cf and csj represent
∂t Dt
the compositions of the two phases and thus must sum to 1. Eq. (12)
∂ ((1 − φ)ρs as ) Ds as states that the diffusive intraphase fluxes must sum to zero and (13)
+ ∇ · ((1 − φ)ρs as vs ) = (1 − φ) ρs − as , states that the total rate of interphase mass transfer from solid to
∂t Dt
(4) liquid is given by the sum of the mass exchanges from solid to liquid
of the individual components.
where Df /Dt and Ds /Dt are Lagrangian derivatives following the
fluid and solid, respectively. It follows that
∂     3 MOMENTUM
φρ f a f + (1 − φ)ρs as + ∇ · φρ f a f v f + (1 − φ)ρs as vs
∂t Derivations of the equations governing conservation of momen-
Df af Ds as tum are much more involved than those governing conservation
= φρ f + (1 − φ) ρs − a, (5)
Dt Dt of mass and have been discussed in detail by many other authors
where a = as − a f . These expressions are useful in writing the (McKenzie 1984; Scott & Stevenson 1984; Fowler 1985; Schmel-
conservation equations to come in a more compact form, as through- ing 2000; Bercovici et al. 2001; Ricard et al. 2001; Bercovici &
out expressions will be cast in terms of Lagrangian derivatives. Ricard 2003; Simpson et al. 2010a,b). Only a brief outline of their
The mass conservation eqs (1) and (2) can be rewritten in La- derivation is given here.
grangian form as Conservation of momentum for the slow creeping flow of the two
phases is
1 Df φ Df (1/ρ f )   
∇ · vf = − + ρf + , (6)
φ Dt Dt φρ f ∇ · φσ f + φρ f g = F, (14)


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 701

∇ · ((1 − φ) σ s ) + (1 − φ) ρs g = −F, (15) These laws appear somewhat asymmetric due to an assumption that
the matrix is much more viscous that the melt. (14), (24), (25) and
where F is the interphase force per unit volume, representing the
(27) then lead to the usual Darcy’s law for the melt,
force one phase exerts on the other and g is the acceleration due
to gravity. σ f and σs are the stress tensors for the two phases. It   kφ  
φ v f − vs = − ∇ P − ρf g , (29)
is assumed that locally components within the same phase feel the μ
same stress, but that different phases feel different stresses. Hence
where the permeability kφ (a function of porosity) is related to the
only two momentum equations are needed (one for each phase),
drag coefficient by d = μφ 2 /kφ . (16) and (25)–(28) lead to a total
rather than the 2n equations that are needed to describe conservation
conservation of momentum equation
of mass (one for each phase and 2(n−1) for the components).   
Eqs (14) and (15) can be summed to give the total conservation    2
∇ P = ∇ · ηφ ∇vs + ∇vsT + ∇ ζφ − ηφ ∇ · vs + ρg,
of momentum equation 3
∇ · σ + ρg = 0, (16) (30)

where the average stress tensor and density are defined by which resembles the equation governing compressible Stokes flow,
where ηφ and ζφ can be interpreted as effective shear and bulk vis-

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


σ = φσ f + (1 − φ) σ s , (17) cosities for the two phase mixture (which are also porosity depen-
dent). More general phenomenological laws than those in (25)–(28)
ρ = φρ f + (1 − φ) ρs . (18) were developed by Bercovici & Ricard (2003) to preserve material
invariance and these laws are outlined briefly in Appendix A (also
Mechanical pressures for the two phases are defined in the standard
see discussion in Simpson et al. 2010a).
way by
1
pf = − tr σ f , (19)
3 4 E N E RG Y

1 When considering the thermodynamics of two phase flow, as we


ps = − tr σ s , (20) must when considering conservation of energy, a key difficulty is the
3
notion of pressure and in particular the difference between ‘thermo-
with a mean mechanical pressure given by
dynamic’ and ‘mechanical’ definitions of pressure. The thermody-
1 namic pressure appears in definitions of thermodynamic potentials
p=− tr σ = φp f + (1 − φ) ps . (21)
3 for example, in the relationship between enthalpy and internal en-
The stress tensors can be split into isotropic and deviatoric (trace- ergy, H = U + pV . The mechanical pressures are defined by minus
free) parts as one third the trace of the stress tensor, as in (19), (20) and (21). Even
for a compressible single phase viscous fluid there is difference be-
σ f = − pf I + τ f , (22) tween these two definitions of pressure. This difference depends on
the bulk viscosity and the divergence of the velocity field and disap-
σ s = − ps I + τ s , (23) pears in equilibrium. For a two phase system, the situation is much
where τ f and τ s are the deviatoric stress tensors and I is the identity less clear as there are multiple mechanical pressures and potentially
tensor. multiple thermodynamic pressures.
A simple phenomenological law for the interphase force F is In what follows all thermodynamic potentials for both phases are
given by (Drew 1983) defined with P, the interface pressure, as the appropriate thermo-
  dynamic pressure. There is some justification for this in the work
F = d v f − vs − P∇φ, (24) of Sramek et al. (2007) where it was shown that this leads to a
where d is a drag coefficient associated with resistance of motion particularly natural characterisation of phase change. For the phe-
of the two phases past each other and P is the interface pressure, nomenological laws considered, the difference between mechanical
which produces a net force if porosity spatially varies. Throughout and thermodynamic pressures only depends on the divergences of
this work we will neglect surface tension, but a generalisation of the velocity fields, in a way analogous to the compressible single
the above law to encompass surface tension effects can be found in phase case. Using P for the thermodynamic pressure is also iden-
Bercovici et al. (2001). tical to the assumption made by McKenzie (1984) where the fluid
To complete the set of equations, phenomenological laws are pressure (identical to the interface pressure, 25) was chosen as the
needed to determine the pressure differences P− pf and P− ps common thermodynamic pressure for the two phases. Nevertheless,
and the deviatoric stress tensors τ f and τ s . To be consistent with thermodynamic pressure remains a thorny aspect of two phase flow
the simplified two phase flow theory of McKenzie (1984), the phe- theories and deserves further careful study.
nomenological laws must take the form Conservation of total internal energy is (McKenzie 1984; de
Groot & Mazur 1984)
P − p f = 0, (25)
Df u f Ds u s
φρ f + (1 − φ) ρs − u
ζφ Dt Dt
P − ps = ∇ · vs , (26)  
1−φ = Q − ∇ · q + ∇ · φσ f · v f + (1 − φ) σ s · vs + φρ f v f · g
+ (1 − φ) ρ v · g, (31)
s s
τ f = 0, (27)
where u f and u s are the internal energies per unit mass of the
 
ηφ 2 two phases, Q is the rate of internal heat production (e.g. from
τs = ∇vs + ∇vsT − (∇ · vs ) I . (28)
1−φ 3 radioactivity), q is the diffusive heat flux and the remaining terms


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
702 J. F. Rudge, D. Bercovici and M. Spiegelman

on the right hand side are sources of energy due to work. Using the enthalpy equation (37) can be written as an entropy equation,
the momentum eqs ((14), (15) and (24), the energy equation can be Df sf Ds ss
simplified to T φρ f + T (1 − φ) ρs − T s
Dt Dt
 j 
Df u f Ds u s = Q−∇ ·q+ +
j
μ f ∇ · J f + μsj ∇ · Jsj +  j μ j ,
φρ f + (1 − φ) ρs − u = Q − ∇ · q − P∇ · v + ,
Dt Dt j
(32) (39)
where we have used conservation of components (9), (10). The
where is the viscous dissipation,
above expression can be written as an entropy balance,
 2  
= d vf − vs + φ P − p f ∇ · v f + (1 − φ) (P − ps ) ∇ · vs Df sf Ds ss
φρ f + (1 − φ) ρs − s = −∇ · j + σ, (40)
+ φτ f : ∇vf + (1 − φ)τ s : ∇vs . Dt Dt
(33)
where j is the entropy flux and σ is the entropy production. Compar-
With the simplified phenomenological laws (25)–(28), the viscous ing (39) and (40), we see that the entropy flux j is related to fluxes
dissipation can be written as of heat and components by
μφ 2  2

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


q  J f μ f + Jsj μsj
j j
= vf − vs + ζφ (∇ · vs )2 j= − (41)
kφ T T
 2 j
ηφ 2
+ ∇vs + ∇vs − I (∇ · vs ) .
T (34) and the entropy production σ is given by
2 3  
1  j j
Here squares represent a dot product for vectors a2 = a · a and a σ = Q + − j · ∇T +  j μ j − J f · ∇μ f − Jsj · ∇μsj .
T
double dot product for second rank tensors A2 = A : A. A more j
(42)
general expression for the viscous dissipation can be found in Sec-
tion A using the phenomenological laws of Bercovici & Ricard The second law of thermodynamics requires that σ ≥ 0 (also known
(2003). as the Clausius–Duhem inequality).
Conservation of energy (32) can be rewritten using conservation
of mass (8) as 5 E Q UAT I O N S O F S TAT E
Df u f Ds u s
φρ f + (1 − φ) ρs − u Equations of state need to be prescribed for the two phases. In
Dt Dt
 theory, this could be done using the internal energy, by specifying
D f (1/ρ f ) Ds (1/ρs ) j
functions u i (si , ρi , ci ) for the two phases. The temperature, pressure
= Q − ∇ · q − P φρ f + (1 − φ) ρs
Dt Dt and chemical potentials could then be derived from these
 functions
j j
using the Gibbs relation du i = T dsi − Pd(1/ρi ) + j μi dci or
−(1/ρ) + . (35)
∂u i ∂u i j ∂u i
T = , P =− , μi = . (43)
In applications it is useful to rewrite the energy equation in terms of ∂si ∂(1/ρi ) ∂ci
j

different thermodynamic potentials. For example, specific enthalpy j


However, thermodynamic data is not usually given in (si , ρi , ci )
satisfies
co-ordinates, but rather in (P, T , cij ) co-ordinates and it is helpful to
P Di h i Di u i 1 Di P Di (1/ρi ) re-express the equations of state. Such co-ordinates are also useful
hi = ui + , = + +P , (36)
ρi Dt Dt ρi Dt Dt since we are assuming a common temperature T and interface pres-
sure P for both phases. The co-ordinate change uses the following
where the subscript i refers to the phase, i= s, f . As discussed
standard partial derivatives in (P, T , cij ) co-ordinates,
above, the definition of specific enthalpy used here is in terms of the
interface pressure P. Using (36), conservation of energy (31) can ∂(1/ρi ) ∂(1/ρi )
αi = ρi , βi = −ρi , (44)
be written as an enthalpy equation ∂T ∂P
Df h f Ds h s ∂si ∂h i
φρ f + (1 − φ) ρs − h Ci = T = , (45)
Dt Dt ∂T ∂T
Df P Ds P
=φ + (1 − φ) + Q − ∇ · q + , (37) αi is the thermal expansion coefficient, βi is the isothermal com-
Dt Dt pressibility and Ci is the specific heat capacity at constant pressure.
which is the form of the energy equation used in the enthalpy method The dependence on composition is captured by introducing partial
(Katz 2008). We will return to this enthalpy equation in Section 6 specific quantities as
to write a temperature equation.  c j
j j 1
hi = ci h i , = i
, (46)
j
ρi j ρi
j

4.1 Entropy j
where 1/ρi is the partial specific volume of component j in
Perhaps the most important rewriting of the energy equation is as phase i and h ji is the corresponding partial specific enthalpy. Us-
an equation for entropy. Since ing (44)–(46), we may then write equations of state in terms of
 Lagrangian derivatives
1 Di P  j Di ci
j
j j Di h i Di si
h i = T si + μi ci , =T + + μi , Di (1/ρi ) Di T Di P  ρi Di ci
j
j
Dt Dt ρi Dt j
Dt ρi = αi − βi + , (47)
j
j ρi
(38) Dt Dt Dt Dt


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 703

1 − αi T Di P  j Di ci
j The temperature eq. (53) closely resembles the standard tempera-
Di h i Di T
= Ci + + hi . (48) ture equation for single phase flow. The two phase nature of the
Dt Dt ρi Dt j
Dt
flow appears in the averaging of specific heat capacities, thermal
Of course, the quantities αi , βi , etc. may be functions of P, T and expansivities and velocities and through the latent heat term. The
cij but are often assumed constant for simplicity. Equations of state multicomponent nature of the flow appears in the terms involving
j
are also needed to describe the chemical potentials μi and these are diffusive fluxes of components, the different latent heats for the
discussed later in Section 7. different components and the potential for composition dependent
properties.

6 T E M P E R AT U R E E Q UAT I O N S
7 CHEMICAL POTENTIALS
Using (47) we may rewrite the mass conservation eqs (6) and (7) as
To have a full description of the thermodynamics, we need to relate
Df P  j − ∇ · Jf
j chemical potentials to temperature, pressure and composition. It is
1 Df φ Df T
∇ · vf = − + αf − βf + , an unfortunate fact that when working with chemical potentials we
φ Dt Dt Dt j
φρ f
j have to deal with both mass fractions (denoted by c j ) and mole

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


(49)
fractions (denoted by x j ). It is straightforward to convert between
1 Ds φ Ds T Ds P   j + ∇ · Jsj the two sets of variables:
∇ · vs = + αs − βs − , M jx j Mc j
1 − φ Dt Dt Dt j (1 − φ)ρs
j
cj = , xj = , (58)
(50) M Mj
where the derivatives of composition have been removed using con- where M j are the molar masses of each component (kg mol−1 ) and
servation of components (9), (10). The corresponding equation for M is the mean molar mass, given by
the mean velocity (8) becomes ⎛ ⎞−1
  cj
Df T Ds T Df P Ds P M= M jx j = ⎝ ⎠ . (59)
∇ · v = φα f + (1 − φ)αs − φβ f − (1 − φ)βs Mj
Dt Dt Dt Dt j j


 ∇ · Jf
j
∇ · Js j
Another way of expressing these relationships is by
−  j (1/ρ j ) + + .
ρ
j
ρs
j    
j f cj = M jx j , x j = c j /M j , (60)
 j
Using (48), the conservation of energy eq. (37) can be rewritten where {·} refers to normalising to unit sum, a = a j / k a k .
as The chemical potentials are related to temperature, pressure and
Df T Ds T Df P composition by the standard relationships (Spear 1993; Anderson
φρ f C f + (1 − φ) ρs Cs − φα f T 2005)
Dt Dt Dt
Ds P  ν
− (1 − φ) αs T μsj = μsj◦ (P, T ) + R j T log γsj xsj , (61)
Dt
   ν
 j j
= Q+ −∇ ·q +  j h j − J f · ∇h f − Jsj · ∇h sj , j j◦ j j
μ f = μ f (P, T ) + R j T log γ f x f , (62)
j (51)
where again the derivatives of composition have been removed using where γ j are activity coefficients and x j are molar concentrations.
conservation of components (9), (10). q is an alternative definition The activity model used here (γ j x j )ν is a non-ideal one-site sub-
of heat flux, related to the original fluxes by stitution model suitable for minerals such as olivine: for activity
 j j models for more complex assemblages the reader is encouraged to
q = q − J f h f + Jsj h sj . (52) consult the textbooks (e.g. Chapter 7 of Spear 1993). The activity
j coefficients γ j = 1 for an ideal solution and ν is the number of
lattice sites per formula unit (e.g. ν = 2 for olivine (MgFe)2 SiO4 ).
The temperature eq. (51) can also be written in terms of averaged
R j is the specific gas constant for component j, R j = R̃/M j ,
quantities,
where R̃ is the universal gas constant ( R̃ = 8.314472 J K−1 mol−1 )
∂T ∂P and M j is the molar mass. This is consistent with the earlier def-
ρC + ρCv · ∇T − T α − T αv · ∇ P
∂t ∂t inition of the chemical potential as being per unit mass (the units
  
= Q + − ∇ · q −
j j
 j L j + J f · ∇h f + Jsj · ∇h sj , (53) of μ are J kg−1 ). The chemical potentials per mole are given by
μ̃sj = μs M j , μ̃ f = μ f M j (with units J mol−1 ).
j
j
j
The differences in chemical potentials, μ j = μsj − μ f satisfy
j
where L j = −h j = h f − h sj is the latent heat (enthalpy of fusion)
for melting of component j and the overbars represent averages as μ j = −R j T log K j + R j T log Q j , (63)

ρC = φρ f C f + (1 − φ)ρs Cs , (54) where K j are the equilibrium constants, functions only of temper-
ature and pressure, defined by
ρCv = φρ f C f v f + (1 − φ)ρs Cs vs , (55) −R j T log K j = μsj◦ − μ f
j◦
(64)
and Q j are the activity ratios, defined by
α = φα f + (1 − φ)αs , (56)
ν
γsj xsj
Q =
j
. (65)
αv = φα f v f + (1 − φ)αs vs . (57) j j
γf xf


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
704 J. F. Rudge, D. Bercovici and M. Spiegelman

(63) can be rewritten as describes the temperature dependence


 
Qj μ j ∂ log K j h j
= exp . (66) = j 2, (74)
Kj RjT ∂T R T
In equilibrium μ j = 0 and Q j = K j . where h j is the change in enthalpy for melting of pure component
It is often more desirable to work with concentration ratios rather j. The pressure dependence is
than activity ratios. If we define K xj and Q jx by ∂ log K j (1/ρ j )
=− . (75)
γf 
j
1/ν xsj ∂P RjT
K xj = j
Kj , Q xj = j
, (67) It is helpful to look at a simplified form of these dependencies. For
γs xf
example, if we assume that h j is independent of temperature and
where Q jx is a molar concentration ratio, then Q jx = K xj in equilib- pressure, then a suitable approximate expression is (Bradley 1962)
rium. We can relate Q jx and K xj to the chemical potential differences  
h j 1 1
by log K = − j
j
− j . (76)
R T Tm (P)
 j 1/ν  
Q xj Q μ j

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


= = exp . (68) Tmj (P) is the melting temperature of the pure component as a func-
Kx
j
Kj νRjT tion of pressure. Recall that h j ≡ −L j , where L j is the latent
heat. In terms of K xj , (76) can be rewritten as
  h j  1 1

j
log K xj = log γ f /γsj − − . (77)
7.1 Solidus and liquidus surfaces νRj T j
Tm (P)
Since Q jx = K xj in equilibrium, the equilibrium molar compositions The function Tmj (P) satisfies the Clapeyron equation,
x j s(eq) and x j f (eq) satisfy
 j  j dTmj T j (1/ρ j )
j j = m , (78)
xs(eq) = K xj x f (eq) , xs(eq) = 1, x f (eq) = 1. (69) dP h j
The permissible values of x j s(eq) and x j f (eq) describe the solidus and which could formally be integrated to determine Tmj (P). However, it
liquidus surfaces, respectively. is often easier to use an approximate parameterised form for Tmj (P),
If K x is a function only of temperature and pressure (as is the such as Simon’s equation,
case for an ideal solution and will be assumed from here on), the   j
j P 1/b
two surfaces can be described separately as follows. The solidus is Tmj (P) = Tm0 1 + j , (79)
a
given by those molar compositions x j s(eq) satisfying
for some coefficients a j and b j . This is the approach taken here.
 xs(eq)
j
 j
j
= 1, xs(eq) =1 (70)
j Kx
8 P H E N O M E N O L O G I C A L L AW S F O R
and the liquidus is given by those molar compositions x j f (eq) satis- INTERPHASE MASS TRANSFER
fying To complete the governing equations, phenomenological laws are
  j
j
K xj x f (eq) = 1, x f (eq) = 1. (71) required describing the fluxes of mass and heat. In this section, we
j focus on the phenomenological laws for interphase mass transfer
and the corresponding discussion for intraphase vector fluxes can be
Eqs (70) and (71) embody the Gibbs’ phase rule: For a two-phase found in Section B. The simplest closure is to assume local thermo-
n-component system the phase rule states that there are n thermody- dynamic equilibrium (e.g. Ribe 1985a; Hewitt & Fowler 2008; Katz
namic degrees of freedom in equilibrium. The n degrees of freedom 2008; Tirone et al. 2009). This adds n algebraic equations to the
could be T, P and n − 2 of the components of xsj . In a binary system j
system, for example, of the form xsj = K x xf (see 69) and constrains
n = 2 and the equilibrium compositions can be completely specified the solid and liquid to lie on the solidus and liquidus on the phase
by T and P. The solidus is given by (70), diagram. The interphase mass transfers are then implicitly deter-
1 2 mined. However, in this work we are interested in disequilibrium
xs(eq) xs(eq)
+ = 1, 1
xs(eq) + xs(eq)
2
=1 (72) effects and thus do not assume local thermodynamic equilibrium.
K x1 (P, T ) K x2 (P, T )
According to the theory of non-equilibrium thermodynamics (de
and the liquidus by (71), Groot & Mazur 1984), linear phenomenological laws can be ob-
K x1 (P, T )x 1f (eq) + K x2 (P, T )x 2f (eq) = 1, x 1f (eq) + x 2f (eq) = 1. tained by examining the expression for entropy production. For
example, the part of the entropy production due to transfer of com-
(73) ponents between phases is given in (42) as
In each case there are two simultaneous equations for two unknowns,  μ j
which can be solved uniquely. These expressions are used to calcu- σ = j . (80)
late the solidus and liquidus surfaces for olivine in Fig. 1. j
T

This expression defines a natural set of conjugate thermodynamic


7.2 Temperature and pressure dependence forces (μ j /T ) and fluxes ( j ) and suggests linear phenomeno-
of equilibrium constants logical laws of the form
   k 1/ν
The temperature and pressure dependence of the equilibrium con- μk Q
j = E jk = E jk ν R k log k
, (81)
stants K j (P, T ) need to be prescribed. The van’t Hoff equation k
T k
K


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 705

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


Figure 1. Examples of isobaric disequilibrium melting of olivine. Each panel is plot of temperature T against composition c. The grey lines in the background
show the olivine binary phase loop (the equilibrium solidus and liquidus curves). c = 1 is pure forsterite, c = 0 is pure fayalite. The dotted line indicates the
bulk composition, c = 0.41. Red curves show the solid composition, blue curves show the liquid composition. Each panel is for a particular style of melting
(either type I or type II) for the range of Damköhler numbers indicated. In each case, dark shading indicates small Damköhler number, light shading indicates
large Damköhler number. (a) and (b) are examples of type I melting, (c) and (d) are examples of type II melting and (e) and (f) are the same for both type I and
type II melting (see discussion in text).

for some matrix of coefficients E jk . Onsager’s reciprocal relations Here we propose a simple non-linear law for interphase mass
state that E jk is a symmetric matrix, E jk = E k j and the second transfer that can be thought of as a generalisation of fractional
law ensures that E jk is positive semi-definite. More precisely, the melting. We introduce the law first and then explain why it may be
theory states that any scalar flux can depend on any scalar force reasonable. The law we propose is
in the entropy production and this could include a dependence of 
 j on velocity field divergences. However, we will neglect such j = E jk Z k , (82)
k
a dependence here and assume the interphase mass transfers only
depend on differences in chemical potential (see Sramek et al. 2007,
 1/ν

Kk K xk x kf
for further discussion). Z k = ν Rk 1 − = ν Rk 1 − , (83)
Far from equilibrium, linear laws may be a poor descrip- Qk xsk
tion of the kinetics. Indeed, this is well known in the context which is very similar to (81). Again E jk is a matrix of coefficients,
of chemical reactions, where the law of mass action (based on which we will assume is symmetric and positive semi-definite (see
products of activities) is generally a more useful phenomeno- Subsection 8.3 for further discussion of the second law). Z k are the
logical law far from equilibrium than linear laws based on the thermodynamic forces which are now given by a non-linear law in
affinities (differences in chemical potential). This motivates the terms of the equilibrium constants K k and activity ratios Q k . The
exploration of laws for interphase mass transfer that are non- dependence on the ratio K k /Q k is found in many kinetic theories,
linear in the chemical potential differences, but that must still such as transition state theory (Anderson 2005; Lasaga 1998). Near
satisfy the constraints of non-equilibrium thermodynamics near equilibrium this non-linear phenomenological law reduces to the
equilibrium. linear law of (81), as Z k ≈ μk /T near equilibrium.


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
706 J. F. Rudge, D. Bercovici and M. Spiegelman

To make the link with fractional melting, it is useful to express j j xsj


the fluxes in a different way, separating out the mass flux associated  j = c , x = j
, (89)
Kx
with phase change () from the fluxes of components that occur
without phase change (Baker & Cahn 1971; Caroli et al. 1986; where (88) constrains the solid to lie on the solidus and (89) states
j
Hillert 2006). Key to this alternative representation is the fact that that the melt produced is in equilibrium with the solid. x represents
j
the symmetric positive semi-definite matrix E jk can be uniquely the molar concentrations of the infinitesimal melt produced and c
j j
decomposed as the corresponding mass concentrations (x and c are related by 58).
In (89) phase change is the only form of mass transfer; comparison
j
E jk = λ c ck + G jk , (84) with (85) shows that the J j (the fluxes of components without phase
j change) are all zero.
where λ > 0, c
is a well-defined composition (i.e. it has unit sum
We would like to relate the laws given by (88)–(89) to those of
and positive entries) and G jk is a positive semi-definite matrix with
j (85)–(87). It is clear that to reproduce fractional melting the coeffi-
zero row and column sum. The subscripts  on λ and c are to
cients G jk should be zero, but is less clear how the coefficients λ
emphasise their association with the phase change, which is clearer j
and c should be chosen. Eq. (89) motivates choosing the compo-
when the phenomenological laws (82) and (83) are rewritten using
sition associated with phase change as
(84) as

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


 j 
j xs
j
 j = c + J j , (85) x = j
(90)
Kx
 where {·} again refers to normalising to unit sum. Note that normal-
 = λ ck Z k , (86)
k
isation is required, since xsj /K xj is not guaranteed to be a valid com-
position (i.e. sum to 1) as the solid is not necessarily constrained
 to the solidus for general disequilibrium phenomenological laws.
Jj = G jk Z k . (87) The phenomenological law describing phase change (86) and (83)
k
becomes
Eq. (85) splits the mass flux of component j into

j twoj parts: a part  K xk x kf
that occurs due to phase change (c , since j c = ) and a  = λ ck ν R k 1 − (91)
xsk
part that occurs without phase change (J j , since j J j = 0). The k
flux due to phase change is controlled by the coefficients λ (a rate

j
constant for phase change) and c , whereas the flux of components  xk K xk x kf
that occurs without ∝ λ s
1− (92)
phase change is controlled by the coefficients K xk xsk
G jk since j k G jk Z k = 0 . This decomposition into fluxes with k

and without phase change is useful because the different mass fluxes
 
may be controlled by quite different physical mechanisms and thus  xk
∝ λ s
−1 (93)
occur at different kinetic rates. k
K xk
Whether written as (82) or (85)–(87), there remains a set of
n(n + 1)/2 coefficients to be prescribed, either in the form of the and thus the constraint that the solid lies on the solidus is recovered
n(n + 1)/2 independent entries of the matrix E jk or as λ , the as λ → ∞. Hence fractional melting is recovered  with
 the coeffi-
j
j
n − 1 independent entries of c and the n(n − 1)/2 independent cient choice of λ → ∞, G jk = 0 and x = xsj /K xj . Indeed, the
jk
entries of G . Either set can be specified and the other set is then main reason for choosing the particular non-linear form in (83) is to
determined. In this work, we will focus on two natural choices for exactly recover the solidus constraint, which is only approximately
j
the composition c and will refer to the two laws as type I and type recovered with thelinear law. We will refer to any melting law which
j
II melting. One of these laws generalises melting laws which leave has x = xsj /K xj as type II melting. A similar generalisation can
the solid composition unchanged (type I melting) and the other be made for fractional crystallisation and is discussed in Section C.
generalises fractional melting (type II melting).
8.2 Solid invariant melting (Type I melting)
j
8.1 Fractional melting (Type II melting) Another natural choice for the composition x is the solid compo-
j
sition itself, x = xsj , which we will refer to as type I melting. If the
During fractional melting the solid remains on the solidus and each
coefficients G jk are zero, then this style of melting keeps the solid
infinitesimal increment of melt produced is in equilibrium with
composition fixed. If λ → ∞, this style of melting constrains the
the solid. Once the infinitesimal melt has formed it is chemically
liquid to the liquidus.
isolated from the solid and no further chemical exchange occurs (al-
The two styles of melting we consider here (type I and type II)
though heat is still exchanged as thermal equilibrium is assumed).
are not the only styles of melting one could consider: they simply
We will refer to this style of fractional melting as thermally equi-
represent two natural choices for the composition associated with
librated fractional melting (which differs from the incrementally
melting based on the solid composition. There is plenty of scope
isentropic fractional melting described by Asimow et al. (1997) and
for further exploration of these laws.
Stolper & Asimow (2007), which does not involve thermal equili-
bration).
The laws for fractional melting can be written as (Spiegelman & 8.3 The positivity of entropy production
Elliott 1993; Spiegelman 1996)
The second law of thermodynamics demands that the entropy pro-
 xk
s
= 1, (88) duction σ is positive. For a linear phenomenological law, such as
k
K xk that given by (81), it is straightforward to derive conditions on the


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 707
  
matrix of phenomenological coefficients E jk such that positive en- L1 1 1
tropy production is assured. Since the entropy production (80) for a K x1 = exp − 1 , (99)
ν R1 T Tm (P)
linear law is simply
  
1   jk L2 1 1
σ = 2 E μ j μk , (94) K x2 = exp − 2 , (100)
T j k ν R2 T Tm (P)

a positive entropy production is assured for all μk if and only if   1   2


P 1/b P 1/b
E jk is a positive semi-definite matrix. Tm1 (P) = Tm0
1
1+ 1 , Tm2 (P) = Tm0
2
1+ 2 ,
With non-linear phenomenological laws the situation is not so a a
straightforward: even if E jk is a positive definite matrix, it is not (101)
necessarily the case that the resulting entropy production is positive    
and other constraints must be placed on the coefficients E jk . For the K1 K2
Z 1 = ν R 1 1 − x1 , Z 2 = ν R 2 1 − x2 , (102)
melting laws given by (82) and (83), the entropy production (80) Qx Qx
can be written in terms of chemical potentials as  
     = λ c Z 1 + (1 − c )Z 2 , (103)

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


μk μ j
σ = E jk ν R k 1 − exp − k , (95)
νR T T  
j k J = λJ Z 2 − Z 1
, (104)
or more simply as
    1 = c  − J, (105)
E ∗ jk 1 − e−d d j ,
k
σ = (96)
j k
 2 = (1 − c )  + J, (106)
∗ jk
where d = μ /ν R T and E = ν R R E . If (96) could be
j j j 2 j k jk
where the composition variable x refers to the molar concentration
linearised 1 − e−d ≈ d k , then the entropy production is assured
k
of component 1 (forsterite): the molar concentration of component
to be positive if E ∗ jk is a positive semi-definite matrix. But for 2 (fayalite) is 1− x. Numerical values for model olivine parameters
the general non-linear law, having E ∗ jk be a positive semi-definite
can be found in Table 1. Note that the kinetic rates λ , λ J > 0 to
matrix is not enough to guarantee positive entropy production. satisfy the second law. λ and λ J could depend on many variables
There are certain simple forms for E ∗ jk that will guarantee a such as temperature or porosity, but for simplicity, here λ and λ J
positive entropy production for the non-linear law. For example, will be assumed constant.
if E ∗ jk is diagonal, with positive entries on the diagonal, then the For type I melting,
entropy production is positive because the function f (x) = x(1−
e−x ) is always positive. If E ∗ jk = λc j ck for some composition c j x = xs , (107)
and constant λ > 0, then the entropy production will be positive
j whereas for type II melting,
provided all the
d k have −dthe same sign, since the two quantities
k
j j
j c d and k c (1 − e ) are then either both positive or both xs /K x1
x = . (108)
negative. Having all the d j the same sign is a common occurrence xs /K x1 + (1 − xs )/K x2
and is true for the example calculations throughout this work. For
Consider heating a piece of olivine at fixed pressure such that
a binary system the entropy production is always positive for ar-
its temperature rises at a constant rate. Let F be the mass fraction
bitrary positive semi-definite matrices E ∗ jk when the d j have the
of melt (initially zero). The evolution of F and composition cf are
same sign, because matrices of the form in (97) also always give
described by
a positive entropy production (recall the decomposition in 84). For
more general cases, the positivity of the entropy production with dF 
= , (109)
this non-linear law can not be taken for granted. If the d j differ in dt ρ0
sign the entropy production may go negative with the present non-
linear law: this implies other non-linear laws are needed to describe d(Fc f ) 1
such circumstances and must be developed. = , (110)
dt ρ0
where ρ0 is the initial density of the unmelted solid. In non-
9 I S O B A R I C B I N A RY M E LT I N G dimensional form this simple system can be described in terms
of temperature as
As a concrete example in which to explore the behaviour of these
phenomenological laws, we now consider the problem of isobaric dF
= , (111)
melting of olivine. Our model olivine is a complete solid solution of dT
two components, forsterite (Mg2 SiO4 ) and fayalite (Fe2 SiO4 ), with
a simple binary loop phase diagram. For a binary system the matrix d(Fc f )
= 1 , (112)
G jk has only one free parameter and takes the form dT
   
1 −1 K x1 R2 K2
G jk = λ J , (97) Z1 = 1 − , Z2 = 1 − x2 ,
−1 1 (113)
Q 1x R1 Qx
where λ J > 0. In full, the phenomenological laws for melting are
 
thus  = Da c Z 1 + (1 − c )Z 2 , (114)
xs 1 − xs  
Q 1x = , Q 2x = , (98)
xf 1− xf J = Da J Z 2 − Z 1 , (115)


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
708 J. F. Rudge, D. Bercovici and M. Spiegelman

Table 1. Approximate parameters for the forsterite–fayalite binary system, (MgFe)2 SiO4 . There are two atoms of (Mg,Fe) per formula unit and thus ν = 2.
The latent heats are chosen to match the phase diagram using the parametrisation of Bradley (1962), although here the specific heat capacities are assumed
identical for solid and liquid. The melting temperature as a function of temperature has been parametrised by Simon’s equation. Thermal expansion coefficients
are assumed identical for the two components.
j = 1: Forsterite j = 2: Fayalite Notes
Formula Mg2 SiO4 Fe2 SiO4
Latent heat of fusion Lj 8.71 × 105 J kg−1 5.17 × 105 J kg−1 Bradley (1962)
Melting temperature T j m0 2163 K 1478 K at 0.1 MPa, Bradley (1962)
Simon’s equation coefficient aj 10.83 GPa 15.78 GPa Bottinga (1985), Poirier (2000)
Simon’s equation coefficient bj 3.70 1.59 Bottinga (1985), Poirier (2000)
Molar mass Mj 0.1406934 kg mol−1 0.2037736 kg mol−1
Specific gas constant Rj 59.096 J K−1 kg−1 40.803 J K−1 kg−1
3222 kg m−3 4392 kg m−3
j
Solid density ρs Deer et al. (1992)
3053 kg m−3 4026 kg m−3
j
Melt density ρf Calculated from Clapeyron relation
2.6 × 10−5 K−1 2.6 × 10−5 K−1
j
Solid thermal expansion coefficient αs Fa at 298-1123 K, Ahrens (1995)

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


4.8 × 10−5 K−1 4.8 × 10−5 K−1
j
Melt thermal expansion coefficient αf Fa at 1573-1873 K, Agee (1992)
Specific heat capacity Cj 1.3 × 103 J K−1 kg−1 1.3 × 103 J K−1 kg−1 Fo at 1500 K, Ahrens (1995)

 1 = c  − J, (116) Fig. 1(c) is perhaps the most important of the panels in Fig. 1
and demonstrates the transition from batch to fractional melting,
where the temperature has been non-dimensionalised by a typical showing type II melting for large Da and a range of Da J . The
temperature scale T , taken here to be the difference between the large Da constrains the solid to the solidus (the constraint imposed
melting points of the two components T = Tm1 − Tm2 . The mass by type II melting), but the liquid composition deviates from the
exchanges  and  1 are non-dimensionalised on T /(ρ0 dT /dt), liquidus as DaJ is reduced. The end member case of DaJ = 0
where d T /d t is the (assumed constant) rate of temperature increase. (dark shading) is essentially the fractional melting path (e.g. Maaløe
The behaviour of this simple system is determined by two non- 1984). The corresponding melt production for this case is shown
dimensional Damköhler numbers, in Fig. 2, which can be determined from Fig. 1(c) using the lever
λ ν R 1 T λ J ν R 1 T rule.
Da = , Da J = , (117)
ρ0 (dT /dt) ρ0 (dT /dt)
which are ratios of the kinetic rates to the rate at which the temper-
ature increases.
Fig. 1 gives examples of the different styles of disequilibrium
melting that occur as the two Damköhler numbers are varied for
the two types of melting. If both Damköhler numbers are large,
Da → ∞ and Da J → ∞, the kinetic rates are much faster than
the rate at which the temperature increases and thus the melting
occurs in equilibrium (batch melting). In equilibrium, the solid
composition lies on the solidus and the liquid composition lies on
the liquidus. Equilibrium melting begins when the bulk composition
intersects the solidus and completes when the bulk composition
intersects the liquidus. To avoid a singularity near the onset of
melting, the calculations of Fig. 1 assume there is initially a very
small melt fraction (F = 10−10 ) present before melting begins with
a composition in equilibrium with the solid.
Fig. 1(a) shows the effect of reducing Da J while keeping Da
large for type I melting. Since Da is large the liquid composition
remains on the liquidus (the constraint imposed by type I melting).
But the solid composition deviates the solidus: in the end-member
case of DaJ = 0 (dark shading), the solid composition is invariant
with changing temperature (since  j = csj in this limit). With
small DaJ and large Da , the lever rule (conservation of mass)
implies that there is little melting until the bulk composition comes
close to crossing the liquidus.
Fig. 1(b) shows the effect of then reducing Da while keeping
DaJ = 0 for type I melting. With Da large, the situation is the
same as seen in Fig. 1a with the solid invariant and the liquid on the
liquidus. The solid remains invariant as Da is varied (since DaJ
= 0 and thus  j = csj ) and for the end member case of Da = Figure 2. (a) Mass fraction of melt F as a function of temperature T for type
Da J = 0 (dark shading) there is no mass transfer whatsoever and II isobaric melting with Da = 105 and varying DaJ (see Fig. 1c, and also
both liquid and solid composition remain invariant with increasing Fig. 5 of Asimow et al. 1997). (b) Corresponding isobaric melt productivity
temperature. dF/dT .


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 709

Fig. 1(d) shows the effect of then reducing Da while keeping dvs − P̃
DaJ = 0 for type II melting. With Da large, the situation is the = , (124)
dz ζφ + 43 ηφ
same as seen in Fig. 1(a) with the solid on the solidus and the liquid
following the fractional melting path. Similar to Fig. 1(b), for the dT dP
end member case of Da = Da J = 0 (dark shading) there is no ρCv = T αv − 1 L 1 − 2 L 2 , (125)
dz dz
mass transfer whatsoever and both liquid and solid composition
remain invariant with increasing temperature. where the viscous dissipation term has been neglected in the
Figs 1(e) and (f) demonstrate situations that are essentially inde- energy eq. (125) (see Asimow 2002, for justification). The quantity
pendent of the choice of c and thus the same for both type I and P̃ is a compaction pressure, defined by −σ zz = P + P̃. cf and cs are
type II melting. Fig. 1e has Da J large and thus there is the approxi- the concentrations by mass of component 1: the concentrations of
mate constraint that Z 1 = Z 2 and hence  = λ Z 1 , independent of component 2 are given by c2f = 1 − cf , c2s = 1 − cs . Constitutive laws
c . With Da large the situation is essentially batch melting, while for viscosities and permeabilities are assumed to take the simplified
the other end member of Da = 0 (dark shading) has no melting, forms
 
although the solid and liquid can still exchange components without 4 4
phase change. Since the initial melt fraction is exceedingly small ζφ + ηφ = ζ0 + η0 φ −m , (126)
3 3

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


(F = 10−10 ) the solid composition is buffered and is invariant as the
temperature is increased, while the liquid composition undergoes a kφ = k0 φ n , (127)
notable shift.
Fig. 1(f) has Da = 0 and thus no melting ( = 0). Hence where the typical choices of the exponents are m = 1 (Simpson
 1 = − 2 = λ J (Z 2 − Z 1 ) independent of the choice of c . For et al. 2010a) and n = 3 (Kozeny–Carman equation). The above eqs
large Da J the situation is the same as that seen in Fig. 1(e), with an (118)–(127) together with the phenomenological laws for melting
invariant solid composition and a shifting liquid composition. For (98)–(106) form a closed set of governing equations.
the other end member case of Da = Da J = 0 (dark shading) there Most of the boundary conditions for this problem are at the onset
is once again no mass transfer whatsoever, with both liquid and solid of melting: the incoming material there has zero porosity, density
composition remaining invariant with increasing temperature. ρ0 , upwelling velocity v 0 and composition c0 . T and P are chosen
such that the system is initially at the onset of melting (either T or
P can be set and the other is then determined from the solidus). The
1 0 1 - D B I N A RY M E LT I N G C O L U M N final boundary condition concerns the momentum equations and
we choose d P̃/dz = 0 on the top boundary (a free flux condition
One of the simplest problems combining compaction and melting Spiegelman (1993a); see Sramek et al. (2007) and Hewitt & Fowler
is the 1-D steady state melting column. The problem was first ad- (2008) for further discussion of possible boundary conditions).
dressed using the McKenzie (1984) equations by Ribe (1985a) and Conservation of mass places strong constraints on 1-D steady
has since been studied by many other authors (Asimow & Stolper state melting (Ribe 1985a; Spiegelman & Elliott 1993; Asimow &
1999; Spiegelman & Elliott 1993; Sramek et al. 2007; Hewitt & Stolper 1999). On summing (118) and (119) and integrating, we
Fowler 2008; Katz 2008). Here we investigate the effects of dis- have
equilibrium on this problem. Again we consider melting of pure
olivine, although it should be noted that real mantle melting in- φρ f v f + (1 − φ)ρs vs = ρ0 v0 . (128)
volves a multiphase assemblage of a number of minerals. Olivine is Similarly, from (120) and (121),
the most abundant mantle mineral by mass, but it is not the mineral
that melts most during real mantle melting. Nevertheless it pro- φρ f c f v f + (1 − φ)ρs cs vs = ρ0 c0 v0 . (129)
vides a simple binary test system for studying the behaviour of the The governing eqs (118)–(127) can be simplified by introducing the
equations. degree of melting F, which can be defined as a ratio of melt flux to
In 1-D, with no diffusion of heat or components, the conservation incoming mass flux,
eqs (1), (2), (9), (10), (29), (30) and (53) are
φρ f v f
d   F= , (130)
φρ f v f = , (118) ρ0 v0
dz
where F = 0 at the onset of melting and F = 1 once melting is
complete. It follows from (128)–(130) that
d
((1 − φ) ρs vs ) = −, (119)
dz Fc f + (1 − F)cs = c0 . (131)
In the case of equilibrium melting, the conservation of energy equa-
d  
φρ f v f c f =  1 , (120) tion implies that total entropy flux is conserved along the column
dz and a similar equation to (131) can be written for entropy (i.e. Fsf
+ (1− F) ss = s 0 , Asimow & Stolper 1999). The isentropic nature
d of the equilibrium process simplifies the analysis greatly (Asimow
((1 − φ) ρs vs cs ) = − 1 , (121)
dz et al. 1997; Asimow & Stolper 1999; Stolper & Asimow 2007).
However, when there is disequilibrium between the two phases,
dP d P̃ there is the potential for entropy production and we cannot make
= −ρg − , (122)
dz dz this simplification here.
In terms of F, the governing eqs (118)–(127) are
d P̃ μ   dF 
= φ v f − vs − (1 − φ) ρg, (123) = , (132)
dz kφ dz ρ0 v0


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
710 J. F. Rudge, D. Bercovici and M. Spiegelman

dm 1 Table 2. Non-dimensional parameters used in the 1-D melting column


= , (133) calculations of Fig. 4 (defined in 140–144).
dz ρ0 v0
Adiabatic parameter A 0.012
dP d P̃ Buoyancy parameter B 49,000
= −ρg − , (134) Compaction parameter C 260
dz dz
Expansivity ratio rα 1.8
Density ratio rρ 0.98
d P̃ μ Specific gas constant/ specific heat capacity R1 0.091
= q − (1 − φ) ρg, (135)
dz kφ Specific gas constant/ specific heat capacity R2 0.063
Stefan number St 1 0.98
− P̃ Stefan number St 2 0.58
dvs
= , (136) Damköhler number Da 50
dz ζφ + 43 ηφ Damköhler number DaJ 0 or 50

  dT
FC f + (1 − F)Cs Table 3. Dimensional parameters used in the 1-D melting column calcula-
dz

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


  tions of Fig. 4.
αf αs dP 1 L 1 2 L 2
=T F + (1 − F) − − , (137) Acceleration due to gravity g 9.8 m s−2
ρf ρs dz ρ0 v0 ρ0 v0 Permeability coefficient k0 10−7 m2
where m = Fcf is the mass of component 1 in the fluid phase.
Permeability exponent n 3
The notation q = φ v f − vs is used to signify the Darcy flux. Fluid viscosity μ 1 Pa s
Conservation of mass (128)–(131) implies Bulk viscosity coefficient ζ0 + 43 η0 1019 Pa s
Bulk viscosity exponent m 1
ρ0 v0 (1 − F) ρ0 v0 ρ0 v0 Upwelling velocity v0 50 mm yr−1
vs = , q=F + (1 − F) − vs , (138)
ρs (1 − φ) ρf ρs Bulk composition c0 0.9

m c0 − m
cf = , cs = . (139) A is an adiabatic parameter, which is the product of the adiabatic
F 1− F gradient and the column length scale (sometimes termed the dis-
Formally the densities ρs and ρ f could vary as a function of tem- sipation number). B is a buoyancy parameter, which is essentially
perature, pressure and composition. In practice, these variations are a ratio of percolation velocity to upwelling velocity, although the
slight (i.e. much less than the difference in density between the two φ dependence is missing. C is a compaction parameter, a ratio of
phases) and can be neglected with the exception of the adiabatic compaction length squared to column length squared. The inverse
term in the energy equation. Thus we approximate ρs and ρ f as of C is sometimes referred to as a melt retention number (Tackley
constants in the equations to follow. Since we assume initial zero & Stevenson 1993). rα is a ratio of expansivities and rρ a ratio of
porosity, ρ0 = ρs . We will assume that the specific heat capacities densities. R1 and R2 are ratios of specific gas constants to specific
of solid and fluid are constant and equal and thus on the left hand heat capacity. St1 and St2 are Stefan numbers, which are a ratio of
side of (137), FCf + (1− F) Cs = C. latent heat to sensible heat for the two components. Finally, Da
and DaJ are Damköhler numbers, which are a ratio of reaction rates
to upwelling rates.
10.1 Non-dimensionalisation Values of the non-dimensional parameters used in the calcula-
For numerical solution and further analysis it is helpful to non- tions are given in Table 2, based on dimensional parameters in
dimensionalise the equations. If z is a typical length scale (e.g. the Tables 1 and 3. The adiabatic parameter A is small, which justifies
height of the melting column) and T is a typical temperature the neglect of the adiabatic term by some authors (Sramek et al.
scale (e.g. the difference in melting temperatures (T = Tm1 − Tm2 ), 2007). The buoyancy parameter B is very large which indicates that
then we can non-dimensionalise as z = z  z, T = T  T, P = the melt velocities will be much larger than the mean upwelling ve-
P  ρs gz, P̃ = P̃  (ζ0 + 43 η0 )v0 /z,  =   ρs v0 /z, V = locity. The density ratio rρ is close to 1, which some authors exploit
V  v0 , q = q  v0 , etc. The following non-dimensional parameters by making a Boussinesq approximation (Ribe 1985a), although we
control the behaviour of the system: do not do this here.
  The Damköhler numbers used in the calculations are designed
αs gz k0 ρg k0 ζ0 + 43 η0 to demonstrate the effects of disequilibrium rather than necessar-
A= , B= , C= , (140)
C μv0 μ(z)2 ily being realistic. The Damköhler numbers depend on the melting
kinetics at small scales, which are poorly understood. A crude or-
αf ρf der of magnitude estimate can be obtained by assuming that the
rα = , rρ = , (141)
αs ρs melting kinetics are controlled by diffusion into the solid grains.
A typical estimate of the diffusion coefficient for Fe–Mg exchange
ν R1 ν R2 in olivine is ∼ 10−16 m2 s−1 (Chakraborty 1997). For a typical
R1 = , R2 = , (142)
C C grain size of 1 mm, this implies a kinetic rate of ∼10−10 s−1 . For
a 60 km melting column and an upwelling rate of 50 mm yr−1 ,
L1 L2 this implies a Damköhler number ∼4 000. This is large and sug-
St1 = , St2 = , (143)
CT CT gests that the two phases should be close to equilibrium throughout
the melting column. However, the disequilibrium calculations we
λ ν R 1 z λ J ν R 1 z perform use somewhat smaller Damköhler numbers so that the ef-
Da = , Da J = . (144) fects of disequilibrium can be shown.
ρs v0 ρs v0


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 711
 
The non-dimensional parameters A, B and C are related to the  = Da c Z 1 + (1 − c )Z 2 , (160)
natural length scales
 
C J = Da J Z 2 − Z 1 , (161)
δa = = A−1 z, (145)
αg
  1 = c  − J, (162)
kφ (ζφ + 43 ηφ )
δc = = φ (n−m)/2 C 1/2 z, (146)  2 = (1 − c )  + J, (163)
μ

 1/2 xs /K x1
v0 (ζφ + 43 ηφ ) C x = . (164)
δr = = φ −m/2 z, (147) xs /K x1 + (1 − xs )/K x2
ρg B
where δa is the adiabatic length, δc is the compaction length Here T 1m0 and T 2m0 are the non-dimensional melting tempera-
(McKenzie 1984) and δr is the reduced compaction length (Ribe tures and the Simon’s equation coefficients a 1 and a 2 are non-
1985b). dimensionalised on the appropriate pressure scale.

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


In non-dimensional form, the governing equations are
dF 10.2 Zero compaction length approximation
= , (148)
dz
Perhaps the most important simplification to these equations that can
dm be made is to assume C = 0, that is, zero compaction length (Ribe
= 1 , (149) 1985a; Spiegelman 1993a,b) (termed the Darcy approximation by
dz
Sramek et al. 2007). Formally, this is a singular perturbation of these
equations, and by setting C = 0 the compaction boundary layers
dP C d P̃
= −1 + (1 − rρ )φ − (1 − rρ ) , (150) are neglected. However, the problem is then a more straightforward
dz B dz initial value problem, with boundary conditions only specified at the
onset of melting. Essentially, (150), (151) and (152) are replaced by
d P̃ φ −n q B
= − (1 − φ) , (151) dP
dz C C = −1 + (1 − rρ )φ, (165)
dz
dvs
= −φ m P̃, (152)
dz 0 = φ −n q − B (1 − φ) . (166)
  Combining (166) with (154) leads to an algebraic equation for the
dT rα dP
=A F +1− F T −  1 St1 −  2 St2 , (153) porosity φ (Ribe 1985a; Spiegelman & Elliott 1993)
dz rρ dz
F
with Bφ n (1 − φ)2 + (1 − F)φ − (1 − φ) = 0. (167)

1− F F m c0 − m
vs = , q= + 1 − F − vs , cf = , cs = .
1−φ rρ F 1− F The zero compaction length approximation is the leading order
(154) outer solution for the full problem and is a good approximation for
B sufficiently small and at points far away from the boundaries. This
At the onset of melting (say z = 0) we have F = 0, m = 0, vs = 1
is essentially the approximation used in the original study of Ahern
and T and P at some given values on the solidus with composition
& Turcotte (1979). With this approximation, the disequilibrium only
c0 . At the surface (z = 1) we have d P̃/dz = 0. This is a two point
affects the porosity φ through changes in the degree of melting F.
boundary value problem. The governing equations are completed
The matrix is usually assumed sufficiently permeable that the
by the non-dimensional phenomenological laws for melting from
porosity φ remains small throughout the melting column (φ  1).
(98)–(106),
With small porosities, (165) and (166) can be approximated by
xs 1 − xs
Q 1x = , Q 2x = , (155) dP
xf 1− xf = −1, (168)
dz
  
St1 1 1 q = Bφ n ,
K x1 = exp − 1 , (156) (169)
R1 T Tm (P)
that is, the fluid pressure is approximately lithostatic and the fluid
   is simply driven by its buoyancy. The 1-D melting column problem
St2 1 1
K x2 = exp − 2 , (157) then simply consists of solving for T, F and m as a function of
R2 T Tm (P)
pressure P. For large values of the Damköhler number the problem
 1/b1   2 becomes stiff and a stiff ODE solver was used to calculate numerical
P P 1/b solutions. Examples of such solutions are shown in Fig. 3. In the
Tm1 (P) = 1
Tm0 1+ 1 , Tm2 (P) = 2
Tm0 1+ 2 ,
a a limit of infinite Damköhler number (equilibrium), the equations
(158) become differential-algebraic in nature, but are still amenable to
  solution by stiff ODE solvers.
K x1 R2 K x2 Fig. 3 provides examples of near-equilibrium (Da = 104 , Da J =
Z1 = 1 − , Z2 = 1− , (159)
Q 1x R1 Q 2x 104 ) and near-fractional (Da = 104 , Da J = 0) melting (dotted


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
712 J. F. Rudge, D. Bercovici and M. Spiegelman

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


Figure 3. Type II decompression melting for Da = 10, Da J = 10 (solid blue line) and Da = 10, Da J = 0 (solid red line). Melting starts at 10 GPa
(higher than for real mantle melting—chosen for emphasis of the differences) with a bulk composition of 0.9. There is a notable reactive boundary layer at the
onset of melting. Dotted lines show near-equilibrium (blue, Da = 104 , Da J = 104 ) and near-fractional (red, Da = 104 , Da J = 0) paths for comparison.
(a) Temperature versus pressure. Dashed line shows the solid adiabat. (b) Degree of melting versus pressure. (c) dT /dP versus pressure. Dashed line shows
the solid adiabatic gradient. (d) Melt productivity (−dF/dP) versus pressure. Near-equilibrium melting has a slightly higher productivity than near-fractional
melting.

lines). There is very little difference between the two cases, with For fractional melting,
near-fractional melting showing a slightly lower melt productiv-    
μ j
ity (−dF/dP) and lower temperature drop (dT /dP) than near-  =
j j
c ,
j
c 1 − exp − =0 (171)
equilibrium melting. In both cases, melt productivity increases with j
νRjT
decreasing pressure, which is expected. The increase in melt pro-
ductivity with decreasing pressure has been discussed in some detail Expanding the sum for small values of μ j /ν R j T , we have
by Asimow et al. (1997), although it should be noted that here the  j  μ j (μ j )2

latent heats are assumed constant while Asimow et al. (1997) have c − + · · · =0 (172)
j
T νRjT2
latent heats increasing with T (they assumed a constant entropy
jump). and thus
The difference between fractional and equilibrium melting is
 j μ j j
c (μ j )2
much less pronounced here than seen in other studies, such as σ = c ≈ j
T νR j T2
those which consider incrementally isentropic fractional fusion j
(Asimow et al. 1997; Stolper & Asimow 2007). The thermally   2
j Qx
equilibrated fractional melting considered here is fundamentally ≈ ν R j c −1 > 0. (173)
different from incrementally isentropic fractional fusion because j
Kx
here the two phases are always assumed to be in thermal equilib-
Since the two phases are not in chemical equilibrium for fractional
rium and thus at the same temperature. In incrementally isentropic
melting, there is a chemical potential difference between the two
fractional fusion each increment of melt is produced in an isen-
phases and thus a positive entropy production according to the
tropic step and then removed from the system. However, if each
above expression. However, in practice this entropy production is
increment of melt were brought to surface pressures isentropically
very slight and the sum in the above is usually much less than the
then each increment would be at a different temperature and thus
latent heat.
not in thermal equilibrium with the other increments. Thus these
It should also be noted that in the case of a single component,
two types of fractional melting will have slightly different melt
incrementally isentropic fractional fusion differs from equilibrium
productivities.
melting, but thermally equilibrated fractional melting is identical
There is a small entropy production associated with the thermally
to equilibrium melting. For a single component, if the two phases
equilibrated fractional melting considered here. Recall that the part
have the same thermodynamic pressure and the same temperature
of the entropy production due to interphase transfer is
then they must be in thermodynamic equilibrium with one another.
For two or more components the two phases can differ in chemical
 μ j
σ = j . (170) composition whilst having the same temperature and pressure and
j
T thus are not necessarily in equilibrium.


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 713

As the Damköhler numbers are reduced, the kinetics of melting the onset of melting associated with zero porosities, initial condi-
becomes slower and the reactive boundary layers, which occur on tions are chosen with a small initial degree of melting F0 = 5×10−5
the onset of melting, grow. An example of these boundary layers with initial porosity φ0 and initial upwelling rate v 0 such that there
can be seen in Fig. 3 (solid lines). The structure of these boundary is initially no separation of melt from residue that is, zero initial
layers is fairly intuitive: The temperature gradient initially follows Darcy flux q 0 = 0. This is somewhat unphysical (Sramek et al.
the solid adiabatic gradient and there is initially no melting. As 2007), but is a simple way to remove the initial singularity. The zero
the degree of disequilibrium grows and the kinetics get faster the compaction length solution is also shown on the figure (dotted line)
temperature gradient transitions to that of equilibrium melting and and was used as an initial guess for the full numerical solver, which
the productivity transitions to the equilibrium melt productivity. is a standard two point boundary value problem solver.
For the single component case, the reactive boundary layers can Similar to Fig. 3, Fig. 4 shows near-equilibrium (Da =
be treated analytically and this is done in Section D. The boundary 50, Da J = 50) and near-fractional (Da = 50, Da J = 0) scenarios.
layer thickness for a single component is given by As in the earlier calculations, the differences between the two cases
is slight. The most notable difference is in the fluid composition cf ,
ρs v0 Cs T02 ν RCs T02 z
l = = , (174) as would be expected. Also as expected, the zero compaction length
λ L 2 L 2 Da approximation is a good approximation for most of the melting col-

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


where T 0 is the temperature at the onset of melting. The non- umn, with the exception of the bottom boundary. At the top of the
dimensional quantity ν RCs T02 /L 2 is similar for both pure forsterite melting column, the zero compaction length solution is exact, as a
(0.95) and pure fayalite (0.87) (Table 1) and is close to 1. Thus result of the free flux boundary condition applied there.
for the binary case, an approximate boundary layer thickness of There is a reactive boundary layer at the onset of melting and
l ≈ z/Da is expected and indeed this is what is seen in Fig. 3 this is visible in Fig. 4 in the build up of compaction pressure at
where Da = 10: The pressure drop across the boundary layer is the base of the column. From the Damköhler number, the expected
on the order of 1 GPa, which is a tenth of the total pressure drop reactive boundary layer thickness is around 1.2 km. The structure of
over the column. It should be noted that the value of Da = 10 in the boundary layer can be seen more clearly in Fig. 5 which shows a
Fig. 3 is chosen simply to demonstrate the boundary layer structure; zoomed in view of the first 5 km of the column. The zero compaction
realistic Damköhler numbers are likely to be much larger. length approximation provides a reasonable estimate of the structure
of the reactive boundary layer, but there are differences. The most
notable difference is in the first 0.5 km, where there is a further
10.3 Full numerical solution inner boundary layer. The structure of this inner boundary layer is a
Fig. 4 shows a full numerical solution of a 1-D column without the result of the finite porosity, zero Darcy flux initial conditions used
approximations of the previous section. To avoid the singularities at to avoid the initial singularity and is unphysical. Further asymptotic

Figure 4. 1-D binary melting column calculations. Blue line is near-equilibrium (Da = 50, Da J = 50) and red line is near-fractional (Da = 50, Da J = 0).
Dotted line shows the zero compaction length approximation for the near equilibrium case. From left to right: temperature T, fluid pressure P, degree of melting
F, solid composition cs , fluid composition cf , porosity φ, solid velocity V , fluid velocity v, compaction pressure P̃. The difference between near-equilibrium
and near-fractional is most notable in the fluid composition cf , whereas in the other variables there is little difference.


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
714 J. F. Rudge, D. Bercovici and M. Spiegelman

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


Figure 5. A zoomed-in view of the reactive boundary layer of Fig. 4. The boundary layer structure is clearest in the derivatives: dT /dz, dP/dz, dF/dz and
dφ/dz are shown here, along with the compaction pressure P̃.

analysis is needed to explore the singularity at the onset of melting This channelling instability relies on having at least two compo-
for finite kinetic rate, as was done for infinite kinetic rate by Sramek nents and two dimensions. Current models of the instability have
et al. (2007) and Hewitt & Fowler (2008), but we do not attempt used somewhat ad hoc laws for interphase mass transfer: With
this here. the framework presented here connections could be made to real
The boundary layer structure for dT /dz and dF/dz is largely phase diagrams, and melting can occur in a self-consistent energy
similar to that of dT /dP and dF/dP in Fig. 3: the temperature gra- conserving manner rather than as an ad hoc imposed function of
dient transitions from the solid adiabat value to equilibrium melting depth. Some further analytical work on reaction infiltration may be
value and the melting rate starts off at zero and increases to the possible, such as a linear stability analysis of the two component
equilibrium melt productivity. The porosity gradient dφ/dz initially equations, but most future work will need detailed numerics.
increases in a similar manner to dF/dz, but then decreases: Ini-
tially, there is no separation of melt from matrix, but as the porosity
AC K N OW L E D G M E N T S
increases, permeability increases. Melt then separates from the ma-
trix and porosity increases at a lower rate. A similar effect occurs We thank Yanick Ricard and Harro Schmeling for their thorough
in the compaction pressure profile: P̃ changes sign as we go from a reviews. This work was supported by NSF grants OCE-0452457,
situation where melt is initially locked to the matrix (so the matrix EAR-0610138 and EAR-0537599 and also by funding from Trinity
must dilate as a result of melting, P̃ < 0) to one where melt can College, Cambridge.
flow freely and the matrix can compact ( P̃ > 0). The compaction
pressure is directly related to the rate of melting and the build up in
compaction pressure reflects the increase in the rate of melting. REFERENCES
It should be reiterated that the Damköhler numbers used in the Agee, C.B., 1992. Thermal expansion of molten Fe2 SiO4 at high pressure,
above calculations are demonstrative rather than necessarily real- Geophys. Res. Lett., 19, 1173–1176.
istic. The reactive length scale for mantle melting is around 10 m Aharonov, E., Whitehead, J.A., Kelemen, P.B. & Spiegelman, M., 1995.
(Aharonov et al. 1995) and so in practice the reactive boundary Channeling instability of upwelling melt in the mantle, J. geophys. Res.,
layers at the onset of melting are negligible in extent. Nevertheless, 100, 20 433–20 450.
reactive boundary layers are an important feature of the disequilib- Ahern, J.L. & Turcotte, D.L., 1979. Magma migration beneath an ocean
ridge, Earth planet. Sci. Lett, 45, 115–122.
rium equations.
Ahrens, T.J., (ed.) 1995. Mineral Physics and Crystallography : a Handbook
of Physical Constants, AGU, Washington, D.C., ISBN 0-87590-852-7.
Anderson, G.M., 2005. Thermodynamics of Natural Systems, Cambridge
1 1 C O N C LU S I O N S University Press, Cambridge.
Asimow, P.D., 2002. Steady state mantle-melt interaction in one dimension:
The main outcome of this work is a framework for studying disequi- II. Thermal interactions and irreversible terms, J. Petrol., 43, 1707–1724.
librium two phase multicomponent flow that generalises the familiar Asimow, P.D. & Stolper, E.M., 1999. Steady-state mantle-melt interactions
batch and fractional models of melting. Two simple melting prob- in one dimension: I. Equilibrium transport and melt focusing, J. Petrol.,
lems have been addressed which give a flavour of the behaviour 40, 475–494.
of the equations, but the more interesting problems that need to be Asimow, P.D., Hirschmann, M.M. & Stolper, E.M., 1997. An analysis of
tackled next are time-dependent, two or three dimensional, with two variations in isentropic melt productivity, Phil. Trans. R. Soc. Lond. A,
355, 255–281.
or more components and require a more detailed numerical study.
Baker, J. & Cahn, J.W., 1971. Solidification, p. 23, Am. Soc. Metals, Metals
Existing geodynamic codes already model equilibrium two phase
Park, Ohio.
multicomponent flow (e.g. Katz 2008) and hopefully only small Bercovici, D. & Ricard, Y., 2003. Energetics of a two-phase model of litho-
modifications of these codes will be needed to address disequilib- spheric damage, shear localization and plate-boundary formation, Geo-
rium. phys. J. Int., 152, 581–596.
Perhaps the most important problem to revisit is the reaction in- Bercovici, D., Ricard, Y. & Schubert, G., 2001. A two-phase model for com-
filtration instability (Aharonov et al. 1995; Spiegelman et al. 2001). paction and damage 1. General theory, J. geophys. Res., 106, 8887–8906.


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 715

Bottinga, Y., 1985. On the thermal compressibility of silicate liquids at high Spiegelman, M., 1993a. Flow in deformable porous media. Part 1 Simple
pressure, Earth planet. Sci. Lett, 74, 350–360. analysis, J. Fluid Mech., 247, 17–38.
Bradley, R.S., 1962. Thermodynamic calculations on phase equilibra in- Spiegelman, M., 1993b. Flow in deformable porous media. Part 2 Numerical
volving fused salts, part II, Solid solutions and applications to olivines, analysis - the relationship between shock waves and solitary waves, J.
Am. J. Sci., 260, 550–554. Fluid Mech., 247, 39–63.
Caroli, B., Caroli, C. & Roulet, B., 1986. Interface kinetics and solidification Spiegelman, M., 1996. Geochemical consequences of melt transport in 2-D:
of alloys: a discussion of some phenomenological models, Acta Metall., the sensitivity of trace elements to mantle dynamics, Earth planet. Sci.
34, 1867–1877. Lett, 139, 115–132.
Chakraborty, S., 1997. Rates and mechanisms of Fe-Mg interdiffusion in Spiegelman, M. & Elliott, T., 1993. Consequences of melt transport for
olivine at 980◦ − 1300◦ C, J. geophys. Res., 102, 12 317–12 331. U-series disequilibrium in young lavas, Earth planet. Sci. Lett, 118, 1–20.
Connolly, J.A.D. & Podladchikov, Y.Y., 1998. Compaction-driven fluid flow Spiegelman, M., Kelemen, P.B. & Aharonov, E., 2001. Causes and conse-
in viscoelastic rock, Geodinamica Acta, 11, 55–84. quences of flow organization during melt transport: the reaction infiltra-
de Groot, S.R. & Mazur, P., 1984. Non-equilibrium Themodynamics, Dover, tion instability in compactible media, J. geophys. Res., 106, 2061–2077.
Mineola, NY, ISBN 978-0486647418. Sramek, O., Ricard, Y. & Bercovici, D., 2007. Simultaneous melting
Deer, W.A., Howie, R.A. & Zussman, J., 1992. An Introduction to the Rock- and compaction in deformable two-phase media, Geophys. J. Int., 168,
forming Minerals - 2nd edition, Prentice Hall, Harlow. 964–982.

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


Drew, D.A., 1983. Mathematical modeling of two-phase flow, Ann. Rev. Stolper, E. & Asimow, P., 2007. Insights into mantle melting from graphical
Fluid Mech., 15, 261–291. analysis of one-component systems, Am. J. Sci., 307, 1051–1139.
Fowler, A.C., 1985. A mathematical model of magma transport in the as- Tackley, P.J. & Stevenson, D.J., 1993. A mechanism for spontaneous self-
thenosphere, Geophys. Astophys. Fluid Dyn., 33, 63–96. perpetuating volcanism on the terrestrial planets, in Flow and Creep in
Hewitt, I.J. & Fowler, A.C., 2008. Partial melting in an upwelling mantle the Solar System: Observations, Modeling and Theory, pp. 307–321, eds
column, Proc. Roy. Soc. A, 464, 2467–2491. Stone, D.B. & Runcorn, S.K., Kluwer, Dordrecht.
Hillert, M., 2006. An application of irreversible thermodynamics to diffu- Tirone, M., Ganguly, J. & Morgan, J.P., 2009. Modeling petrological geody-
sional phase transformations, Acta Mater., 54, 99–104. namics in the earth’s mantle, Geochem. Geophys. Geosyst., 10, Q04012,
Katz, R.F., 2008. Magma dynamics with the enthalpy method: benchmark doi:10.1029/2008GC002168.
solutions and magmatic focusing at mid-ocean ridges, J. Petrol., 49,
2099–2121.
Kelemen, P.B., Hirth, G., Shimizu, N., Spiegelman, M. & Dick, H.J.B., A P P E N D I X A : T H E R E L AT I O N S H I P S
1997. A review of melt migration processes in the adiabatically upwelling BETWEEN PRESSURES
mantle beneath oceanic spreading ridges, Phil. Trans. R. Soc. Lond. A,
355, 283–318. A detailed description of the relationships between interface pres-
Lasaga, A.C., 1998. Kinetic Theory in the Earth Sciences, Princeton Uni- sure, fluid pressure and solid pressure was given by Bercovici &
versity Press, Princeton, NJ. Ricard (2003) and Ricard (2007), and is reiterated in outline below.
Liang, Y., 2003. Kinetics of crystal-melt reaction in partially molten sili- Bercovici & Ricard (2003) assumed the interface pressure to be a
cates: 1. Grain scale processes, Geochem. Geophys. Geosyst., 4, 1045, linear combination of the fluid and solid pressures,
doi:10.1029/2002GC000375.
Maaløe, S., 1984. Fractional crystallization and melting within binary sys- P = (1 − ω) p f + ωps , (A1)
tem with solid solution, Am. J. Sci., 284, 272–287.
McKenzie, D., 1984. The generation and compaction of partially molten where ω quantifies the partitioning of surface energy. ω = 0 cor-
rock, J. Petrol., 25, 713–765. responds to the case described by McKenzie (1984). In terms of
Poirier, J.-P., 2000. Introduction to the Physics of the Earth’s Interior, pressure differences, (A1) can be written as
Cambridge University Press, Cambridge.
P − p f = ωp, (A2)
Ribe, N.M., 1985a. The generation and composition of partial melts in the
earth’s mantle, Earth planet. Sci. Lett., 73, 361–376.
Ribe, N.M., 1985b. The deformation and compaction of partial molten P − ps = −(1 − ω)p, (A3)
zones, Geophys. J. R. astr. Soc., 83, 487–501.
where p ≡ ps − p f is the mechanical pressure difference between
Ricard, Y., 2007. Physics of mantle convection, in Treatise on Geophysics,
chapter 7.02, pp. 31–87, ed. Schubert, G., Elsevier, Amsterdam. solid and fluid. The viscous dissipation (33) then becomes
Ricard, Y., Bercovici, D. & Schubert, G., 2001. A two-phase model  2  
= d vf − vs + p φω∇ · v f − (1 − φ)(1 − ω)∇ · vs
for compaction and damage 2. Applications to compaction, deforma-
tion, and the role of interfacial surface tension, J. geophys. Res., 106, +φτ : ∇v + (1 − φ)τ : ∇v . (A4)
f f s s
8907–8924.
Schmeling, H., 2000. Partial melting and melt segregation in a convecting
The above form of the viscous dissipation suggests a linear phe-
mantle, in Physics and Chemistry of Partially Molten Rocks, pp. 141–178, nomenological law of the form
eds Bagdassarov, N., Laporte, D. & Thompson, A.B., Kluwer, Dordrecht.  
p = B φω∇ · vf − (1 − φ)(1 − ω)∇ · vs , (A5)
Scott, D.R. & Stevenson, D.J., 1984. Magma solitons, Geophys. Res. Lett.,
11, 1161–1164. provided there is no coupling with the other scalar thermodynamic
Simpson, G., Spiegelman, M. & Weinstein, M.I., 2010a. A multiscale model fluxes (Sramek et al. 2007). The form of the phenomenological
of partial melts:1. Effective equations, J. geophys. Res., 115, B04410, coefficient B is unknown and will depend on parameters such as
doi:10.1029/2009JB006375. porosity. If we write the coefficient B = ζφ /(1 − ω − φ)2 such that
Simpson, G., Spiegelman, M. & Weinstein, M.I., 2010b. A multiscale model
the phenomenological law is written as
of partial melts: 2. Numerical results, J. geophys. Res., 115, B04411,  
doi:10.1029/2009JB006376. φω∇ · vf − (1 − φ)(1 − ω)∇ · vs
Spear, F.S., 1993. Metamorphic Phase Equilibria and Pressure- p = ζφ , (A6)
(1 − ω − φ)2
Temperature-Time Paths, Mineralogical Society of America, Washington,
D.C., ISBN 0939950340. then ζφ can be identified with the effective bulk viscosity used by
Spera, F.J. & Trial, A.F., 1993. Verification of the Onsager reciprocal rela- McKenzie (1984). When ω = 0, eqs (A2) and (A3) with (A6) are
tions in a molten silicate solution, Science, 259, 204–206. identical to (25) and (26). The advantage of (A2), (A3) and (A6) is


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
716 J. F. Rudge, D. Bercovici and M. Spiegelman

that they are clearly symmetric: they are invariant under the change This suggests linear phenomenological laws of the form
φ ↔ 1−φ, ω ↔ 1−ω, p f ↔ ps , v f ↔ vs (assuming the effective
∇T  B kf ∇T μkf  B k ∇T μk
bulk viscosity ζφ obeys this same symmetry). q = −A − − s s
, (B5)
T 2 T T
In Bercovici et al. (2001) the symmetric form k k
 
K 0 μ f + μs
B= (A7) ∇T  L f f ∇T μkf  L f s ∇T μks
jk jk
φ(1 − φ) j
J f = −C f
j
− − , (B6)
T 2 T T
is suggested for the phenomenological coefficient B, where K 0 is a k k
constant, and μ f and μs are the true (rather than effective) viscosities
of fluid and solid. The corresponding deviatoric stress tensors are ∇T  L s f ∇T μkf  L jk ∇T μk
jk

given by Jsj = −Csj − − ss s


. (B7)
T2 T T
  k k
2 
τ f = μ f ∇v f + ∇vTf − ∇ · vf I , (A8) However, the forces and fluxes here are not independent, as there
3
are linear relations between them. The diffusive mass fluxes of
  components are constrained by conservation of mass through (12)

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


2
τ s = μs ∇vs + ∇vsT − (∇ · vs ) I . (A9)  j 
3 Jf = Jsj = 0, (B8)
When the fluid is much less viscous than the solid, μ f  μs , these j j

reduce to (27) and (28), provided ηφ → μs (1 − φ) in this limit. and the forces are constrained by Gibbs–Duhem equations for the
Another way of writing the phenomenological laws is in terms two phases,
of an effective pressure pe , which can be defined as the difference  j j 1
between mean mechanical pressure p and interface pressure P (cf. c f ∇μ f = −s f ∇T + ∇ P, (B9)
Connolly & Podladchikov 1998). The effective pressure pe is lin- j
ρf
early related to the pressure difference p by
 1
pe ≡ p − P = (1 − ω − φ)p. (A10) csj ∇μsj = −ss ∇T + ∇ P, (B10)
j
ρs
Thus a phenomenological law for effective pressure is
 
φω∇ · v f − (1 − φ)(1 − ω)∇ · vs which can be rewritten using (B3) as
pe = ζφ . (A11) 
(1 − ω − φ) j j
ρ f c f ∇T μ f = ∇ P, (B11)
When ω = 0, pe = −ζφ ∇ · vs and when ω = 1, pe = −ζφ ∇ · v f . j
Further discussion of the pressure relationships can be found in

Simpson et al. (2010a). ρs csj ∇T μsj = ∇ P. (B12)
j

APPENDIX B: VECTOR The right hand side in the above is often neglected (see de Groot &
P H E N O M E N O L O G I C A L L AW S Mazur 1984, Chapter 11), but this assumes mechanical equilibrium.
When there exist linear relationships between the forces and
The derivation of intraphase phenomenological laws for fluxes of
fluxes there are two ways to proceed. Either one can use the linear
heat and mass follows directly from the derivation for single phase
relationships to define a new set of independent fluxes and forces,
fluids which is described in detail by de Groot & Mazur (1984) and
or one can keep the existing fluxes and forces and acknowledge
a brief summary is given here. It is useful to use the alternative
that there are constraints on the coefficients. Without independent
definition of heat flux given in (52). The entropy flux (41) can then
fluxes and forces, the phenomenological coefficients are not unique,
be written as
however they can always be chosen such that the Onsager reciprocal
q  j j relations hold (see de Groot & Mazur 1984, chapter 6). In this case,
j= + J f s f + Jsj ssj , (B1)
T j the Onsager reciprocal relations are
j j
j
where sf and ß j are the specific entropies per components. The Bf = Cf , Bsj = Csj , (B13)
corresponding entropy production (42) can be written as
q  j j
jk
Lff = Lff,
kj jk
L f s = Ls f ,
kj
L ssjk = L kssj . (B14)
Tσ = Q + − · ∇T +  j μ j − J f · ∇T μ f − Jsj · ∇T μsj
T j For the sake of simplicity, it is helpful to assume that a lot of these
(B2) cross couplings are negligible and reduce the above to simplified
since phenomenological laws, for example,

j j j
μf j hf j μf q = −k T ∇T (B15)
T∇ = ∇T μ f − ∇T = ∇μ f − ∇T (B3)
T T T
j j j 
n−1
and μ f = h f − T s f . ∇T refers to gradients at fixed temperature. j
J f = −φρ f
jk
D f ∇ckf , j = 1, 2, . . . , n − 1 (B16)
The entropy production due to vector fluxes is given from (B2) k=1
as
q  j ∇T μ f j
∇T μsj 
n−1
σ =− 2
· ∇T − Jf · − Jsj · . (B4) Jsj = −(1 − φ)ρs Dsjk ∇csk , j = 1, 2, . . . , n − 1 (B17)
T j
T j
T
k=1


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS
Disequilibrium multicomponent mantle melting 717


n−1
j
For small deviations in temperature from equilibrium (i.e. rapid
Jnf = − Jf (B18) kinetics), the above law can be linearised as
k=1
λ L
= (T − Tm (P)) , (D2)
T02

n−1
Jns = − Jsj , (B19) where T 0 is the temperature at the onset of melting. A natural scale
k=1 for temperature differences during melting is L/C and we may write
where kT is the effective thermal conductivity of the mixture and L
jk
Df and Dsjk are intra-phase effective diffusion coefficients for the T = T0 + , (D3)
C
melt and the matrix. Eq. (B15) is Fourier’s law of heat flow and
 
(B16) and (B17) are the appropriate generalisations of Fick’s law of dTm  dTm  L z
Tm = T0 + (P − P ) = T − ρ g z = T0 − ,
diffusion. The word ‘effective’ should be emphasised here—the dif- dP T0 dP T0
0 0 s
C lm
fusion coefficients are not necessarily the same as for a single phase
(D4)
fluid and can be controlled by phenomena such Taylor dispersion.
Moreover, the diffusion might not be isotropic as is assumed in the assuming an approximately linear Clapeyron slope. lm is a natural

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


laws above and a more general treatment would have second rank length scale for temperature changes during melting, given by
diffusivity tensors. It should also be noted that due to Onsager’s re- L
ciprocal relations, there are only n(n−1)/2 independent coefficients lm = . (D5)
jk ρs gC dTm /dP|T0
in each of Df and Dsjk . In particular, note that it is not possible to
jk
choose Df and Dsjk as diagonal matrices, unless the diffusivities are The phenomenological law for melting can then be written as
   
all the same. For a more detailed exposition of the constraints on λ L 2 z ρs v0 z
diffusion coefficients and the coupling between heat flow and mass =  + =  + (D6)
C T02 lm l lm
transport (the Soret and Dufour effects), the reader is encouraged to
consult the textbooks (see Lasaga 1998; de Groot & Mazur 1984). where l is a reactive length scale,
Onsager’s reciprocal relationships were verified experimentally for ρs v0 C T02
molten silicates by Spera & Trial (1993). l = . (D7)
λ L 2
The ratio of the two length scales lm and l defines a Damköhler
A P P E N D I X C : C RY S TA L L I S AT I O N number,
Although our main focus in this work is on generalisations of frac- lm λ L 3
Da∗ = = 2 . (D8)
tional melting, it should be noted that fractional crystallisation can l ρs v0 C T02 dTm /dP|T0
2

be described in a similar way. The phenomenological laws (82)–(83)


Note that this differs from the Damköhler numbers defined in the
have to be adjusted slightly to read
 main text which are based on scaling by the column height.
j = E jk Z k , (C1) The steady state 1-D column equations can be non-
k dimensionalised using z = z  lm ,  =   ρs v0 /lm . The non-
dimensional governing equations are, using (132), (137) and (D6),

 1/ν

Qk xsk dF
Z = νR
k k
− 1 = νR k
−1 . (C2) = , (D9)
Kk K xk x kf dz
j j j
Type I crystallising has x = x f and type II crystallising has x = d
 j j = −A∗ (1 + St∗ ) − , (D10)
K x xs . It should be noted that related phenomenological laws for dz
interphase mass transfer were proposed by Liang (2003) based on
averaging a grain scale model. Two regimes were identified, and the  = Da∗ ( + z) , (D11)
terminology used here is based on Liang (2003). Regime I of Liang where
(2003) (diffusion-in-melt-limited dissolution) is essentially type I
αs T0 L
melting and regime II (diffusion-in-solid-limited precipitation) is A∗ = , St∗ = . (D12)
essentially type II crystallising. ρs C dTm /dP|T0 C T0
A∗ is the adiabatic parameter and St∗ is a Stefan number. At the
onset of melting F = 0 and  = 0. The boundary layer structure
APPENDIX D: SINGLE COMPONENT is apparent when the equations are rescaled using  = (Da∗ )−1 (a
M E LT I N G small parameter for rapid kinetics) with z = y, F =  f,  = θ .
A number of simple analytical expressions exist for the single com- To leading order in ,
ponent 1-D melting column. In this section, we will make the zero df
compaction length approximation, assume pressure can be treated = , (D13)
dy
as approximately lithostatic (dP/dz = −ρs g) and assume constant
densities except for the adiabatic gradient term. The phenomeno- dθ
logical law for melting is = −A∗ − , (D14)
    dy
L 1 1
 = λ ν R 1 − exp − . (D1)
νR T Tm (P)  = θ + y, (D15)


C 2010 The Authors, GJI, 184, 699–718
Geophysical Journal International 
C 2010 RAS
718 J. F. Rudge, D. Bercovici and M. Spiegelman

with f (0) = 0 and θ (0) = 0. These integrate to give where Fm is defined to be the degree of melting for equilibrium
melting (Da∗ → ∞). Thus there is small degree of superheating
θ(y) = −y + (1 − A∗ ) (1 − e−y ), (D16)
(T >Tm ) and a slightly lower degree of melting (F< Fm ) that exists
  throughout the column as a result of the finite kinetics.
f (y) = (1 − A∗ ) y − 1 + e−y , (D17) The boundary layer structure outlined above is very similar
with derivatives to that found in the reaction infiltration instability problem de-
dθ scribed by Aharonov et al. (1995). However, there is an impor-
= −1 − (1 − A∗ ) (1 − e−y ), (D18) tant difference between the Aharonov et al. (1995) problem and
dy
the single component melting described above: For reaction infil-
  tration, differences in concentration cause interphase mass trans-
df
= (1 − A∗ ) 1 − e−y . (D19) fer, whereas here interphase mass transfer is caused by differences
dy in temperature. Concentration perturbations are advected with the
In dimensional form, the above can be written as fluid velocity in the Aharonov et al. (1995) problem, whereas

 temperature perturbations here travel more slowly, at the mean
dT αs gT0 dTm  αs gT0  
=− − ρs g  − 1 − e−z/l , (D20) upwelling velocity (v). As a result, the single component melt-

Downloaded from https://academic.oup.com/gji/article/184/2/699/593112 by guest on 07 September 2020


dz C dP T0 C ing equations are stable and do not have a reaction infiltration
instability.


dF C dTm  αs gT0   The stability can be demonstrated as follows. In 1-D, total con-
= ρs g  − 1 − e−z/l , (D21) servation of mass is
dz L dP T0 C
∂ρ ∂
eqs (D20) and (D21) are fairly intuitive: The gradient in temperature + (ρv) = 0. (D26)
∂t ∂z
starts off at the solid adiabatic gradient and ends up at the Clapeyron
Porosities are typically small (φ  1) and thus ρ ≈ ρs and ρv ≈
slope and the melt productivity starts off at zero and ends up at
ρs v0 . Neglecting the adiabat, conservation of energy is then
the same productivity as for equilibrium batch melting (Asimow
et al. 1997). The boundary layer thickness is controlled by the ∂T ∂T
ρs C + ρs v0 C = −L , (D27)
reactive length l which shrinks as the reaction rate increases or the ∂t ∂z
upwelling rate slows. which in non-dimensional variables is
The dimensional temperature and degree of melting profiles are

∂ ∂
  + = −Da ( + z) . (D28)
dTm  dTm  αs gT0   ∂t ∂z
T = T0 − ρs g z + ρ g − l 1 − e−z/l ,
dP T0 dP T0
s
C Where an appropriate time scale for non-dimensionalising is t 0 =
(D22) lm /v 0 . Solutions to this equation can be sought in terms of pertur-

bations about the steady state,  = (z) + εeimz+σ t . Neglecting

C dTm  αs gT0    boundary layers, a suitable steady state is
F= ρs g − z − l 1 − e−z/l . (D23)
L dP T0 C 1
(z) = −z + , (D29)
Da
Outside the boundary layer (z  l ), these can be written as
 and the growth rate of perturbations is
L dFm 
T = Tm (z) + l , (D24)
C dz T0 σ = −Da − im. (D30)

 Hence any perturbations in temperature will simply decay away.


dFm 
F = Fm (z) − l , (D25)
dz T0


C 2010 The Authors, GJI, 184, 699–718

Geophysical Journal International 


C 2010 RAS

You might also like