Geophys. J. Int. (2006) 167, 1461–1481 doi: 10.1111/j.1365-246X.2006.03131.x
GJI Tectonics and geodynamics
Models of large-scale viscous flow in the Earth’s mantle with
constraints from mineral physics and surface observations
Bernhard Steinberger
1
and Arthur R. Calderwood
2
1
Center for Geodynamics, NGU, N-7491 Trondheim, Norway. E-mail: bernhard.steinber[email protected]
2
Sauder School of Business, University of British Columbia, Vancouver, BC, Canada V6T 1Z2
Accepted 2006 July 5. Received 2006 June 20; in original form 2005 August 16
SUMMARY
Modelling the geoid has been a widely used and successful approach in constraining flow and
viscosity in the Earth’s mantle. However, details of the viscosity structure cannot be tightly
constrained with this approach. Here, radial viscosity variations in four to five mantle layers
(lithosphere, upper mantle, one to two transition zone layers, lower mantle) are computed with
the aid of independent mineral physics results. A density model is obtained by converting
s-wave anomalies from seismic tomography to density anomalies. Assuming both are of ther-
mal origin, conversion factors are computed based on mineral physics results. From the density
and viscosity model, a model of mantle flow, and the resulting geoid and radial heat flux profile
are computed. Absolute viscosity values in the mantle layers are treated as free parameters and
determined by minimizing a misfit function, which considers fit to geoid, ‘Haskell average’
determined from post-glacial rebound and the radial heat flux profile and penalizes if at some
depth computed heat flux exceeds the estimated mantle heat flux 33 TW. Typically, optimized
models do not exceed this value by more than about 20 per cent and fit the Haskell average well.
Viscosity profiles obtained show a characteristic hump in the lower mantle, with maximum
viscosities of about 10
23
Pa s just above the D

layer— several hundred to about 1000 times the
lowest viscosities in the upper mantle. This viscosity contrast is several times higher than what
is inferred when a constant lower mantle viscosity is assumed. The geoid variance reduction
obtained is up to about 80 per cent—similar to previous results. However, because of the use
of mineral physics constraints, a rather small number of free model parameters is required, and
at the same time, a reasonable heat flux profile is obtained. Results are best when the lowest
viscosities occur in the transition zone. When viscosity is lowest in the asthenosphere, variance
reduction is about 70–75 per cent. Best results were obtained with a viscous lithosphere with a
few times 10
22
Pa s. The optimized models yield a core-mantle boundary excess ellipticity sev-
eral times higher than observed, possibly indicating that radial stresses are partly compensated
due to non-thermal lateral variations within the lowermost mantle.
Key words: geoid, heat flow, mantle convection, mantle viscosity, mineralogy, tomography.
1 INTRODUCTION
Mantle rheology is still one of the rather poorly known properties of
the Earth. It is widely agreed upon that, below a brittle lithospheric
layer, mantle rocks behave, on timescales of thousands or millions
of years, like a highly viscous fluid. Viscous flow in the Earth’s
mantle is presumably the principal way how the Earth transports
heat through the bulk mantle, and an underlying cause for gravity
and geoid undulations, tectonic plate motions as well as stresses in
the Earth’s lithosphere.
Research efforts to determine mantle viscosity can be broadly
dividedinto three areas: (i) mineral physics,(ii) post-glacial rebound
and (iii) large-scale mantle flow.
Determining viscosity from mineral physics alone is difficult, be-
cause different deformation mechanisms—diffusion creep and dis-
location creep—may play a role. For either mechanism, the effec-
tive viscosity depends on a number of factors—such as temperature,
grain size, water content, etc., and many of these are poorly known.
Measurement of post-glacial rebound has been the classical
method of determining mantle viscosity ever since the canonical
value of 10
21
Pa s was established by Haskell (1935). Newer results
(e.g. Mitrovica 1996; Lambeck & Johnston 1998) confirm this, and
additionally also indicate a viscosity increase with depth; however,
they also show that post-glacial rebound cannot resolve details of
mantle viscosity structure, and is particularly insensitive to viscosity
below about 1400 km depth.
C
2006 The Authors 1461
Journal compilation
C
2006 RAS
1462 B. Steinberger and A. R. Calderwood
Mantle flow can be computed, if density field, viscosity structure,
etc., are known, and comparison of computed advected heat flux,
plate motions, geoid, stresses, etc., with observations can help to as-
sess the ‘success’ of the model, and thus help to constrain viscosity
and other properties on which the results depend. For simplicity and
computational efficiency, the assumption of radial viscosity varia-
tions only is frequently made. In this case, density and flow field
can be expanded in spherical harmonics, and computed separately
for each degree and order, using a formalism developed by Hager
& O’Connell (1979, 1981). This formalism was extended by Ricard
et al. (1984) and Richards & Hager (1984) to the computation of
the geoid. Since then, the geoid (which is extremely well known
compared to other quantities that can be used) has been taken as a
constraint to mantle viscosity and flow in numerous publications.
These essentially show that large parts of the geoid can be explained
based on viscous flow models with radial viscosity variations only.
Thoraval & Richards (1997) review this body of publications; how-
ever, they also show that the geoid alone cannot give a very tight
constraint on mantle flow and the quantities on which it depends,
such as viscosity and density anomalies, and the only robust result
is, that a substantial viscosity increase with depth is required to fit
the geoid data. Including lateral variations in viscosity was found
not improve the fit to the geoid (Zhang & Christensen 1993). More
recently,
ˇ
Cadek & Fleitout (2003, 2006), however, found that the
fit can be improved by including lateral viscosity variations in the
top 300 km, and close to the core–mantle boundary (CMB). Also
on a regional scale, lateral viscosity variations appear to be im-
portant in determining the mantle flow field (Albers & Christensen
2001).
Because of the limitations of each individual method, the need to
jointly fit several observations and incorporate other data to con-
strain viscosity and other properties that determine flow in the
Earth’s mantle became apparent. Mitrovica & Forte (1997) jointly
fit the geoid and post-glacial rebound observables, and confirm a
significant increase of viscosity with depth. Pari & Peltier (1995)
use heat flow constraints in addition to the geoid. Other quanti-
ties considered include plate motions (e.g. Ricard & Vigny 1989;
Lithgow-Bertelloni & Richards 1998; Becker & O’Connell 2001;
Conrad & Lithgow-Bertelloni 2002, 2004), dynamic surface topog-
raphy (Lithgow-Bertelloni & Silver 1998; Kaban et al. 1999; Pari
& Peltier 2000; Panasyuk & Hager 2000; Steinberger et al. 2001;
ˇ
Cadek & Fleitout 2003), CMB topographyand ellipticity (Forte et al.
1993, 1995) and lithospheric stresses (Ricard et al. 1984; Bai et al.
1992; Steinberger et al. 2001; Lithgow-Bertelloni & Guynn 2004)
or a combination of several of these (Mitrovica & Forte 2004). How-
ever, none of these quantities is as accurately known as the geoid.
In order to optimize the fit to these various observations, a large
number of parameters can be adjusted, and the task of mantle flow
modelling with observational constraints can thus become rather
complex. In order to reduce the number of parameters it is, therefore,
useful to consider constraints from mineral physics as well.
Here we will derive flow models making these simplifying as-
sumptions: (1) Both lateral density and seismic velocity variations
are due to temperature variations; the conversion factors between
these variations depend only on depth, hence tomography models
can be converted to density models. (2) Mantle viscosity only de-
pends on depth. We will then use an adiabatic thermal profile with
boundary layers and results from mineral physics to derive viscos-
ity and conversion factors as a function of depth. We will derive
relative viscosity variations in mantle layers with approximately
constant mineralogy (upper mantle, one or two layers in transition
zone, lower mantle) and leave absolute values as free parameters. In
addition, we will keep the lithospheric viscosity as a free parameter.
This is necessary in order to compensate for the fact that the viscous
rheology used here is not appropriate for the lithosphere. A more
realistic treatment is difficult, we still lack a detailed knowledge of
lithospheric rheology, and self-consistent models of plate tectonics
are only beginning to emerge. The optimized lithospheric viscos-
ity obtained in that way represents an effective viscosity along plate
boundaries, where most of the lithospheric deformation occurs. This
is much less than whatis generally thought to be appropriate for plate
interiors. We will then use density models inferred from seismic to-
mography in combination with viscosity models to compute mantle
flow. The geoid is computed from density anomalies and deforma-
tion of boundaries caused by the flow. The radial heat flux profile
is computed from the flow and density variations converted back to
temperature variations. Our optimization is done by minimizing a
misfit function that is computed based on
(i) the difference between predicted and observed geoid
(ii) compatibility of viscosity structure with post-glacial rebound
results
(iii) compatibility of radial heat flux profile with observations
Plate velocity predictions as well as predictions of lithospheric
stress and dynamic topography turn out to be rather insensitive
to variations of viscosity with depth (Becker & O’Connell 2001;
Steinberger et al. 2001); therefore, we do not use these in our opti-
mization. For models with lateral viscosity variations, though, Con-
rad & Lithgow-Bertelloni (2006) recently showed that deeply pen-
etrating continental roots increase the magnitude of shear tractions
that mantle flow exerts on the base of Earth’s lithosphere by a factor
of 25, compared to a 100-km-thick lithosphere.
Obviously, there are also other uncertainties than absolute viscos-
ity values, which are free parameters of our optimization. We will
treat those by modifying other model parameters relative to the ref-
erence model and discussing how results change. Model parameters
are listed in Table 1.
In the next chapter, we will discuss how we derive the viscos-
ity and scaling factor profiles from mineral physics. After this, the
computation of mantle flow, geoid and advected heat flux, and how
the misfit function is constructed, is explained. We will then present
results. We will also discuss some a posteriori predictions of quan-
tities that were not used in the optimization—surface motion and
CMB excess ellipticity (Mathews et al. 2002). This will point to-
wards shortcomings of this work, and future improvements.
In particular, there has been recently increasing evidence
(e.g. Masters et al. 2000; Trampert et al. 2004; Ishii & Tromp 2004)
that probably not all seismic velocity anomalies are due to tempera-
ture anomalies, as assumed here. The approach taken here is to test
how well observations can be fit under the assumption that seismic
velocity anomalies are caused by temperature anomalies, and how
we have to choose modelling parameters in order to obtain an op-
timum fit. We test a range of s-wave tomographic models (Becker
& Boschi 2002; Ritsema & van Heijst 2000; Masters et al. 2000;
Grand 2002; M´egnin & Romanowicz 2000; Su et al. 1994; Gu et al.
2001). This approach actually yields additional evidence for chem-
ical heterogeneities, as will be explained in more detail below, and
discussed further elsewhere (Steinberger & Holme 2006).
2 MINERAL PHYSICS THEORY
In this chapter we discuss (if adopted from elsewhere) or derive
parameters used in our model. They are assumed either constant or
depth dependent and listed alphabetically in Table 1. We will first
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1463
Table 1. Model parameters. For constant parameters, values are given for reference model, followed by alternative values (model number
in brackets), followed by range covered in contour plots, unless they are allowed to vary in the optimization.
Symbol Value Depth (range) Name and/or comment
a
0
2.9; 3.5 (M2); 2–4 >660 km Thermal expansivity coefficient (eq. 17)
a
1
0.9 >660 km Thermal expansivity coefficient (eq. 17)
b 1.4; 0 (M4) >660 km Specifies depth dependence of α (eq. 15)
0 <660 km
C
p
1250 J kg
1
K
1
Specific heat
d
D

200 km; 100–350 km >2541 km Thickness of bottom thermal boundary layer
d
lith
100 km <100 km Thickness of top thermal boundary layer
(ln v
s,0
/∂T )
p
Fig. 5 See Section 2.5
dμ
0
dT
27 MPa K
1
>660 km Temperature derivative of shear modulus at p = 0
F
l
0.5; 0–1 <220 km Conversion factor reduction relative to Fig. 6
g 12; 30; 20 (M2,5); 4–70 >660 km Relates H and T
m
(eq. 10), thus influences
scaling factor profile through eq. (22)
g
(= g/n) 12; 20 (M5); 4–20 >660 km Determines steepness of viscosity profile (eq. 11)
H Fig. 1 Activation enthalpy
n 1; 2.5 (M2); 1–3.5 >660 km Stress exponent
3.5 <660 km
p Pressure
q
i
Allowed to vary in optimization Determine absolute viscosity (eq. 19)
Q See Section 2.5 Seismic Q-factor
R 8.3144 J K
1
mol
1
Universal gas constant
T Fig. 2 (Laterally averaged) temperature
T
0
285 K 0 km Surface temperature
T
CMB
3500 K; 3000–4000 K 2891 km Temperature at core–mantle boundary
T
lm,0
See Section 2.5 >660 km Lower mantle potential surface temperature
T
m
Figs. 1, 2 Melting temperature
T
um,0
1613 K <400 km (Upper mantle) potential surface temperature
α Fig. 3 Thermal expansivity
γ From PREM gravity
ρ 0.3 · 10
3
MPa K
1
· kg m
3
660 km Product of Clapeyron slope and density jump
0.5 · 10
3
MPa K
1
· kg m
3
400 km at phase boundary
δ
T 0
5.5 Specifies depth dependence of α (eq. 15)
˜η Fig. 4 Normalized viscosity
μ From PREM Shear modulus
ρ From PREM Actual density
ρ
0
See Section 2.3 Density extrapolated to zero pressure
discuss the viscosity law in general (Section 2.1.1) and restricted
to only radial viscosity variations (Section 2.1.2). This is followed
by a discussion of the parameters involved. These include the stress
exponent n, activation enthalpy H, melting temperature T
m
and
the factors g and g
(Section 2.2), as well as a number of further
parameters that determine the radial profiles of laterally averaged
temperature
T and thermal expansivity α (derived jointly in Sec-
tion 2.3), as well as the viscosity scaling factors (Section 2.4) which
are treated as free parameters in the optimization. We further dis-
cuss parameters affecting the relation between seismic velocity and
temperature anomalies (Section 2.5) and the factor F
l
used to adjust
the relation between seismic velocity and density variations in the
uppermost mantle.
2.1 Mantle viscosity
Our numerical flow model will only consider radially varying, New-
tonian viscosity, however, in order to derive the appropriate viscosity
profile, and in order to make our derivation extendable to more re-
alistic rheologies, we keep it rather general.
2.1.1 General rheological model
We adopt the frequently used approach of assuming a power-law
rheology, where the relation between strain rate ˙ and stress σ (more
specifically, the square root of the second invariant of the respective
tensors) is of the form
˙= C
1
σ
n
exp
H
RT
, (1)
whereby H is activation enthalpy, R is the universal gas constant,
T is temperature and C
1
is a constant. Solving the equation for σ
gives
σ = ˙
1
n
·
1
C
1
n
1
exp
H
nRT
. (2)
With the usual definition of viscosity η, it follows
η =
σ
= ˙
1
n
1
·
1
2C
1
n
1
exp
H
nRT
. (3)
In order to avoid the singularity η −→ for ˙ −→ 0 we replace
˙
1
n
1
with (
˙
2
˙
2
(z)
+ ˙
2
0
)
1n
2n
·(˙
2
(z))
1n
2n
. Here ˙
2
(z) is the laterally
averaged second invariant of the strain rate tensor which only de-
pends on the radial coordinate z, and ˙
0
is a number smaller than 1.
T is split up into a laterally averaged part
T (z) which only depends
on z, and the ‘temperature anomaly’ δT:
1
T
=
1
T + δT
=
1
T
δT
T (T + δT )
. (4)
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1464 B. Steinberger and A. R. Calderwood
We can then write viscosity as
η = η
i
· V
rT
(z) · V
rs
(z) · V
lT
· V
ls
, (5)
V
lT
= exp
HδT
nRT(T +δT )
and V
ls
=
˙
2
˙
2
(z)
+ ˙
2
0
1n
2n
,
(6)
are the lateral variation of viscosity due to lateral temperature and
strain rate variations and are not considered in our flow computation.
We continue now with discussing η
i
, V
rT
and V
rs
which determine
the radial viscosity profile.
2.1.2 Radial viscosity profile—equations
η
i
are called ‘anchor viscosities’ and are adjusted such as to mini-
mize the misfit between model predictions and observations as de-
scribed below. They are determined independently for various depth
ranges for which different phase assemblages occur (upper mantle,
one to two layers in transition zone, lower mantle).
V
rT
(z) = exp
H(z)
nRT(z)
, (7)
is the radial variation of viscosity due to radial temperature and
pressure variations at constant strain rate and
V
rs
(z) = (˙
2
(z))
1n
2n
, (8)
is the radial variation of viscosity due to strain rate variations.
Similarly, it can be shown that for constant stress η
exp(H/(RT)) and for constant dissipation rate η exp(2H/((n +
1)RT)). In other words, effective viscosity can be expressed in the
form
η exp
rH
RT
, (9)
with r =1/n for constant strain rate, r = 2/(n +1) for constant dis-
sipation rate and r = 1 for constant stress. For Newtonian viscosity
(n = 1) it is r = 1 in all cases. Christensen (1983) showed for 2-D
numerical experiments that the properties of non-Newtonian flow
with n = 3 can be closely imitated by Newtonian flow with activa-
tion enthalpy reduced by a factor r = 0.3–0.5. Thus the appropriate
viscosity dependence appears to be somewhere between constant
strain rate and constant dissipation rate. Here we use the constant
strain rate formulation, that is, eq. (7) and V
rs
=1. We consider this
appropriate for the purpose of this paper, for the following reasons:
(i) For the lower mantle, we will explicitly discuss the depen-
dence of results on the reduction factor r (expressed in terms of g
).
Our reference case is with n = 1 (Newtonian viscosity), however,
given the results of Christensen (1983), our results are probably
applicable to the case of non-Newtonian viscosity as well.
(ii) For the mantle above 660 km it will turn out that the result-
ing optimized viscosity profile is mainly determined by the anchor
viscosities, and results are rather insensitive to the factor r. In par-
ticular, we will show that results remain rather similar regardless
of whether variations in viscosity due to variations in temperature
and activation enthalpy in the upper mantle and transition zone lay-
ers are considered or whether constant viscosity is assumed within
these layers.
We will now discuss the stress exponent n and profiles H(z) and
T (z).
2.2 Stress exponent, activation enthalpy
and melting temperature
2.2.1 Stress exponent
The appropriate value for the stress exponent n is not well known,
since solid-state flow in the mantle can be achieved through both
dislocation creep and diffusion creep. For dislocation (or power-law)
creep, n 3.5 is usually considered appropriate, whereas n = 1 for
diffusion creep. Power-law creep is favoured for high stresses, large
grain sizes, high temperatures and low pressures, whereas diffusion
creep for low stresses, small grain sizes, low temperatures and high
pressures. It is thought that both mechanisms may contribute to flow
in the upper mantle, with composite viscosity intermediate, but lab-
oratory studies favour dislocation creep in the shallow upper mantle
(Ranalli 1995; Schubert et al. 2001). Further evidence for disloca-
tion creep in the upper mantle comes from geodynamic modelling
(van Hunen et al. 2005). Ranalli (1995) concludes on p. 390 that ‘if
there is no thermal boundary layer (TBL) between upper and lower
mantle (mantle-wide convection) power-law creep is predominant
in the lower mantle’. On the other hand, the fact that the lower mantle
is nearly isotropic, has been interpreted such that diffusion creep is
the dominant deformation mechanism (Karato et al. 1995), at least
above the D

layer. As reference case will use eq. (7) with n = 3.5
above 660 km and n = 1 below.
2.2.2 Activation enthalpy
Activation enthalpy H is the sum of activation energy plus pressure
times activation volume. Kohlstedt and Goetze (1974) determined
activation energy 525 kJ mol
1
for dislocation (power-law) creep in
dry olivine. Activation volume is more uncertain. Above 660 km,
we will use the continuous line in Fig. 1 as reference case profile
2500
2000
1500
1000
500
0
depth [km]
0 200 400 600 800
activation enthalpy [kJ mol
–1
]
0 1000 2000 3000 4000 5000 6000 7000 8000
temperature [K]
Figure 1. Solid line—upper scale: activation enthalpy profile based on
Calderwood (1999) used in the upper mantle. dashed line—lower scale:
lower mantle melting temperature profile used, intermediate between the
curves determined for MgSiO
3
perovskite (Wang 1999) and MgO (Zerr &
Boehler 1994). The two scales differ by a factor gR = 100 J mol
1
K corre-
sponding to g = 12. In this way, the curve for the lower mantle can be also
used with the upper scale and upper mantle viscosity law to determine lower
mantle viscosity and vice versa.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1465
for activation enthalpy. It is based on Calderwood (1999) and very
similar to a H(z) profile given by Ranalli (1995).
For zero depth, it equals the value determined by Kohlstedt &
otze (1974). Its increase with depth corresponds to an activa-
tion volume of about 12 cm
3
mol
1
at depth 100 km, decreasing to
10 cm
3
mol
1
at depth 660 km—within the range of experimental
results. Since upper mantle viscosities are mainly determined by the
anchor viscosities, and rather insensitive to the activation enthalpy
profile, we shall not discuss it in more detail.
2.2.3 Melting temperature
The activation enthalpy profile is mainly important for determining
the viscosity profile in the lower mantle. Weertman & Weertman
(1975) give an empirical relation
H = gRT
m
, (10)
whereby g is a dimensionless constant, and T
m
is melting tempera-
ture. Then eq. (7) becomes
F
rT
(z) = exp
g
· T
m
(z)
T (z)
, (11)
with g
= g/n. In the reference case for the lower mantle, we will
use eq. (11) with g
=12, the arithmetic mean of values determined
by Yamazaki & Karato (2001) for silicon diffusion in MgSiO
3
per-
ovskite and MgO. However, we will consider other values as well,
and emphasize that our reference value may also correspond to ef-
fective viscosity for power-law creep with larger g.
Experimental values exist for the the melting curves of lowerman-
tle constituents MgSiO
3
perovskite (Wang 1999) and MgO (Zerr
& Boehler 1994). Yamazaki & Karato (2001) consider using the
melting curves of these constituents—which both look similar—
appropriate. Here, a melting curve intermediate between those two
curves (see Fig. 1) is used.
2.3 Mantle temperature and thermal expansivity
The radial temperature profile is also required to derive viscosity
as a function of depth. We assume here the temperature profile is
adiabatic except for TBLs. This is not standard practice in the geo-
dynamical literature—the adiabatic gradient is generally removed
because it does not influence the dynamics directly. However, it
does influence the mineral physics constraints on material proper-
ties, which is why it is included here. A number of further parame-
ters are involved here. These can be grouped into those determining
the adiabatic temperature gradient (thermal expansivity α(z), grav-
ity γ (z), specific heat C
p
and mantle potential surface temperature
T
um,0
; Section 2.3.1), further parameters determining thermal ex-
pansivity (a
0
, a
1
, b, δ
T 0
, ρ
0
, ρ; symbols explained in Section 2.3.2),
the product of Clapeyron slope and density jump ρ determining
jumps in the adiabatic temperature profile across phase boundaries
(Section 2.3.3), and parameters T
CMB
, T
0
, d
D

and d
lith
defining
thermal structure of TBLs (Section 2.3.4).
2.3.1 Adiabatic temperature profile
An adiabatic temperature profile can be computed by integration of
d
T
dz
=
T (z) ·α(z) · γ (z)/C
p
(z). (12)
Starting point is mantle potential surface temperature (i.e. extrapo-
lation of the mantle adiabat to the surface) T
um,0
.WeuseT
um,0
=
1340
C = 1613 K based on decompression melt studies of MORBs
(White & McKenzie 1995; Iwamori et al. 1995). The gravity profile
γ (z) can be computed from the Earth’s radial density distribution and
is, therefore, known rather accurately. For heat capacity, we adopt C
p
1250 J kg
1
K
1
(e.g. Stacey 1992; Schubert et al. 2001), which
is also considered to be known rather accurately. Since α also de-
pends on temperature, the radial profiles for α and
T are determined
jointly.
2.3.2 Thermal expansivity
The relation between thermal expansivity and density (along
isotherms) can be expressed in the form
ln α
ln ρ
T
=−δ
T
. (13)
In the upper mantle, a constant δ
T
= 5.5 is used here as reference
case. Chopelas & Boehler (1989) experimentally determined δ
T
=
5.5 ± 0.5. In this case, integration yields
α(p, T ) = α
0
(T )
ρ( p, T )
ρ
0
(T )
δ
T
, (14)
whereby α
0
(T ) and ρ
0
(T ) are thermal expansivity and density as
a function of temperature at zero pressure. Experimental results
exist for both α
0
(T ) and ρ( p, T)
0
(T ). We follow here Schmeling
et al. (2003) where explicit formulae and original references were
given. Their treatment is simplified, in that they do not consider the
effect of phase transitions. Therefore, we also use the profile derived
by Calderwood (1999) for a pyrolite mineral phase assemblage for
comparison.
In the lower mantle, depth dependence of δ
T
may play a role. The
relation between δ
T
and ρ was found to be
δ
T
= δ
T 0
ρ
0
(T )
ρ( p, T )
b
, (15)
with δ
T 0
5.5 and b 1.4 (Anderson et al. 1992; Schubert et al.
2001). In this case, integration along isotherms yields
α(p, T ) = α
0
(T )exp
δ
T 0
b
1
ρ
0
(T )
ρ( p, T )
b

. (16)
We will use this equation in the lower mantle in the reference case,
however, we will additionally show results with constant δ
T
= 5.5
in the entire mantle, in order to assess how large the effect of the
depth dependence of δ
T
on our results is. We use
α
0
(T ) = (a
0
+ a
1
T/1000 K) ·10
5
K
1
, (17)
with a
0
= 2.9 and a
1
= 0.9 in the reference case. This is inter-
mediate between various experimental results and ab initio calcu-
lations for MgSiO
3
perovskite, the main lower mantle constituent
(Oganov et al. 2001). For magnesiow¨ustite (MgO), another ma-
jor lower mantle constituent, α
0
(T ) is probably similar but slightly
higher (e.g. Duffy & Anderson 1989, fig. 5). Because of the con-
siderable uncertainty, we consider cases for higher or lower α
0
as
well.
Actual density in the lower mantle is reasonably well known,
and ρ
0
can, for example, be determined by extending the PREM
(Dziewonski & Anderson 1981) lower mantle density profile to the
surface. However, the PREM profile is approximately adiabatic, and
not isothermal. Hence, the PREM value has to be corrected for adia-
batic temperature difference. Transitionof the majorite phase, which
constitutes about 30 per cent of mantle material, is not abrupt but
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1466 B. Steinberger and A. R. Calderwood
Figure 2. Computed adiabatic temperature profiles (dashed lines), and temperature profiles with thermal boundary layers (solid lines). Black lines: referenc
e
case; orange lines: computation using thermal expansivity profile of Calderwood (1999) (see Fig. 3); brown line: computation with constant δ
T
, all other
assumptions as in reference case. Labels indicate model numbers (Table 2) here and in the following figures. The dotted line shows again the assumed lower
mantle melting temperature.
occurs gradually between about 660 and 730 km (e.g. Akaogi & Ito
1999). Therefore, within this depth range, we use a linear superposi-
tion α(z) =α
lm
(z)+0.3 ·(α
um
(z)α
lm
(z))·(z 730 km)/(660 km–
730 km) whereby α
um
and α
lm
are determined with eqs (14) and (16),
respectively.
2.3.3 Temperature jumps at phase boundaries
Temperature jumps at phase boundaries are smeared out due to dif-
fusion, but the jump between adiabatic profiles above and below
phase boundaries is
T
L
= Q
L
/C
p
= ρT
pb
/
ρ
2
pb
C
p
. (18)
Hereby is Q
L
the latent heat release per unit mass, is the Clapey-
ron slope, ρ is the density jump across the phase boundary, T
pb
is
the average temperature below and above the phase boundary, and
ρ
pb
is the average density above and below the phase boundary. Two
phase transitions at depths 400 and 660 km are considered here. In
the reference case, for depth 400 km, ρ = 0.5 · 10
3
MPa K
1
·
kg m
3
is used, based on Akaogi et al. (1989) for a pyrolite man-
tle with 60 per cent olivine content, for depth 660 km, ρ =
0.3 · 10
3
MPa K
1
· kg m
3
is used, as given by Akaogi & Ito
(1999). Besides the spinel–perovskite transition, this value also in-
cludes the effects of the majorite–perovskite transition, which occurs
at a similar depth with positive Clapeyron slope.
2.3.4 Temperatures at the top and bottom of the mantle
At the top and bottom of the mantle are two TBLs with larger tem-
perature gradient. Temperature at the CMB is T
CMB
=4000 ±600 K
according to Boehler (1996) and Schubert et al. (2001). The thick-
ness of the TBL is estimated to be about 200 km (Schubert et al.
2001), but it may be thicker, if there are chemical variations at the
base of the mantle. We use T
CMB
= 3500 K, bottom TBL thick-
ness d
D

= 200 km, surface temperature T
0
= 285 K and top TBL
thickness d
lith
= 100 km in the reference case. For the difference
between adiabatic and actual temperature profile at distance x from
a thermal boundary with total non-adiabatic temperature drop T
0
and thickness d we use T = T
0
· [1 erf (x/d)].
2.3.5 Temperature and thermal expansivity profiles
Resulting temperature profiles are shown in Fig. 2. The effect of
using different thermal expansivity profiles, and of phase boundaries
are both rather small.
Corresponding thermal expansivity profiles are shown in Fig. 3.
The black ‘reference profile’ features a decrease from 2.5 ·
10
5
K
1
below 670 km to 1.0 · 10
5
K
1
at the base of the man-
tle, in agreement with Schubert et al. (2001). For comparison, the
profile of Calderwood (1999), derived from a pyrolite mineral phase
assemblage and thermal expansivities of individual phases is also
shown. It agrees with the reference profile qualitatively, but de-
creases somewhat less with depth in the lower mantle. Differences
Figure3. Thermal expansivity profiles for the same cases as shown in Fig. 2.
Additionally, the dashed line shows the profile with α
0
reduced by 0.6 ·
10
5
K
1
, the dotted line with α
0
increased by the same amount.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1467
do not exceed 20 per cent. Thermal expansivity decreases more
strongly with depth, if a constant δ
T
is assumed. The dashed line is
obtained, if α
0
(T ) = (2.3 + 0.9 T /1000 K) · 10
5
K
1
is used in-
stead, the dotted line for α
0
(T ) = (3.5 +0.9 T /1000 K) ·10
5
K
1
.
Both values are still within the range of results proposed (Oganov
et al. 2001). All profiles show an overall decrease with depth, but an
increase with depth across the 660 km discontinuity. Qualitatively,
this jump corresponds to α
0
being larger for lower mantle materials
than for upper mantle materials (see e.g. Duffy & Anderson 1989,
fig. 5).
2.4 Normalized radial viscosity profile—results
We can now express
η(z) = q
i
· η
0
· ˜η(z). (19)
η
0
is a constant scaling viscosity. ˜η(z) shown in Fig. 4 is the normal-
ized viscosity profile proportional to F
rT
(z) from eqs (7) and (11),
but adjusted by adding a constant to H in the lower mantle such that
the jump in H(z)/nR between upper and lower mantle is removed.
This adjustment is of no consequence to the results, and merely
serves to make the factors q
i
more easily interpretable. Without the
adjustment, ˜η(z) would increase by a factor 2000 from above to
below 660 km because of the different viscosity law assumed above
and below. This jump would be smaller for a lower value of g
, and
approximately removed for g
=8. Our results will show that such a
lower value of g
(corresponding to a less steep profile in the lower
mantle) also increases the fit to the geoid. Factors q
i
in individual
layers (lithosphere, upper mantle, one to two layers transition zone,
lower mantle) are treated as free parameters in the optimization dis-
cussed below. Figuratively speaking, the optimization consists of
shifting corresponding parts of the curves in Fig. 4 to the left or
right. ˜η(z) is only shown below depth 70 km, because the mecha-
nisms discussed here are not appropriate to model deformation of the
lithosphere, and lithospheric viscosity is treated as a free parameter.
It is shown for the same three cases as in Fig. 2. All profiles shown
Figure 4. Black and brown lines: non-optimized, normalized viscosity pro-
files for for the same cases as shown in Fig. 2. Additionally, the violet line
shows the profile computed with g
= 20 instead, and the orange line shows
the profile of Calderwood (1999). Dashed lines are adiabatic profiles, solid
lines are with thermal boundary layers. During the optimization, parts of
the profiles are shifted left or right relative to each other, yielding optimized
profiles as in Figs 9, 13, 14 and 15.
feature a characteristic hump in the lower mantle, like the profiles
derived by Ranalli (1995) for whole-mantle flow. The curvature of
the hump depends on the curvature of the assumed melting curve.
For example, the profile derived by Calderwood (1999) in a similar
manner is less curved in most of the lower mantle. The height of this
hump strongly depends on the value of g
and also depends on man-
tle potential surface temperature. Variations in the shape of the ˜η(z)
profile in the lower mantle as a result of changing potential surface
temperature by ±100 K can be closely mimicked by appropriate
(small) changes in g
and are, therefore, not considered separately.
Whereas in the reference case, the viscosity increase from below
660 km to the maximum in the lower mantle is about a factor 40,
it is about a factor 100 for the brown line, although the only dif-
ference is a slightly smaller temperature increase towards the base
of the mantle, and it is even about a factor 500 for the violet line
with g
= 20. We, therefore, discuss the effect of g
on our model,
and which range of g
gives best results. Additionally, we discuss
the effect of d
D

and T
CMB
, corresponding to the thickness of the
basal mantle layer where viscosity is decreased, and to the total vis-
cosity decrease, and which combinations of T
CMB
and d
D

yield
optimal results. Steinberger & Holme (2006) extend this discussion
to include the effect of a basal layer with chemical variations.
2.5 Relation between seismic velocity variations
and temperature variations
If variations in seismic s wave speed are of thermal origin, they
relate to temperature variations through
δT = (δv
s
/v
s
(z))/( ln v
s
/∂T )
p
. (20)
δv
s
/v
s
(z) are relative seismic wave speed variations, for which a
number of global tomography models exist. (ln v
s
/∂T )
p
is the
partial derivative of the logarithm of seismic s wave speed with re-
spect to temperature at constant pressure. Following Karato (1993),
it can be written as
ln v
s
T
p
=−
ln v
s,0
T
p
+ F(α) ·
Q
1
π
·
H
RT
2
, (21)
whereby the first term is the anharmonic part and the second one
the anelastic part. Q is the seismic Q-factor, and we set F(α) = 1.
With eq. (10) this becomes
ln v
s
T
p
=−
ln v
s,0
T
p
+ F(α) ·
Q
1
π
·
gT
m
T
2
, (22)
in the lower mantle.
2.5.1 Anelastic part
We consider two Q profiles here: The first one follows Anderson &
Hart (1978) except that a continuous instead of a stepwise profile is
used. It is similar in shape to the logarithm of the non-dimensional
viscosity profiles ˜η(z) in Fig. 4, which makes intuitively sense, since
both anelasticity and viscosity are caused by similar mechanisms.
Because of this similarity, weuse the Anderson & Hart (1978) profile
rather than a newer one. For the second profile, we choose the shape
to exactly coincide with log(˜η(z)), hence we choose a, b and c such
that Q
2
(z) = a + bz + clog(˜η(z)) agrees with Anderson & Hart
(1978) at depths 250 km, 2400 km and the base of the mantle. Both
profiles for the anelastic part are shown in Fig. 5. In the reference
case, weusean anelastic part that is thearithmetic average for the two
cases described. Q is also expected to depend on temperature, hence
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1468 B. Steinberger and A. R. Calderwood
Figure 5. Profiles of ( ln v
s
/∂T )
p
, and contributing parts. Red line:
anharmonic part. Green lines: anelastic part. The dotted line is for the
Q-
profile of Anderson & Hart (1978), the dashed line for Q
2
(z). Black line:
anharmonic plus anelastic part in the reference case (arithmetic average of
two green lines). Orange line: profile derived by Calderwood (1999). The
blue line uses g = 30 for the anelastic part (eq. 22) instead. Brown line:
same case as in Figs 2 to 4. Other model assumptions are as in the reference
case.
a more accurate treatment should also consider lateral variations of
Q, for which models exist, at least in the upper mantle (Gung &
Romanowicz 2004).
2.5.2 Anharmonic part
For (ln v
s,0
/∂T )
p
we adopt the profile given by Goes et al. (2004) in
the upper mantle. It is based on values for individual phases and the
mantle phase diagram of Ita & Stixrude (1992) for a pyrolite mantle,
using a geotherm similar to the reference case geotherm in Fig. 2.
In the lower mantle, (ln v
s,0
/∂T )
p
is recomputed to ensure con-
sistency with our profile of α(z). We follow Duffy & Anderson
(1989) in assuming that radial derivatives of shear modulus μ for
different temperatures are related through
dμ
dz
(
T (z) ±T (z)) =
dμ
dz
(
T (z))(1 ±α(z)T (z)). (23)
This is based on the observation that in the PREM (Dziewonski &
Anderson 1981) lower mantle approximately
d ln
dμ
dp
d ln ρ
≈−1, (24)
and the implicit assumption that this also holds along isobars.
μ(
T (z) ± T (z)) is determined by integrating eq. (23) along
adiabats, while computing temperature and thermal expansion co-
efficients along the same adiabats. Starting point is
μ(T
lm,0
± T
0
) = μ(T
lm,0
) ±T
0
·
dμ
0
dT
, (25)
at z =0. Based on parameters compiled by Cammarano et al. (2003)
and the pyrolite phase diagram of Ita & Stixrude (1992) it is es-
timated
dμ
0
dT
= 27 MPa K
1
for lower mantle mineralogy. T (z)
and α(z) are computed for the entire mantle with the lower man-
tle formulation described above. ‘Lower mantle potential surface
temperature’ T
lm,0
is found iteratively by matching the previously
determined lower mantle temperatures.
dμ
dz
(T (z)) and μ(T
lm,0
) are
determined from the parameters of PREM layer 4, which comprises
most of the lower mantle. Thus
∂μ
T
p
μ(
T (z) +T (z)) μ(T (z) T (z))
2T (z)
, (26)
is found, and with this (ln v
s,0
/∂T )
p
can be computed (Fig. 5).
2.5.3 Sum of anharmonic and anelastic part
For the sum of anharmonic and anelastic part, agreement between
our reference case and Calderwood (1999) is good in the upper
mantle (Fig. 5), and would be even better if Q
2
(z) was used, but
our reference profile is lower in the lower mantle: There, we choose
the factor g = 12 in the relation between activation enthalpy and
melting temperature. This value is much lower than previous es-
timates of g = 20–40 (Karato 1993) and thus leads to a smaller
anelastic contribution: The blue curve where g = 30 is used for
computation of the anelastic part (eq. 22) is higher in the lower man-
tle. The brown curve corresponds to somewhat lower temperatures
in the lowermost mantle (see Fig. 2) and is, therefore, also some-
what higher in the lowermost mantle. On the other hand, because
of the different way of computation, we compute a larger anhar-
monic part than Karato (1993) in the lower mantle. In our model,
the anharmonic part is larger below 660 km than above 660 km, be-
cause of larger values of
dμ
0
dT
(Cammarano et al. 2003). Differences
for anharmonic and anelastic parts between our results and those
of Karato (1993) partly compensate each other, such that the sum
curve we obtain in our reference model is similar to that of Karato
(1993).
2.6 Relation between seismic velocity variations and
density variations
The scaling factor ( ln ρ/ln v
s
)
p
is then obtained by dividing
thermal expansivity through (ln v
s
/∂T )
p
(Fig. 6). In the refer-
ence case, it is similar to Calderwood (1999) in the upper man-
tle, but, corresponding to differences in Figs 3 and 5 somewhat
larger in the lower mantle. Again, the overall shape is similar to that
of Karato (1993). For the profiles derived here, the scaling factor
somewhat decreases with depth in the lower mantle, whereas the
Figure 6. Profiles of scaling factors ( ln ρ/ ln v
s
)
p
for the same cases
as in Figs 3 and 5. Additionally, the dotted blue line is with a
0
= 3.5 and
g = 30.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1469
profile derived by Calderwood (1999) is almost constant over most
of the lower mantle, due to a smaller decrease of thermal expan-
sivity with depth. In the case with constant δ
T
(brown line), the
two effects of stronger decrease of thermal expansivity with depth
(Fig. 3) and less decrease of the (ln v
s
/∂T )
p
profile (Fig. 5)
combine to a much stronger decrease of scaling factor with depth.
A lower scaling factor in the lower mantle is obtained if either
alowerα
0
(black dashed line) or a larger g (blue lines) is as-
sumed. Changing potential surface temperature shifts the scaling
factor profile (less than about ±10 per cent change for ±100 K po-
tential temperature change), and similar changes would also occur
if a different Q-factor profile was used. Because we obtain simi-
lar changes also by changing F
l
, a
0
and/or g, changes in poten-
tial surface temperature and Q-factor profile are not considered
separately.
In the reference model, the scaling factor is arbitrarily reduced,
compared Fig. 6, by a factor F
l
= 0.5 in the uppermost 220 km,
in order to account for the fact that density variations within the
lithosphere are probably only partly of thermal origin. Similarly,
other studies set density anomalies at lithospheric depths to zero
(Lithgow-Bertelloni & Silver 1998; Becker et al. 2003; Behn et al.
2004). However, we also consider the effect of varying F
l
on our
results and discuss which values give best results. In most cases,
and unless mentioned otherwise, we use tomography model smean
(Becker & Boschi 2002) for seismic velocity variations. Other to-
mography models are also considered.
3 OPTIMIZATION PROCEDURE
3.1 Mantle flow computation
We can now combine a scaling factor profile as derived in the pre-
vious section with an s-wave tomography model to obtain a density
model. With a viscosity profile as described above, mantle flow can
be computed. Equations of viscous flow are solved with a spectral
method originally developed by Hager & O’Connell (1979, 1981):
Only radial viscosity variations are considered, thus the equations
decouple in the spherical harmonic domain and can be solved sep-
arately for each spherical harmonic. In the reference case, and if
not mentioned otherwise, spherical harmonic expansion is done up
to l
max
= 15, but cases with l
max
= 31 are considered as well. The
original method was modified to consider depth-dependent gravity
(from PREM), and to include the effects of compressibility (as de-
scribed in Steinberger 2000, following Panasyuk et al. 1996), and
of deflections δz of phase boundaries at depth 400 and 660 km in
thermal equilibrium. The latter effect was included as sheet mass
anomalies
δz · ρ =
ρ
ρ
2
pb
γ
pb
α
pb
· δρ , (27)
whereby δρ is the density anomaly as inferred from the tomogra-
phy model, at the depth of the phase transition. ρ
pb
, γ
pb
and α
pb
are density, gravity and thermal expansivity at the phase bound-
ary. Values for ρ were given in the previous chapter. For ρ
pb
,
the average value between above and below the phase boundary
(from the PREM model) is used, for α
pb
, we use 2.079 ·10
5
K
1
at
660 km and 2.378 ·10
5
K
1
at 400 km—approximate average val-
ues between above and below the phase boundary from the reference
case in the previous chapter.
3.2 Variables in optimization
The optimization consists of finding the minimum of a misfit func-
tion in parameter space. Parameter space consists of the viscosity
factors q
i
, i =1 ...n
vis
in eq. (19). n
vis
is the number of layers where
viscosity is considered to vary independently. In most cases, we use
n
vis
= 4, in which case i = 1 corresponds to the lithosphere (0–
100 km), i = 2 the upper mantle (100–400 km), i = 3 the transition
zone (400–660 km) and i = 4 the lower mantle (below 660 km). In
some cases, the upper part of the transition zone (400–520 km) and
its lower part are considered separately. In this case, it is n
vis
=5, i =
3 corresponds to the upper part of the transition zone, i =4 its lower
part and i = 5 the lower mantle. Besides these parameters allowed
to vary within each optimization, variations are also considered for
a number of further parameters. These are kept constant within
each optimization, and the optimization is performed for several
sets of parameter values, and several 2-D grids in parameter space.
These further parameters include a
0
, b, g, g
, T
CMB
, d
D

and F
l
(see Table 1).
3.3 Haskell constraint
The Haskell constraint states that the logarithmic average of vis-
cosity, weighted with an appropriate sensitivity kernel equals about
10
21
Pa s. We express it in the form
CMB
120 km
log
10
η(r)
1Pas
· K (r) dr/
CMB
120 km
K (r) dr = 21, (28)
and use the sensitivity kernel for the Angerman River site given by
Mitrovica (1996) for K(r) (Fig. 7). If we choose η
0
such that, in
eq. (19) η
0
· ˜η(z) satisfies the Haskell constraint, we find that η(z)
also does if 1.67 ·q
2
+1.33 ·q
3
+q
4
0 in the case of one transition
zone layer, and 1.67 ·q
2
+ 0.67 ·q
3
+ 0.66 ·q
4
+q
5
0 in the case
of two. We, therefore, choose the penalty function
P
1
= (1.67 · q
2
+ 1.33 ·q
3
+q
4
)
2
, (29)
in the first case, and a corresponding function in the second.
2500
2000
1500
1000
500
0
depth [km]
0 2 4 6 8 10 12
sensitivity kernel
Figure7. Kernel showing the sensitivity of post-glacial rebound to viscosity
at various depths. From Mitrovica (1996) for Angerman River.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1470 B. Steinberger and A. R. Calderwood
3.4 Heat flux constraint
For each density field and corresponding flow field, the advected
heat flux is computed by integrating radial velocity v
r
times heat
anomaly δT ·C
p
·ρ =δ ln v
s
/( ln v
s
/∂T )
p
·C
p
·ρ over spherical
surfaces at given depths:
(r) = C
p
· ρ/( ln v
s
/∂T )
p
·
S
δ ln v
s
· v
r
dA. (30)
For this computation present-day surface plate motions (DeMets
et al. 1990) are used as boundary condition. This curve should ap-
proximately fall between theoretical steady-state profiles for an in-
ternally heated mantle and a basally heated mantle. Here we imple-
ment this constraint by penalizing if (r) >
bh
or (r) <
ih
(r).
For
bh
we use a constant value 33 TW representing our estimate
for the mantle heat flux (Calderwood 1999), which is global heat
flux minus heat produced in the crust. Pollack et al. (1993) compute
a global heat flux 44 TW. By subtracting estimated heat production
in the continental crust from it, Schubert et al. (2001) estimate a
mantle heat flux 37 TW. More recently, Hofmeister (2005) com-
puted a smaller global heat flux 31 ± 1 TW, corresponding to only
about 24 TW mantle heat flux. However, another recent analysis
(Wei & Sandwell 2006) yielded 42–44 TW.
ih
(r) has been modi-
fied from the theoretical steady-state profile for an internally heated
mantle with constant heat production rate per unit mass in order to
account for the fact that conduction is the dominant heat transport
mechanism close to the surface (see Fig. 8). We use the penalty
function
P
2
=
max
(r)
bh
bh
, 0
2
+ max
ih
(r)(r)
bh
, 0
2
2891 km
dr. (31)
Obviously, the assumption that heat fluxat anydepth does not change
with time, which is implicitly made here, is not exactly satisfied for
the Earth. (r) is time variable and the present-day surface heat flux
may be anomalously high or low. Hence we do not require P
2
= 0
exactly.
Figure 8. Optimized viscosity profiles (left) and advected heat flux profiles (right), for cases where not all constraints are imposed. Orange lines: only heat
flux and Haskell constraint. Violet lines: only geoid and Haskell constraint. Green lines: only geoid and heat flux constraint. Red lines: viscosity profile not
based on mineral physics but constant in lithosphere, upper mantle, transition zone and lower mantle. Continuous lines are with parameters as in the reference
case, dotted lines as in model 2.
bh
and
ih
(r) are shown as grey dashed lines. The heat flux profiles should approximately lie in between.
3.5 Geoid constraint
For each density field and corresponding flow field, the inferred
geoid is computed; flow deforms boundaries, and both the density
anomalies that drive flow and the deformed boundaries contribute
to the geoid. The inferred geoid is, therefore, quite sensitive to the
assumed viscosity structure. The fit between computed and actual
geoid is expressed in terms of the penalty function
P
3
=
Var(Predicted–Observed)
Var(Observed)
. (32)
1 P
3
is referred to as (geoid) variance reduction. For the geoid
computation, stress-free boundaries at the Earth’s surface and CMB
are assumed in the reference case; however, other boundary condi-
tions at the Earth’s surface are considered as well.
3.6 Misfit function
We optimize the model in parameter space by minimizing a misfit
function using a downhill simplex method (Press et al. 1986). This
method finds a local, not necessarily global minimum in parameter
space. In most cases, we will either minimize
1
= c
1
· P
1
+ c
2
· P
2
+ P
3
, (33)
(misfit criterion 1) or require the Haskell constraint to be fit exactly,
and minimize
2
= c
2
· P
2
+ P
3
, (34)
under the additional constraint P
1
= 0 (misfit criterion 2, corre-
sponding to criterion 1 with c
1
−→ ). Cases indicated with a
star in Table 2 use criterion 2 (except for M17, M17b in Section 4.9
which use a modified criterion), all others (including those in Figs 11
and 12, but except for cases M19 and M19b in Table 1 and Sec-
tion 4.2) use criterion 1. In most cases, we will use c
1
= 1, c
2
= 4. We find that results remain very similar with both criteria,
and hence do not strongly depend on c
1
, as the Haskell constraint
is always fit very well. The value for c
2
is discussed in the next
section.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1471
Table 2. List of numerical models. The ‘comments’ column indicates differences to the reference model, except for ‘b’ models, where it
indicates the difference to models without ‘b’. VR is geoid variance reduction, P
2
heat flow misfit (eq. 31). Numbers in brackets indicate
either results with expansion up to l
max
= 31 (M1–M3b), or with model 2 parameters (M7–21). Stars indicate cases where an exact fit
to the Haskell constraint has been prescribed. UM-upper mantle, TZ-transition zone, LM-lower Mantle.
Model # comment VR [per cent] P
2
M1
Reference case 75.9 (74.7) 0.0165 (0.0133)
M1b
Lowest viscosity in asthenosphere 71.6 (70.5) 0.0125 (0.0090)
M2
a
0
= 3.5, g = 30 79.7 (78.6) 0.0133 (0.0099)
M2b Lowest viscosity in asthenosphere 76.0 (75.4) 0.0109 (0.0078)
M3
Based on Calderwood (1999) 79.5 (78.5) 0.0122 (0.0093)
M3b Lowest viscosity in asthenosphere 75.3 (74.6) 0.0105 (0.0076)
M4
δ
T
= 5.5 constant 65.7 0.0044
M4b
Lowest viscosity in transition zone 52.8 0.0140
M5
g
= 20 15.5 0.0356
M6
η = const in UM, TZ; model 2 parameters 79.6 0.0138
M7 (M7b) η = const in UM, TZ, LM 81.2 (80.8) 0.0163 (0.0143)
M8
(M8b
) Independent η in upper, lower TZ 80.0 (79.7) 0.0164 (0.0131)
M9
(M9b
) Tomography model S20RTS 74.2 (76.1) 0.0129 (0.0114)
M10
(M10b
) Tomography model SB4L18 47.9 (64.6) 0.0175 (0.0052)
M11
(M11b
) Tomography model Grand (2002) 67.7 (69.0) 0.0250 (0.0296)
M12
(M12b
) Tomography model SAW24B16 64.7 (74.3) 0.0187 (0.0154)
M13
(M13b
) Tomography model S12WM13 44.3 (56.0) 0.0188 (0.0086)
M14
(M14b
) Tomography model S362D1 55.9 (69.3) 0.0116 (0.0066)
M15
(M15b
) Fixed surface 38.4 (45.8) 0.0138 (0.0194)
M16 (M16b) Plates moving half free speed 11.2 (19.1) 0.0082 (0.0052)
M17
(M17b
) Plates free speed, tractions at surface, v
rms
penalty 24.1 (31.0) 0.0068 (0.0130)
M18
(M18b) Prescribed plate motions 13.9 (14.8) 0.0029 (0.0034)
M19 (M19b) Only heat flux and Haskell constraint 3140.3 (1923.6) 0.0000 (0.0001)
M20
(M20b
) Only geoid and Haskell constraint 78.0 (79.7) 0.1133 (0.0134)
M21 (M21b) Only geoid and heat flux constraint 78.0 (79.5) 0.0173 (0.0138)
4 RESULTS
4.1 Overview
We first show optimized viscosity and heat flux profiles (Fig. 8), as
well as variance reduction and heat flow misfit P
2
for cases where
the constraints are not all imposed simultaneously (M7, 7b, 19–21b
in Table 2). This includes one case with constant viscosities in dif-
ferent mantle layers instead of profiles based on mineral physics.
We then go on to imposing all constraints simultaneously and show
Figure 9. Optimized viscosity profiles (left) and advected heat flux profiles (right). Colour and type of lines correspond to Figs 4 and 6.
bh
and
ih
(r) are
shown as grey dashed lines. Per diagram there are two black, orange and blue lines: one, with higher heat flux values in the upper mantle, was computed with
l
max
= 31, the other with l
max
= 15.
results for our reference model (M1; Figs 9 and 10), further ‘pre-
ferred’ models (M2 with a
0
=3.5 and g =30, and M3 with profiles
from Calderwood (1999)) and further models (M4, M4b and M5)
with stronger depth dependence of either lower mantle viscosity
or thermal expansivity (Fig. 9). We then systematically vary pa-
rameters a
0
, g, g
, n = g/g
, T
CMB
, d
D

and F
l
within bounds as
listed in Table 1, and present contour plots of optimized variance
reduction (Fig. 11) and heat flow misfit P
2
(Fig. 12) as a function
of two of these parameters. Subsequently, we consider results with
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1472 B. Steinberger and A. R. Calderwood
observed
predicted
-100 -80 -60 -40 -20 0 20 40 60 80 100
geoid in m
Figure 10. Above: geoid computed for the reference case. below: observed geoid.
the further constraint that viscosity is lowest in the asthenosphere
(M1b, M2b, M3b), with constant viscosity in the upper mantle and
transition zone (M6) viscosity in the upper and lower part of the
transition zone varying independently (M8, 8b) (all in Fig. 13), dif-
ferent tomography models (M9–14b; Fig. 14), and different surface
boundary conditions (M15–18b; Fig. 15).
4.2 Cases where not all constraints are imposed
simultaneously
Discussing model cases where not all constraints have been im-
posed helps clarifying which features of the results are due to which
constraints. Since the geoid fit is only affected by relative viscosity
variations, we need at least one additional constraint for the abso-
lute viscosity level—either the Haskell or the heat flux constraint.
If we impose geoid and Haskell constraint (use misfit criterion 2
with c
2
= 0) we obtain a geoid variance reduction of 78 per cent or
80 per cent with other assumptions as in the reference case (M20)
or in case M2 (M20b). Resulting viscosity structure (violet lines in
Fig. 8, left-hand side panel) is very similar in both cases. For M20,
the computed heat flux is too large in the lower mantle (continuous
violet line in Fig. 8, right-hand side panel), whereas for M20b (dot-
ted violet line) it matches the observation-based estimate of mantle
heat flux quite well, even without imposing any heat flux constraint.
If we impose instead geoid and heat flux constraint (use misfit cri-
terion 1 with c
1
=0 and c
2
=4), we obtain heat flux profiles (green
lines in Fig. 8, right-hand side panel) similar to M20b. The opti-
mized viscosity structure in case M21b (dotted green line in Fig. 8,
left-hand side panel) remains very similar to case M20b, that is, the
Haskell constraint is matched well in M21b even without impos-
ing it. In case M21 (continuous green line), viscosity is somewhat
higher than in case M20 by approximately a factor 1.6 throughout
the mantle. Geoid variance reduction (78 per cent in M21 and 79
per cent in M21b) remains similar to cases M20 and M20b, hence
with this relative weighting of geoid and heat flux constraint (c
2
=
4) the latter is not overly constraining.
We find that both Haskell and heat flux constraint can be simul-
taneously fit almost perfectly (M19 and M19b); therefore, there is
no benefit in looking at cases with only one of them imposed. Re-
sulting viscosity and heat flux profiles if we impose Haskell and
heat flux constraint (i.e. minimize
1
in eq. (33) with c
1
= 1,
c
2
=4, and without the term P
3
) are shown as continuous (M19) and
dotted (M19b) orange lines in Fig. 8. However, there is a huge misfit
between predicted and observed geoid. Thus, the geoid constraint
is the most essential one.
Finally, with constant viscosities in upper mantle, transition zone
and lower mantle we are able to obtain geoid variance reduc-
tion 81 per cent and reasonable heat flux profiles (cases M7 and
M7b), with lower mantle viscosity of about 3 · 10
22
Pa s. This is
similar to results obtained in previous work. The main objective
of this paper is to find out whether we can replace the assump-
tion of layers with constant viscosities (in particular in the lower
mantle) with viscosity profiles based on mineral physics without
substantially degrading the fit to the geoid, heat flux and Haskell
constraints.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1473
Figure 11. Geoid variance reduction as a function of a
0
and g (but g
=12 constant) (top left), of g
and n =g/g
(top right), of g = g
and T
CMB
(centre right),
of d
D

and T
CMB
(centre left), and of d
D

and F
l
(bottom left); other parameters as in the reference case. These are cross-sections through a 5-D parameter
space intersecting along the grey lines S1, S2, S3 and S4. For each point, the model has been optimized by minimizing the misfit function P
1
+ 4P
2
+ P
3
in
a 4-D parameter space. The thick black line divides regions where the lowest viscosity occurs in the transition zone from where it occurs in the asthenosphere
(indicated by the letter ‘a’). Location of models M1, M2 and M5 is also indicated.
4.3 Reference case
Fig. 10 compares the predicted geoid with the observed one in the
reference case (M1). The variance reduction is 76 per cent. This is
slightly less than in cases M20 and M21 above—as more constraints
are imposed simultaneously, the fit to the geoid is somewhat reduced.
The optimized viscosity model and corresponding advected heat
flux profiles for cases given in Figs 4 and 6 are shown in Fig. 9; for
the reference case (black lines), the difference between the lowest
viscosity (in the transition zone) and the highest viscosity at the base
of the mantle is about a factor 700. With the relative weighting of
geoid and heat flux constraint c
2
= 4 variance reduction decreases
by only a few per cent compared to the case without any heat flux
constraint. At the same time the heat flux constraint is probably
fit well enough considering possible time variations in heat flux
mentioned above. We will hence maintain this relative weighting in
what follows.
4.4 Dependence on thermal expansivity and relation
between s wave speed and temperature
We can achieve an even higher variance reduction with a somewhat
higher α and larger g in eq. (22) while keeping g
=12 in the viscosity
law eq. (11), resulting in higher values of (ln v
s
/∂T )
p
and lower
values of ( ln ρ/ ln v
s
)
p
. For example, the model with a
0
= 3.5
and g = 30 (M2; blue dotted lines) gives a variance reduction of
80 per cent. Also, the heat flux profile now fits even better, as it
stays below 34 TW throughout the mantle. We saw above for cases
M20b and M21b that with these parameters a good fit to the heat
flux or Haskell constraint is obtained even without applying those
constraints. Hence it is not surprising that results stay very similar
if, in case M2, we apply these constraints simultaneously.
Also, if both profiles for thermal expansivity and ( ln v
s
/∂T )
p
are adopted from Calderwood (1999), a variance reduction of
80 per cent is achieved along with a suitable heat flux profile
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1474 B. Steinberger and A. R. Calderwood
Figure 12. Misfit P
2
, as defined in eq. (31), of the heat flux profile for the same cross-sections and models as in Fig. 11, where further explanations are given.
(M3; orange lines). However, heat flux in the uppermost few 100 km
below the lithosphere is under-predicted in all three models dis-
cussed so far—and this is in fact a general problem with all models
that give otherwise a good fit. Likely explanations for this misfit
are discussed in the last section. This misfit is somewhat less with
l
max
= 31. In this case geoid variance reduction is slightly reduced
to 75, 79 and 78 per cent, respectively, for the three cases M1, M2
and M3 considered thus far.
The dependence of variance reduction and misfit function on g
used in the velocity-temperature relation eq. (22) and a
0
, but keeping
g
= 12 in the viscosity law eq. (11) is further illustrated in the top
left panels of Figs 11 and 12: If a
0
and g are both increased relative to
the reference case g =12, both geoid variance reduction and the heat
flux profile are somewhat improved. For g =12 the highest variance
reduction is obtained with somewhat smaller a
0
3 whereas for
g =42 itis obtained with somewhatlarger a
0
4; both combinations
yield values of ( ln ρ/ ln v
s
)
p
0.3 in the lower mantle, which
appears to work best for achieving a good fit to the geoid.
The better fit of the heat flux profile for larger g is a result of
the higher values of ( ln v
s
/∂T )
p
in the lower mantle, resulting
in lower temperature anomalies inferred from seismic tomography,
thus lower inferred heat flux. Probably the improved variance reduc-
tion for larger g is due to the same reason; once the heat flux con-
straint becomes less difficult to satisfy, the optimization can rather
focus on optimizing the fit to the geoid.
In the case of a stronger decrease of α with depth (no depth
dependence of δ
T
, that is, b =0; M4; brown lines in Fig. 9) variance
reduction is reduced to 66 per cent. Both the resulting optimum
viscosity structure with a rather high transition zone viscosity and
heat flux profile look somewhat unrealistic. If, in this case, the lowest
viscosity is restricted to occur in the transition zone, profiles look
more realistic, but variance reduction in this case (M4b) is further
reduced to 52 per cent. Thus, the smaller decrease of α with depth in
the reference case appears more appropriate to achieve good results.
This is probably partly due to the fact that a constant δ
T
leads to less
temperature increase with depth, hence a steeper viscosity profile.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1475
4.5 Dependence on lower mantle viscosity
A steeper viscosity profile is also obtained with a larger value of
g
. With g
= 20 (M5; violet lines in Fig. 9), we obtained a vari-
ance reduction of only 15 per cent and the heat flux profile also
looks unrealistic. The dependence of variance reduction and heat
flux misfit on the steepness of the viscosity profile is further illus-
trated in the centre right panels of Figs 11 and 12: The viscosity
increase through the lower mantle becomes larger with increasing
g
, the viscosity drop at the base of the mantle becomes larger with
increasing T
CMB
. An increase of either of them relative to the ref-
erence case leads to substantial deterioration of both the predicted
geoid and heat flux profile, but decrease does not yield substantial
improvement. In fact, low values of g tend to also give a larger misfit
of the heat flow profile; this arises, as the viscosity hump in the lower
mantle is then less high, and thus the heat flux predicted tends to be
too high in part of the lower mantle.
The top right panels of Figs 11 and 12 show that the deteriorating
fit with increasing g and g
in the centre right panels is in fact due to
the viscosity profile; if we leave g
constant in the viscosity relation
eq. (11) but increase g =g
n in eq. (22) through increasing n, the fit
rather tends to improve—for reasons already discussed for the top
left panels.
There is also a trade-off between the assumed temperature at
the base of the mantle and the thickness of the thermal boundary
layer d
D

at its base, as shown in the centre left panels of Figs 11
and 12: For d
D

350 km, the highest geoid variance reduction
is obtained with T
CMB
3000 K, for d
D

200 km with T
CMB
3500 K (reference case), andfor d
D

125 km with T
CMB
4000 K.
The smallest heat flow misfit tends to occur with somewhat smaller
thermal boundary layer thicknesses and/or CMB temperatures, as
this gives smaller predicted heat flux in the lowermost mantle, which
is too high in the reference model.
4.6 Dependence on upper mantle velocity
to density scaling
In the reference model, the scaling factor is arbitrarily reduced by
a factor F
l
= 0.5 in the uppermost 220 km, in order to account
Figure 13. Optimized viscosity profiles (left) and heat flux profiles (right). Black, orange and dotted blue lines (two per diagram) correspond to Figs 9, but
with the restriction that lowest viscosities are to occur in the asthenosphere. Light green and dark green dotted lines are for the same models as black and blue
dotted lines, but with viscosities in the upper and lower transition zone independently varying.
bh
and
ih
(r) are shown as grey dashed lines. Red dotted lines
are for the same case as blue dotted, but with constant viscosities 100–410 and 410–660 km.
for the fact that density variations within the lithosphere are prob-
ably only partly of thermal origin. The bottom left panel of Fig. 11
shows that this is in fact an appropriate choice; the highest variance
reduction occurs for intermediate values of F
l
around 0.5, but the
fit to the geoid does not strongly depend on F
l
. On the other hand,
the bottom left panel of Fig. 12 shows that a better fit to the heat
flow profile is obtained with larger F
l
, because this increases the
computed heat flow below the lithosphere, thus reducing the misfit
there, particularly in cases where the lowest viscosity occurs in the
asthenosphere.
4.7 Dependence on upper mantle and transition
zone viscosity
The resulting optimized viscosity structure remains very similar
in all three cases included in Fig. 9 that give good results—the
reference case, model 2 and the case with profiles from Calderwood
(1999); it shows a pronounced viscosity decrease in the transition
zone—about one order of magnitude relative to above or below.
In order to assess whether the result that the lowest viscosity in
our reference model occurs in the transition zone is of significance,
we also include an optimization restricted to models with transition
zone viscosities higher than in the upper mantle. The variance re-
duction is then slightly reduced to 72 per cent (71 per cent) with
reference case parameters (M1b), 76 per cent (75 per cent) with
model 2 parameters (M2b), and 75 per cent (75 per cent) with the
profiles from Calderwood (1999) (M3b) (numbers in brackets with
l
max
= 31). The heat flux profile remains very similar, and the vis-
cosity maximum in the lower mantle remains similar at about, or
slightly below 10
23
Pa s for all three models (Fig. 13). We find a
general tendency that models with lowest viscosity in the transition
zone tend to yield somewhat higher variance reduction than models
with the lowest viscosity in the asthenosphere (Fig. 11). It is hard
to judge whether this slightly reduced variance reduction warrants
attaching significance to the result that the lowest viscosity occurs
in the transition zone. The best fit of the heat flux profile is obtained
for cases where the lowest viscosity occurs in the asthenosphere,
since this allows for a higher heat flux in the uppermost few 100 km
below the lithosphere, thus reducing the misfit there (Fig. 12).
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1476 B. Steinberger and A. R. Calderwood
Results stay very similar, if the viscosity is kept constant in the
upper mantle, and constant in the transition zone (Fig. 13, red dotted
lines). This confirms that results—as stated in Section 2.2.2—do
not strongly depend on the activation enthalpy profile in the upper
mantle. Furthermore, we consider the case that viscosity in the upper
and lower part of the transition zone, that is, above/below 520 km,
varies independently; in the reference case, this leads to an increase
in variance reduction to 80 per cent, with a rather strong viscosity
contrast between the upper and lower part of the transition zone. In
model 2, variance reduction remains 80 per cent, and the optimized
viscosity structure remains very similar. We, therefore, regard the
constraints of our method as insufficient to resolve details of the
viscosity structure such as the difference between the upper and
lower part of the transition zone. We also computed results without
consideration of phase boundaries, for an incompressible mantle,
and for the two different Q profiles; in all cases results stay very
similar, and are therefore not shown.
4.8 Dependence on tomography model
Fig. 14 shows results for different tomography models with model
2 parameters. In most cases, the smallest misfit occurs with the
lowest viscosity in the transition zone, only for SB4L18 (M10b) the
smallest misfit occurs with the lowest viscosity in the asthenosphere.
With reference case parameters, this happens with SB4L18 (M10),
Grand’s model (M11), SAW24B16 (M12) and S362D1 (M14). The
resulting maximum viscosity in the lower mantle is very similar for
all models—around, or slightly below 10
23
Pa s. There are quite
large differences between the heat flux profiles for various models.
Largely these are due to the differences in rms amplitudes versus
depth among the tomography models. For example, Grand’s model
(blue lines) has higher rms amplitudes than S20RTS (red line) in
the lowermost mantle, but lower amplitudes between 1500 and
300 km depth. The model smean (black) is an average of Grand’s
model (blue), S20RTS (red) and SB4L18 (green) and gives a heat
flux profile that is approximately constant through much of the man-
tle. Geoid variance reductions and heat flow misfit P
2
of optimized
models are listed in Table 2. There is a general tendency for better
Figure 14. Optimized viscosity profiles (left) and heat flux profiles (right) for different tomography models. Other assumptions are as in model 2. Black: smean
(Becker & Boschi 2002)—identical to the blue dotted line in Fig. 9; red: S20RTS (Ritsema & van Heijst 2000); green: SB4L18 (Masters et al. 2000); blue:
model by Grand (2002); violet: SAW24B16 (M´egnin & Romanowicz 2000); orange: S12WM13 (Su et al. 1994); brown: S362D1 (Gu et al. 2001).
bh
and
ih
(r) are shown as grey dashed lines.
fit with model 2 parameters, which gives smaller density anomalies.
In particular, the difference is pronounced for those models with
comparatively large amplitudes away from the boundaries, that is,
where geoid kernels are larger—such as SB4L18. Such models give
generally lower variance reduction, as they would presumably re-
quire even lower conversion factors from seismic velocity to density
anomalies in order to yield a good fit. Models with lower amplitudes
(S20RTS, Grand) tend to give better fit and the difference between
the fits with reference case parameters and model 2 parameters is
smaller. However, with no individual tomography model we were
able to exceed the variance reductions obtained with the average
model smean.
4.9 Dependence on upper boundary condition
Upper boundary condition for the computation of the geoid in the
models discussed thus far was a free surface. Since this is not entirely
appropriate for a surface broken into plates not well characterized
by viscous rheology, we show in Fig. 15 results for different sur-
face boundary conditions. For fixed surface, variance reduction of
the optimized models with reference case parameters (M15, grey
lines) is 38 per cent, with model 2 parameters (M15b, black lines)
it is 46 per cent. With plates moving freely according to the forces
acting on them due to mantle flow, with tractions applied at the
base of the plates at depth 100 km (see Steinberger et al. 2001,
for a detailed description of the methodology) we obtain a negative
variance reduction—also plates are moving too fast compared to
observations. Thus, we go on to a case with plates moving at half
the speed of free plate motions. In this case (M16, green lines with
reference case parameters; M16b, red lines with model 2 parame-
ters), calculated rms plate motions v
rms,c
are 5.2 cm yr
1
(M16) and
5.3 cm yr
1
(M16b), similar to the observed v
rms,o
= 4.4 cm yr
1
.
For the difference between predicted and observed plate motions,
rms values are 3.8 cm yr
1
in both cases, corresponding to plate
motion variance reductions of 25 per cent (M16) and 23 per cent
(M16b). Geoid variance reduction remains low at 11 per cent
(M16) and 19 per cent (M16b). A similar geoid variance reduc-
tion is obtained with prescribed plate motions (13.9 per cent and
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1477
Figure 15. Optimized viscosity profiles (left) and heat flux profiles (right) for different surface boundary conditions. Fixed surface for reference case (grey)
and model 2 (black); plates moving at half the speed of free plate motions, for tractions applied at the base of the lithosphere at 100 km depth for refere
nce
case (green) and model 2 (red); free plate motions for tractions applied at the surface, with additional penalty added to misfit function if computed rms surface
plate motions differs from observed one for reference case (blue) and model 2 (orange).
bh
and
ih
(r) are shown as grey dashed lines.
14.8 per cent in cases M18 and M18b). This rather poor model per-
formance arises, because plate-like surface motions do not naturally
arise with a viscous rheology, and if they are artificially imposed,
artefacts occur in the computed geoid, which degrade the fit. The
optimized models in this case show rather low viscosities in the up-
per 100 km as in this way artefacts are kept small. Using stresses
at depth 100 km for computing plate motions, if the model does
not contain a high-viscosity lithosphere above, also appears inap-
propriate. We, therefore, go on to a case where we apply stresses
at the surface. For this case, optimized models have very high vis-
cosity in the uppermost 100 km and very slow plate motion; result-
ing sublithospheric viscosity and heat flux profiles are very simi-
lar to the fixed surface case shown. In the last case we, therefore,
add
P
4
= (1 v
rms,c
/v
rms,o
)
2
, (35)
to the misfit function
2
in eq. (34), that is, penalizing if calculated
and observed rms plate motions differ. Then we obtain for reference
case parameters (M17, blue lines) a geoid variance reduction of 24
per cent, v
rms,c
is 4.0 cm yr
1
and plate motion (i.e. surface velocity)
variance reduction 56 per cent. For model 2 parameters (M17b,
orange lines) corresponding numbers are 31 per cent, 3.8 cm yr
1
and 58 per cent. Viscosities in the upper 100 km are a few times
10
22
Pa s—as with a free surface in the three cases discussed that
give a high geoid variance reduction—the reference case M1, model
2 (M2) and the case M3 with profiles from Calderwood (1999). For
these three cases, v
rms,c
is 3.3, 5.0 and 4.6 cm yr
1
, surface velocity
variance reduction is 28, 5 and 4 per cent. Thus, these models with
free surface give approximately correct rms surface velocities even
though these are not used as constraints; lithospheric viscosities (to
be interpreted as effective viscosity along plate boundaries) required
to obtain a good fit to the geoid are in the same range as those
required to obtain appropriate plate speeds. These results do not
change substantially if we penalize a misfit of rms plate motions
as described. The smaller surface velocity variance reduction in
models M1, M2 and M3 can be attributed to the fact that here surface
velocities are not plate-like.
5 DISCUSSION AND CONCLUSIONS
5.1 Resulting viscosity profiles
We have derived here mantle viscosity profiles with a combina-
tion of mineral physics and mantle dynamics. They feature a strong
viscosity increase with depth in the lower mantle up to about
10
23
Pa s or slightly less above the D

layer, several hundred times
more than the viscosity low in the upper mantle. This is a larger vis-
cosity contrast than previously inferred from the geoid constraint
in global geodynamics (e.g. Thoraval & Richards 1997). We obtain
this larger contrast because our resulting viscosity models have a
characteristic hump in the lower mantle; preferred models feature a
viscosity increase of about a factor 30 from below the 660 km dis-
continuity to the viscosity high in the lower mantle, and a viscosity
decrease of somewhat more than a factor 100 from the viscosity high
to the CMB. Models with much larger viscosity increase through the
lower mantle tend to give a much higher misfit and are, therefore,
considered less likely. Models with much larger viscosity decrease
in the lowermost mantle also tend to give much higher misfit, unless
the thermal boundary layer at the base of the mantle is much thinner.
Because the highest viscosities are restricted to the deep mantle, they
are also compatible with results from post-glacial rebound, which is
rather insensitive to viscosities at that depth.
The combination of mineral physics, geoid and heat flux con-
straints applied here was motivated by the fact that the geoid alone
does not place strong constraints on mantle viscosity structure
(Thoraval & Richards 1997). However, we found that our model
still leaves considerable uncertainties; for example, we cannot reli-
ably distinguish whether the lowest viscosities occur in the astheno-
sphere or in the transition zone; generally, a somewhat better model
fit occurs with lowest viscosities in the transition zone. This was
e.g. also found by Marquart (2006). A possible explanation could
be the presence of water in the transition zone (Smyth et al. 2004;
Kavner 2003). Such a viscosity low serves to somewhat decouple
flow in the upper and lower mantle, which appears to improve the
fit to the geoid. However, this can also be achieved by narrow lay-
ers of low viscosity (Panasyuk & Hager 2000; Mitrovica & Forte
2004).
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1478 B. Steinberger and A. R. Calderwood
If we require the lowest viscosities to occur in the asthenosphere,
then best-fit viscosity models look similar to the profile obtained
by Mitrovica & Forte (2004) without mineral physics constraints. A
difference is the narrow layer of low viscosity above depth 670 km
which occurs in the model of Mitrovica & Forte (2004), and which
is not possible with our parametrization.
With alternative surface boundary conditions (fixed or plate-like)
the best-fit models have the lowest viscosity in the asthenosphere.
Any indication of a viscosity low in the transition zone could, there-
fore, merely reflect the inappropriate surface boundary condition
and neglect of lateral variations in rheology. However, this will need
to be addressed by future studies.
5.2 Fit to the geoid
Our best-fit models with viscosity and scaling factor profiles based
on mineral physics give a geoid variance reduction of about 80 per
cent. This is similar to values achieved by previous authors, how-
ever, we were able to obtain this result in a parameter space that is
rather small, as we use mineral physics to pre-constrain the viscosity
structure and conversion factor from seismic velocities to densities.
A variance reduction close to 80 per cent could be achieved for a
range of modelling parameters, but not be substantially exceeded.
With constant viscosities in upper mantle, transition zone and
lower mantle, we obtain a slightly higher geoid variance reduction
(81 per cent). However, we want to replace ad hoc assumptions
(such as constant viscosities) by models based on mineral physics.
We regard the outcome that this only leads to a slightly worse fit to
the geoid, while at the same time a realistic heat flux profile (see
discussion below) is maintained as quite satisfactory.
A geoid variance reduction of slightly more than 80 per cent ap-
pears to be the approximate limit that can be achieved with the inter-
pretation of seismic wave speed anomalies purelyin terms of thermal
anomalies, and without consideration of lateral viscosity variations.
Like Thoraval & Richards (1997), we obtained the best result with
a free-slip surface boundary condition—probably, because this is
the boundary condition that is consistent with the purely viscous
rheology without lateral variations that we also use to model the
lithosphere; plate-like surface velocities without lateral variations
in rheology in the lithosphere lead to artefacts in the computed
geoid.
The good fit to the geoid obtained is indicative that at depths
where the geoid kernels (showing the sensitivity of the geoid to den-
sity variations) are comparatively large—away from the surface and
CMB—seismic anomalies are primarily due to temperature anoma-
lies. This leaves the regions close to the surface and close to the
CMB as potential regions for large chemical variations: Our finding
that the fit to the geoid is somewhat improved, if the conversion fac-
tor from seismic velocities to densities is reduced by about a factor
0.5 in the upper 220 km, may be additional evidence that part of
the density anomalies in the uppermost mantle are due to compo-
sitional, rather than thermal anomalies. A more detailed treatment
of the shallowest mantle would need to consider the non-linear de-
pendence of seismic velocity on temperature anomalies; the same
temperature variation causes larger s wave speed variation at lower
temperature than at higher temperature (Cammarano et al. 2003).
On the other hand, surface velocity variance reduction obtained
is highest with plate-like surface motions. This points towards the
need of combining mineral physics constraints with geodynamic
models that include lateral variations in rheology—particularly in
the uppermost mantle (
ˇ
Cadek & Fleitout 2003).
5.3 Heat flux
In addition to a high geoid variance reduction, our best-fit models
also give heat flux profiles that are compatible with the observation-
based mantle heat flux. Heat flux profiles are quite variable for
different density models—based on different mantle tomography
models—used, but our best-fit models have a roughly constant heat
flux through much of the mantle, corresponding to a mantle largely
heated from below. The rather small fraction of plume heat flux
compared to total surface heat flux has frequently been taken as
evidence for a largely internally heated mantle. However, much
higher values of plume flux have recently been inferred with the
help of seismic tomography (Nolet et al. 2005). Experimental re-
sults (Jellinek et al. 2003) have shown that a significant fraction of
core heat flux can be drawn into large-scale upwellings and is thus
not observed at hotspots. A large part of core heat flux may be due
to warming of slabs settling on the CMB (Yoshida & Ogawa 2005;
Mittelstaedt & Tackley 2006). If seismic velocity anomalies are
due to temperature anomalies, the skewness of s-wave tomography
models suggests significant basal heating (Yanagisawa & Hamano
1999).
Predicted heat flux is generally too low in the uppermost few
hundred km. This may be due to a number of reasons:
(i) The scaling factor in the uppermost 220 km has been arbi-
trarily reduced by a factor F
l
= 0.5; the misfit becomes smaller for
larger F
l
(Fig. 12).
(ii) Heat flow in the lithosphere is dominated by diffusion, which
is not included in our computation, and in many places the litho-
sphere is probably considerably thicker than the 100 km assumed
here.
(iii) Probably a substantial fraction of heat flux occurs through
subduction of lithospheric slabs, and those are not fully resolved
in our model (expansion up to degree 31), especially in the upper
mantle. Heat flow misfit P
2
for expansion up to degree l
max
= 31 is
less than for l
max
= 15. The largest differences occur in the upper
part of the mantle. Most density models used here to drive mantle
flow are expanded to degree 31 or less, and we expect that P
2
would
be further reduced if density and flow models were expanded to
larger l
max
.
(iv) Further heat flux at shorter wavelength may occur through
small-scale convection, primarily in the regions with lowest vis-
cosity. Since the remaining heat flux misfit is primarily in the depth
region 100–400 km, small-scale convection could explain part of the
misfit, especially if the lowest viscosities also occur in that depth
range.
In our reference model, heat flux in the lower mantle tends to be
somewhat too high. Better results are obtained in model 2, where
temperature anomalies are less. While the amplitude of tomography
models is currently not well constrained and the somewhat lower
amplitudes required to obtain a better heat flux profile are probably
within the range of uncertainties, this might also be an additional
indication that some of the anomalies observed by seismic tomog-
raphy are due to compositional, rather than temperature variations
(Trampert et al. 2004).
5.4 Lowermost mantle
Lower boundary condition for the flow computations is a free sur-
face at the CMB. For the optimized models we compute CMB to-
pography and in particular excess ellipticity, which is constrained
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1479
by observations (Mathews et al. 2002). In the reference case, com-
puted CMB excess ellipticity is eight times as large as observed, for
both model 2 and the case with profiles from Calderwood (1999) it
is seven times. We find generally that models that match heat flux
profile and geoid well tend to imply CMB excess ellipticity several
times larger than observed. We can reduce it to values compatible
with observations if we choose sufficiently low viscosities of the
lowermost mantle. However, in this case both the heat flux profile
and geoid fit deteriorate to an unacceptable level. Within the model
space considered, we were not able to obtain an acceptable fit to the
geoid, heat flux profile and CMB excess ellipticity at the same time;
for example, in the centre right panels of Figs 11 and 12 accept-
able CMB excess ellipticity is only found in the top right corner,
where both geoid and heat flux do not fit well. Results do not signif-
icantly change even if the excess ellipticity misfit is penalized in the
optimization. Among the models considered we found the largest
variation in predicted CMB excess ellipticity to be due to choosing
different density models; among the tomography models included if
Fig. 14, with other assumptions as in the reference case and model
2, predicted CMB topography varies between three and eight times
the observed value. Forte et al. (1995) showed, using an older to-
mography model, that it is possible to modify the tomography model
to fit excess ellipticity without significantly degrading the fit to the
seismological data, and it remains to be tested whether this is still
possible for the more recent tomography models considered here.
The discrepancy may, however, also be due to chemical variations at
the base of the mantle; if they occur close to the CMB they would be
largely isostatically compensated and thus not strongly contribute
to the geoid. On the other hand, stresses from overlying mantle flow
could be largely compensated due to these chemical variations, and
the topography at the CMB could be reduced to values compatible
with the observed excess ellipticity (Steinberger & Holme 2006).
Our preferred models have a temperature at the lower boundary of
3500 K, and a TBL thickness of 200 km. 3500 K is within the uncer-
tainties of CMB temperatures inferred from experiments (Boehler
1996), but at the lower end. A possible explanation for this rather
low temperature is again that there is chemically distinct material
at the base of the mantle which is not, or only partly, entrained in
flow in the overlying mantle, and the preferred temperature of about
3500 K does not actually correspond to CMB temperature, but tem-
perature around the top of the chemically distinct material. The in-
ferred high proportion of basal heating would not only include heat
from the core but also from the chemically distinct layer. However,
if dislocation creep is the dominant mechanism in the D

layer, the
same viscosity drop could result from a larger temperature increase
in the D

layer. Steinberger & Holme (2006) extend this geody-
namic model to consider chemical variations within the lowermost
300 km of the mantle. In this case, the optimal TBL thickness
shifts towards somewhat higher values around 300 km. Implications
of the model presented here for the lowermost mantle are further
complicated by the occurrence of a phase boundary recently found
(Murakami et al. 2004). As its Clapeyron slope is currently not
well known, it is, at present, difficult to implement into geodynamic
models.
ACKNOWLEDGMENTS
Funding for this research was provided in part by the Deutsche
Forschungsgemeinschaft, Project Number STE 907/7–1. We thank
Saskia Goes and an anonymous reviewer for extensive and de-
tailed comments and suggestions, and Harro Schmeling and Karen
Niehuus for discussions and further comments. This input helped
to considerably clarify the paper. Figures were prepared using GMT
(Wessel & Smith 1998).
REFERENCES
Akaogi, M. & Ito, E., 1999. Calorimetric study on majorite-perovskite tran-
sition in the system Mg
4
Si
4
O
12
–Mg
3
Al
2
Si
3
O
12
: transition boundaries
with positive pressure-temperature slopes, Phys. Earth planet. Inter., 114,
129–140.
Akaogi, M., Ito, E. & Navrotsky, A., 1989. Olivine—modified spinel—
spinel transitions in the system Mg
2
SiO
4
–Fe
2
SiO
4
: calorimetric mea-
surements, thermochemical calculations, and geophysical application,
J.
geophys. Res., 94, 15671–15 685.
Albers, M. & Christensen, U.R., 2001. Channelling of plume flow beneath
mid-ocean ridges, Earth planet. Sci. Lett., 187, 207–220.
Anderson, D.L. & Hart, R.S., 1978. Q of the Earth, J. geophys. Res., 83,
5869–5882.
Anderson, O.L., Oda, H. & Isaak, D., 1992. A model for the computation of
thermal expansivity at high T and high compression: MgO as an example,
Geophys. Res. Lett., 19, 1987–1990.
Bai, W., Vigny, C., Ricard, Y. & Froidevaux, C., 1992. On the origin of
deviatoric stresses in the lithosphere, J. geophys. Res., 97, 11 729–11 737.
Becker, T.W. & Boschi, L., 2002. A comparison of tomographic and geody-
namic mantle models, Geochem. Geophys. Geosyst., 3, 2001GC000168.
Becker, T.W. & O’Connell, R.J.,2001.Predicting plate velocitieswithmantle
circulation models, Geochem. Geophys. Geosyst., 2, 2001GC000171.
Becker, T.W., Kellogg, J.B., Ekstr¨om, G. & O’Connell, R.J., 2003. Compar-
ison of azimuthal seismic anisotropy from surface waves and finite-strain
from global mantle-circulation models, Geophys. J. Int., 155, 696–714.
Behn, M., Conrad, C. & Silver, P.G., 2004. Detection of upper mantle flow
associated with the African superplume, Earth planet. Sci. Lett., 224,
259–274.
Boehler, R., 1996. Melting temperature of the Earth’s mantle and core:
Earth’s thermal structure, Ann. Rev. Earth Planet. Sci., 24, 15–40.
ˇ
Cadek, O. & Fleitout, L., 2003. Effects of lateral viscosity variations in
the top 300 km on geoid, dynamic topography and lithospheric stresses,
Geophys. J. Int., 152, 566–580.
ˇ
Cadek, O. & Fleitout, L., 2006. Effect of lateral viscosity variations in the
core-mantle boundary region on predictions of the long-wavelength geoid,
Stud. Geophys. Geod., 50, 217–232.
Calderwood, A.R., 1999. Mineral physics constraints on the temperature
and composition of the Earth’s mantle, PhD thesis, University of British
Columbia, Dept. of Earth and Ocean Sciences, Canada.
Cammarano, F., Goes, S., Vacher, P. & Giardini, D., 2003. Inferring upper
mantle temperatures from seismic velocities, Phys. Earth planet. Inter.,
138, 197–222.
Christensen, U.R., 1983. Convection in a variable-viscosity fluid: Newtonian
versus power-law rheology, Earth planet. Sci. Lett., 64, 153–162.
Chopelas, A. & Boehler, R., 1989. Thermal expansion measurements at
very high pressure, systematics, and a case for a chemically homogeneous
mantle, Geophys. Res. Lett., 16, 1347–1350.
Conrad, C.P. & Lithgow-Bertelloni, C., 2002. How mantle slabs drive plate
tectonics, Science, 298, 207–209.
Conrad, C.P. & Lithgow-Bertelloni, C., 2004. The relative importance of
‘slab pull’ and ‘suction’ as forces that drive plate motions, J. geophys.
Res., 109, B10407.
Conrad, C.P. & Lithgow-Bertelloni, C., 2006. Influence of continental roots
and asthenosphere on plate-mantle coupling, Geophys. Res. Lett., 33,
L05312.
DeMets, C., Gordon, R.G., Argus, D.F. & Stein, S., 1990. Current plate
motions, Geophys. J. Int., 101, 425–478.
Duffy, T.S. & Anderson, D.L., 1989. Seismic velocities in mantle minerals
and the mineralogy of the upper mantle, J. geophys. Res., 94, 1895–1912.
Dziewonski, A.M. & Anderson, D.L., 1981. Preliminary Reference Earth
Model, Phys. Earth planet. Inter., 25, 297–356.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
1480 B. Steinberger and A. R. Calderwood
Forte, A.M., Dziewonski, A.M. & Woodward, R.L., 1993. Aspherical struc-
ture of the mantle, tectonic plate motions, nonhydrostatic geoid, and to-
pography of the core-mantle boundary, in Dynamics of the Earth’s Deep
Interior and Earth Rotation,Vol. 72, pp. 135–166, eds Le Mou¨el, J.-L.,
Smylie, D.E. & Herring, T., Geophys. Mon. Ser., AGU, Washington, DC.
Forte, A.M., Mitrovica, J.X. & Woodward, R.L., 1995. Seismic-geodynamic
determination of the origin of excess ellipticity of the core-mantle bound-
ary, Geophys. Res. Lett. 22, 1013–1016.
Goes, S., Cammarano, F. & Hansen, U., 2004. Synthetic seismic signature
of thermal mantle plumes, Earth planet. Sci. Lett., 218, 403–419.
Grand, S.P., 2002. Mantle shear-wave tomography and the fate of subducted
slabs, Philos. Trans. R. Soc. London, Ser. A, 360, 2475–2491.
Gu, Y.J., Dziewonski, A.M., Su, W. & Ekstr¨om, G., 2001. Models of the
mantle shear velocity and discontinuities in the pattern of lateral hetero-
geneities, J. geophys. Res., 106, 11 169–11 199.
Gung, Y. & Romanowicz, B., 2004. Q tomography of the upper mantle us-
ing three-component long-period waveforms. Geophys. J. Int., 157, 813–
830.
Hager, B.H. & O’Connell, R.J., 1979. Kinematic models of large-scale man-
tle flow, J. geophys. Res., 84, 1031–1048.
Hager, B.H. & O’Connell, R.J., 1981. A simple global model of plate dy-
namics and mantle convection, J. geophys. Res., 86, 4843–4867.
Haskell, N.A., 1935. The motion of a fluid under a surface load, 1, Physics,
6, 265–269.
Hofmeister, A.M., 2005. Earth’s heat flux revised and linked to chemistry,
Tectonophysics, 395, 159–177.
Ishii, M. & Tromp, J., 2004. Constraining large-scale mantle heterogeneity
using mantle and inner-core sensitive normal modes, Phys. Earth planet.
Int., 146, 113–124.
Ita, J.J. & Stixrude, L., 1992. Petrology, elasticity, and composition of the
mantle transition zone, J. geophys. Res., 97, 6849–6866.
Iwamori, H., McKenzie, D. & Takahashi, E., 1995. Melt generation by isen-
tropic mantle upwelling, Earth planet. Sci. Lett., 134, 253–266.
Jellinek, A.M., Gonnermann, H.M. & Richards, M.A., 2003. Plume cap-
ture by divergent plate motions: implications for the distribution of
hotspots, geochemistry of mid-ocean ridge basalts, and estimates of the
heat flux at the core-mantle boundary, Earth planet. Sci. Lett., 205, 361–
378.
Kaban, M.K., Schwintzer, P. & Tikhotsky, S.A., 1999. A global isostatic
gravity model of the Earth. Geophys. J. Int., 136, 519–536.
Karato, S., 1993. Importance of anelasticity in the interpretation of seismic
tomography, Geophys. Res. Lett., 20, 1623–1626.
Karato, S., Zhang, S. & Wenk, H.R., 1995. Superplasticity in Earth’s lower
mantle: evidence from seismic anisotropy and rock physics, Science, 270,
458–461.
Kavner, A., 2003. Elasticity and strength of hydrous ringwoodite at high
pressure, Earth planet. Sci. Lett., 214, 645–654.
Kohlstedt, D.L. & G¨otze, C., 1974. Low-stress high-temperature creep in
olivine single crystals, J. geophys. Res., 79, 2045–2051.
Lambeck, K. & Johnston, P., 1998. The viscosity of the mantle: evidence
from analyses of glacial-rebound phenomena, in The Earth’s Mantle, pp.
461–502, ed. Jackson, I., Cambridge University Press, Cambridge.
Lithgow-Bertelloni, C. & Richards, M.A., 1998. The dynamics of Cenozoic
and Mesozoic plate motions, Rev. Geophys., 36, 27–78.
Lithgow-Bertelloni, C. & Silver, P.G., 1998. Dynamic topography, plate
driving forces and the African superswell, Nature, 395, 269–272.
Lithgow-Bertelloni, C. & Guynn, J.H., 2004. Origin of lithospheric stress
field, J. geophys. Res., 109, B01408.
Marquart, G., 2006. Inferring mantle viscosity and s-wave-density conver-
sion factor from new seismic tomography and geoid data, Geophys. J. Int.,
submitted.
Masters, G., Laske, G., Bolton, H. & Dziewonski, A.M., 2000. The relative
behavior of shear velocity, bulk sound speed, and compressional velocity
in the mantle: implications for chemical and thermal structure, in Earth’s
deep interior. Mineral physics and tomography from the atomic to the
global scale,Vol. 117, pp. 63–87, eds Karato, S., Forte, A.M., Liebermann,
R.C., Masters, G. & Stixrude, L., Geophys. Mon. Ser., AGU, Washington,
DC.
Mathews, P.M., Herring, T.A. & Buffett, B.A., 2002. Modeling of nutation
and precession: new nutation series for nonrigid Earth and insights into
the Earth’s interior, J. geophys. Res., 107, doi:10.1029/2001JB000390.
egnin, C. & Romanowicz, B., 2000. The shear velocity structure of the
mantle from the inversion of body, surface, and higher modes waveforms,
Geophys. J. Int., 143, 709–728.
Mittelstaedt, E. & Tackley, P.J., 2006. Plume heat flow is much lower than
CMB heat flow. Earth planet. Sci. Lett., 241, 202–210.
Mitrovica, J.X., 1996. Haskell [1935] revisited, J. geophys. Res., 101, 555–
569.
Mitrovica, J.X. & Forte, A.M., 1997. Radial profile of mantle viscosity:
results from the joint inversion of convection and postglacial rebound
observables, J. geophys. Res., 102, 2751–2769.
Mitrovica, J.X. & Forte, A.M., 2004. A new inference of mantle viscosity
based upon joint inversion of convection and glacial isostatic adjustment
data, Earth planet. Sci. Lett., 225, 177–189.
Murakami, M., Hirose, K., Kawamura, K., Sata, N. & Ohishi, Y., 2004.
Post-perovskite phase transition in MgSiO
3
, Science, 304, 855–858.
Nolet, G., Karato, S.-i. & Montelli, R., 2005. Plume fluxes from seismic
tomography, Geophys. Res. Abs., 7, 04221.
Oganov, A.R., Brodholt, J.P. & Price, G.D., 2001. Ab initio elasticity and
thermal equation of state of MgSiO
3
perovskite, Earth planet. Sci. Lett.,
184, 555–560.
Panasyuk, S.V. & Hager, B.H., 2000. Inversion for mantle viscosity profiles
constrained by dynamic topography and the geoid, and their estimated
errors, Geophys. J. Int., 143, 821–836.
Panasyuk, S.V., Hager, B.H. & Forte, A.M., 1996. Understanding the effect
of compressibility on geoid kernels, Geophys. J. Int., 124, 121–133.
Pari, G. & Peltier, W.R., 2000. Subcontinental mantle dynamics: a further
analysis based on the joint constraints of dynamic surface topography and
free-air gravity. J. geophys. Res., 105, 5635–5662.
Pari, G. & Peltier, W.R., 1995. Heat flow constraint on mantle convection, J.
geophys. Res., 100, 12731–12 751.
Pollack, H.N., Hurter, S.J. & Johnson, J.R., 1993. Heat loss from the
Earth’s interior: analysis of the global data set, Rev. Geophys., 31, 267–
280.
Press, W.H., Flannery, B.P., Teukolski, S.A. & Vetterling, W.T., 1986. Nu-
merical Recipes—The Art of Scientific Computing (FORTRAN version),
Cambridge University Press, Cambridge.
Ranalli, G., 1995. Rheology of the Earth, 2nd edn, Chapman & Hall, London.
Ricard, Y. & Vigny, C., 1989. Mantle dynamics with induced plate tectonics,
J. geophys. Res., 94 17 543–17 559.
Ricard, Y., Fleitout, L. & Froidevaux, C., 1984. Geoid heights and litho-
spheric stresses for a dynamic Earth, Ann. Geophys., 2, 267–286.
Richards, M.A. & Hager, B.H., 1984. Geoid anomalies in a dynamic Earth,
J. geophys. Res., 89, 5987–6002.
Ritsema, J. & van Heijst, H.J., 2000. Seismic imaging of structural hetero-
geneity in Earth’s mantle: evidence for large-scale mantle flow, Sci. Progr.,
83, 243–259.
Schmeling, H., Marquart, G. & Ruedas, T., 2003. Pressure- and temperature-
dependent thermal expansivity and the effect on mantle convection and
surface observables, Geophys. J. Int., 154, 224–229.
Schubert, G., Turcotte, D.L. & Olson, P., 2001. Mantle Convection in the
Earth and Planets, Cambridge University Press, Cambridge.
Smyth, J.R., Holl, C.M., Frost, D.J. & Jacobsen, S.D., 2004. High pressure
crystal chemistry of hydrous ringwoodite and water in the Earth’s interior,
Phys. Earth planet. Inter., 143–144, 271–278.
Stacey, F.D., 1992. Physics of the Earth, 3rd edn, John Wiley & Sons, New
York, NY.
Steinberger, B., 2000. Plumes in a convecting mantle: models and observa-
tions for individual hot-spots, J. geophys. Res., 105, 11 127–11 152.
Steinberger, B. & Holme, R., 2006. Mantle flow models with core-
mantle boundary constraints: evidence for chemical heterogeneities in
the lowermost mantle, J. geophys. Res., submitted (available online at
www.geodynamics.no/steinberger/steinberger
and holme.pdf).
Steinberger, B., Schmeling, H. & Marquart, G., 2001. Large-scale litho-
spheric stress field induced by global mantle circulation, Earth planet.
Sci. Lett., 186, 75–91.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS
Large-scale mantle flow and mineral physics 1481
Su, W., Woodward, R.L. & Dziewonski, A.M., 1994. Degree 12 model of
shear velocity heterogeneity in the mantle, J. geophys. Res., 99, 6945–
6980.
Thoraval, C. & Richards, M.A., 1997. The geoid constraint in global geody-
namics: viscosity structure, mantle heterogeneity models and boundary
conditions, Geophys. J. Int., 131, 1–8.
Trampert, J., Deschamps, F., Resovsky, J. & Yuen, D., 2004. Probabilistic
tomography maps chemical heterogeneities throughout the lower mantle,
Science, 306, 853–856.
van Hunen, J., Zhong, S., Shapiro, N.M. & Ritzwoller, M.H., 2005. New
evidence for dislocation creep from 3-D geodynamic modeling of the
Pacific upper mantle structure, Earth planet. Sci. Lett., 238, 1–270.
White, R.S. & D. McKenzie, 1995. Mantle plumes and flood basalts, J.
geophys. Res., 100, 17543–17 585.
Wang, Z.W., 1999. The melting of Al-bearing perovskite at the core-mantle
boundary. Phys. Earth planet. Inter., 115, 219–228.
Weertman, J. & Weertman, J.R., 1975. High temperature creep of rock and
mantle viscosity. Annu. Rev. Earth Planet. Sci., 3, 293–315.
Wei, M. & Sandwell, D., 2006. Estimates of heat flow from Cenozoic seafloor
using global depth and age data, Tectonophysics, 417, 325–335.
Wessel, P. & Smith, W.H.F., 1998. New, improved version of the Generic
Mapping Tools released, EOS Trans. AGU, 79, 579.
Yamazaki, D. & Karato, S.-I., 2001. Some mineral physics constraints on the
rheology and geothermal structure of Earth’s lower mantle, Am. Mineral.,
86, 385–391.
Yanagisawa, T. & Hamano, Y., 1999. ‘Skewness’ of S wave velocity in the
mantle, Geophys. Res. Lett. 26, 791–794.
Yoshida, M. & Ogawa, M., 2005. Plume heat flow in a numerical model
of mantle convection with moving plates, Earth planet. Sci. Lett. 239,
276–285.
Zerr, A. & Boehler, R., 1994. Constraints on the melting temperature of
the lower mantle from high-pressure experiments on MgO and magne-
siow¨ustite, Nature, 371, 506–508.
Zhang, S. & Christensen, U., 1993. Some effects of lateral viscosity varia-
tions on geoid and surface velocities induced by density anomalies in the
mantle, Geophys. J. Int., 114, 531–547.
C
2006 The Authors, GJI, 167, 1461–1481
Journal compilation
C
2006 RAS