Vous êtes sur la page 1sur 23

Physics of the Earth and Planetary Interiors 129 (2002) 43–65

The influences of composition-sand temperature-dependent


rheology in thermal-chemical convection on
entrainment of the D-layer
B. Schott a,∗ , D.A. Yuen b , A. Braun c
a
Theoretical Geophysics, Utrecht University, P.O. Box 80021, 3508 TA Utrecht, The Netherlands
bDepartment of Geology and Geophysics, Minnesota Supercomputing Institute, University of Minnesota,
1200 Washington Avenue South, Minneapolis, MN 55415-1227, USA
c GeoForschungsZentrum Potsdam, Section 1.2, Division 1, Potsdam, c/o DLR Oberpfaffenhofen, 82234 Wessling, Germany
Received 7 November 2000; received in revised form 3 May 2001; accepted 8 May 2001

Abstract
The entrainment dynamics in the D -layer are influenced by multitudinous factors, such as thermal and compositional
buoyancy, and temperature- and composition-dependent viscosity. Here, we are focusing on the effect of compositionally
dependent viscosity on the mixing dynamics of the D -layer, arising from the less viscous but denser D -material. The marker
method, with one million markers, is used for portraying the fine scale features of the compositional components, D -layer
and lower-mantle. The D -layer has a higher density but a lower viscosity than the ambient lower-mantle, as suggested
by melting point systematics. Results from a two-dimensional finite-difference numerical model including the extended
Boussinesq approximation with dissipation number Di = 0.3, show that a D -layer, less viscous than the ambient mantle
by 1.5 orders of magnitude, cannot efficiently mix with the lower-mantle, even though the buoyancy parameter is as low as
Rρ = 0.6. However, very small-scale schlieren structures of D -layer material are entrained into the lower-mantle. These
small-scale lower-mantle heterogeneities have been imaged with one-dimensional wavelets in order to delineate quantitatively
the multiscale features. They may offer an explanation for small-scale seismic heterogeneity inferred by seismic scattering in
the lower-mantle. © 2002 Elsevier Science B.V. All rights reserved.
Keywords: D -layer; Entrainment; Seismic scattering; Thermochemical convection

1. Introduction Cleary, 1974). Since then D -layer and lower-mantle


heterogeneity was found on all length scales. Models
Almost 30 years ago, lower-mantle seismic hetero- of seismic wave propagation have shown, that these
geneity close to the core mantle boundary (CMB) was PKP precursors can be explained by either small-scale
detected from travel time anomalies due to the seis- CMB topography undulations or by deep lower-mantle
mic phase precursor PKP (Cleary and Haddon, 1972; heterogeneity (e.g. Bataille and Flatte, 1988). How-
ever, recent seismological studies using PKP and
∗ Corresponding author. Tel.: +31-30-253-5076;
PKKP precursors indicate a uniform distribution of
fax: +31-30-253-5030.
E-mail addresses: bert@geo.uu.nl (B. Schott),
small-scale low-amplitude scatterers throughout the
davey@krissy.msi.umn.edu (D.A. Yuen), braun@gfz-postdam.de mantle (Hedlin et al., 1997; Shearer et al., 1998).
(A. Braun). Moreover, the P-wave reflectivity of the D -layer was

0031-9201/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 3 1 - 9 2 0 1 ( 0 1 ) 0 0 2 3 4 - 5
44 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

explained by heterogeneities with scale lengths in has been around for a long time and extension has
the range of 10–100 km (Reasoner and Revenaugh, been made to three-dimensional space (Schmalzl et al.,
1999), and recent investigations by (Cormier, 1999; 1996), where it was shown that due to both geometry
Cormier, 2000) on seismic scattering further under- of the flow and the dynamics, mixing might be less ef-
scored the multiscale nature of the heterogeneities in ficient than in two dimensions. Up to now, mixing pro-
the deep mantle. These observed small-scale seismic cesses in mantle convection have often been restricted
velocity perturbations can be due to small-scale com- to thermal convection, even with variable viscosity
positional heterogeneity resulting from incomplete studies (Ten et al., 1997; van Keken and Ballantine,
convective mantle mixing, because of long mixing 1999). In this study, we will investigate mixing effi-
time-scales for fluids having a large lateral viscosity ciency in the presence of both thermal-chemical con-
contrast (Spence et al., 1987; Manga, 1996). Seismic vection and temperature- and composition-dependent
scattering in the lowermost mantle is also attributed viscosity.
to scattering at partial melt, when the scatterers Our intention is not to cover a large parameter space,
fall into the ultralow velocity zone (Garnero et al., but to draw attention to the complicated phenomenol-
1993; Garnero and Helmberger, 1995; Garnero and ogy coming along with the mixing dynamics of differ-
Helmberger, 1998). Therefore, the question is, where ent mantle materials. We are presenting results from a
do these small-scale compositional heterogeneities two-component numerical model of thermal-chemical
come from. In an attempt to address this question, convection in two-dimensional Cartesian geometry to
we will study with a marker–convection model the monitor the structural evolution of an initially strat-
plume driven mixing behavior of the Earth’s D -layer ified D -layer. An ultrahigh resolution of ‘chemical’
(Dahm, 1934; Dahm, 1936) with the lower-mantle. markers is used to resolve the entrainment structures
It has long been suggested (Davies and Gurnis, arising from the interaction of compositionally denser
1986; Hansen and Yuen, 1988), that the D -layer is D -layer material with lower-mantle convection, to
chemically distinct from the overlying mantle, but monitor the evolution of mixing and finally to inves-
its origin is still debatable. Possible mechanisms tigate the mixing efficiency from chemical buoyancy
seem to be: (1) chemical reactions between iron and composition-dependent rheology of the D -layer.
core and silicate mantle (Knittle and Jeanloz, 1986; The dominant mode of mantle convection to-
Song and Ahrens, 1994), (2) the breakdown of (Mg, day consists of plate-scale flow interacting with a
Fe)SiO3 -perovskite (Saxena et al., 1996; Mao et al., strongly time-dependent mantle circulation. The ap-
1997), (3) the accumulation of (parts of) subducted parent lag between the surface plate configuration and
slabs (Grand et al., 1997; van der Hilst et al., 1997) at the lower-mantle was explained by a large viscos-
the base of the mantle, and possibly any combination ity increase in the lower mantle or by some layered
of these mechanisms, which finally would lead to the mantle convection (Chase, 1979). Slabs subducting
formation of an oxide-rich D -layer (Anderson, 1998; in the upper mantle may not be blocked temporar-
Ruff and Anderson, 1980). ily by endothermic phase transition (Christensen and
Formation of the D -layer by the segregation of Yuen, 1984). This would seriously exert a dramatic
FeO from the ambient lower-mantle has been mod- effect on the flow structures and cause a triggering
eled by double-diffusive thermochemical convection of lower-mantle plumes (Honda et al., 1993), like
(Nauheimer et al., 1996), and by the segregation of the one causing the Pacific superswell (Cazenave and
eclogite from deeply subducted slabs, using a marker Thoraval, 1994; Maruyama, 1994).
technique for tracing the different components, within Another prominent lower-mantle upwelling, the hot
the framework of the Boussinesq approximation plume under Africa (Ritsema et al., 1998), may be
(Christensen and Hofmann, 1994). An additional neg- located there, because of higher mantle temperatures
ative compositional buoyancy equivalent to Rρ = 1.5 due to the continent’s blanketing effect, amplified by
was needed to stabilize the forming D -layer at the its immobility with respect to the hotspot reference
CMB. frame since the opening of the Atlantic about 100 Ma
The use of markers to portray mixing processes ago, as suggested by the regular appearance of the
(Olson et al., 1984; Hoffman and McKenzie, 1985) seafloor magnetization (Muller et al., 1997).
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 45

Because of the manifold interactions between plates function Ψ with either constant or variable viscos-
and global mantle circulation, we have not included ity η and both thermally and compositionally in-
plates or subducting slabs in our model, to avoid un- duced buoyancy forces (Eq. (1));
necessary complications and to focus on the dynamics 2. energy, resulting in an equation for the heat trans-
in the deep mantle. port, and including the effects of viscous and adi-
The entrainment structure modeled, here, may well abatic heating (Eq. (2));
describe those plumes in the deep mantle produced by 3. composition, by applying a marker approach to
various forms of focussing processes (Hansen et al., keep track of the high- and low-density components
1993), which may be responsbile for the megaplumes being advected by the convective velocity field.
like under the Pacific (Cazenave and Thoraval, 1994)
The elliptic PDE for the streamfunction Ψ is solved
or under Africa (Ritsema et al., 1998).
over an equidistant 121 × 121 FD-grid in the limit of
In Section 2 we give a model description of the
Pr → ∞ by the Cholesky decomposition method. The
thermal-chemical convection employed. This then is
extended-Boussinesq approximation (e.g. Christensen
followed by a comparison between the style of plume
and Yuen, 1985) is used. Its non-dimensional form is
convection near the CMB for constant and vari-
able viscosity, in which the composition-dependent  2   2 
∂ ∂2 ∂ Ψ ∂ 2Ψ ∂ 2η ∂ 2Ψ
viscosity will be emphasized. We then apply the − η − +4
∂z 2 ∂x 2 ∂z 2 ∂x 2 ∂x∂z ∂x∂z
wavelet method to analyse the multiscale nature
∂T ∂C
of the thermal-chemical plumes developed in the = −Ra − Rc (1)
lower-mantle. Section 6 gives a summary and conclu- ∂x ∂x
sions drawn from these simulations. where the viscosity η = η(C, T , z) or η = constant.
The equation for the heat transport becomes
2. Thermal-chemical convection—model  
∂T Φ
description + (
v · ∇)T = κ T + Di − vz (T + T0 )
∂t Ra
The problem to be studied is the mixing and entrain- (2)
ment characteristics of a dense D -layer material with
where T denotes the temperature, which is non-dimen-
the ambient lower-mantle. Our lower-mantle model
sionalised by the scaling temperature Tscale = 1000 K.
of thermochemical convection in two-dimensional
Eq. (2) is solved on a grid with four times higher
Cartesian geometry is covering a depth of 2000 km.
spatial resolution (481 × 481 grid points). The surface
The aspect ratio is one, because of the high resolution
temperature T0 = 0.273 is included and the ther-
required to resolve the fine structures presented in the
mal diffusivity κ is constant. v is the velocity field
following sections. We are using the finite-difference
and vz is the vertical velocity component. C is the
program FDCON, developed by Schmeling (Wein-
chemical field representing a binary system consist-
berg and Schmeling, 1992), which combines a
ing of D -layer material (high density) and ambient
finite-difference and marker technique to solve the
lower-mantle (low density) substance. x and z are the
below mentioned partial differential equations (PDE)
horizontal position and the depth.
governing multi-component thermochemical convec-
In Eq. (1) the thermal and the chemical Rayleigh
tion. FDCON was successfully checked in benchmark
numbers are defined as Ra = ρ0 gα T h3 /κη0 and
tests (e.g. Blanckenbach et al., 1989; van Keken
Rc = ρ gh3 /κη0 , respectively, and Di = αgh/cp =
et al., 1997) and applied to several multi-component
0.3 is the dissipation number. The dissipation function
problems (e.g. Schott and Schmeling, 1998; Walzer
Φ is defined by the relation Φ = 2ηėij ėij (van den
and Hendel, 1997). Our approach, based on a
Berg and Yuen, 1997).
finite-difference and marker technique, calls for the
As an equation of state we use ρ(T , C) = ρ0 (1 −
solution of three conservation equations for
αT + C ρ/ρ0 ), where ρ0 is the reference density,
1. mass and momentum, combined together as an el- ρ is the density difference due to the chemical com-
liptic partial differential equation for the stream- position and α is the thermal expansivity. T is the
46 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

temperature difference between the bottom and the top lar C = 0 points within the C = 1 domain, which
surface, h is the distance between bottom and top, re- are purely numerical. This type of interpolation of the
spectively and η0 = 1021 Pa s is the scaling viscosity C-field by counting markers within grid squares works
of the mantle, the acceleration of gravity g points in well as long as
√the grid spacing dx is larger than ap-
the z-direction. proximately 2 2dm, where dm is the average marker
FDCON is using a marker approach with a spacing. The C-field can therefore easily be interpo-
fourth-order Runge–Kutta integration in time com- lated on a 361×361 grid from 1000×1000 irregularly
bined with a predictor-corrector method. The markers but ‘almost uniformly’ distributed markers.
are carrying the values C = 0 or 1, representing the
different materials by their densities and rheologies,
to minimize numerical diffusion. The markers are 3. Boundary and initial conditions
advected with the flow, and for each time-step the
effective composition C is calculated from the sur- The square computational domain has free-slip
rounding markers at each finite difference grid point. boundaries at the top and bottom boundaries, including
The density and rheology is then chosen according to reflecting boundary conditions at the vertical side-
the actual C value (Weinberg and Schmeling, 1992; walls. The thermal boundary conditions are isother-
van Keken et al., 1997). The exact percentage of man- mal top (T = 0) and bottom (T = 1), developing a
tle to D -material is given by the value of the C-field, thermal boundary layer at the CMB over which the
which comes from measuring the relative fraction of temperature increases by ≈1000 K. Heat flux through
the two types of markers within a local grid area. A the side boundaries is set to zero.
pure mantle material is given by C = 0, while C = 1 Initially the model area is isothermal (T = 0) with a
represents an area occupied only by D -material. hot bottom boundary (T = 1) and a thermal perturba-
The C-field, which exists on a grid with spatial tion in the left lower corner. Therefore, cold-looking
resolution dxdz, is interpolated from the irregularly temperatures like T = 0.1 are representing material,
distributed markers by counting the markers within a that is 100 K hotter than the surrounding material and
square of dimensions dxdz centered at each grid point the area, where T ≥ 0.1 will be used to define thermal
(i, j ). When the resolution of the grid, on which the plumes and boundary layers.
C-field is interpolated, becomes too fine, then it can The D -layer material is initially stratified along the
happen, that there are marker-free squares within a bottom boundary, the CMB, with a constant thickness
C = 1 domain, getting the value C = 0 by default. of 200 km (Hansen and Yuen, 1988; Sidorin et al.,
In this case, the interpolation produces a few singu- 1999) (Fig. 1).

Fig. 1. Initial conditions illustrated by the marker distribution with the streamlines overlaid (left panel) and the isothermal temperature
field (right panel). D -markers appear black in the left panel. Ambient lower-mantle markers are not shown.
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 47

Table 1 In Fig. 2, left panel, single markers are visible in the


List of the modelsa compositional plume head (at least in a zoom-in) due
Model η η Rρ Ra Di to a change in marker spacing from the internal fluid
C0 Constant – 0 106 0.3 deformation. In this case, it would not help maters by
C1 Constant – 0.4 106 0.3 increasing the size of the markers. They would simply
C2 Constant – 0.6 106 0.3 overlap each other, producing anything but a black
C3 Constant – 0.8 106 0.3 unstructured area.
V3.1 Variable 40 0.4 106 0.3
V3 Variable 30 0.4 106 0.3
V2 Variable 30 0.6 106 0.3 4.1. Constant viscosity
V1 Variable 30 0.8 106 0.3
a Symbols are explained in the text.
We first show how a constant viscosity plume with-
out any chemical buoyancy, Rρ = 0, would behave as
a reference model with other more complicated plumes
to follow. We also show the development of the mark-
4. Results ers, originating in the D -layer. This starting plume is
thermally driven and is entirely composed of pristine
The runs we have conducted are based on either con- D -material. The material boundary between D -layer
stant viscosity (Section 4.1) or variable viscosity (Sec- and the ambient lower-mantle is nearly parallel to the
tion 4.2), which depends on composition, temperature, T = 0.1 isotherm, in the starting plume scenario, as
and depth. We have varied the buoyancy parameter depicted in Fig. 2. As already stated, the T = 0.1
Rρ = Rc/Ra = ρ/(ρ0 α T ) from no chemical isotherm is used for defining the thermal plume.
buoyancy, Rρ = 0, to strongly chemically buoyant, In Fig. 3, we show the effects from a small amount
Rρ = 0.8. An overview of the models is given in of chemical buoyancy, Rρ = 0.4, due to the denser
Table 1. Here, we want to draw the reader’s attention D -material. The thermal plume (area in which T ≥
to the technical point, that it is very difficult to display 0.1) becomes now larger than the compositional plume
all of the markers, because there are so many of them. filled with the D -material. The compositional buoy-
We cannot see individual markers any longer because ancy decelerates the rising plume and thermal diffu-
of their extreme dense spacing. In Fig. 1, left panel, sion is more effective in removing heat from the plume
the D -markers are shown in black, resulting in a dark due to the longer time of plume evolution.
D -layer, while the ambient lower-mantle markers are This effect becomes more prominent with a larger
not at all shown. chemical buoyancy of Rρ = 0.6 (Fig. 4). In this case,

Fig. 2. In a constant viscosity model without any compositional buoyancy (Rρ = 0) the material boundary between D -layer and
lower-mantle is almost isothermal in a developing plume (Ra = 106 , model C0).
48 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 3. In a constant viscosity model with an increased compositional buoyancy of Rρ = 0.4 of the D -layer material, the thermal plume
fills a larger area than the compositional plume (Ra = 106 , model C1).

the thermal plume speeds away upward from the com- of internal convection inside the D -layer but their
positional plume, giving rise to a situation in which resolution was not as dense as the one used here. This
the thermal plume has a bigger head than that of the type of small-scale secondary convection cells has
compositional plume upon reaching the surface. been suggested for the foot of the D -layer under the
With an additional influx of denser material at Hawaiian hotspot from seismic studies (Russell et al.,
Rρ = 0.8, hill-like structures are formed (Hansen and 1998). These constant viscosity results are shown here
Yuen, 1988, 1989), since the really dense material in order to compare with the results from a more com-
cannot be entrained into the thermal plume any longer. plicated rheology involving a large viscosity contrast
This scenario of little entrainment is displayed in between the D -layer and the lower-mantle.
Fig. 5. The D -layer now thickens, giving rise to the
interesting possibility for internal convection to occur 4.2. Variable viscosity
in the D -layer, which can be clearly discerned inside
the D -layer in the right panel of Fig. 5. Hansen and We now introduce a more realistic rheologi-
Yuen, 1989 also found some evidence for such a type cal model, where the viscosity η(C, T , z) depends

Fig. 4. With a compositional buoyancy of the D -layer material of Rρ = 0.6 the thermal plume is ascending faster than the compositional
plume, leading to a larger thermal plume head than the compositional one (Ra = 106 , model C2).
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 49

Fig. 5. A compositional buoyancy of Rρ = 0.8 of the D -layer material does not allow that material to enter the thermal plume any more.
However, there is little entrainment (Ra = 106 , model C3).

on composition, temperature, and depth, with the to 0.8. Small-scale secondary convection is devel-
denser material having a weaker rheology be- oped ever more with an increase of Rρ , as we can
cause of a lower melting point (Weertman, 1970; see from the sequence of time-frames from Figs. 6–8
Karato, 1997). This viscosity takes the following (Rρ = 0.4) to Fig. 9 (Rρ = 0.6). The latter portrays a
form: flat D -layer scenario due to the localized strength of
chemical buoyancy. The effect of the low-viscosity in
η(C, T , z) the D -layer, caused by the combined efforts of ther-
 
(γ a + λbz)C + (a + bz)(1 − C) mal and compositional dependence of the rheology is
= η0 exp (3) to inhibit mixing in the D -layer by virtue of the large
T + T0
lateral variations in the viscosity field (Manga, 1996).
where a = 1, b = 0.3, and T0 = 0.273, resulting in Therefore, from Fig. 6 even for a modest chemi-
a viscosity contrast of η = η(C = 0, T , z)/η(C = cal buoyancy strength of Rρ = 0.4, very much less
1, T , z) ≈ 30 for γ = λ = 0.3 and η ≈ 40 for D -layer material is entrained into the lower-mantle,
γ = 0.1, λ = 0.3, respectively; with the D -layer as compared to the isoviscous case, shown in
being the weaker material (η(C = 1, T , z)) than the Fig. 3.
ambient lower-mantle (η(C = 0, T , z)). Almost half In the case of a moderately strong chemical buoy-
of the viscosity drop is due to the temperature in- ancy (Rρ = 0.4, Fig. 6), this dense layer can be swept
crease in the hot thermal boundary layer and plumes, aside by the strong lower-mantle flow to form lumps or
and the viscosity is increasing by a factor of 4.5 over hills, which can reach a height of around 500 km above
the depth of the mantle due to increasing pressure. the CMB, when they are hot enough. This resembles
This range of viscosity contrast between lower-mantle the giant plume-like structures in the lower-mantle
and D -material is on the conservative side of the under the Pacific Ocean and Africa (Cadek et al.,
estimates of Yamazaki and Karato, 2001, who sug- 1984).
gest a viscosity drop of 1–3 orders of magnitude in Subsequently, when they have lost sufficient amount
the D -layer. Nearly a quarter of a century ago, it of heat, these hills would collapse. This kind of mo-
was recognized, that a constant Newtonian viscos- tion in the deep lower-mantle, in which the lump is
ity is too much an oversimplification in this strongly always connected to the CMB, produces extremely
time-dependent problem (Jones, 1977). small-scale entrainment of the D -material into
We then increase incrementally the strength of the the lower-mantle, which leads to the production of
chemical buoyancy from Rρ = 0.4–0.6 and finally schlieren structures above the CMB. The expression
50 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 6. D -layer dynamics with variable viscosity visualized by the chemical composition of the two-material system—the lower-mantle and
the D -layer (black markers)—with stream-lines overlaid. A not too high compositional buoyancy of Rρ = 0.4 allows the D -material to
rise up to 500 km above the CMB (t = 346 Ma). However, these D -hills collapse, when having lost sufficient heat, producing small-scale
schlieren structures due to the compositional viscosity contrast of η = 30 (t = 366 Ma, model V3).
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 51

Fig. 7. Thermal evolution of the initially horizontally stratified isothermal lower-mantle—D -layer system with Rρ = 0.4 and η = 30
(model V3).
52 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 8. Evolution of the D -layer with a viscosity contrast of η = 40, Rρ = 0.4 (model V3.1). Again only the markers of the D -material
are shown, not to obfuscate the tiny schlieren structures developing within ≈200 Ma after an initial transient period of D -layer internal
convection.
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 53

Fig. 9. D -layer internal convection is dominant at higher compositional buoyancy of Rρ = 0.6 and a viscosity contrast of η = 30
(model V2). Thermal plumes starting from the D -hills are entraining little D -material into the ambient mantle. Composition is shown
by D -markers in the left, and temperature in the right panels.

‘schlieren’ is the one used in, e.g. schlieren photog- 5. Analysis of the schlieren structure using
raphy, a common technique to visualize flow patterns wavelets
in experimental fluid dynamics (van Dyke, 1982). We
will analyze these schlieren structures with wavelets The structures in thermal-chemical convection have
in Section 5. A wavelet analysis of a whole time se- a distinct multiscale character because of the dif-
ries would lie beyond the scope of this paper, where ferences in the scales of the chemical and thermal
we analyze two representative snapshots, to illustrate forcings and in the thermal and chemical viscosity
typical multiscale features of the evolved structures. gradients. Wavelets represent a recently developed
With increasing value of Rρ in the D -layer of up tool in mathematics, which allows one to analyze
to Rρ = 0.6 (Fig. 9), the dense D -layer becomes multiscale structures in a readily understandable for-
more and more stably stratified with thermal plumes mat (Mallat, 1998). These schlieren or tendril-like
issuing from the top of the hill, but entraining very structures have been analyzed by interpolating a con-
little D -material (Figs. 8–10). This scenario of large tinuous compositional field C either on a 121 × 121
values of Rρ has already been studied by Hansen and grid or on a 361 × 361 grid from 1000 × 1000 mark-
Yuen (1989), Hansen and Yuen (1990) and Tackley ers, thus yielding a spatial resolution of 16.5 km or
(1998b). 5.5 km, respectively. These continuous C-fields are
54 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 10. A high compositional buoyancy of Rρ = 0.8 is causing a flat D -layer (η = 30, model V1). The layer internal convection cells
have aspect ratios between 1.4 and 2 with an average of 1.6. Very little D -material is entrained into the ambient mantle. Composition is
shown by D -markers in the left, and temperature in the right panels.

displayed in Figs. 11 and 12, they have been taken Due to computational efficiency, all profiles were
from the last time-steps in Figs. 6 and 8. We are extended by zero leading to 512 intervals with a spac-
analyzing two different, however typical, schlieren ing of 5.5 km (Fig. 13. This particular spacing allows
structures at two different spatial resolutions in order a minimum resolvable wavelength of 11 km, which
to demonstrate the sensitivity of the analysis to spatial is comparable to the length-scale of heterogeneities,
resolution. as observed by scattered waves in PKIKP/PKP stud-
First, we will focus on the high resolution grid ies (Cormier, 2000). The maximum wavelength is
(Fig. 12). For extracting the relevant scale informa- 5632 km corresponding to twice the extended pro-
tion from the grid, a suitable way is to transform file length. Before transforming the profiles, a linear
one-dimensional spatial profiles, using both Fourier trend was removed for avoiding aliasing effects. The
and wavelet methods. The mixing schlieren structure horizontal profiles are located in depths ranging from
shown in Fig. 12 is now analyzed along the nine hori- 600 km in the lower-mantle (profile X1) down to
zontal profiles X1 to X9. Analysis was done along the 1900 km near the CMB (profile X9). The normalized
horizontal and vertical profiles, however, the results Fourier amplitude spectra are shown in Fig. 14.
are the same for both sets of profiles. Therefore, we Obviously, the spectral energy is transferred from
are focusing on the horizontal profiles only. the short wavelengths in profiles crossing the top of
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 55

Fig. 11. Chemical field of the last time-step shown in Fig. 6 interpolated on a 121 × 121 grid, yielding a spatial resolution of 16.5 km.
The locations of the examined horizontal profiles across the composital schlieren structure are indicated by lines.

the mixing structure (profiles X2 and X3) to longer transformation cannot identify precisely the source
wavelengths near the CMB (profile X9). The shortest of the energy, it can only determine the sum of the
wavelength is associated with the entrainment of the power, while losing the phase information needed for
D -material. locating the position of the heterogeneities. Addition-
For multiscale features, such as shown in the en- ally, the advantage of a wavelet analysis as compared
trainment process, a wavelet analysis is a more suit- to a windowed Fourier transform comes from its be-
able tool for analyzing the spatial profiles. Fourier ing scale-independent, because wavelets with short
56 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 12. Chemical field of the last time-step shown in Fig. 8 interpolated on a 361 × 361 grid, yielding a spatial resolution of 5.5 km. The
locations of the examined profiles, X1 to X9, across the composital schlieren structure are shown by dashed lines.

windows are used for short wavelengths and long wavelet ψx by shifting and scaling. The shifting pa-
windows for long wavelengths. Hence, there are no rameter χ is responsible for the location, and s is the
aliasing effects due to high and low frequency contri- scaling parameter, stretching (s > 1) or compressing
butions. We will take again one-dimensional profiles (s < 1) the mother wavelet. We have chosen a com-
and determine the dominant modes of variability plex Morlet wavelet (Morlet et al., 1982) for the anal-
and how these modes would change along profile ysis to minimize the space–wavenumber uncertainty
(Fig. 13). This technique has been used by Simons of χ k ≥ 1/4π . Here, the scaling parameter s can
and Hager (1997) to examine post-glacial rebound. be regarded as a wavelength, although for most other
We have used the continuous wavelet transform wavelets, s has a more general meaning and is called
(CWT) (Daubechies, 1992) and a one-dimensional ‘scale’. This kind of Morlet wavelet has been used
Morlet wavelet (Goupillaud et al., 1984) for analyz- recently in analyzing time-series in mantle convec-
ing the 9 profiles shown in Fig. 11 and the 18 profiles tion (Vecsey and Matyska, 2001). A complex Morlet
shown in Fig. 12, respectively. Generally,
 a CWT wavelet is a plane wave modulated by a Gaussian
can be written like Wψ (s, χ ) = w(x)ψs,χ (x) dx, function and takes the form
√ ψs,χ (x) represents the wavelet ψs,χ (x) =
where
ψs,χ (x) = π −1/4 (sl)−1/2 e−i2π(1/s)(x−χ)
(1/ s)ψ((x − χ )/s) (s > 0, χ ∈ R). A set of
wavelets ψs,χ (x) is created from a so-called mother ×e−1/2((x−χ)/sl)
2
(4)
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 57

Fig. 13. Horizontal profiles of the chemical field. Their location is shown in Fig. 12.
58 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 14. Fourier amplitude spectra of the nine horizontal profiles, for location see Fig. 12. Each spectrum is normalized to its maximum.
The shaded lines indicate how the energy is moving from short to longer wavelength. The profiles starting at the top of the mixing structure
in 600 km depth (top) down to the CMB (bottom).


Fig. 15. The nine selected horizontal profiles of Fig. 13 were analyzed using a continuous wavelet transform with a Morlet wavelet of length
six. Each of the nine panels is divided into four diagrams, showing in the upper left—the wavelet-scalogram; upper right—the normalized
Fourier amplitude spectrum; lower left—the analyzed profile; lower right—the real part of the Morlet wavelet. The wavelet-scalogram
shows the color-coded space–wavelength distribution of the analyzed profile. Red colors indicate high-amplitudes, blue colors indicate
low-amplitudes, values below 15% of the maximum are left white. The color-coded plots allow to identify both the wavelength of the
individual signal as well as its location in the spatial profile, since the spatial and wavelength axes of the original profile, the Fourier
spectrum and the wavelet transform are identical, respectively. The upper left panel is representing the wavelet transform of a single
schlieren structure of thickness 5.5 km, here, the advantages of the wavelet transform become obvious, since the Fourier transform is not
able to detect neither the wavelength nor the location of the signal. The upper three panels crossing the top of the schlieren structure
show wavelengths from 11 to 40 km on the one hand, and from 200 to 400 km on the other. The middle three panels show wavelengths
ranging from 100 to 700 km. Finally, the lower three panels indicate wavelengths around 1000 km representing the thickness of the base
of the evolving D -structure. Thus, we identify three different levels of schlieren thickness, 6–20 km at the top, 50–350 km in the middle
and round 500 km at the base of the model.
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 59

Fig. 15.

An additional parameter l was introduced to allow for uncertainty principle, the greater the number of oscil-
an arbitrary shift of the wavelet transform resolution in lations, the lower would be the spatial resolution and
favor of spatial or wavenumber resolution for a fixed vice versa. The sampling in space and wavenumber is
scale s without changing the central wavenumber. We then determined by
have determined l, the length of the wavelet, by the √ √
spatial and wavenumber resolution, thus, l represents sl 2 s 2
χ = √ , k = , = (5)
the number of oscillations of the wavelet. From the 2 4π sl s 4π l
60 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Fig. 16. The nine selected horizontal profiles of Fig. 11 were analyzed using a continuous wavelet transform with a Morlet wavelet of
length six. Grouping of the panels is analog to that in Fig. 15, where the panels are explained in the figure caption. The upper left panel
is representing the wavelet transform of a single schlieren structure of thickness 16.5 km. The upper three panels crossing the top of the
schlieren structure show a maximum amplitude for wavelengths around 100 km, with a significant contribution for wavelengths down to
some 10 km. The middle three panels show a maximum amplitude for wavelengths around 1000 km, with a significant contribution for
wavelengths down to 100 km. Finally, the lower three panels indicate wavelengths around 1000 km representing the thickness of the base
of the evolving D -structure. The lower right panel indicating an amplitude shift towards shorter wavelength. Thus, we identify three
different levels of schlieren thickness, 10–50 km at the top, 50–500 km in the middle and round 500 km at the base of the model.
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 61

By inspecting at the third equation in Eq. (5), we can the small-scale schlieren at the top of the mixing
see that a logarithmic equidistant sampling in χ is ap- structure may explain the existence of small-scale
propriate. For maintaining a balance between the spa- scatterers, found in seismic wave inversions (Frey-
tial and the wavenumber resolution, a length of l = bourger et al., 1999; Cormier, 2000).
6 was chosen. A linear fit was again removed from
the profiles and each profile was then operated on by
the Morlet wavelet. The numerical implementation of 6. Conclusions and discussion
the CWT is a discrete convolution between the spa-
tial profile and the discretized wavelet at all scales. The purpose of our work on compositionally de-
For interpretation, the wavelet transform was interpo- pendent rheology in thermal-chemical convection
lated in the space–wavenumber domain and the am- is to draw attention to the complicated D -layer
plitudes were color-coded. Figs. 15 and 16 present the entrainment structures most likely existing in the
wavelet-scalograms of the nine respective horizontal lower-mantle, which have the potential of explaining
profiles (Figs. 11 and 12) together with the spatial pro- the observed seismic scattering. Here, we are not
file, its Fourier spectrum and the real part of the Morlet concerned with subduction and the overall mantle
wavelet used. The L2-normalized wavelet-scalogram mixing processed (e.g. Gurnis, 1986a; Olson et al.,
in the upper left box of each panel of Figs. 15 and 16 1984; Christensen and Hofmann, 1994).
reveal clearly the high-amplitudes in red colors and The results of our numerical thermal-compositional
the low-amplitudes in blue colors. A cut-off at 15% of convection model show, that a low viscous D -layer
the maximum is introduced for suppressing the lowest with a rather small density increased of 5% may
amplitudes in order to enhance the clarity. not be entrained into thermal plumes starting from
Three different levels of tendril wavelength/thickness the D -layer. A similar behavior has been found for
can be identified in both schlieren structures multi-layer Rayleigh–Taylor instabilities occurring in
(Figs. 15 and 16): the upper three panels contain magmatic diapirs (Cruden et al., 1995). However, the
high-amplitudes for wavelengths between 11 and partially entrained D -layer material can form distinct
40 km at 5.5 km resolution (Fig. 15, 361 × 361 grid, lumps, which do not need to have a hill-like shape
corresponding profiles X1 to X3 in Fig. 12) and with a smooth surface (Tackley, 1998a), but can have a
wavelengths between some 10 and 100 km at 16.5 km complex tree- or polyp-like structure. Similar complex
resolution (Fig. 16, 121 × 121 grid, corresponding structures have been modeled by Hansen and Yuen
profiles in Fig. 11). Panels in the middle indicate (2000), however, their model could not resolve the
wavelength between 100 and 700 km at 5.5 km reso- small-scale structure of branches or tentacles reaching
lution (Fig. 15, 361 × 361 grid, corresponding profiles out into the lower-mantle (Fig. 12). The heterogeneity
X4 to X6 in Fig. 12) and wavelengths between 100 is developing the shape of a large ‘blob’ connected to
and 1000 km at 16.5 km resolution (Fig. 16, 121 × 121 long thin tendrils, which lengthen exponentially with
grid, corresponding profiles in Fig. 11). The lower time, while the ‘blob’ deforms much slower (Gurnis,
three panels show highest values at wavelengths of 1986b). Similar ‘blobs’ may, therefore, explain some
about 1000 km on both grids (Figs. 15 and 16). aspects of heterogeneity in the Earth’s mantle (Becker
Thus, the corresponding schlieren thickness is et al., 1999). For these tiny schlieren structures we
5–20 km at the top of the mixing structure, 50–500 km find scaling diameters in the order of a few to 10 km,
in the middle, and 500 km at the CMB. This result which is too small to be resolved in numerical mod-
remains the same for both schlieren structures and els of double-diffusive thermal-chemical convection
holds for both spatial resolutions examined. Such en- (Hansen and Yuen, 2000). The double-diffusive con-
trainment pattern typically evolve in thermochemical vection approach also inhibits the development of
convection, when a large viscosity difference between small-scale mixing (Montague and Kellogg, 2000).
the fluids is included. It also confirms the usefulness Schlieren structures of diameters down to about
of the wavelet analysis as a tool for examining and 5 km are robust features in our 2000 km × 2000 km
comparing the characteristics of multiscale structures large model and subsequent analysis, when 1000×
at different spatial resolution. We note especially that 1000 markers are used, because a 5 km thick ‘tentacle’
62 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

is defined by about three markers along its diame- Models of mantle mixing should principally be stud-
ter. Additional schlieren structures with even smaller ied in three dimensions (Ferrachat and Ricard, 1998),
diameters could be made visible by an increasing because it is difficult to scale the mixing behavior from
number of markers, while already detected structures 2D to 3D, and mixing in 2D is more efficient than
would become smoother, but would not disappear. in 3D (Schmalzl et al., 1996). However, the need of
We suggest, that these tiny schlieren structures can very high spatial resolution and many markers to track
explain the recently proposed uniform (Hedlin et al., the materials with high accuracy over long time spans
1997) or slightly depth dependent (Shearer et al., is not easily achievable in 3D today. Computationally
1998; Cormier, 2000) distribution of small-scale faster 2D models of thermal-chemical convection re-
low-amplitude scatterers throughout the mantle with main still a versatile tool to study basic mixing prop-
typical scale-lengths around 10 km. Due to thermal erties.
diffusion pronounced small-scale thermal perturba- For the choice Di = 0 (incompressible convec-
tions cannot survive on the long time-scale our mod- tion) an isothermal mantle is the only reasonable ini-
els are covering, but seismic velocity perturbations tial condition, that is gravitationally stable. For the
on length-scales of 200–1000 km may be of thermal choice Di > 0 (‘compressible’ convection) all initial
origin (Fig. 7). conditions between an isothermal mantle and an adi-
The observed ‘non-decaying’ power spectrum of abatic temperature gradient are gravitationally stable.
scattered seismic energy attributed to heterogeneity While the choice of an adiabatic temperature gradi-
significantly departs from that one expected from ther- ent is essential from the viewpoint of thermodynam-
mal variations in a convecting mantle. The entrain- ics, we have chosen the isothermal initial condition to
ment of small-scale (10–100 km) schlieren structures make the different results easier to compare. We are
is a likely explanation for such a ‘non-decaying’ power therefore overestimating the thermal buoyancy forces
spectrum (Cormier, 2000). for Di > 0. However, preliminary results for starting
A maximum of about 40% of the initially strati- with an adiabatic temperature gradient show, that even
fied D -layer material can be ‘mixed’ into the am- very little additional compositional buoyancy of the
bient lower-mantle within a model time of 475 Ma D -layer, as low as Rρ ≈ 0.2, are enough for stabiliz-
(Fig. 6), if we are counting the C = 1 markers (rep- ing the D -layer at the CMB. This is in full agreement
resenting the D -material) having a depth shallower with the results of Tackley (1998b).
than 500 km above the CMB. However, most of the Values of Rρ greater than one are typically needed
D -material is forming into large lumps, which are all to stabilize the D -layer at the CMB when the Boussi-
the time staying close to the CMB. We, therefore, con- nesq approximation is used together with constant ma-
clude, that these lumps can still serve today as a dis- terial properties, such as the thermal expansivity and
tinct chemical reservoir in the Earth’s mantle, similar viscosity (Christensen and Hofmann, 1994; Davaille,
to the deep-mantle blobs (Becker et al., 1999). 1999; Montague and Kellogg, 2000). The Boussinesq
While the small-scale schlieren cause seismic scat- approximation is, therefore, not appropriate at all for
tering (Cormier, 2000), the bulk of the D -layer shows asessing the stability of the D -layer.
a strong longer–wavelength topographical undulation, We have not considered the additional stirring of
which is reflected by the short-length scale horizontal the D -layer by other processes such as by subduct-
variations in the topography, due to a P-wave reflec- ing slabs, because recent numerical models show, that
tor attributed to the top portion of the D -layer. Such cold dense downwellings sinking into the D -layer
variations of the depth of the D -reflector between 150 do not necessarily trigger instability (thermochemical
and 200 km have been found beneath the southwest- plumes) in their surrounding, by either sweeping the
ern Pacific (Yamada and Nakanishi, 1998) and north- D -layer aside or by increasing the temperature gra-
ern Siberia (Freybourger et al., 1999). Under northern dient, but may lead to the formation of highly compli-
Siberia PdP travel times have also been explained with cated structures of layered thermochemical convection
a 7 km thick 3% low velocity lamella lying approx- (Hansen, 2000). Moreover, subducting slabs may not
imately 300 km above the CMB (Freybourger et al., reach the CMB, but hover in the mid-lower-mantle,
1999). due to the decreasing thermal expansivity with in-
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 63

creasing pressure (Chopelas and Boehler, 1992; Cazenave, A., Thoraval, C., 1994. Mantle dynamics constrained
Kesson et al., 1998; Kaneshima and Helffrich, 1999). by degree 6 surface topography, seismic tomography and geoid:
However, this picture may change from slab to slab inference on the origin of the South Pacific superswell. Earth
Planet. Sci. Lett. 122 (1/2), 207–219.
(Cizkova et al., 1998) because of the rheological de- Chase, C.G., 1979. Subduction, the geoid, and lower mantle
pendence on the subducting velocity (Karato et al., convection. Nature 282, 464–468.
2001). Such additional complications are beyond this Chopelas, A., Boehler, R., 1992. Thermal expansivity of the lower
study, where we are focusing on the style of entrain- mantle. Geophys. Res. Lett. 19 (19), 1983–1986.
ment of D -material into the lower-mantle. Once Christensen, U.R., Hofmann, A.W., 1994. Segregation of subducted
the D -material is entrained by the plumes, it will oceanic-crust in the convective mantle. J. Geophys. Res. Solid
Earth 99 (B10), 19867–19884.
certainly be further mixed into the mantle by fluid
Christensen, U.R., Yuen, D.A., 1984. The interaction of a
dynamical processes stirring the mantle. subducting lithospheric slab with a chemical or phase boundary.
J. Geophys. Res. 89, 4389–4402.
Christensen, U.R., Yuen, D.A., 1985. Layered convection induced
Acknowledgements by phase transitions. J. Geophys. Res. 90, 10291–10300.
Cizkova, H., Cadek, O., Slancova, A., 1998. Regional correlation
analysis between seismic heterogeneity in the lower mantle and
We thank discussions with U. Christensen, U. subduction in the last 180 million years: implications for mantle
Hansen, S. Karato, H. Paulssen, P. Tackley, W. Wang, dynamics and rheology. Pure Appl. Geophys. 151 (2–4), 527–
and D. Yamazaki, and particularly H. Schmeling and 537.
an anonymous reviewer for their excellent reviews, Cleary, J.R., 1974. The D -region. Phys. Earth Planet. Inter. 9,
13–27.
that helped us improving the manuscript. Special
Cleary, J.R., Haddon, R.A.W., 1972. Seismic wave scattering near
thanks to H. Schmeling for kindly providing his code the core-mantle boundary: a new interpretation of precursors
FDCON, and to Swedish NFR for supporting Bertram to PKP. Nature 240, 549–551.
Schott with a Post-Doc stipend, which made this re- Cormier, V.F., 1999. Anisotropy of heterogeneity scale lengths in
search possible. Support for this research also comes the lower mantle from PKIKP precursors. Geophys. J. Int.
from the geophysics program of NSF. All calculations 136 (2), 373–384.
Cormier, V.F., 2000. D as a transition in the heterogeneity
have been done on the SGI-Origin2000 computers of
spectrum of the lowermost mantle. J. Geophys. Res. 105 (B7),
the Minnesota Supercomputing Institute (MSI). 16193–16205.
Cruden, A.R., Koji, H., Schmeling, H., 1995. Diapiric basal
entrainment of mafic into felsic magma. Earth Planet. Sci.
References Lett. 131 (3/4), 321–340.
Dahm, C.G., 1934. A study of dilational wave velocity in the
Anderson, D.L., 1998. The EDGES of the mantle. In: Gurnis, Earth, as a function of depth, based on a comparison of the P,
M., Wysession, M.E., Knittle, E., Buffett, B.A. (Eds.), P and PcP phases. Ph.D. dissertation, St. Louis Univesity, St.
The Core-Mantle Boundary Region: Geodynamics, Vol. 28. Louis.
American Geophysical Union, pp. 255–271. Dahm, C.G., 1936. Velocity of P and S waves. Bull. Seism. Soc.
Bataille, K., Flatte, S.M., 1988. Inhomogeneities near the Am. 26, 157–171.
core-mantle boundary inferred from short-period scattered PKP Daubechies, I., 1992. Ten lectures on wavelets. In: Proceedings
waves recorded at the global digital seismograph network. J. of the CBMS–NSF Regional Conference Series in Applied
Geophys. Res. 93, 15057–15064. Mathematics, Vol. 61. Philadelphia, SIAM Press, Philadelphia,
Becker, T.W., Kellogg, L.H., O’Connell, R.J., 1999. Thermal p. 357.
contraints on the survival of primitive blobs in the lower mantle. Davaille, A., 1999. Simultaneous generation of hotspots and
Earth Planet. Sci. Lett. 171, 351–365. superswells by convection in a heterogeneous planetary mantle.
Blanckenbach, B., Busse, F., Christensen, U., Cserepes, L., Gunkel, Nature 402, 756–760.
D., Hansen, U., Harder, H., Jarvis, G., Koch, M., Marquart, Davies, G.F., Gurnis, M., 1986. Interaction of mantle dregs with
G., Moore, D., Olson, P., Schmeling, H., Schnaubelt, T., 1989. convection: lateral heterogeneity at the core-mantle boundary.
A benchmark comparison for mantle convection codes. J. Geophys. Res. Lett. 13, 1517–1520.
Geophys. Int. 98 (1), 23–38. Ferrachat, S., Ricard, Y., 1998. Regular vs. chaotic mantle mixing.
Cadek, O., Yuen, D.A., Steinbach, V., Chopelas, A., Matyska, C., Earth Planet. Sci. Lett. 155 (1/2), 75–86.
1984. Lower mantle thermal structure deduced from seismic Freybourger, M., Krüger, F., Achauer, U., 1999. A 22◦ long seismic
tomography, mineral physics and numerical modeling. Earth profile for the study of the top of D . Geophys. Res. Lett.
Planet. Sci. Lett. 121 (3/4), 385–402. 26 (22), 3409–3412.
64 B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65

Garnero, E.J., Grand, S.P., Helmberger, D.V., 1993. Low P-wave Karato, S., 1997. Phase transformation and rheological properties
velocity at the base of the mantle. Geophys. Res. Lett. 20 (17), of mantle minerals. In: Crossley, D., Soward, A.M. (Eds.),
1843–1846. Earth’s Deep Interior. Gordon and Breach, New York.
Garnero, E.J., Helmberger, D.V., 1995. A very slow basal layer Karato, S., Riedel, M.R., Yuen, D.A., 2001. Rheological structure
underlying large-scale low-velocity anomalies in the lower and deformation of subducted slabs in the mantle transition
mantle beneath the Pacific-evidence from core phases. Phys. zone: implications for mantle circulation and deep earthquakes.
Earth Planet. Inter. 91 (1–3), 161–176. Phys. Earth Planet. Inter., in press.
Garnero, E.J., Helmberger, D.V., 1998. Further structural Kesson, S.E., Gerald, J.D.F., Shelley, J.M., 1998. Mineralogy and
constraints and uncertainties of a thin laterally varying dynamics of a pyrolite lower mantle. Nature 393, 252–254.
ultralow-velocity layer at the base of the mantle. J. Geophys. Knittle, E., Jeanloz, R., 1986. High-pressure metallization of FeO
Res. 103 (B6), 12495–12509. and implications for the Earth’s core. Geophys. Res. Lett. 13,
Goupillaud, P., Grossmann, A., Morlet, J., 1984. Cycle-octave and 1541–1544.
related transforms in seismic signal analysis. Geoexploration Mallat, S., 1998. A Wavelet Tour of Signal Processing. Academic
23, 85–102. Press, Boston.
Grand, S.P., van der Hilst, R.D., Widiyantoro, S., 1997. High Manga, M., 1996. Mixing of heterogeneities in the mantle: effect
resolution global tomography: a snapshot of convection in the of viscosity differences. Geophys. Res. Lett. 23, 403–406.
Earth. Geol. Soc. Am. Today 7, 1–7. Mao, H.K., Shen, G.Y., Hemley, R.J., 1997. Multivariable
Gurnis, M., 1986a. The effects of chemical density differences dependence of Fe–Mg partitioning in the lower mantle. Science
on convective mixing in the Earth’s mantle. J. Geophys. Res. 278 (5346), 2098–2100.
91 (B11), 11407–11419. Maruyama, S., 1994. Plume tectonics. J. Geol. Soc. Jpn. 100 (1),
Gurnis, M., 1986b. Stirring and mixing in the mantle by plate-scale 25–49.
flow: large persistent blobs and long tendrils coexist. Geophys. Montague, N.L., Kellogg, L.H., 2000. Numerical models of a
Res. Lett. 13 (13), 1474–1477. dense layer at the base of the mantle and implications for the
Hansen, U., 2000. Evolution of structures by double-diffusive geodynamics of D . J. Geophys. Res. 105 (B5), 11101–11114.
convection at the core-mantle boundary. EOS Trans. AGU Morlet, J., Arens, G., Fourgeau, E., Giard, D., 1982. Wave
81 (48), F907. propagation and sampling theory. Part II. Sampling theory and
Hansen, U., Yuen, D.A., 1988. Numerical simulations of complex waves. J. Geophys. 47 (2), 222–236.
thermal-chemical instabilities and lateral heterogeneities at the Muller, R.D., Roest, W.R., Royer, J., Gahagan, L.M., Sclater, J.G.,
core-mantle boundary. Nature 334, 237–240. 1997. Digital isochrons of the world’s ocean floor. J. Geophys.
Hansen, U., Yuen, D.A., 1989. Dynamical influences from Res. 102 (B2), 3211–3214.
thermal-chemical instabilities at the core-mantle boundary. Nauheimer, G., Fradkov, A.S., Neugebauer, H.J., 1996. Mantle
Geophys. Res. Lett. 16 (7), 629–632. convection behavior with segregation in the core-mantle
Hansen, U., Yuen, D.A., 1990. Nonlinear physics of boundary. Geophys. Res. Lett. 23 (16), 2061–2064.
double-diffusive convection in geological systems. Earth Sci. Olson, P.L., Yuen, D.A., Balsiger, D.S., 1984. Mixing of passive
Rev. 29 (1–4), 385–399. heterogenieties by mantle convection. J. Geophys. Res. 89,
Hansen, U., Yuen, D.A., 2000. Extended-Boussinesq thermal- 425–436.
chemical convection with moving heat sources and variable Reasoner, C., Revenaugh, J., 1999. Short-period P wave constraints
viscosity. Earth Planet. Sci. Lett. 176 (3/4), 401–411. on D reflectivity. J. Geophys. Res. 104 (B1), 955–961.
Hansen, U., Yuen, D.A., Kroening, S.E., Larsen, T.B., Ritsema, J., Ni, S., Helmberger, D.V., Cortwell, H.P., 1998.
1993. Dynamical consequences of depth-dependent thermal Evidence for strong shear velocity reduction and velocity
expansivity and viscosity on mantle circulations and thermal gradients in the lower mantle beneath Africa. Geophys. Res.
structure. Phys. Earth Planet. Inter. 77 (3/4), 205–223. Lett. 25 (23), 4245–4248.
Hedlin, M., Shearer, P., Earle, P., 1997. Seismic evidence for Ruff, L.J., Anderson, D.L., 1980. Core formation, evolution, and
small-scale heterogeneity throughout the earth’s mantle. Nature convection: a geophysical model. Phys. Earth Planet. Inter. 21,
387 (6629), 145–150. 181–202.
Hoffman, N.R.M., McKenzie, D.P., 1985. The destruction of Russell, S.A., Lay, T., Garnero, E.J., 1998. Seismic evidence for
geochemical heterogeneities by differential fluid motions during small-scale dynamics in the lowermost mantle at the root of
convection. Geophys. J. R. Astron. Soc. 82, 163–206. the Hawaiian hotspot. Nature 396 (6708), 255–258.
Honda, S., Balachandar, S., Yuen, D.A., Reuteler, D., 1993. Saxena, S.K., Dubrovinsky, L.S., Lazor, P., Cerenius, Y., Haggkvist,
Three-dimensional mantle dynamics with an endothermic phase P., Hanfland, M., Hu, J., 1996. Stability of perovskite (MgSiO3 )
transition. Geophys. Res. Lett. 20 (3), 221–224. in the Earth’s mantle. Science 274, 1357–1359.
Jones, G.M., 1977. Thermal interaction of the core and the mantle Schmalzl, J., Houseman, G.A., Hansen, U., 1996. Mixing in
and long-term behavior of the geomagnetic field. J. Geophys. vigorous, time-dependent three-dimensional convection and
Res. 82 (11), 1703–1709. application to Earth’s mantle. J. Geophys. Res. 101, 21847–
Kaneshima, S., Helffrich, G., 1999. Dipping low-velocity layer in 21858.
the mid-lower mantle: evidence for geochemical heterogeneity. Schott, B., Schmeling, H., 1998. Delamination and detachment of
Science 283 (5409), 1888–1891. a lithospheric root. Tectonophysics 296 (3/4), 225–247.
B. Schott et al. / Physics of the Earth and Planetary Interiors 129 (2002) 43–65 65

Shearer, P.M., Hedlin, M.A.H., Earle, P., 1998. PKP and PKKP in lubricating mantle flow. Earth Planet. Sci. Lett. 151 (1/2),
precursor observations: implications for small-scale structure 33–42.
of the deep mantle and core. In: Gurnis, M., Wysession, M.E., van der Hilst, R., Widiyantaro, S., Engdahl, E.R., 1997. Evidence
Knittle, E., Buffett, B.A. (Eds.), The Core-Mantle Boundary for deep mantle circulation from global tomography. Nature
Region: Geodynamics, Vol. 28. American Geophysical Union, 386, 578–584.
pp. 37–55. van Dyke, M., 1982. An Album of Fluid Motion. Parabolic Press,
Sidorin, I., Gurnis, M., Helmberger, D.V., 1999. Evidence for Stanford, CA.
a ubiquitous seismic discontinuity at the base of the mantle. van Keken, P.E., Ballantine, C.J., 1999. Dynamical models of
Science 286, 1326–1331. mantle volatile evolution and the role of phase transitions and
Simons, M., Hager, B.H., 1997. Localization of the gravity field temperature-dependent rheology. J. Geophys. Res. 104 (B4),
and the signature of glacial rebound. Nature 390, 500– 7137–7152.
504. van Keken, P.E., King, S.D., Schmeling, H., Christensen, U.R.,
Song, X., Ahrens, T.J., 1994. Pressure-temperature range of Neumeister, D., Doin, M., 1997. A comparison of methods for
reactions between liquid iron in the outer core and mantle the modeling of thermochemical convection. J. Geophys. Res.
silicates. Geophys. Res. Lett. 21, 153–156. 102 (B10), 22477–22495.
Spence, D.A., Sharp, P.W., Turcotte, D.L., 1987. Buoyancy-driven Vecsey, L., Matyska, C., 2001. Wavelet spectra and chaos in
crack propagation: a mechanism for magma migration. J. Fluid thermal convection modelling. Geophys. Res. Lett. 28 (2),
Mech. 174, 135–153. 395–398.
Tackley, P.J., 1998a. Self-consistent generation of tectonic plates Walzer, U., Hendel, R., 1997. Tectonic episodicity and convective
in three-dimensional mantle convection. Earth Planet. Sci. Lett. feedback mechanism. Phys. Earth Planet. Inter. 100 (1–4),
157, 9–22. 167–188.
Tackley, P.J., 1998b. Three-dimensional simulations of mantle Weertman, J., 1970. The creep strength of the Earth’s mantle.
convection with a thermochemical CMB boundary layer: D . Rev. Geophys. Space Phys. 8, 145–168.
In: Gurnis, M., Wysession, M.E., Knittle, E., Buffett, B.A. Weinberg, R.F., Schmeling, H., 1992. Polydiapirs—multi-
(Eds.), The Core-Mantle Boundary Region: Geodynamics, Vol. wavelength gravity structures. J. Struct. Geol. 14 (4), 425–436.
28. American Geophysical Union, pp. 231–253. Yamada, A., Nakanishi, I., 1998. Short-wavelength lateral variation
Ten, A., Yuen, D.A., Podladchikov, Yu., Larsen, T.B., Pachepsky, of a D P-wave reflector beneath the southwestern Pacific.
E., Malvesky, A.V., 1997. Fractal features in mixing of Geophys. Res. Lett. 25 (24), 4545–4548.
non-Newtonian and Newtonian mantle convection. Earth Yamazaki, D., Karato, S., 2001. Some mineral physics constraints
Planet. Sci. Lett. 146, 401–414. on the rheology and geothermal structure of Earth’s lower
van den Berg, A.P., Yuen, D.A., 1997. The role of shear heating mantle. Am. Mineral. 86, 385–391.

Vous aimerez peut-être aussi