Earth and Planetary Science Letters 312 (2011) 348–359 Contents lists available at SciVerse ScienceDirect Earth and Planetary Science Letters journal homepage: www.elsevier.com/locate/epsl Bent-shaped plumes and horizontal channel flow beneath the 660 km discontinuity Nicola Tosi a, b,⁎, David A. Yuen c, d a Department of Planetary Physics, German Aerospace Center (DLR), Berlin, Germany Department of Planetary Geodesy, Technical University Berlin, Germany c Department of Earth Sciences, University of Minnesota, Minneapolis, USA d Minnesota Supercomputing Institute, University of Minnesota, Minneapolis, USA b a r t i c l e i n f o Article history: Received 1 August 2011 Received in revised form 23 September 2011 Accepted 9 October 2011 Available online 20 November 2011 Editor: Y. Ricard Keywords: mantle plumes phase transitions temperature dependent viscosity thermal expansivity lattice thermal conductivity a b s t r a c t Recent high-resolution seismic imaging of the transition zone topography beneath the Hawaiian archipelago shows strong evidence for a 1000 to 2000 km wide hot thermal anomaly ponding beneath the 660 km boundary west of Hawaii islands [Q. Cao et al. Seismic imaging of transition zone discontinuities suggests hot mantle west of Hawaii. Science (2011), 332, 1068–1071]. This scenario suggests that Hawaiian volcanism may not be caused by a stationary narrow plume rising from the core–mantle boundary but by hot plume material first held back beneath the 660 km discontinuity and then entrained under the transition zone before coming up to the surface. Using a cylindrical model of iso-chemical mantle convection with multiple phase transitions, we investigate the dynamical conditions for obtaining this peculiar plume morphology. Focusing on the role exerted by pressuredependent thermodynamic and transport parameters, we show that a strong reduction of the coefficient of thermal expansion in the lower mantle and a viscosity hill at a depth of around 1800 km allow plumes to have enough focused buoyancy to reach and pass the 660 km depth interface. The lateral spreading of plumes near the top of the lower mantle manifests itself as a channel flow whose length is controlled by the viscosity contrast due to temperature variations ΔηT. For small values of ΔηT, broad and highly viscous plumes are generated that tend to pass through the transition zone relatively unperturbed. For higher values (102 ≤ ΔηT ≤ 103), we obtain horizontal channel flows beneath the 660 km boundary as long as 1500 km within a timescale that resembles that of Hawaiian hotspot activity. This finding could help to explain the origin of the broad hot anomaly observed west of Hawaii. For a normal thermal anomaly of 450 K associated with a lower mantle plume, we obtain activation energies of about 400 kJ/mol and 670 kJ/mol for ΔηT = 102 and 103, respectively, in good agreement with values based on lower mantle mineral physics. If an increase of the thermal conductivity with depth is also included, our model can predict both long channel flows beneath the 660 km discontinuity and also values of the radial velocity of local blobs sinking in the lower mantle of around 1 cm/yr, in good agreement with those inferred for the sinking rate of cold remnants of the subducted lithosphere. © 2011 Elsevier B.V. All rights reserved. 1. Introduction According to the traditional plume hypothesis (see e.g. Campbell, 2007, for a recent review), volcanism at Hawaii and Emperor chain is attributed to a narrow plume that rises from the core–mantle boundary (CMB) and remains essentially stationary for at least 100 Myr below the Pacific plate. This model is supported by geochemical (e.g. Hofmann, 1997) and seismic observations according to which a distinct thermal anomaly beneath Hawaii can be tracked deep into the lower mantle (e.g. Montelli et al., 2006; Wolfe et al., 2009). However, the seismic evidence of deep rooted plumes is controversial and considered as not conclusive (e.g. Boschi et al., 2006). On the other hand, seismic studies have been conducted that, underneath Hawaii, do not show any anomalous decrease in the wave speed nor in the thickness of the transition ⁎ Corresponding author. Department of Planetary Physics, German Aerospace Center (DLR), Berlin, Germany. E-mail address: [email protected] (N. Tosi). 0012-821X/$ – see front matter © 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.epsl.2011.10.015 zone nor are compatible with high temperatures or extensive amounts of melt (e.g. Deuss, 2007; Ritsema et al., 2009; Tauzin et al., 2008). To explain intra-plate volcanism then, different dynamical mechanisms have been invoked that are controlled by upper mantle processes (Anderson, 2010) such as small-scale convection (Ballmer et al., 2010, 2011) or shear-driven generation of upwellings (Conrad et al., 2011). While still advocating the deep plume hypothesis, on paleomagnetic grounds (Tarduno et al., 2003), Tarduno et al. (2009) propose the prominent bend in the Hawaiian hotspot track to be a consequence either of large-scale deep mantle circulation advecting the plume conduit (Steinberger et al., 2004), the so-called mantle-wind, or of the motion of the base of the plume (Hansen et al., 1993) or of the capture, and subsequent release, of the plume head by a migrating ridge. Yet another interesting point of view is offered by recent highresolution seismic imaging of the transition zone topography that has displayed a great complexity in the shape of phase transitions boundaries beneath Hawaii (Cao et al., 2011). In particular, Cao and co-authors reported an unexpected correlation of the 410, 520 and N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 660 km boundaries about 2000 km west of the Hawaiian hotspot, with the three interfaces exhibiting marked downward deflections with respect to their reference depths. While the phase transitions from olivine to wadslyite at 410 km and from wadslyite to ringwoodite at 520 km have positive Clapeyron slopes and are indeed expected to occur anomalously deep in a region of high temperatures such as beneath a hotspot, the transition from ringwoodite to perovskitepericlase at 660 km has negative Clapeyron slope and is expected instead to take place at shallower depths. These authors interpreted the downward deflection of the 660 km boundary as a plausible evidence for an exothermic post-garnet transition (e.g. Hirose, 2002; Yu et al., 2011) forming the main seismic contrast and being caused by a temperature anomaly as high as 450 K with a broad lateral extent that ponds beneath the transition zone. Moving closer to the hotspot, updoming of the 660 km interface is observed, consistently with the expected deformation of the post-spinel phase boundary. From a geodynamic point of view, this scenario suggests that, despite the presence of an exothermic post-garnet transition, the top of the lower mantle can act as a barrier to an upwelling. Its deep mantle source would then not be aligned with the present-day location of the hotspot which in turn could represent the surface expression of complex pathways taken by the upwelling itself in the upper mantle. In our previous works, we followed the latest theoretical and experimental results on constraining thermodynamic properties of minerals at lower mantle conditions (e.g. de Koker, 2010; Goncharov et al., 2010; Katsura et al., 2009; Tang and Dong, 2010). In particular, we focused on the combined effects of strongly depth-dependent thermal expansivity and conductivity in global plane-layer models of mantle convection with multiple phase transitions (Tosi et al., 2010), including the one from perovskite to post-perovskite (Murakami et al., 2004; Oganov and Ono, 2004). We showed that despite the propensity of post-perovskite to destabilize the bottom thermal boundary layer (Matyska and Yuen, 2006; Nakagawa and Tackley, 2004), low thermal expansivity (Katsura et al., 2009) and high thermal conductivity values at depth (de Koker, 2010; Hofmeister, 2008; Ohta, 2010; Tang and Dong, 2010) allow for the formation of long-lived, large-scale thermal and thermo-chemical anomalies atop the CMB, with interesting consequences for the long-term stability of thermo-chemical piles (Dziewonski et al., 2010; Torsvik et al., 2010). Furthermore, we showed that strongly depth-dependent coefficients of thermal expansion and conduction turn out to be important ingredients for the formation of a regional plume-fed asthenosphere with low viscosity in the vicinity of hot upwellings in oceanic mantle (Yamamoto et al., 2007; Yuen et al., 2011). A plume morphology resembling that hypothesized by Cao et al. (2011) was observed, though not analyzed in detail, in numerical models of axisymmetric mantle convection with a modest resolution by Yuen et al. (1996, 2007) and by Yuen et al. (2011) in Cartesian models. Farnetani and Samuel (2005) and Samuel and Bercovici (2006) observed similar structures in the framework of regional, thermo-chemical plume dynamics. In this work we aim at studying the conditions under which this type of plumes can be obtained in connection with the key-role played by depth-dependent thermodynamic parameters. Moreover, as according to Cao et al. (2011) the hot material ponding near the top of the lower mantle can have a large lateral extent (about 1000 and possibly up to 2000 km), we will use this new piece of seismic information to place constraints on the mantle rheology. We will show that the magnitude of the lateral viscosity contrast due to temperature plays an important role in regulating the amount of lateral spreading of plume material beneath the 660 km boundary. Finally, our findings will be reinforced by demonstrating that, in our preferred models, the velocity of sinking slab remnants in the lower mantle matches well the constraint of about 1 cm/yr recently derived by van der Meer et al. (2010) by correlating cold remnants of subducted lithosphere, as seen in a tomographic model, with the corresponding tectonically-induced orogenic belts. 349 2. Model description 2.1. Governing equations and numerical method We have solved the conservation equations of mass, linear momentum and thermal energy with multiple phase transitions in a full cylindrical annulus using the extended Boussinesq approximation (e.g. King et al., 2010): ∂j vj ¼ 0; ð1Þ ! 3 X Rbl ¼ Ra αT− Γ δ ; −∂j p þ ∂j η ∂j vi þ ∂i vj Ra l jr l¼1 ð2Þ 3 X DT Di Rb DΓ ¼ ∂j κ∂j T þ Diαvr ðT þ T 0 Þ þ Ф þ Di l l γl ðT þ T 0 Þ þ R: Dt Ra Ra Dt l¼1 ð3Þ Symbols and parameters are explained in Table 1. The continuity Eq. (1) describes the conservation of mass for an incompressible fluid. In the momentum Eq. (2), where an infinite Prandtl number is assumed, the first and second terms on the right hand side account for the buoyancy forces due to temperature differences and phase transitions, respectively. Three phase transitions are considered: at 410 km depth from olivine (ol) to spinel (sp) (l=1), at 660 km depth from spinel to perovskite (pv) (l=2) and from perovskite to post-perovskite (ppv) (l=3). The effect of the l-th transition is taken into account through the traditional approach of Christensen and Yuen (1985) using a phase function Γl: Γl ¼ 1 z−zl ðT Þ 1 þ tanh ; 2 w ð4Þ where z is the depth, w is the width of the phase transition and zl(T) is defined as 0 0 zl ðT Þ ¼ zl þ γl T−T l ; ð5Þ where γl, zl0 and Tl0 are the Clapeyron slope, transition depth and transition temperature. In the thermal energy Eq. (3), heat advection is balanced by heat diffusion, adiabatic heating/cooling, viscous dissipation, latent heat release/absorption due to phase changes and internal heating due to radioactive sources. Since the models have been ran to steadystate, secular cooling has been neglected. Eqs. (1)–(3) have been solved with the 2D–3D spherical code Gaia (Huettig and Breuer, 2011). Gaia is based on a finite volume formulation with second-order accuracy in space and time. It uses the SIMPLE algorithm (Patankar, 1980) to enforce the incompressibility constraint (1) and a fully implicit three-levels time scheme to solve the advection–diffusion Eq. (3) (Harder and Hansen, 2005). This study is different from our previous works that were conducted in Cartesian geometry with a large aspect-ratio of 10 (Tosi et al., 2010; Yuen et al., 2011). Here we have adopted a cylindrical domain consisting of a full annulus. As discussed by Yuen et al. (2011), the choice of this geometry is appropriate in models with depthdependent thermal expansivity in which thermal plumes play a prominent role. In fact, in Cartesian models, the large size of the CMB surface allows for the formation of a large number of plumes that can cause an excessive increase of the mantle temperature. The smaller ratio of the inner to outer surface along with the increased degrees of freedom characteristic of the cylindrical geometry help to deliver smaller and more realistic temperatures. The use of full cylindrical or spherical domains in connection with free-slip boundary conditions can induce in the solution a spurious component of pure rotation of the mantle as a whole since the velocity field is determined only up to a rigid body rotation. To remove this pure 350 N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 Table 1 Non-dimensional variables and numbers with the corresponding scaling and dimensional parameters employed in this study. Note that while Eqs. (4) and (5) are written in nondimensional form, here, for clarity, the dimensional values of the reference depth, reference temperature and Clapeyron slope of the phase transitions are reported. Symbol/ definition Description Scaling/ numerical value t z vj vr p T T0 = Ts/ΔT Φ = η(∂jvi + ∂ivj)∂ivj η α k k κ¼ ρ0 cp Time Depth Velocity component Radial velocity Dynamic pressure Temperature Surface temperature (non-dimensional) Viscous dissipation d2κs− 1 d κsd− 1 κsd− 1 ηsκsd− 2 ΔT 0.85 ηsκs2d− 4 Viscosity Thermal expansivity Thermal conductivity ηs αs ks Thermal diffusivity κs Thermal Rayleigh number 1.9 × 107 l-th phase Rayleigh number Rayleigh number for ol–sp transition Rayleigh number for sp–pv transition Rayleigh number for pv–ppv transition – 1.1 × 107 1.4 × 107 2.7 × 106 Dissipation number 0.68 v′ ¼ v−ω r: ð7Þ The surface and CMB are treated as free-slip and isothermal boundaries. At the CMB, the temperature is set to 3800 K consistently with recent estimates obtained by combining seismological and mineral physics models (Kawai and Tsuchiya, 2009; van der Hilst et al., 2007). As initial condition we employ a randomly perturbed adiabatic profile with a potential temperature of 1600 K along with 100 km thick thermal boundary layers at the top and bottom. To discretize the model domain, we use 256 equally spaced radial shells corresponding to a radial resolution of about 11 km. Each shell is then subdivided into 2749 portions in such a way that the lateral resolution of the grid equals the radial resolution at the mid-mantle depth. A total of more than 7 × 105 grid points is then employed for all calculations. Resolution tests conducted using 384 shells and about 1.2 × 106 grid points have not shown any significant difference in the results. 3 ρ α s ΔTd g 0 Ra ¼ 0 ηs κ s where V is the shell volume and r the position vector. Then we subtract the rotational motion from the solution vector v to obtain a corrected velocity v′ (Kameyama et al., 2008): 3 δρ d g 0 Rbl ¼ l ηs κ s Rb1 Rb2 Rb3 α s dg 0 Di ¼ cp 2.2. Rheology and thermodynamic parameters We have assumed a Newtonian rheology with viscosity dependent on depth and temperature as follows: 2 R¼ d g0 ρ0 cp ηs αs ks κs Ts ΔT w T10 T20 T30 z10 z20 z30 γ1 γ2 γ3 δρ1 δρ2 δρ3 H0 H0 d cp κ s ΔT Internal heating number 20 Mantle depth Gravity acceleration Reference density Heat capacity Reference viscosity Reference thermal expansivity Reference thermal conductivity Reference thermal diffusivity Surface temperature Temperature drop across the mantle Phase transition width Reference temperature for ol–sp transition Reference temperature for sp–pv transition Reference temperature for pv–ppv transition Reference depth of ol–sp transition Reference depth of sp–pv transition Reference depth of pv–ppv transition Clapeyron slope for ol–sp transition (Bina and Helffrich, 1994) Clapeyron slope for sp–pv transition (Bina and Helffrich, 1994) Clapeyron slope for pv–ppv transition (Hirose, 2006) Density jump for ol–sp transition (Steinbach and Yuen, 1995) Density jump for sp–pv transition (Steinbach and Yuen, 1995) Density jump for pv–ppv transition (Oganov and Ono, 2004) Internal heating rate 2890 km 9.8 m s−2 4500 kg m−3 1250 J kg−1 1022 Pa s 3 × 10− 5 K− 1 3.3 W m−1 K−1 5.9 × 10− 7 m 2s−1 300 K 3500 K 25 km 1760 K ηðz; T Þ ¼ η1 ðzÞη2 ðT Þ; 1870 K 3600 K ð8Þ where the depth-dependent part η1(z) is defined as in Hanyk et al. (1995): 2 : η1 ðzÞ ¼ 1 þ 214:3 z exp −16:7ð0:7−zÞ ð9Þ Eq. (9) implies a viscosity maximum in the mid lower mantle (solid black line in Fig. 1) (Morra et al., 2010; Tosi et al., 2010) which is consistent with modeling studies of the dynamic geoid (Čadek and van den Berg, 1998; Ricard and Wuming, 1991), postglacial rebound (Mitrovica and Forte, 2004; Tosi et al., 2005) and with molecular dynamics simulations of self-diffusion of MgO-periclase (Ito and Toriumi, 2010). 410 km 660 km 2890 km 3 MPa K−1 − 2.5 MPa K−1 13 MPa K−1 273 kg m−3 342 kg m−3 67.5 kg m−3 5.6 × 10− 12 W kg−1 rotation from the solution, we calculate at each time step the average angular velocity ω of the cylindrical shell: 1 rv ω¼ ∫ dV; V V ‖r‖2 ð6Þ Fig. 1. Non-dimensional vertical profiles of the thermal conductivity (blue), thermal expansivity (red), decadic logarithm of the viscosity according to Eq. (9) (solid black) and according to Eq. (8) with ΔηT = 10 (dashed black), ΔηT = 102 (dashed-dotted black) and ΔηT = 103 (dotted black) calculated for a typical mantle geotherm. N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 For the temperature-dependent part we have employed a standard linearization of the Arrhenius law 351 3. Results 3.1. Global convection patterns η2 ðT Þ ¼ expð−βT Þ; ð10Þ where β = log(ΔηT) regulates the magnitude of the viscosity contrast due to temperature variations. In Fig. 1 we have plotted the viscosity profiles obtained from Eq. (8) using an adiabat with potential temperature of 1600 K, 100 km thick thermal boundary layers and β = 2.3 (dashed black line), 4.6 (dash-dotted black line) and 6.9 (dotted black line), corresponding to changes in viscosity by one, two or three orders of magnitude respectively, i.e. ΔηT = 10, 102 and 103. Admittedly, Eq. (10) does not allow us to account for the viscosity gradient in the upper mantle and for the strong viscosity localizations that would be present if a proper plastic rheology for plate-like behavior were considered (e.g. (Tackley, 2000)). Nevertheless, it represents a good approximation of the Arrhenius law in the lower mantle where the average temperature is higher. Beside the phase transitions at 410 and 660 km depth, all models also include the transition from pv to ppv near the CMB. The current knowledge of ppv rheology is extremely uncertain (Karato, 2010). Nevertheless, as experimental (Hunt et al., 2009) and theoretical evidence (Ammann et al., 2010) suggest that ppv may be significantly weaker than pv, we have assumed the viscosity of the ppv phase to be one order of magnitude lower than that of the surrounding pv phase. With this assumption, the ratio between the maximum lower mantle viscosity and the average viscosity of the D″ layer can become quite large. Such a viscosity structure is in accord with a recent study of Nakada and Karato (in press). In fact, by analyzing the decay time of the Chandler wobble and time-dependent tidal deformations in the framework of viscoelastic deformation models, Nakada and Karato estimated the viscosity in the bottom 300 km of the mantle to be 2–3 orders of magnitude lower than the average lower mantle viscosity and up to 4 orders of magnitude lower in the bottom 100 km. Note that even though the viscosity of ppv can influence the stability of the bottom thermal boundary layer (Tosi et al., 2010), our results are not affected significantly by the assumption of a weaker ppv. In those models where the thermal expansivity α depends on depth, we have assumed the Anderson-Grüneisen parameter to be 5 in the upper mantle and 6.5 in the lower mantle. These values, which are consistent with experiments conducted on olivine (Chopelas and Boehler, 1992) and perovskite (Katsura et al., 2009), imply an overall decrease of α of about one order of magnitude throughout the mantle (red line in Fig. 1). The temperature dependence of the thermal expansivity, which is known to be mostly relevant for the heat transport and distribution of flow velocities in the upper mantle (Ghias and Jarvis, 2008; Schmeling et al., 2003), is here neglected. Although no conclusive agreement has been reached on the values that the lattice thermal conductivity k takes on at lower mantle pressures, it is generally well accepted that it increases steadily both across the upper (e.g. (Xu et al., 2004)) and lower mantle (e.g. (Goncharov et al., 2010; Ohta, 2010)). Ab-initio calculations of the thermal conductivity of MgO periclase yield values of the total lattice conductivity near the CMB that range from 6 Wm− 1 K − 1 (de Koker, 2010) to 10 Wm − 1 K − 1 (Tang and Dong, 2010), the latter being also consistent with experiments conducted on perovskite (Ohta, 2010). Nevertheless, values as high as 20–30 Wm− 1 K− 1 have also been proposed (Hofmeister, 2008). In our simulations in which k depends on depth, we have assumed for simplicity a linear increase across the mantle by a factor of 3, corresponding to k ~ 10 Wm − 1 K − 1 at the CMB (blue line in Fig. 1). Although under high pressure conditions of the deep mantle a T − 1 dependence of the lattice conductivity on temperature is likely to be present (de Koker, 2010; Ohta, 2010), for simplicity it is here neglected. We start showing in Fig. 2 long-term snapshots of the temperature field. Fig. 2d shows a model where both the thermal expansivity α and thermal conductivity k are constant and the temperature viscosity contrast ΔηT is 102. The temperature distribution is that characteristic of convection at relatively high Rayleigh numbers with thin and highly unstable thermal boundary layers (TBL). Cold downwellings interact with the 410 and 660 km discontinuities and are often deflected horizontally by the endothermic phase transition at 660 km depth that tends to retard slabs from sinking into the lower mantle. Despite the relatively high viscosity imposed in the mid-lower mantle (Fig. 1), hot upwellings are short-lived and loose their thermal signature before reaching the upper mantle. Models that feature a constant thermal expansivity and conductivity but different values of ΔηT (not shown here) exhibit a very similar behavior. The effects induced by depth-dependent thermal expansivity can be readily observed in the top line of Fig. 2 which shows global temperature snapshots obtained using three different temperature viscosity contrasts: ΔηT = 10 (Fig. 2a), 102 (Fig. 2b) and 103 (Fig. 2c). A thermal expansivity that decreases strongly with depth promotes the formation of large-scale plumes (Hansen et al., 1993; Tosi et al., 2010). In fact, as soon as a thermal instability forms at the bottom TBL and starts to rise, it acquires buoyancy during its ascent because of the growing thermal expansivity with height. The opposite argument applies to downwellings that lose their buoyancy while sinking into the mantle because of the reduction of α with the depth. As a consequence, and also because of the presence of a viscosity hill, they stagnate in the mid mantle causing an overall cooling of the lower mantle. The effects of employing different values of ΔηT can be readily seen by comparing panels a, b and c. For small viscosity contrasts (ΔηT =10, panel a), plumes are thick and highly viscous. Thus, they tend to pass the endothermic phase transition at 660 km depth essentially unperturbed, although occasionally short channeling of the flow beneath the transition zone can take place as for the plume portrayed in the bottom right part of panel a. For ΔηT =102 (panel b), the interaction of plumes with the 660 km phase change is now much more pronounced. Plumes have a reduced buoyancy, being less viscous and thinner. Locally layered convection generally causes them to travel horizontally along the top part of lower mantle cells giving rise to pronounced flow channeling beneath the 660 km discontinuity. As an alternative mechanism, tilting of the plume conduit by large-scale mantle flow is also observed (Kerr and Mériaux, 2004; Richards and Griffiths, 1988; Skilbeck and Whitehead, 1978). A more significant layering is observed when ΔηT is raised to 103 (panel c). A few plumes still retain enough buoyancy to reach the surface, normally after interacting and being horizontally deflected by the spinel–perovskite boundary. However, since plumes are thin and have a low viscosity, they often fail to develop completely, as large-scale mantle flow tends to shear and diffuse them. It is well known that the presence of internal heat sources influences the planform of convection by increasing the importance of downwellings over that of upwellings (e.g. (van den Berg et al., 2002)). Nevertheless, with a strong decrease of the thermal expansivity, the role of plumes is so prominent that the convection pattern is qualitatively unaltered by the presence of internal heating. This can be readily seen by comparing the models shown in Fig. 2b (without internal heating) and 2e (with internal heating), which look very similar, apart from an expected increase of the average temperature. From a quantitative viewpoint, the volume-averaged mantle temperature is 1750 K in the first case and 1840 K in the second one. Furthermore, plume buoyancy fluxes are not much dissimilar in the two cases: 1200 kg/s and 850 kg/ s are typical values obtained in the two simulations shown in Fig. 2b and e, respectively. In the following we will then consider for simplicity only basally heated situations. 352 N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 Fig. 2. Long-term snapshots of the temperature distribution for six models. All models are basally heated except for (e) in which also internal heating is considered. Top line: constant thermal conductivity, depth-dependent thermal expansivity and ΔηT = 10 (a), 102 (b) and 103 (c). Bottom line: ΔηT = 102 and constant thermal conductivity and expansivity (d), constant thermal conductivity and depth-dependent thermal expansivity (e), depth-dependent thermal conductivity and expansivity (f). To highlight the effects due to variable thermal conductivity, in Fig. 2f we show a snapshot from a model that features both depthdependent thermal expansivity and conductivity along with a temperature viscosity contrast of 102. This plot can be directly compared with Fig. 2b in which k is constant. The morphology of plumes and the way these are bent by the 660 km boundary and form long channel flows is very similar. However, with an increasing thermal conductivity that favors heat transport by conduction at depth, plumes become thicker and stronger. They rise through the lower mantle in an essentially vertical fashion and tilting of the plume conduit due to largescale mantle flow is not observed. 3.2. Channel flows beneath the 660 km discontinuity: morphology As described in the previous section, models that do not include a depth-dependent thermal expansivity fail to deliver plumes that are buoyant enough to pass through the transition zone and reach the surface. Therefore, as our goal is to characterize mantle upwellings that interact clearly with the transition zone, from now on we will focus only on models that incorporate a depth-dependent thermal expansivity, with the possible addition of depth-dependent thermal conductivity. In Fig. 3 we show contour plots of the temperature anomaly δT ¼ T−T ðrÞ (where T ðr Þ is the instantaneous, radially averaged temperature profile) for six different models, while in Fig. 4 we show the corresponding viscosity distribution. Since in all models the viscosity of ppv is assumed to be lower than that of pv by one order of magnitude, in Fig. 4 the occurrence of the ppv phase can be readily seen. Figs. 3 and 4 portray the typical behavior of upwellings interacting with the 660 km depth discontinuity, which depends on the three parameters α, k and ΔηT. The thermal conductivity is constant in the top line (panels a, b, c), while it increases with depth in the bottom line (panels d, e, f). In the first column (panels a, d), the temperature viscosity contrast ΔηT is set to 10, while in the second (b, e) and third columns (c, f) is set to 102 and 103, respectively. In order to emphasize the differences in the amount of hot material that accumulates near the top of the lower mantle, the snapshots shown in Figs. 3 and 4 have been selected for the maximum offset between the upper and lower mantle branches of the plume (see also Figs. 5 and 6). Quantitatively, the excess temperature associated with upwellings does not present large differences among the models shown here. Plumes are typically characterized by a compact hot core close to the CMB where δT = 1000–1250 K. Near the top of the lower mantle, temperature anomalies of ~500 K are observed that decrease to ~250 K in the upper mantle. Plume size and dynamics are strongly influenced by the value employed for the viscosity contrast due to temperature. For ΔηT = 10 (panels a and d in Figs. 3 and 4), plumes are only slightly deflected by the 660 km discontinuity. As their viscosity is close to that of the ambient mantle, they do not deform easily and the negative buoyancy arising from the upward deflection of the spinel–perovskite boundary is not sufficient to make hot rising material accumulate at the top of the lower mantle. For values of ΔηT of 102 and 103 (panels b, e and c, f in both Figs. 3 and 4), plumes have a significantly lower viscosity with respect to the surroundings, which facilitates their deformation. The increased effectiveness of this transition in hindering the motion of rising material causes plumes to assume a bent-shaped form. They tend to flow preferentially beneath the phase boundary, following a horizontal track which generally extends over several hundred kilometers, in N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 353 Fig. 3. Effect of the temperature viscosity contrast ΔηT on plume-channel flows beneath the 660 km discontinuity. The color scale indicates temperature variations with respect to the average radial profile. Top line: constant thermal conductivity and depth-dependent thermal expansivity with ΔηT =10 (a), ΔηT =102 (b) and ΔηT =103 (c). Bottom line: depth-dependent thermal conductivity and expansivity with ΔηT =10 (d), ΔηT =102 (e) and ΔηT =103 (f). Dashed lines indicate the actual shape of the boundaries of the olivine-spinel and spinel–perovskite phase transitions. some cases over 1000 km or more (Fig. 3b and e). Unlike the stagnant slab flows, which are sluggish, these hot channel flows are rapid, with horizontal velocities beneath the 660 km boundary as high as 15 cm/yr. From a qualitative point of view, the effects due to the presence of depth-dependent thermal conductivity can be easily recognized by comparing the top and bottom panels of Fig. 3. Using a depthdependent k promotes a slightly hotter mantle with plumes that tend to be thicker than their counterparts obtained with a constant conductivity (Tosi et al., 2010). Nevertheless, the manner in which plume motion is affected by the phase transitions remains essentially unchanged. The temperature distribution portrayed in Fig. 3e is particularly noteworthy (see also Fig. 6 for a plot of the time-evolution). Here a cold downwelling splashes on a hot plume channel flowing beneath the 660 km boundary and causes the formation and rise of a secondary plume. This behavior reveals that the plume channels observed in our models can also serve as local boundary layers and hence as a source of secondary thermal instabilities. 3.3. Channel flows beneath the 660 km discontinuity: formation and evolution We analyze now more closely the formation of channel flows and discuss quantitatively the effects due to variable conductivity. To this end, we show, in Figs. 5 and 6 respectively, the temporal evolution of plumes observed in a model with and without depth-dependent thermal conductivity obtained using ΔηT = 102. In both cases the way the channel flow forms is similar. First, a large-sized plume head reaches the transition zone. Its thinned tip flows through the upper mantle, helped by the exothermic transition at 410 km depth. As more and more hot material is supplied from the bottom, upward distortion of the 660 km boundary becomes more and more significant and causes the plume to bend. The convection becomes locally layered. Hence, instead of flowing in a vertical fashion, the rising plume flows now preferentially along the top of the lower mantle while advecting its upper mantle branch (Fig. 9). At the same time, the base of the plume near the CMB moves in the opposite direction. For the two plumes shown in Figs. 5 and 6, and generally for the majority of those generated in models that feature the same parametrizations, the horizontal channel flows beneath the 660 km boundary reach a considerable length, over 700 km. Hot plume material with an excess temperature as high as 500 K can eventually become spread laterally over long distances of the order of 1000 to 1500 km. While the plumes shown in Figs. 5 and 6 do not exhibit principle qualitative differences both in their shape and in the way they form, they are characterized instead by a remarkably different temporal evolution. With constant thermal expansivity (Fig. 5), the time interval measured from the moment the plume hits the 660 km discontinuity until the resulting channel flow reaches its maximal length is about 250 Myr, while it is much shorter (~100 Myr) for the case with variable thermal conductivity (Fig. 6). In fact, large values of k near the CMB favor a diffusive regime for the coldest portions of the mantle (Tosi et al., 2010). This amount of additional heat in turn powers the advection of plumes which tend to flow faster. Thus the lifetimes of the hot channel flows below the transition zone are influenced by the thermal conductivity of the deep mantle, which is a somewhat surprising finding. 3.4. Flow velocities in the lower mantle As long as a variable thermal expansivity is used, long plume channel flows are obtained both using constant and variable conductivity 354 N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 Fig. 4. Viscosity distribution for the same models plotted in Fig. 3. Fig. 5. Time-evolution of plume-channel flow for a model with constant thermal conductivity, depth-dependent thermal expansivity and ΔηT = 102. The color scale indicates temperature variations with respect to the averaged radial profile. Dashed lines indicate the actual shape of the boundaries of the olivine–spinel and spinel–perovskite phase transitions. N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 355 Fig. 6. As Fig. 5 but for a model with depth-dependent thermal conductivity and expansivity and ΔηT = 102. Fig. 7. Time-evolution of sinking slab remnants and a rising plume in the lower mantle for a model with constant thermal conductivity, depth-dependent thermal expansivity and ΔηT = 102. Top line: temperature variations with respect to the averaged radial profile. Bottom line: distribution of the radial velocity. Dashed lines indicate the shape of the boundaries of the olivine–spinel and spinel–perovskite phase transitions. 356 N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 even though different time-scales characterize their evolution. In order to establish which is the preferable model, we have looked at characteristic mantle flow velocities taken from our simulations. In Fig. 7 we show three snapshots, taken from a simulation in which k is constant and ΔηT = 102, that portray the motion of typical remnants of cold downwellings that sink in the lower mantle (top line) along with the corresponding distribution of radial flow velocity (bottom line). As a clear consequence of mass conservation, the nearly vertical downgoing motion of subducted material is accompanied by the simultaneous generation of a plume that rises with a very similar speed (~0.5 cm/yr). This indicates that, in first approximation, values of the average radial flow velocity can be considered as representative of both down- and upwellings. Under this assumption, we show in Fig. 8 a summary of vertical profiles of the radial (red lines) and tangential velocities (blue lines) for nine different models. In the top line both α and k are constant, in the middle line only α is depth-dependent while in the bottom line both α and k are depth-dependent. The three columns correspond to values of ΔηT of 10, 102 and 103, respectively. As a reference, we have also plotted the value of 1.2 cm/yr (dashed black line) recently estimated by (van der Meer et al., 2010) for the sinking rate of ancient slab remnants. This value was obtained by identifying in a global tomographic model a large number of oceanic plates subducted in the lower mantle and linking them to the present-day mountain ranges to which they are likely to have given rise. On the one hand, using constant thermodynamic parameters (panels a–c), a characteristic arc-shaped profile with a peak in the mid lower mantle and no evidence of the presence of phase transitions is obtained for the radial velocity. On the other hand, panels d–i, in which either α (d–f) or both α and k (g–i) are depthdependent, clearly show partial layering of the mantle. In fact, local minima and maxima near the 660 km boundary are observed for the radial and horizontal velocity, respectively. With ΔηT = 102 (Fig. 8b), a relatively good match with the value proposed by van der Meer et al. (2010) is observed, although this model is not able to generate ponds of hot material beneath the transition zone as hypothesized by Cao et al. (2011). While this can be easily achieved using depth-dependent thermal expansivity only, panels d–f of Fig. 8 also show that the radial velocity profiles obtained for these models clearly underestimate the reference value of 1.2 cm/yr no Fig. 8. Summary of profiles of the average absolute radial (red) and horizontal velocity (blue) for different models. Top line (a, b, c): constant thermal expansivity and conductivity. Mid line (d, e, f): depth-dependent thermal expansivity and constant conductivity. Bottom line (g, h, i): depth-dependent thermal expansivity and conductivity. Left column (a, d, f): ΔηT = 10. Mid column (b, e, h): ΔηT = 102. Right column (c, f, i): ΔηT = 103. The green box highlights the two models whose radial velocity profile best approximates the value of 1.2 cm/yr proposed by (van der Meer et al., 2010) for the rate of sinking velocity of slabs remnants in the lower mantle (black dashed lines). N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 matter what value of temperature viscosity contrast is used. The addition of depth-dependent thermal conductivity plays a fundamentally important role in regulating mantle flow velocities (see Section 3.3). Indeed if thermal expansivity and conductivity are considered together (Fig. 8g–i) and, in addition, the parameter ΔηT is chosen between 10 2 and 103, also a good match to the value of radial velocity predicted by van der Meer et al. (2010) can be attained (Fig. 8h and i). It is difficult to obtain these values of the radial flow velocity on the basis of Stokes' flow analysis using only a depth-dependent viscosity and depthdependent thermal expansivity. 4. Discussion and conclusions Recent seismic imaging investigation with multiply reflected S-waves has revealed a broad, hot, pond-like structure beneath the transition zone west of Hawaii that may represent a plausible source for the volcanism at Hawaii itself and perhaps other islands (Cao et al., 2011). This finding, albeit controversial (e.g. Wolfe et al., 2009), has allowed us to infer the existence of a channel flow under the 660 km discontinuity by using a numerical model of mantle convection in a 2D cylindrical geometry. Such a flow, when applied to the context of the central Pacific, would connect a lower mantle plume located west of Hawaii and a conduit-like plume in the upper mantle that fuels Hawaiian volcanism. We have found that this behavior is largely influenced by the depthdependent properties of the lower mantle, such as the thermal expansivity and thermal conductivity. The importance of the depthdependent expansivity on plumes has been known for a long time (Zhao and Yuen, 1987). In fact, a strong decrease of the expansivity with depth allows plumes to maintain their buoyancy and speeds them up on their upward passage. It is essential that plumes remain highly buoyant when they reach 660 km in order that they interact strongly with the negative Clapeyron slope phase transition so that the accompanying upward distortion of the phase boundary is significant (Schubert et al., 1975). This helps to stop the plumes and bend them at the same time, thus generating the channel flow we observe. In the framework of our models, without any depth-dependence of α, it is difficult to generate ponds of hot material beneath the phase boundary. However, it is important to note that this condition on the thermal expansivity can be not so strict in a wider context. The numerical models of Farnetani and Samuel (2005) and Samuel and Bercovici (2006) show in fact that accumulation of plume material beneath the transition zone can be easily achieved with a constant α if chemical heterogeneities are also accounted for. The importance of depth-dependent conductivity is qualitatively smaller than that of thermal expansivity. The plumes and channel flows look very similar both using constant and depth-dependent k. However, the important differences are in the faster flow velocities and shorter lifetimes of the channel flow in the cases in which k increases with the depth. Our models show that the characteristic time scale of a plume from its encounter with the phase transition at the top of the lower mantle to its eventual rise to the surface and subsequent advection in the upper mantle can range between approximately 100 Myr to 250 Myr, depending on whether a depth-dependent k is taken into account or not. From the viewpoint of global mantle dynamics, with large values of the thermal conductivity in the deep mantle, our models predict radial flow velocities which are in good agreement with those estimated recently by van der Meer et al. (2010). The range of lateral viscosity contrast between the hot plume and the adjacent mantle is also quite critical in maintaining this delicate configuration of channel flow and plume in dynamical equilibrium. On the one hand, for a small value of the temperature viscosity contrast (ΔηT = 10), our models predict large-scale plumes that, because of their large viscosity, tend to pass through the phase transition undergoing little deformation and thus failing to generate ponds of hot material near the top of the lower mantle. On the other hand, for larger values 357 of ΔηT, we obtain channel flows with a large lateral extent that could explain the origin of the broad anomaly observed west of Hawaii. In particular, when both α and k are depth-dependent, a typical channel flow thermal anomaly of 450 K, as observed in our models and as proposed by (Cao et al., 2011), implies an activation energy of about 190 kJ/mol for ΔηT = 10, 400 kJ/mol for ΔηT = 102 and 670 kJ/mol for ΔηT = 10. These values have been calculated from the Arrhenius law assuming that a given value of ΔηT corresponds to the mantle/plume viscosity contrast for a plume with an excess temperature of 450 K. Note that the latter two values bound the estimate of 500 kJ/mol proposed by Yamazaki and Karato (2001) for the lower mantle, which makes our preferred models also consistent with mineral physics estimates. The goal of this study was not to provide a detailed quantitative model for the Hawaiian hotspot. Nevertheless, when both α and k are depth-dependent, typical buoyancy fluxes that we obtain are approximately 1900 kg/s if ΔηT = 102 (Fig. 3e) and 4700 kg/s if ΔηT = 103 (Fig. 3f). These values compare reasonably well with estimates based on the magnitude of the swell that predict buoyancy fluxes in the range of 4000–6000 kg/s (Vidal and Bonneville, 2004). Thus, despite the simplified 2D geometry, our models do not imply unrealistically strong plumes. Fig. 9 schematically shows the channel flow scenario which leads to the buildup of the large pond of hot material beneath the transition zone. The plume advances with the growth of the pond. This mechanism can be thought of as a new form of plume advection being different from the large-scale mantle wind and plume capture by a ridge proposed by Tarduno et al. (2009). The length d of the pond being around 1000 to 1500 km is an important constraint of this model and could be tested using other data sources besides seismic imaging such as petrological, paleomagnetic and plate reconstruction data acting in concert. In this connection, another site where the channel flowplume scenario may be present is the Hainan plume (Lei et al., 2009). However, here the situation is more complex than for Hawaii because of the influence of the neighboring Philippine subducting slab. In a sense, there is a complementary relationship between channel flow and stagnant slab in the transition zone, both having been detected by seismology. The two phenomena represent opposite ends of the spectrum but they both yield important information about the physical properties of the mantle, such as the rheology. Research devoted to the detection and characterization of plume channel flows beneath the transition could represent an interesting counterpart of the research devoted to stagnant slabs which has been a major focus in Japan for over a decade (Fukao et al., 2009; Suetsugu et al., 2010). Fig. 9. Schematic diagram (not drawn to scale) illustrating the formation mechanism of a bent-shaped plume with the accompanying channel flow below the transition zone. Time t0: a highly buoyant plume reaches the top of the lower mantle. Time t1: after being retarded and slightly deformed by the 660 km phase transition, the plume reaches the surface at a location x1. Time t2: hot material supplied from below is held back by the phase transition and flows preferentially horizontally while the upper mantle branch of the plume is advected towards a new position x2. Time t3: horizontal motion beneath the transition zone continues until the tip of the plume at the surface reaches the position x3 and the channel flow attains its maximal horizontal length d. 358 N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 Our models are characterized by a series of simplifications that should be addressed in the future. We should include the phase changes in the garnet system and test the effects of the majorite–perovskite transition. If exothermic (Hirose, 2002), this transition would speed up the hottest portion of the plumes, reducing the propensity to local layering and to the formation of bent-shaped plumes. Nevertheless, the exact nature of the Clapeyron slope is still fraught with some uncertainty from both theoretical and experimental standpoints (Wentzcovitch et al., 2010). The temperature dependence of α and the influence of the T− 1 term in the phonon part of the thermal conductivity should be ultimately included in global mantle convection models as much as a plate-like rheology from which a more realistic convective behavior of the upper mantle and lithosphere is to be expected. Eventually the plume morphology that we have discussed in this work should be studied in a regional, three-dimensional context. Acknowledgments We are grateful to Maxim Ballmer and an anonymous reviewer for their thorough and constructive reviews and to the editor Yanick Ricard for his comments. We thank discussions with Maarten DeHoop. This work has been supported by the Czech Science Foundation (project P210/11/1366), the Helmholtz Association through the Research Alliance “Planetary Evolution and Life”, the Deutsche Forschungs Gemeinschaft (grant number TO 704/1-1) and the NSF Geophysics Program. D.A. Yuen was supported by the Senior Visiting Professorship Program of the Chinese Academy of Sciences. References Ammann, M.W., Brodholt, J.P., Wookey, J., Dobson, D.P., 2010. First-principles constraints on diffusion in lower-mantle minerals and a weak D″ layer. Nature 465, 462–465. doi:10.1038/nature09052. Anderson, D.L., 2010. Hawaii, boundary layers and ambient mantle — geophysical constraints. J. Pet.. doi:10.1093/petrology/egq068 Ballmer, M.D., Ito, G., van Hunen, J., Tackley, P.J., 2010. Small-scale sublithospheric convection reconciles geochemistry and geochronology of ‘Superplume’ volcanism in the western and south Pacific. Earth Planet. Sci. Lett. 290, 224–232. doi:10.1016/j.epsl.2009.12.025. Ballmer, M.D., Ito, G., van Hunen, J., Tackley, P.J., 2011. Spatial and temporal variability in hawaiian hotspot volcanism induced by small-scale convection. Nat. Geosci. 4, 457–460. doi:10.1038/ngeo1187. Bina, C., Helffrich, G., 1994. Phase transitions Clapeyron slopes and transition zone seismic discontinuity topography. J. Geophys. Res. 99, 15853–15860. Boschi, L., Becker, T.W., Soldati, G., Dziewonski, A.M., 2006. On the relevance of Born theory in global seismic tomography. Geophys. Res. Lett. 33. doi:10.1029/2005GL025063. Čadek, O., van den Berg, A.P., 1998. Radial profiles of temperature and viscosity in the Earth's mantle inferred from the geoid and lateral seismic structure. Earth Planet. Sci. Lett. 164, 607–615. doi:10.1016/S0012-821X(98), 00244-1. Campbell, I.H., 2007. Testing the plume theory. Chem. Geol. 241, 153–176. doi:10.1016/j. chemgeo.2007.01.024. Cao, Q., van der Hilst, R.D., de Hoop, M.V., Shim, S.H., 2011. Seismic imaging of transition zone discontinuities suggests hot mantle west of Hawaii. Science 332, 1068–1071. doi:10.1126/science.1202731. Chopelas, A., Boehler, R., 1992. Thermal expansivity in the lower mantle. Geophys. Res. Lett. 19, 1983–1986. Christensen, U.R., Yuen, D.A., 1985. Layered convection induced by phase transitions. J. Geophys. Res. 90, 10291–10300. Conrad, C.P., Bianco, T.A., Smith, E.I., Wessel, P., 2011. Patterns of intraplate volcanism controlled by asthenospheric shear. Nat. Geosci. 4, 317–321. doi:10.1038/ngeo1111. de Koker, N., 2010. Thermal conductivity of MgO periclase at high pressure: implications for the D″ region. Earth Planet. Sci. Lett. 292, 392–398. doi:10.1016/j.epsl.2010.02.011. Deuss, A., 2007. Seismic observations of transition zone discontinuities beneath hotspot locations. In: Foulger, G.T., Jurdy, D.M. (Eds.), Plates, Plumes and Planetary Processes. The Geological Society of America, pp. 121–136. Dziewonski, A.M., Lekic, V., Romanowicz, B.A., 2010. Mantle anchor structure: an argument for bottom up tectonics. Earth Planet. Sci. Lett. 299, 69–79. doi:10.1016/j. epsl.2010.08.013. Farnetani, C., Samuel, H., 2005. Beyond the thermal plume paradigm. Geophys. Res. Lett. 32. doi:10.1029/2005GL022360. Fukao, Y., Obayashi, M., Nakakuki, T., Utada, H., Suetsugu, D., Irifune, T., Ohtani, E., Yoshioka, S., Shiobara, H., Kanazawa, T., Hirose, K., 2009. Stagnant slab: a review. Ann. Rev. Earth Planet. Sci. 37, 19–46. doi:10.1146/annurev.earth.36.031207.124224. Ghias, S.R., Jarvis, G.T., 2008. Mantle convection models with temperature- and depthdependent thermal expansivity. J. Geophys. Res. 113. doi:10.1029/2007JB005355. Goncharov, A.F., Struzhkina, V.V., Montoya, J.A., Kharlamova, S., Kundargi, R., Siebert, J., Badro, J., Antonangeli, D., Ryerson, F.J., Mao, W., 2010. Effect of composition, structure, and spin state on the thermal conductivity of the Earth's lower mantle. Phys. Earth Planet. Inter. 180, 148–153. doi:10.1016/j.pepi.2010.02.002. Hansen, U., Yuen, D.A., Kroening, S.E., Larsen, T.B., 1993. Dynamical consequences of depth-dependent thermal expansivity and viscosity on mantle circulations and thermal structure. Phys. Earth Planet. Inter. 77, 205–223. Hanyk, L., Moser, J., Yuen, D.A., Matyska, C., 1995. Time-domain approach for the transient responses in stratified viscoelastic Earth. Geophys. Res. Lett. 22, 1285–1288. doi:10.1029/95GL01087. Harder, H., Hansen, U., 2005. A finite-volume solution method for thermal convection and dynamo problems in spherical shells. Geophys. J. Int. 161, 522–532. Hirose, K., 2002. Phase transitions in pyrolitic mantle around 670-km depth: implications for upwelling of plumes from the lower mantle. J. Geophys. Res. 107. doi:10.1029/ 2001JB000597. Hirose, K., 2006. Postperovskite phase transition and its geophysical implications. Rev. Geophys. 44. doi:10.1029/2005RG000186. Hofmann, H.W., 1997. Mantle geochemistry: the message from oceanic volcanism. Nature 385, 219–229. Hofmeister, A.M., 2008. Inference of high thermal transport in the lower mantle from laser-flash experiments and the damped harmonic oscillator model. Phys. Earth Planet. Inter. 170, 201–206. doi:10.1016/j.pepi.2008.06.034. Huettig, C., Breuer, D., 2011. Regime classification and planform scaling for internally heated mantle convection. Phys. Earth Planet. Inter. 186 (3–4), 111–124. doi:10.1016/j. pepi.2011.03.01. Hunt, S.A., Weidner, D.J., Li, L., Wang, L., Walte, N.P., Brodholt, J.P., Dobson, D.P., 2009. Weakening of calcium iridate during its transformation from perovskite to postperovskite. Nat. Geosci. 2, 794–797. doi:10.1038/ngeo663. Ito, I., Toriumi, M., 2010. Silicon self-diffusion of MgSiO3 perovskite by molecular dynamics and its implication for lower mantle rheology. J. Geophys. Res. 115. doi:10.1029/ 2010JB000843. Kameyama, M., Kageyama, A., Sato, T., 2008. Multigrid-based simulation code for mantle convection in spherical shell using Yin-Yang grid. Phys. Earth Planet. Int. 171, 19–32. doi:10.1016/j.pepi.2008.06.025. Karato, S., 2010. The influence of anisotropic diffusion on the high-temperature creep of a polycrystalline aggregate. Phys. Earth Planet. Inter. 183, 468–472. doi:10.1016/j. pepi.2010.09.001. Katsura, T., Yokoshi, S., Kawabe, K., Shatskiy, A., Manthilake, M.A.G.M., Zhai, S., Fukui, H., Hegoda, H.A.C.I., Yoshino, T., Yamazaki, D., Matsuzaki, T., Yoneda, A., Ito, E., Sugita, M., Tomioka, N., Hagiya, K., Nozawa, A., Funakoshi, K., 2009. P–V–T relations of MgSiO3 perovskite determined by in situ X-ray diffraction using a large-volume high-pressure apparatus. Geophys. Res. Lett. 36. doi:10.1029/2008GL035658. Kawai, K., Tsuchiya, T., 2009. Temperature profile in the lowermost mantle from seismological and mineral physics joint modeling. Proc. Natl. Acad. Sci. 106, 22119–22123. doi:10.1073/pnas.0905920106. Kerr, R.C., Mériaux, C., 2004. Structure and dynamics of sheared mantle plumes. Geochem. Geophys. Geosyst. 5. doi:10.1029/2004GC000749. King, S.D., Lee, C., van Keken, P., Leng, W., Zhong, S., Tan, E., Tosi, N., Kameyama, M., 2010. A community benchmark for 2D Cartesian compressible convection in the Earth's mantle. Geophys. J. Int. 180, 73–87. doi:10.1111/j.1365-246X.2009.04413.x. Lei, J., Zhao, D., Steinberger, B., Wu, B., Shen, F., Li, Z., 2009. New seismic constraints on the upper mantle structure of the Hainan plume. Phys. Earth Planet. Inter. 173, 33–50. doi:10.1016/j.pepi.2008.10.013. Matyska, C., Yuen, D.A., 2006. Lower mantle dynamics with the post-perovskite phase change, radiative thermal conductivity, temperature- and depth-dependent viscosity. Phys. Earth Planet. Inter. 154, 196–207. doi:10.1016/j.pepi.2005.10.001. Mitrovica, J.X., Forte, A., 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. Montelli, R., Nolet, G., Dahlen, F.A., Masters, G., 2006. A catalogue of deep mantle plumes: new results from finite-frequency tomography. Geochem. Geophys. Geosyst. 7. doi:10.1029/2006GC001248. Morra, G., Yuen, D.A., Boschi, L., Chatelain, P., Koumoutsakos, P., Tackley, P.J., 2010. The fate of the slabs interacting with a viscosity hill in the mid-mantle. Phys. Earth Planet. Inter. 180, 271–282. doi:10.1016/j.pepi.2010.04.001. Murakami, M., Hirose, K., Kawamura, K., Sata, N., Ohishi, Y., 2004. Post-perovskite phase transition in MgSiO3. Science 304, 855–858. Nakada, M., Karato, S., in press. Low viscosity of the bottom of the Earth's mantle inferred from the analysis of Chandler wobble and tidal deformation. Phys. Earth Planet. Int. Nakagawa, T., Tackley, P.J., 2004. Effects of a perovskite-post perovskite phase change near core-mantle boundary in compressible mantle convection. Geophys. Res. Lett. 31. doi:10.1029/2004GL020648. Oganov, A.R., Ono, S., 2004. Theoretical and experimental evidence for a post-perovskite phase of MgSiO3 in Earth's D″. Nature 430, 445–448. Ohta, K., 2010. Electrical and thermal conductivity of the Earth's lower mantle. Ph.D. thesis. Tokyo Institute of Technology. Patankar, S.V., 1980. Numerical heat transfer and fluid flow. Hemisphere series on computational methods in mechanics and thermal science. Taylor & Francis. Ricard, Y., Wuming, B., 1991. Inferring viscosity and the 3-D density structure of the mantle from geoid, topography and plate velocities. Geophys. J. Int. 105, 561–572. Richards, M.A., Griffiths, R.W., 1988. Deflection of plumes by mantle shear flow: experimental results and a simple theory. Geophys. J. 94, 367–376. Ritsema, J., Cupillard, P., Tauzin, B., Xu, W., Stixrude, L., Lithgow-Bertelloni, C., 2009. Joint mineral physics and seismic wave traveltime analysis of upper mantle temperature. Geology 37, 363–366. Samuel, H., Bercovici, D., 2006. Oscillating and stagnating plumes in the Earth's lower mantle. Earth Planet. Sci. Lett. 248, 90–105. doi:10.1016/j.epsl.2006.04.037. N. Tosi, D.A. Yuen / Earth and Planetary Science Letters 312 (2011) 348–359 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, 2224–2229. doi:10.1046/j.1365-246X.2003.01949.x. Schubert, G., Yuen, D.A., Turcotte, D.L., 1975. Role of phase transitions in a dynamic mantle. Geophys. J. R. Astron. Soc. 42, 705–735. Skilbeck, J., Whitehead, J., 1978. Formation of discrete islands in linear island chains. Nature 272, 499–501. doi:10.1038/272499a0. Steinbach, V., Yuen, D.A., 1995. The effects of temperature-dependent viscosity on mantle convection with the two major phase transitions. Phys. Earth Planet. Inter. 90, 13–36. doi:10.1016/0031-9201(95)03018-R. Steinberger, B., Sutherland, R., O'Connell, R.J., 2004. Prediction of Emperor-Hawaii seamount locations from a revised model of global plate motion and mantle flow. Nature 430, 167–173. doi:10.1038/nature02660. Suetsugu, D., Bina, C.R., Inoue, T., Wiens, D.A., 2010. Preface: deep slab and mantle dynamics. Phys. Earth Planet. Inter. 183, 1–3. doi:10.1016/j.pepi.2010.10.001 Special issue on deep slab and mantle dynamics. Tackley, P.J., 2000. Self-consistent generation of tectonic plates in time-dependent, threedimensional mantle convection simulations 1. Pseudo-plastic yielding. Geochem. Geophys. Geosyst. 1. doi:10.1029/2000GC000036. Tang, X., Dong, J., 2010. Lattice thermal conductivity of MgO at conditions of Earth's interior. Proc. Natl. Acad. Sci. 107, 4539–4543. doi:10.1073/pnas.0907194107. Tarduno, J.A., Duncan, R.A., Scholl, D.W., Cottrell, R.D., Steinberger, B., Thordarson, T., Kerr, B.C., Neal, C.R., Frey, F.A., Torii, M., Carvallo, C., 2003. The Emperor Seamounts: southward motion of the Hawaiian hotspot plume in Earth's mantle. Science 301, 1064–1069. doi:10.1126/science.1086442. Tarduno, J., Bunge, H.P., Sleep, N., Hansen, U., 2009. The bent Hawaiian-Emperor hotspot track: inheriting the mantle wind. Science 324, 50–53. doi:10.1126/science.1161256. Tauzin, B., Debayle, E., Wittlinger, G., 2008. The mantle transition zone as seen by global Pds phases: no clear evidence for a thin transition zone beneath hotspots. J. Geophys. Res. 113. doi:10.1029/2007JB005364. Torsvik, T.H., Burke, K., Steinberger, B., Webb, S.J., Ashwal, L.D., 2010. Diamonds sampled by plumes from the core–mantle boundary. Nature 466, 352–357. doi:10.1038/ nature09216. Tosi, N., Sabadini, R., Marotta, A.M., Veermersen, L.A., 2005. Simultaneous inversion for the Earth's mantle viscosity and ice mass imbalance in Antarctica and Greenland. J. Geophys. Res. 110. doi:10.1029/2004JB003236. Tosi, N., Yuen, D.A., Čadek, O., 2010. Dynamical consequences in the lower mantle with the post-perovskite phase change and strongly depth-dependent thermodynamic and transport properties. Earth Planet. Sci. Lett. 298, 229–243. doi:10.1016/j.epsl.2010. 08.001. 359 van den Berg, A.P., Yuen, D.A., Allwardt, J.R., 2002. Non-linear effects from variable thermal conductivity and mantle internal heating: implications for massive melting and secular cooling of the mantle. Phys. Earth Planet. Inter. 129, 359–375. van der Hilst, R., De Hoop, M.V., Wang, S.H., Shim, L., Tenorio, P., 2007. Seismo-stratigraphy and thermal structure of Earth's core–mantle boundary region. Science 315, 1813–1817. van der Meer, D.G., Spakman, W., van Hinsbergen, D.J.J., Amaru, M.L., Torsvik, T.H., 2010. Towards absolute plate motions constrained by lower-mantle slab remnants. Nat. Geosci. 3, 36–40. doi:10.1038/ngeo708. Vidal, V., Bonneville, A., 2004. Variations of the hawaiian hot spot activity revealed by variations in the magma production rate. J. Geophys. Res. 109. doi:10.1029/2003JB002559. Wentzcovitch, R.M., Wu, Z., Carrier, P., 2010. First principles quasiharmonic thermoelasticity of mantle minerals. Rev. Mineral. Geochem. 71, 99–128. doi:10.2138/rmg.2010.71.5. Wolfe, C.J., Solomon, S.C., Laske, G., Collins, J.A., Detrick, R.S., Orcutt, J.A., Bercovici, D., Hauri, E.H., 2009. Mantle shear-wave velocity structure beneath the Hawaiian hot spot. Science 326, 1388–1390. doi:10.1126/science.1180165. Xu, Y., Shankland, T.J., Linhardt, S., Rubie, D.C., Langenhorst, F., Klasinski, K., 2004. Thermal diffusivity and conductivity of olivine, wadsleyite and ringwoodite to 20 GPa and 1373 K. Phys. Earth Planet. Inter. 143–144, 321–336. doi:10.1016/j.pepi.2004.03.005. Yamamoto, M., Phipps Morgan, J., Morgan, W.J., 2007. Global plume-fed asthenosphere flow–I: motivation and model development. In: Foulger, G.T., Jurdy, D.M. (Eds.), Plates, Plumes and Planetary Processes. The Geological Society of America, pp. 165–188. Yamazaki, D., Karato, S.I., 2001. Some mineral physics constraints on rheology and geothermal structure of Earth's lower mantle. Am. Mineral. 86, 358–391. Yu, Y.G., Wentzcovitch, R.M., Vinograd, V.L., Angel, R.J., 2011. Thermodynamic properties of MgSiO3 majorite and phase transitions near 660 km depth in MgSiO3 and Mg2SiO4: a first principles study. J. Geophys. Res. 116. doi:10.1029/2010JB007912. Yuen, D.A., Čadek, O., van Keken, P.E., Reuteler, D.M., Kyvalova, H., Schroeder, B.A., 1996. Combined results from mineral physics, tomography and mantle convection and their implications on global geodynamics. In: Boschi, E., Ekström, G., Morelli, A. (Eds.), Seismic Modeling of the Earth's Structure. Editrice Compositori, Bologna, pp. 463–505. Yuen, D.A., Monnereau, M., Hansen, U., Kameyama, M., Matyska, C., 2007. Dynamics of superplumes in the lower mantle. In: Yuen, D.A., Maruyama, S., Karato, S.I., Windley, B.F. (Eds.), Superplumes: Beyond Plate Tectonics. Springer, pp. 239–267. Geophysics and Geodesy. Yuen, D.A., Tosi, N., Čadek, O., 2011. Influences of lower-mantle properties on the formation of asthenosphere in oceanic upper mantle. J. Earth Sci. 22, 143–154. doi:10.1007/s12583011-0166-9. Zhao, W., Yuen, D.A., 1987. The effects of adiabatic and viscous heatings on plumes. Geophys. Res. Lett. 14, 1223–1226.
© Copyright 2025 Paperzz