dynamics of a laminar plume in a cavity: the in uence of

21
Article Volume 14, Number 1 31 January 2013 doi:10.1029/2012GC004383 ISSN: 1525-2027 Dynamics of a laminar plume in a cavity: The inuence of boundaries on the steady state stem structure P. E. van Keken Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, Michigan, USA ([email protected]) A. Davaille Laboratoire FAST, Université Paris-Sud, Orsay, France J. Vatteville LDFG, Institut de Physique du Globe de Paris, Paris, France Laboratoire FAST, Université Paris-Sud, Orsay, France [1] The steady state structure of thermal plumes rising from a small heater is studied in high Prandtl number uids. We show good agreement between laboratory experiments and numerical simulations. We study the effect of the boundaries on the plume development by numerically simulating the plume rise in very large geometries. The thermal structure of the plume axis is similar for all box sizes considered, but the velocity structure changes strongly as box sizes are increased. We show that the effect of the side boundaries becomes unimportant for large aspect ratio, but that the free-slip top boundary has a strong inuence on the velocity structure under all conditions. We show that the use of an outow boundary condition signicantly reduces the inuence of the top boundary. Under these conditions we recover to good precision the theo- retical predictions for plumes rising in an semi-innite half-space. The strong inuence of the boundaries in high Prandtl number uids is important in the interpretation of laboratory experiments and numerical simulation for the dynamics of the Earths mantle. Components: 9,800 words, 15 figures. Keywords: laboratory experiments; geodynamics; mantle plumes; hot spots. Index Terms: 0550 Computational Geophysics: Model verification and validation; 8137 Tectonophysics: Hotspots, large igneous provinces, and flood basalt volcanism; 8121 Tectonophysics: Dynamics: convection currents, and mantle plumes. Received 9 August 2012; Revised 15 November 2012; Accepted 15 November 2012; Published 31 January 2013. van Keken, P. E., A. Davaille, and J. Vatteville (2013), Dynamics of a laminar plume in a cavity: The inuence of boundaries on the steady state stem structure, Geochem. Geophys. Geosyst., 14, 158178, doi:10.1029/2012GC004383. ©2013. American Geophysical Union. All Rights Reserved. 158

Upload: others

Post on 25-Jan-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dynamics of a laminar plume in a cavity: The in uence of

Article

Volume 14, Number 1

31 January 2013

doi:10.1029/2012GC004383

ISSN: 1525-2027

Dynamics of a laminar plume in a cavity: The influenceof boundaries on the steady state stem structure

P. E. van KekenDepartment of Earth and Environmental Sciences, University of Michigan, Ann Arbor, Michigan,USA ([email protected])

A. DavailleLaboratoire FAST, Université Paris-Sud, Orsay, France

J. VattevilleLDFG, Institut de Physique du Globe de Paris, Paris, France

Laboratoire FAST, Université Paris-Sud, Orsay, France

[1] The steady state structure of thermal plumes rising from a small heater is studied in high Prandtl numberfluids. We show good agreement between laboratory experiments and numerical simulations. We study theeffect of the boundaries on the plume development by numerically simulating the plume rise in very largegeometries. The thermal structure of the plume axis is similar for all box sizes considered, but the velocitystructure changes strongly as box sizes are increased. We show that the effect of the side boundariesbecomes unimportant for large aspect ratio, but that the free-slip top boundary has a strong influence on thevelocity structure under all conditions. We show that the use of an outflow boundary condition significantlyreduces the influence of the top boundary. Under these conditions we recover to good precision the theo-retical predictions for plumes rising in an semi-infinite half-space. The strong influence of the boundaries inhigh Prandtl number fluids is important in the interpretation of laboratory experiments and numericalsimulation for the dynamics of the Earth’s mantle.

Components: 9,800 words, 15 figures.

Keywords: laboratory experiments; geodynamics; mantle plumes; hot spots.

Index Terms: 0550 Computational Geophysics: Model verification and validation; 8137 Tectonophysics:Hotspots, large igneous provinces, and flood basalt volcanism; 8121 Tectonophysics: Dynamics: convectioncurrents, and mantle plumes.

Received 9 August 2012; Revised 15 November 2012; Accepted 15 November 2012; Published 31 January 2013.

van Keken, P. E., A. Davaille, and J. Vatteville (2013), Dynamics of a laminar plume in a cavity: The influence ofboundaries on the steady state stem structure,Geochem. Geophys. Geosyst., 14, 158–178, doi:10.1029/2012GC004383.

©2013. American Geophysical Union. All Rights Reserved. 158

Page 2: Dynamics of a laminar plume in a cavity: The in uence of

1. Introduction

1.1. Mantle Convection and Plumes

[2] Slow convection in the Earth’s mantle isexpressed at the surface by creation of plates atmidoceanic ridges and subsequent destruction ofthe plates upon subduction at convergent marginswith related volcanism. Hot spots such as Hawai‘iand Iceland are characterized by midplate volca-nism or excess volcanism at ridges. These hot spotregions appear to be supported by the long-termpresence of hot material below the lithosphere. Thebest fluid dynamical explanation for hotspotsremains the rise of mantle plumes, which are co-lumnar upwellings with vertical speeds that are sig-nificantly higher than they move laterally [e.g.,Morgan, 1971]. These plumes are ubiquitous fea-tures of thermal convection in viscous fluids. Fluiddynamical modeling using laboratory experiments,theoretical arguments, and numerical simulations(for recent reviews, see Ribe et al. [2006], Ito and vanKeken [2007], and Davaille and Limare [2007])demonstrates that the physical characteristics of hot-spots in many cases can be explained by plumesrising from the deep mantle, although shallowmantleorigins have also been suggested [King and Ritsema,2000]. Seismological evidence for low velocityregions below hotspots have also been used to arguefor the existence of plumes [Montelli et al., 2004;Wolfe et al., 2009], although it is likely that applica-tion of standard tomographic techniques is limiteddue to the effects of wave front healing that mayrender plumes invisible in the deeper mantle [Hwanget al., 2011].

[3] In order to improve our understanding of theformation of hotspots in the convecting mantle, it isessential to develop a strong fluid dynamical basisfor studying the development of mantle plumes.Numerous studies have been devoted to that task inthe last 50 years. The temperature is hot and uni-form over the whole core-mantle boundary andseveral plumes are expected to develop from the hotthermal boundary layer there. However, in order tobetter focus on the development of a single plume,a small patch heated at constant temperature can betaken as a proxy. This confined heat source ensuresthat only one plume is generated, and that it evolvesto a quasi steady state. This is the geometry that weshall also adopt in this study.

[4] Turner [1962] was the first to study what hecalled the starting plume, comprising a steadyplume conduit capped by a large buoyant head. He

further suggested that the buoyancy of the headincreases since the head is fed by the stem as aresult of the slower upwards motion of the headcompared with the steady stem below it. Sincethen, much effort has been devoted to understandplume dynamics and to provide scalings for plumeascent velocity [Whitehead and Luther, 1975;Shlien, 1976; Olson and Singer, 1985; Chay andShlien, 1986; Griffiths and Campbell, 1990; Moseset al., 1993; Couliette and Loper, 1995; van Keken,1997; Kaminski and Jaupart, 2003; Rogers andMorris, 2009; Davaille et al., 2011] or for steadystate plume stem structure [Batchelor, 1954; Fujii,1962; Shlien and Boxman, 1979; Tanny and Shlien,1985; Worster, 1986; Moses et al., 1993; Olsonet al., 1993; Couliette and Loper, 1995; Vasquezet al., 1996; Laudenbach and Christensen, 2001;Whittaker and Lister, 2006a, 2006b; Davaille et al.,2011]. Additionally, plume growth by entrainmentof fluid by thermal diffusion, continuous feedingfrom the source, laminar entrainment of surroundingmaterial at the rear of a leading vortical head hasbeen studied [Griffiths and Campbell, 1990; Moseset al., 1993; Couliette and Loper, 1995; Kumagai,2002]. It is well recognized that plume morphologyand time evolution can be complex due to phasechanges, rheological variations, mean ambient shearflow, and compositional effects [e.g., Bercoviciand Mahoney, 1994; van Keken, 1997; Thompsonand Tackley, 1998; Davaille, 1999; Farnetani andSamuel, 2005; Lin and van Keken, 2005; Kumagaiet al., 2008].

[5] Although the different studies generally agreeon the broad picture of plume dynamics, they oftenpropose quantitatively different scalings. One rea-son probably resides in the differences in boundaryconditions adopted by the authors. Theoreticalstudies often need to assume an infinite or semi-infinite fluid, while laboratory experiments gener-ally use rigid bottom and side boundaries anddeformable or rigid top boundary. On the other hand,numerical studies for a long time preferentially pre-scribed free-slip nondeformable boundaries. In somecases, rigid boundaries [e.g., Blankenbach et al.,1989], deformable boundaries [e.g., Schmeling et al.,2008], or open box conditions approximating a semi-infinite fluid [e.g.,Olson et al., 1993] have been usedas well. We nevertheless lack a detailed comparisonof the different assumptions on boundary conditionsto better understand their role on plume dynamics.

[6] In this paper, we will combine laboratorymodels with numerical simulations to quantify theinfluence of the proximity and nature of boundary

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

159

Page 3: Dynamics of a laminar plume in a cavity: The in uence of

conditions on the dynamics of a laminar thermalplume that rises from a small constant heater in anearly constant-viscosity fluid. The experimentalresults obtained in finite domains will be comparedto the theoretical predictions obtained in infinitefluid. This will enable us to discuss the limitationsof the different techniques of investigation formantle plume modeling.

1.2. Goal of the Paper

[7] This paper is one in a sequence of three. Itfollows up on Vatteville et al. [2009] and is acompanion to Vatteville et al. (J. Vatteville et al.,Development of a laminar thermal plume in acavity, paper II, in preparation, Journal of FluidMechanics). This set of papers studies the forma-tion of thermal plumes rising from a small heaterin very viscous fluids. We use a combination oflaboratory experiments (using silicone oils) and nu-merical simulations (using finite element methods).In Vatteville et al. [2009] we directly compared nu-merical simulations with the laboratory experimentsfor a single small box geometry. In paper II weconsider the various stages of the plume formation,from the conductive growth of the boundary layerto the formation of the steady state plume.

[8] In the present paper, we consider the establish-ment of the steady state structure and how theboundaries of the box affect the plume structure.We will first review the theory of the velocity andthermal structure of thermal plumes at high Prandtlnumber Pr (Pr= n/k, where n is the kinematic vis-cosity and k is the thermal conductivity). We willconfirm the good comparison between laboratorymodels and numerical simulations [Vatteville et al.,2009]. We then use the numerical models to scalethe models to much larger domains that wouldresemble laboratory experiments in very large tanks,with a viscous fluid filling a cavity that has no-slip(zero velocity) boundary conditions at the base andthe sides, and a free-slip (zero normal velocity, zerotangential stress) boundary at the top. We willdemonstrate that these boundary conditions play acrucial role in the formation of the plume structure.While the side boundaries become unimportant at asufficiently large aspect ratio of the domain, the topfree-slip boundary condition is important even atvery large box sizes. We will finally show that amodification of the boundary conditions will allowfor a good reproduction of the predicted structurefor a semi-infinite fluid from independent theory[Whittaker and Lister, 2006a].

1.3. Predictions for the Stem Structureof a Laminar Plume at High PrandtlNumber in a Semi-infinite Fluid

[9] Batchelor [1954] predicted that, for a steadystate laminar plume rising from a point source in aninfinite and constant viscosity fluid at infinitePrandtl number, the velocity at the plume axis Vshould be height-independent and scale with atypical velocity V0:

V � V0 ¼ gaQrCpn

� �12

(1)

where g is the gravitational constant, a is the ther-mal expansion coefficient, Q is the power of theplume, r is the density, Cp is the specific heat andn is the kinematic viscosity. The power of theplume is related to the temperature contrast ΔTbetween plume and ambient fluid and the plumevelocity V. It can be found at any depth in thefluid from the horizontal integral

Q ¼ 2prCp

Z 1

0VΔTrdr (2)

where r is the distance from the plume axis.

[10] A typical radius of the plume can be defined bya radius a outside of which the temperature anomalyis zero and inside which the temperature anomaly isquasi-constant. At high Pr the radius grows princi-pally by diffusion and a� kt where t is time. Usingthe definition of a typical length scale L0

L0 ¼ k2rCpngaQ

� �14

¼ k=V0 (3)

we can write a � zL0ð Þ12:Fujii [1962] suggested that

at the center line of a plume at infinite Pr the tem-perature contrast varies as

ΔT r ¼ 0ð Þ ¼ Q

4pkz(4)

where k is the thermal conductivity of the fluid andz is the height above the source.

[11] Using an asymptotic approach, Worster [1986]developed an analytical solution for the velocityand temperature contrast in the plume for fluids thathave high, but not infinite Pr, which introduced anexplicit Pr dependence in (1):

V ¼ f Prð ÞV0 (5)

where f Prð Þ ¼ lne=2pð Þ12 in which e is the solu-

tion to the equation e4ln(1/e2) = 1/Pr in the interval0 < e < e�

14 . In this analysis, the temperature con-

trast at the plume axis is given by (4).

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

160

Page 4: Dynamics of a laminar plume in a cavity: The in uence of

[12] In all cases above, the fluid is assumed to beinfinite and the velocity is independent of the heightabove the source. Whittaker and Lister [2006a]developed a boundary layer theory for a very viscousplume rising from a point source that sits on a planeboundary. This semi-infinite geometry is more rele-vant to studies of plumes in natural or laboratorysettings. Whittaker and Lister [2006a] derived thatthere is a typical height z0 above the heater at whichadvection and diffusion are comparable:

z0 ¼ 32pð Þ12L0 � 10L0 (6)

and that above this height the plume radius and verti-cal velocity in the plume slowly increase with height:

aWL � z0zð Þ12 ln z=z0ð Þ½ ��14 (7)

and

V � V0 ln z=z0ð Þ½ �12 (8)

[13] The temperature in the plume center remainsthe same as (4).

[14] Whittaker and Lister [2006b] provided anextension of this study to that of a finite pointsource (Figure 1). In this case the temperature atthe center of the plume is similar to that of (4) ifit is assumed that there is a virtual point source atsome depth z*< 0 below the actual heater (seealso Shlien and Boxman [1979] and Shlien andBoxman [1981]):

ΔT r ¼ 0ð Þ ¼ CQ

k z� z�ð Þ (9)

and

V ¼ VWL ¼ CvV0 ln z� z�ð Þ=z0ð Þ½ �12 (10)

where we have introduced a proportionality con-stant Cv. The prediction for C remains equal toCw ¼ 1

4p. We will refer to this set of predictions(9 + 10) of centerline velocity and temperature asWL06.

[15] The fluid dynamical experiments described inVatteville et al. [2009] and used here are for finite Pr(5000–50,000) fluids in a confined tank. Whittakerand Lister [2006a] describe that inertial effectsintroduce secondary corrections to their theory,

which become important at a radius of z0Pr lnPrð Þ12.For our laboratory experiments and related numeri-cal simulations this radius is well in excess of 10m,which is obviously much larger than the laboratorydimensions. In the mantle, plumes carrying 20–300GW [Olson et al., 1993] would feel inertialeffects for radii greater than 1025 km. The infinitePr description of [Whittaker and Lister, 2006a]should therefore be appropriate for plumes in thelaboratory as well as in the mantle. Nevertheless,the confined medium introduces important influ-ences of the side and top boundaries. Due to the highPr nature of the fluid these are felt by the plume inits rise and development of steady state structure.This leads to a centerline velocity profile that isquite distinct from (10). It is our main goal in thispaper to evaluate this difference quantitatively andto determine if we can approximate (10) better if(a) we simulate the plume evolution in boxes withvery large aspect ratio (to minimize the effect ofthe side boundaries) and (b) use an open boundarycondition at the top of the model (to minimize theeffect of the top boundary).

2. Experiment Set Up

2.1. Laboratory Experiments

[16] The laboratory set up is the same as that dis-cussed in Vatteville et al. [2009]. The heat sourceconsists of a Peltier element covered by a 3mmthick and 18mm diameter copper disk. It is placedat the center of a tank with a rectangular base ofdimension W2 where W = 150 mm. The tank isfilled with silicone oils (Rhodorsil 47V500 or47V5000) to a height H= 165 mm (Figure 2a). Theupper surface of the fluid is free and deformable. Attime t= 0, a constant electric power P, ranging from0.5 to 3.3W, is applied to the heater. In steady state,this corresponds to heater temperature between 21and 54�C.

[17] The silicone oils have nearly constant proper-ties over the temperature range considered in the

R

u u

zr

a(z)

z*

Figure 1. Modified fromWhittaker and Lister [2006b].Sketch of the steady state geometry of a plume risingfrom a disk heater of radius R. Above a certain heightthe temperature distribution in the plume is similar tothat of a plume rising of a point heater at virtual depth z*.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

161

Page 5: Dynamics of a laminar plume in a cavity: The in uence of

experiments. The viscosity variation is less than afactor of three over the largest temperature rangeconsidered [Vatteville et al., 2009].

[18] The fluid is seeded with 10mm diameter glassspheres and two types of thermochromic liquidcrystals. The thermochromic liquid crystals reflectlight at different temperatures [Davaille et al.,2011]. A vertical cross-section of the tank is illu-minated by a 532 nm laser sheet and images arerecorded every second using a CCD camera. Brightlines on the images of the laboratory experimentscorrespond to the two isotherms 24.1� 1�C and34.8� 1.5�C (see Davaille et al. [2011] for detailson the method and calibration).

[19] The velocity field was calculated by followingthe glass spheres using the particle-image-velocimetrypackage DaVis from LaVision. The resulting spatialresolution and amplitude precision were 2.5mm and0.1mm/s (or 5% of the typical maximum velocity),respectively. Although the tank base is square, weverified that the thermal and velocity fields were axi-symmetric by rotating the vertical laser cross-section[Davaille et al., 2011].

2.2. Numerical Simulations

2.2.1. Model Geometryand Boundary Conditions

[20] The geometry for the numerical model is thatof a cylindrical tank of height H and radius L. Atthe center of the base of the cylinder we model asmall cylindrical heater with radius Rheater = 9 mm

and height hheater = 3 mm. We define the coordinatez as in Figure 1 with z = 0 corresponding to the topof the heater. We will use the coordinate y to indi-cate the distance from the base of the tank so thaty= z+ hheater.

[21] We use two different sets of assumptions forgeometry, boundary conditions, and fluid properties.

[22] In the first case we compare the models directlywith the laboratory results and we mimic the labconditions as closely as possible following Vattevilleet al. [2009]. We use rigid bottom and side bound-aries and a free-slip top surface. The volume of thetank is the same as that in the laboratory experimentswith H= 165 mm and L= 85 mm. We use the mea-sured fluid properties (see Vatteville et al. [2009,Table 1] for the fluids Rhodorsil 47V500 or47V5000) and include the temperature-dependenceof viscosity. The initial fluid temperature is the sameas that in the corresponding laboratory experiments,and we prescribe the measured evolution of thetemperature in the heater as boundary condition inthe heat equation. The bottom and side boundariesare kept at constant (room) temperature. We verifiedfor several powers that the steady state velocityand thermal structure did not depend on the initialheating history, since we obtained the same resultsby applying either a constant heat power or a con-stant temperature difference to the heater [Vatteville,2009].

[23] In the second case, where we investigate thesteady state plume structure for large domains, weuse a large range of values for the height and radiusof the tank. We model the fluid in the tank with the

Laser 532 μm

Optic fiber

Cylindricallense

CCD camera

a) b)

Figure 2. Experimental set-up: (a) laboratory experiments and (b) mesh used for the axisymmetric numerical simu-lations. The numerical resolution is finest around the heater with an average nodal point spacing of 0.2mm.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

162

Page 6: Dynamics of a laminar plume in a cavity: The in uence of

same constant properties of the silicone oil Rho-dorsil 47V500 but assume all properties (includingviscosity) to be independent of temperature. Thefluid initially is at uniform temperature Troom =20�C. At time t= 0 we set the heater instanta-neously to a fixed temperature Theater = 80�C. Wemodel the heater, the base of the model and the sideboundaries as no-slip boundaries. The top of themodel is either modeled as a free-slip boundary(vertical velocity and tangential stress are bothzero) or as an “open” inflow-outflow boundary(normal stress and tangential stress are both zero).The smallest box corresponds to the size of thelaboratory tank used in Vatteville et al. [2009] withheight H= 0.165 m and width L= 0.085 m. Weincrease the height up to H= 1.28 m and the widthto L = 4.8 m. Since we compute the temperature andvelocity fields, we can determine the power Q bynumerical integration of (2). For this 60�C tem-perature difference, the computed power variesfrom 2.44 to 2.47W (1.4%), which shows that thesize of the box has only a very minor influence onthe power of the plume.

2.2.2. Governing Equationsand Numerical Solution

[24] We solve the equations of conservation ofmass, momentum, and energy for an incompress-ible fluid at infinite Pr:

r u ¼ 0 (11)

�rP þ r s ¼ ra T � Troomð Þg (12)

rCp@T

@tþ u rT

� �¼ kr2T (13)

where s = �(r u + r uT) is the stress tensor, � isdynamic viscosity, u is velocity, T is temperature,and g is gravity (|g| = 9.8 m/s2). The nominal valuesfor the fluid properties are r = 973 kg/m3, Cp =1460 J/kg/K, a = 9.4 10� 4 1/K, k = 0.156 W/mK, and � = 0.4865 Pa s.

2.2.3. Numerical Solution

[25] The numerical solution of the governingequations is described in Vatteville et al. [2009].Here we summarize the main points. The equationsare solved using the finite element package Sepran[Cuvelier et al., 1986] in a two-dimensional axi-symmetric geometry. We use a penalty functionmethod for the Stokes and mass conservationequations and streamline upwinding for the heatequation. The finite element mesh is strongly refinednear the heater (with a smallest grid spacing of

0.2mm; Figure 2b). The grid spacing remains highto a distance of 10 cm from the heater beyond whichit gradually decreases. At the symmetry axis the gridspacing increases to 0.5 cm (for the smallest heightsused) or 2 cm (for the largest heights used) at the topof the box. Near the edges of the cylinder, wheretemperature is uniform and velocities are very small,the grid spacing increases to 25 cm for the largestboxes considered. This flexible gridding allows forthe accurate yet efficient solution of the governingequations.

2.2.4. Dimensionless Parameters

[26] Our system can be characterized by a set ofdimensionless parameters, which will be useful tocompare the laboratory, numerical, and mantlecases. Besides Pr, we have the Rayleigh numberthat compares the driving thermal buoyancy forcesto the resisting effects of thermal diffusion andviscous dissipation:

Ra ¼ agΔTR3

kn(14)

where R is the radius of the heater, which is a lengthscale close to the plume radius [e.g., Griffiths,1986]. Here, Ra ranges between 2 102 and7 103. Another definition of Ra, RaRB is obtainedby using the total height of the box for the lengthscale and is commonly used in classical Rayleigh-Bénard convection in which the whole lower sur-face is at constant temperature. A value RaRB ≥106 indicates that the box is much larger than thethermal boundary layers and that plume-shapedinstabilities can develop. Here RaRB is between106 (for the lab experiments with the V5000 oil)and 2 1010 (for the numerical simulations in thehighest box). Using typical mantle values one canargue that RaRB probably ranges between 106 and108.2. The aspect ratio of the box L/H ranges inthe numerical simulation from 0.066 to 29.4, andfrom 0.45 to 0.55 in the laboratory experiments.3.The aspect ratio of the height of the box to theradius of the heater (H/R ranging from 18 to 133)and the ratio of the length of the box to the radiusof the heater (L/R ranging from 8 to 533).

2.3. Plume Evolution to Steady State

[27] The development of the thermal plume isshown in Figure 3 for a typical numerical simula-tion with instantaneous heating with ΔT = 60�C,and in Figure 4 for a typical laboratory experimentwith applied constant power to the heater. The

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

163

Page 7: Dynamics of a laminar plume in a cavity: The in uence of

structure and evolution present similar features inboth cases.

[28] Since the heater has a finite size and extends afew millimeters above the bottom of the box, astrong lateral temperature gradient develops first at

the edge of the heater, and therefore that is wherethe fluid becomes first unstable, leading to depres-sion in the temperature contour lines above theheater axis in the early stages of the hot thermalboundary layer growth (Figures 3a and 3b). As theplume rises through the box the plume head grows

0.02 0 0.02

g) t = 170 s

r (m)

0.02 0 0.02

0.01

0.02

0.03

0.04

0.05

z (m

)

0.02 0 0.02 0.02 0 0.02

0.02 0 0.02

f) t = 140 s

r (m)

c) t = 60 sb) t = 50 sa) t = 40 s

0.02 0 0.02

e) t = 110 s

r (m)

0.02 0 0.020

0.05

0.1

0.15

0.2

0.25

0.3d) t = 80 s

r (m)

z (m

)

20.3

23

26

32

50

Figure 3. Plume evolution in a typical numerical simulation in an isoviscous fluid with similar properties to that ofthe silicone oil 47V500. Box has height H= 33 cm and width L= 60 cm. At time t= 0 s the heater boundary conditionis changed from Troom = 20�C to T = 80�C. The computed steady state power of the heater is 2.5W. The contourvalues (in �C) are indicated in frame e); they represent temperatures at fractions 0.005, 0.05, 0.1, 0.2, and 0.5 of thetemperature contrast between heater and room. In the initial stage the fluid becomes unstable more quickly at the edgesof the heater, leading to the minor temperature inversion above the top of the heater.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

164

Page 8: Dynamics of a laminar plume in a cavity: The in uence of

and cools, while the plume stem that followsestablishes a steady state structure relatively soonafter the plume head has passed (Figures 3e–3g,Figures 4b–4c). The velocity profiles along theplume axis register a maximum, which is firstlocated in the plume head (Figure 5), before jump-ing at the bottom of the plume stem when the plumehead has sufficiently cooled down.

2.4. Comparison Between Laboratoryand Numerical Models and Resolution Tests

[29] A systematic comparison of the laboratoryexperiments and their corresponding numericalsimulations shows a good agreement between thetwo independent techniques [Vatteville et al.,2009]. Numerical and laboratory isotherms coin-cide (Figure 4) and the evolution of the velocityfield is found to be similar (Figure 5). Due to thethickness of isotherms obtained in silicone oils, theagreement on the thermal structure seen here ismainly qualitative. However, the velocity fields aremuch more accurately known and the profiles agreeto within 5%. For the temporal evolution of thevelocity maximum Vm(t) along the plume axis, theagreement reaches 3% over the whole experiment(Figure 5b). The height of this velocity maximum isalso well recovered by the two methods: althoughthe laboratory data are noisier, the envelope of the

points with velocity within 95% of Vm(t) followsthe numerical data (Figure 5a). The good agreementbetween the numerical simulations (performedwithout taking into account the inertial terms) andthe laboratory experiments (at finite Pr) confirmsthat, for the parameter range of our study, inertialeffects have no influence on plume dynamics. Thisresult is consistent with observations alreadyobtained in the case of Rayleigh-Bénardconvection, whereby the convection characteristicsseem independent of Pr as soon as Pr> 100[Krishnamurthi, 1970a, 1970b; Schmalzl et al.,2002; Breuer et al., 2004].

[30] We perform a convergence test for the largestbox. Figure 6 shows the velocity and temperatureprofiles along the plume axis for four differentmeshes. The overlap between the results obtainedusing the nominal resolution and those on the higherresolution meshes demonstrates that the models arewell resolved in velocity and temperature.

3. Characterization of a (Near) SteadyState Thermal Plume With Free SlipTop Boundary

[31] As the plume grows the convective flow inthe box becomes increasingly box-filling. This is

3

2425

30

2425

30

24

25

30

25

t = 300 sec. t = 400 sec. t = 500 sec.

TLCs 23.1°C– 25.1°C

Figure 4. Snapshots of the temperature field at time t= 300 (a), 400 (b) and 500 (c) s in oil 47V5000 at power P=1.4W. The numerical simulation is run with the same properties of the fluid, same geometry, and same boundary con-ditions. In each frame the fuzzy background image provides the photo of the fluid where the thermochromic liquidcrystals within the temperature range from 23.1 to 25.1�C show up in darker shades. The solid lines indicate the con-tours from the corresponding numerical model at 24, 25, and 30�C. Similar to Figure 3 the temperature structure of theplume stem reaches a steady state (as indicated by the dashed line).

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

165

Page 9: Dynamics of a laminar plume in a cavity: The in uence of

illustrated for the laboratory experiment by usingtime-lapse photography (Figure 7). The motion ofthe glass spheres causes streaks demonstrating howthe fluid in the box is convecting. The sideboundaries become important in this case betweenabout 150 and 250 s while the plume is still in itsinitial stage of development. The top boundarybecomes important between 340 and 550 s, whichis long before the plume reaches the top boundary.The qualitative suggestion for the importance ofthe boundaries can be quantified by comparing

how closely the models can reproduce the WL06predictions for a semi-infinite space.

[32] The evolution of the temperature profile alongthe plume axis is shown in Figure 8 for plumes innumerical models with a free-slip top boundary.The two models shown here use the smallest andlargest model domain used in this study. The tem-perature evolution is compared to the prediction(9; shown by the symbols) for the steady stateplume temperature. Aside from the minor devia-tions in the plume head the theoretical predictionfits the plume stem temperature well during thedevelopment of the plume and the evolution towardthe steady state.

[33] This raises an important question: if the pres-ence of a top boundary is relatively unimportant inthe comparison of temperature between numericalmodels and theory, does the same hold for thevelocity profile? Or does the box filling effectshown in Figure 7 indicate a possible problem?

[34] In the remainder of this paper we will consideronly the models with constant viscosity and a

Normal Nel=27674

Coarser Nel=12342

Finer Nel=38273

Balanced Nel=34967

0.0

0.2

0.4

0.6

y (m

)

T (C)

0.000 0.001 0.002 0.003 0.004 0.005

v (m/s)

T v

20 30 40 50 60 70 80

Figure 6. Convergence test for the model in the largestbox (H= 128 cm; L= 480 cm). We show temperature andvelocity for 4 different finite element meshes (Nel is thenumber of elements. The results for the mesh employedhere is in black. Results in red show that for a finer grid(with approx. 1.5 times higher resolution). Results inblue are for a coarser grid. Results for a more balancedgrid (with the same resolution as normal near the plume,but significantly higher resolution away from the plumecompared to the normal grid) are shown in green. Theoverlap of the normal, finer and balanced grids show thatthe models in this paper are well resolved. Only at signif-icantly lower resolution we see a divergence of theresults.

20 40 60 80 100 120 140 1600

0.5

1

1.5

2

2.5

3

VE

LOC

ITY

(m

m/s

)

0

20

40

60

80

100

120

140

160

HE

IGH

T (

mm

)

position of Vmax (lab)position of Vmax (num)isoth 24 num

a)

b)

Vmax to 5% labmax dT/dz num

isoth 24 lab

TIME (sec)

Figure 5. Comparison between numerical and experi-mental data in oil V500 (P= 1.7W). Top: Evolutionthrough time of (i) the position of maximum velocityon axis (dots for lab results, smaller dots to estimate95% of that maximum, green line for numerical results)(ii) position of top of head outlined by isotherm 24�C(blue circles for lab results, green diamonds for numer-ical results), (iii) position of maximum temperaturegradient (black crosses are numerical results) . Bottom:Evolution through time of (a) maximum velocity on axis(dots for lab results, green line for numerical results)(b) velocity of uplift of top of head (green diamondsare numerical isotherm 24�C, black crosses are maxi-mum temperature gradient).

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

166

Page 10: Dynamics of a laminar plume in a cavity: The in uence of

heater that is instantaneously heated to a constanttemperature contrast of 60�.

3.1. Temperature Structurein Quasi Steady State

[35] We first will consider the temperature profilealong the plume vertical axis at t= 1000 s as suffi-ciently close to that of the steady state structure.The plume has reached the top of the box for all

heights considered (H = 16.5, 33, 65, and 128 cm).For each of these heights we consider a radius ofthe box of L = 8.5, 15, 30, 60, 120, 240, and480 cm.

[36] We determine the power Q by numerical inte-gration of (2). The power Q varies again veryslightly, between 2.435 and 2.473W (1.4%).

[37] We determine the constants A and z* for eachsteady state temperature profile at the plume axis by

a) b)

d)c)

e)

23.1-25.1°C

29-31°C

Figure 7. Time-lapse photographs showing the evolution of a laminar plume in oil V5000 with an electric power1.17W. The heater is outlined in red and provides the length scale (18mm). (a) t = 150 s; (b) t= 250 s; (c) t = 340 s;(d) t= 550 s; (e) t= 720 s. As time increases the instability grows from an isolated state to a box filling plume. Red starsshow the position of the rotation center over time. The black arrows show the two temperature ranges were the ther-mochromic liquid crystals become bright. The time-lapse pictures allow to follow the particle trajectories.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

167

Page 11: Dynamics of a laminar plume in a cavity: The in uence of

a nonlinear fit to (9). In this fit we exclude thelowermost (where T> 70�C and topmost (wherez> 0.9H) parts of the temperature profile.

[38] Over the range of assumed steady state pro-files, we find that the computed values z* and Cvary by only a few percent due to geometricalvariations (see Appendix A). z*/Rheater is around2.65 and C is around 0.088. We note that our valuesfor C are systematically higher (by about 9%) thanthe theoretical prediction Cw ¼ 1

4p � 0:07958.

[39] We show in Appendix A that the models alsosatisfy the WL06 prediction for the width of thethermal plume.

3.2. Velocity Profiles Along the Plume Axis

[40] We also compare the axial velocity profile withfree-slip top boundary V to that of the predictionVWL (10), where we determine V0, and z0 from thenumerically integrated power Q and z* from thetemperature fit. We plot V (in red) and VWL (inblack) in the left-hand column of Figure 9 for var-iable H and L. The calculated velocity profile does

not recover the small velocity increase predicted bythe theoretical model. Instead, the velocity tends tostay quasi-constant, or to decrease with increasingheight as the box becomes thinner.

[41] Cv is found where the theoretical prediction(10) and the model solution just touch (Figure 9).Cv ranges between 0.284 and 0.317 (see Appendix A).This leads to slightly variable predictions for themaximum of VWL reached in each experiment(Vmax) that tend for each tank height to reach a pla-teau value as L/H ≥ 2 (Figure 10a). This valueincreases with increasing fluid thickness (Figures 9,10a, and 10b).

[42] The right-hand column of Figure 9 shows theratio V/VWL. The modeled axial velocity fits thetheoretical prediction only for a relatively shortdepth interval (z ≤ zm�H/4) and there is a strongdeviation in the top half of the domain. However,since zm increases with fluid thickness, this leads toan increase of VWL predicted by equation (10), andtherefore explain the increase in Vmax/V0 plateauvalues observed in Figures 10a and 10b, as Hincreases.

(T-Troom)/(Tmax-Troom)

z/H

0.0

0.2

0.4

0.6

0.8

0.0 0.2 0.4 0.6 0.8 1.0

30 s

50 s

70 s

100 s

300 s

200 s

H=1.28 m L=4.8 m

H=0.165 m L=0.085 m

Figure 8. Evolution of a typical plume shown by the non-dimensional temperature contrast plotted at non-dimensional height. The evolution is shown for moderate power in both the smallest box (in blue) and largest box(in red) that we considered in this study. For the smallest box (H= 16.5 cm; L= 8.5 cm) we plot the plume tem-perature every 10 s; for the largest box (H = 4.8m; L= 1.28m) we plot the plume temperature every 100 s. Thebold line represents the plume axis temperature at t= 1000 s which we take here to be sufficiently close to steadystate. The theoretical fit to (9) for both examples is shown by the symbols. For the fit we use the temperature profile at1000 s down from z = 0.9H (top of frame) and up to a temperature of 50�C (thin vertical line). For the largest box wefind z* =�0.02380 and C= 0.08779. The root-mean-square (rms) misfit is 0.13�C with a maximum point-wise misfitof 4%. For the smallest box we find z* =�0.02420 and C= 0.08872. The rms and maximum point-wise misfits are0.10�C and 0.012% resp.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

168

Page 12: Dynamics of a laminar plume in a cavity: The in uence of

3.3. Global Structure of the Flow

[43] From the numerical or laboratory velocityfields we calculated the stream function Ψ from ourvelocity fields using [Tanny and Shlien, 1985]:

Ψ r; zð Þ ¼Z r

0rVz r; zð Þdr (15)

[44] As expected from the laboratory images(Figure 7), the stream function exhibits a maximum(Figure 11), the position of which coincides withthe center of rotation of Figure 7.

[45] Investigation of the streamfunction extremaconfirms that the side boundaries become unim-portant for aspect ratios L/H ≥ 2 (see Appendix A).

[46] Close examination of the evolution of thestream function through time shows the develop-ment of weak cold instabilities from the top ofthe box where the temperature is kept at 20�C(Figure 12). Given the axisymmetric nature of theplume, it means that the hot plume impact, and thesubsequent spreading of its hot head below the topsurface, triggers the descent of a ring of coldmaterial around its rim. This has previously beenreported for classical Rayleigh-Bénard convection[Olson et al., 1988; Androvandi et al., 2011].

[47] This phenomenon causes a long term drift ofthe temperature. The stream function for earliertimes is symmetric but then settles to an asymmet-ric pattern with slightly lower convective vigor andlower zmax. It appears that for the small box ofFigure 12, the steady pattern is reached when t ≥4000 s. This long term drift in the solution makes itclear that all of the results shown above are only ina quasi steady state. The models could be integratedfor significantly longer times but that would imposeto take into account the influence of the thermalboundary condition at the top surface. Althoughthis is potentially relevant for applications ofplumes rising in the silicate mantles of the terres-trial planets, it makes a direct comparison withexisting models and theories difficult. It is moreoptimal to render this thermal top boundary effectless important. In the following section we willdescribe how we achieve this with a change fromthe (lab- and mantle-relevant) free slip and iso-thermal boundary condition to that of an outflowboundary, where material can freely move in andout of the top boundary, as to simulate plumeevolution in a semi-infinite fluid.

4. Evolution to a Steady State ThermalPlume With Open Top Boundary

[48] To simulate an open top boundary, we pre-scribe a top boundary condition for the Stokesequations with zero normal stress and zero tangen-tial stress. For the heat equation we assume a nat-ural boundary condition for a segment out to halfthe width of the box where we expect the flow to beout the box. For larger radii we assume a constanttemperature boundary condition so that the returnflow will bring constant temperature fluid into themodel domain. In this case the heat brought up bythe plume is advected out of the computational

0.2

0.4

0.6

0.8

y/H

V/V0

H=0.165 m

0.2 0.4 0.6 0.8 0.8 0.9 1.0

0.2

0.4

0.6

0.8

y/H

H=0.33 m

0.2

0.4

0.6

0.8

y/H

H=0.65 m

0.2

0.4

0.6

0.8

y/H

H=1.28 m

8.5

8.5

8.5

8.5

15

30

15

15

15

30

≥60

≥30

≥30

≥120 60

VWL

Vfs/VWL

V fs

L(cm)

Figure 9. Comparison of modeled axial velocity V andprediction VWL for variable H and L. In the left columnwe plot V/V0 in red and VWL/V0 in black. The ratioV/VWL is shown in red in the right column. The light greyband indicates where this ratio is >0.95. The dark greyband shows where this ratio is>0.98. The small numberson the curves show the width of the box L (in cm).

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

169

Page 13: Dynamics of a laminar plume in a cavity: The in uence of

domain. The spreading of hot fluid under the topboundary that is seen with the free-slip boundary isabsent and we do not see a long term drift due to theadjustment of this boundary. Instead, we find robuststeady state flow patterns as soon as the plume headhas been advected out of the domain.

4.1. Thermal Structure

[49] In this case we find that the computed powervaries from 2.43 to 2.55W, which represents aslightly larger range than with the free-slip topboundary (2.44–2.47W). The values for�z*/Rheater

and C are very close to those obtained for the free-slip boundary and range from 2.63 to 2.72 (3.3%)and 0.08702 to 0.0893 (2.5%), respectively.

[50] At this stage we do not understand the reasonfor the small but systematic difference between the

theoretical prediction for C and those obtained inour experiments. Using further numerical modelswe have explored whether the heater geometry isresponsible for the slight misfit, but we did not findany significant differences with a heater that had asmaller radius, or with a heater that was alignedflush with the bottom boundary. In fact, once theplume has emerged from the thermal boundarylayer (i.e., z≫ z0) the thermal and velocity structurefollow the same trends as described above, irre-spective of the details of the heat sources, such asradius or thickness.

4.2. Velocity Structure

[51] The streamfunction and centerline velocitiesobtained with the open top boundary are shown inFigures 13 and 14. The side boundary remains

0

L / H

1.0

0.2

0.4

0.6

0.8 H = 16.5 cm

33 cm65 cm

128 cm

10-2 100

L / H101 10210-1

10-2 100 101 10210-1

0

Vm

ax /

V0

Vm

ax /

V0

1.0

0.2

0.4

0.6

0.8

H = 16.5 cm

33 cm65 cm

128 cm

0

50

100

150

H (

cm)

Vmax / V0

1.00.80.7 0.9

b) convergence to WL scalinga) free-slip top boundary

c) open top boundary

8.51530

60120240

VWL

Figure 10. Maximum velocity along the plume axis compared for the two cases with free-slip and open top bound-ary. (a) Maximum velocity Vmax for the free slip configuration normalized by Batchelor velocity scale, as the functionof the domain height (indicated by H. Symbol type indicates the domain width L, identified (in cm) in the legend.(b) Vmax/V0 for the largest domain width (L= 60 through L= 240 cm) for different fluid heights. In blue the openboundary case and in black the free surface case. The red symbols indicate the real depths at which the velocitywas maximum (see Figure 9) in the free surface case with H= 240 cm. The solid black line represents Whittakerand Lister’s scaling (10) with Cv= 0.313. (c) Maximum velocity Vmax for the open top boundary configuration normal-ized by Batchelor velocity scale, as the function of the domain aspect ratio.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

170

Page 14: Dynamics of a laminar plume in a cavity: The in uence of

important for small aspect ratio (H/L ≤ 2). We notehowever an excellent agreement (to within 2%)with the theoretical prediction (WL06) for H/L ≥ 2(Figures 14, 10b, and 10c). There remains only a

small influence of the top boundary in the conduitnear the very top of the domain. This suggests thatthe change to an open boundary is sufficient torender unimportant the strong effects on velocity

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

y (m

)

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 2.1 2.2 2.3 2.4

x (m)

position of streamfunction max.

H=0.165m

H=0.33m

H=0.65m

H=1.28m

Figure 11. Streamfunction for variable H at t = 1000 s. The position of the streamfunction maximum is plotted by thediamonds. The position of this extremum tends towards x= y =H/2 for L/H ≥ 2. For clarity we plot the streamfunctionfor each height in the box with L/H � 2. the streamfunction contours are for 10 equidistant intervals to the streamfunc-tion maximum, with cmax = 8 10� 6, 3 10� 5, 1.1 10� 4, and 3.4 10� 4 for H= 16.5, 33, 65, and 128 cm, respec-tively. The blue box in the lower left corner indicates the size of the smallest box used in this study. The largest boxused in the study is twice as wide as the largest frame.

0.00

0.05

0.10

0.15

y (m

)

0.00 0.05

x (m)

100 s 500 s 1000 s 1500 s 2000 s 2500 s 3000 s 3500 s 4000 s

1000 2000 3000 4000

t (s)

22.0

22.1

22.2

22.3

22.4

avT

(C

)

a) Streamfunction evolution

b) average temperature at top

Figure 12. (a) Evolution until t= 4000 s of the streamfunction (a) in the smallest box (H= 16.5 cm, L= 8.5 cm). Afteran initial apparent steady state with symmetrical streamfunction the pattern becomes asymmetric. (b) average temper-ature in a rectangle in the top left corner of 5mm wide by 2mm high.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

171

Page 15: Dynamics of a laminar plume in a cavity: The in uence of

seen in the models with the free-slip top boundary.In this case, the velocity scale becomes independentof H and is well described by:

V zð Þ ¼ 0:313VWL f L; zð Þ (16)

where f(L,z) describes the influence of the side rigidboundaries. The proximity of the side boundariesare increasing the drag on the plume and decreasingits upwelling velocity. For a cylinder or a sphere ofradius a moving along the axis of a cylinder ofradius L, the drag increase (and therefore the veloc-ity decrease of the cylinder/sphere) is a function ofthe ratio a/L [Happel and Brenner, 1973]). Now,we verified that in the open boundary case, theradius of the thermal plume is also well predictedby aT(z) = 0.480aWL(z), and so increases with z.Hence, aT/L increases with depth, and the velocityshould decrease accordingly. Figure 15 shows thevelocity reduction measured in our experimentsfor the largest fluid depth (e.g., Figure 14) comparedto the predictions for the sphere [Happel andBrenner, 1973, pp. 318] and the cylinder [Happeland Brenner, 1973, pp. 342]. We see that themagnitude of velocity reduction observed for thethermal plume is intermediate between the twotheoretical cases.

5. Discussion

[52] We can now attempt to apply this study tomantle plumes and mantle convection. In doing so,

we have to keep in mind several potential short-comings. First, there is no laterally confined heaterat the core-mantle boundary, which instead is aninterface where the temperature is uniformly hot.However, the presence of denser, compositionallyheterogeneous, material on the core-mantle bound-ary can act to modulate this uniform thermalboundary condition, forming hot “piles” on thecore-mantle boundary. Moreover, these piles couldbe enriched in radiogenic elements, which wouldincrease their heating potential (for a review seeGarnero and McNamara [2008]). Long-livedplumes have been shown to develop preferentiallyon and around those piles [Tackley, 1998; Namikiand Kurita, 1999; Davaille, 1999; Davaille et al.,2002; Jellinek and Manga, 2002; Tan et al., 2002;McNamara and Zhong, 2005]. Our small heater canbe regarded as a simplified mantle “pile.”

[53] Another concern is the nearly constant viscos-ity we consider in our study, while mantle materialhas a strongly temperature-dependent viscosity andmantle plumes are therefore often assumed to bemuch less viscous than the ambient mantle. Thenear-constant viscosity plumes considered heremay nevertheless be appropriate since it has beensuggested that hot plumes viscosity may not besignificantly lower due to the effect of grain sizevariations in the lower mantle. In extreme cases,variations of the grain size could be so large that hotplume could be more viscous than ambient mantlebecause of enhanced grain growth with increasingtemperature [e.g., Solomatov et al., 2002; Korenaga,

0.0

0.2

0.4

0.6

0.8

1.0

1.2

y (m

)

x (m)0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2

position of streamfunction max.

H=0.165

H=0.33

H=0.65

H=1.28

Figure 13. As Figure 11 but now with an open top boundary. The streamfunction contours are for 10 equidistantintervals to the streamfunction maximum, with cmax = 2.3 10� 5, 8.5 10� 5, 3.2 10� 4, and 1.2 10� 3 forH= 16.5, 33, 65, and 128 cm, respectively.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

172

Page 16: Dynamics of a laminar plume in a cavity: The in uence of

2005]. Besides, although seismic tomography is stillstruggling to image plumes, recent results of finitefrequency tomography are showing slow anomaliesof 200 to 400 km width under certain hotspots [e.g.,Montelli et al., 2004], which would be more com-patible with plumes as viscous or more viscous thanthe ambient mantle. Therefore, our simplified con-stant viscosity study can be of interest for mantleplumes.

[54] Given the mantle conditions, we expect plumeswith Rayleigh numbers RaRB around 106� 108,

which is the situation that we encountered in ourshortest boxes (cf. section 2). In that case, our studysuggests that care should be taken when applying theresults for Stokes flow in a half-space to the confinedgeometries of the Earth’s mantle, particularly whenvelocities are considered. The application of theo-retical, laboratory, and numerical simulations shouldbe significantly more robust for temperature in theplumes.

[55] The strong importance of the boundaries hasimportant applications to the interpretation of geo-dynamical results. For example, it is common tostudy the subduction of oceanic lithosphere bymodeling the sinking of dense or cold boundarylayers into a viscous medium. Due to restrictions inlaboratory experiments (where the boundaries areessential) or due to computational limitations thegeometries employed often have rigid or free-slipboundaries that do not allow the fluid to enter orleave through the boundaries. A number of papershave highlighted that the sinking of a slab caninduce a strong toroidal component around the slab[see, e.g., Buttles and Olson, 1998; Piromallo et al.,2006; Jadamec and Billen, 2010]. A study thatemployed outflow boundary conditions (similar tothose in this paper) concluded that the toroidal flowwas significantly reduced [Kneller and van Keken,2008]. Most likely this is due to the fact that with

H=0.165 m

H=0.33 m

H=0.65 m

H=1.28 m

V/V0 Vfs/ VWL

0.2 0.4 0.6 0.8 0.8 0.9 1.0

0.2

0.4

0.6

0.8

y/H

0.2

0.4

0.6

0.8

y/H

0.2

0.4

0.6

0.8

y/H

0.2

0.4

0.6

0.8

y/H

8.515

30

≥240

60

120

8.5

30

15

≥60

8.515

≥30

L(cm)

8.5

30

15

60≥120

Figure 14. As Figure 9 but now for open top boundary.The plume center line velocity for large aspect ratioshows a significantly better match with the theoreticalprediction (WL06).

0 0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

V / VWL

z (m

)

1.2

cylinder (theory)sphere (theory)

model H=128 cm

L=8.5 cm

L=240 cm

L=30 cm

Figure 15. Comparison between V/VWL observed inthe open boundary simulations (solid lines) and thedrag reduction predicted for a sphere (short dash lines[Happel and Brenner, 1973, pp. 318]) or for a cylinder(dot-dashed lines [Happel and Brenner, 1973, pp. 342])for H= 128 cm.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

173

Page 17: Dynamics of a laminar plume in a cavity: The in uence of

open boundaries the fluid below a slab can escape,whereas in a finite box it can only move around theslab. The toroidal flow is then effectively generatedby the non-outflow boundary. In the Earth, thedynamics is more complicated. While the free-slipor rigid boundary conditions confine and influencethe flow, the open boundary conditions do notrepresent the fact that the Earth’s mantle is confinedto a spherical geometry and that slabs and plumesinfluence each other over long distances. The use offull spherical geometry is still computationallyexpensive, particularly if one wishes to accuratelyresolve mantle dynamics at the present-dayconvective vigor of the Earth [Stadler et al., 2010].The use of a cylindrical geometry [van Keken,2001; Gérault et al., 2012] or other two-dimensional approximations to the spherical Earth[Hernlund and Tackley, 2008], where the modelstake into account at least in part the curvature of theEarth and the connectedness of the Earth mantle,are therefore still attractive options.

6. Conclusions

[56] We have performed a series of well-resolvednumerical simulations of a steady state plume risingfrom a localized heat source in a cavity. For modelswith a free-slip top boundary we see significantdifferences with theoretical predictions for thecenterline velocity of plumes rising from the same

heater in an infinite half-space. The effect of theside boundaries is negligible for aspect ratios of twoor larger, but the effect of the free-slip top boundaryconditions affects the plume solution for a signifi-cant part of its height. We can better approximatethese half-space solutions by modifying the freeslip boundary to a stress-free boundary, whichsimulates an open boundary condition. In this casewe find that for significantly high aspect ratio thetheoretical prediction for velocity is matched towithin 2%.

[57] The differences between the theoretical pre-dictions for temperature and those obtained fromthe numerical simulations are significantly lesssensitive to the size of the domain and the choice oftop boundary conditions. In fact, the parametersthat determine the shape of the centerline tempera-ture profile are nearly constant across the range ofmodels.

Appendix A

[58] Here we summarize a few technical points andprovide background information on some of theobservations that we provided in the main paper.

[59] The parameters that govern the fit of the pre-dicted centerline temperature to the theory areprovided in (Figure A1). In general we find that

1.5

2.0

2.5

3.0

−z*

/R

L/H

free slip top

0.080

0.085

0.090

0.095

0.100

C

0.1 0.2 0.5 1 2 5 10 20 0.1 0.2 0.5 1 2 5 10 20

L/H

C=Cw

2.0

2.5

3.0

3.5

−z*

/R

open top

0.080

0.085

0.090

0.095

0.100

C

C=Cw

H=0.165 m H=0.33 m H=0.65 m H=1.28 m

Figure A1. The parameters z* (left) and C (right) as determined from the centerline temperature profile by a nonlin-ear fit to equation 9. While the size of the box and the choice of top boundary conditions have a significant effect onthe velocity, the temperature in the plume is principally insensitive to these geometrical considerations.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

174

Page 18: Dynamics of a laminar plume in a cavity: The in uence of

these values are similar when L/H ≥ 2 but changeslightly as a function of H. This suggests thatthe top boundary has the dominant influence onthe plume structure and that the side boundarybecomes unimportant for aspect ratios larger thantwo. The range of �z*/Rheater, and C are 2.639–

2.711 (2.7%) and 0.08745–0.0889 (1.7%),respectively (Figure A3).

[60] Following Whittaker and Lister [2006a], wedefine the width of the thermal plume as the vari-ance of the temperature distribution:

aT zð Þ2 ¼

Z L

0ΔT r; zð Þr3dr

Z L

0ΔT r; zð Þrdr

: (A1)

[61] Figure A2 shows that the plume thermal radiuscalculated with (A1) is well predicted by WL6model (7) with aT = 0.480aWL for H/L ≥ 0.5.

[62] The values for Cv that are determined by takingthe theoretical prediction (10) where it just touchesto the model solution is shown in Figure 9.

[63] For given domain dimensions (L, H), the valueof the stream function maximum Ψmax is propor-tional to V0 (Figure A4). As the domain aspect ratioL/H increases, we find two very well defined trends(Figure A5).

[64] For L/H ≥ 2, the stream function maximumscales as Ψmax = (0.05� 0.005)V0H

2 and its posi-tion becomes close to xmax’ zmax’H/2 (xmax =(0.53� 0.04)H and zmax = (0.51� 0.04)H). This isthe case where the side boundaries are far enoughfrom the plume for both the stream function maxi-mum value and its position to be independent of thebox radius L.

[65] On the other hand, when L/H ≤ 2, the sideboundaries becomes increasingly important for theplume dynamics and we have (for L/H ≤ 0.5)

10 20 30 400

1

aT (mm)

Non

-dim

ensi

onal

hei

ght z

/H

0 0.5 1

aT / aWL

0.2

0.4

0.6

0.8

8.5

15

3060

120

a) b)

Figure A2. a) Radius of the thermal anomaly calcu-lated with (1) as a function of depth for H = 65 cm forvarious choice of L. b) Radius of the thermal anomalyaT divided the WL06 scale aWL. For L > 30 cm andaway from the top and bottom boundaries, the two radiiare proportional with aT = 0.480aWL. The width of themodel L is indicated (in cm) in frame b). Cyan curve isfor L = 8.5 cm, black for L = 15 cm, green for L = 30 cm,blue for L = 60 cm, red for L = 120 cm.

0.26

0.28

0.30

0.32

0.34

Cv

L / H

10-2 10-1 100 101 102

open symbols: free-slip top boundaryblue symbols: open top boundary

8.51530

60120240

Figure A3. Constant Cv in the velocity scaling (10)determined for the free slip (black symbols) and the openboundary (blue symbols) cases. Domain width L is indi-cated (in cm) in the legend for the different symbols.

100

101

102

10

10

10

V0 . R /

max

Figure A4. Maximum value of the stream function as afunction of the Batchelor velocity scale V0 for the labo-ratory experiments with oils V5000 and oils V500.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

175

Page 19: Dynamics of a laminar plume in a cavity: The in uence of

xmax= (0.54� 0.05)L. The values for zmax are muchmore scattered (Figure A5), especially for small boxes.

Acknowledgments

[66] This work was supported by the French ANRs “BeGDY”(ANR-06-BLAN-0109) and “PTECTO” (ANR-09-BLAN-0142),the program DyETI from INSU/CNRS, and by the US NationalScience Foundation under grant number EAR-0855487. Themanuscript was improved thanks to the comments of Norm Sleep,Shijie Zhong and Paul Tackley.

References

Androvandi, S., A. Davaille, A. Limare, A. Fouquier, andC. Marais (2011), At least three scales of convection in a

mantle with strongly temperature-dependent viscosity, Phys.Earth Planet. Int., 188, 132–141.

Batchelor, G. K. (1954), An Introduction to Fluid Dynamics,Cambridge University Press, Cambridge, UK.

Bercovici, D., and J. Mahoney (1994), Double flood basalts andplume head separation at the 660-kilometer discontinuity,Science, 266, 1367–1369.

Blankenbach, B., et al. (1989), A benchmark comparison formantle convection codes, Geophys. J. Int., 98, 23–38.

Breuer, M., S. Wessling, J. Schmalzl, and U. Hansen (2004),Effect of inertia in Rayleigh-Bénard convection, Phys. Rev.E., 69, 026302.

Buttles, J., and P. Olson (1998), A laboratory model ofsubduction anisotropy, Earth Planet. Sci. Lett., 164,245–262.

Chay, A., and D. J. Shlien (1986), Scalar field measurements ofa laminar starting plume cap using digital processing ofinterferograms. Phys. Fluids, 29, 2358–2366.

Couliette, D. L., and D. E. Loper (1995), Experimental, nu-merical and analytical models of mantle starting plumes,Phys. Earth Planet. Inter., 92, 143–167.

10 10 100

101

1010

100

101

zm

ax /

H

L / H

c)

max

/ V

0.H

2

10 10 100

101

1010

10

10

10

100

L / H L / H

x max

/ H

slope 1

10 10 100

101

1010

10

100

101

a) b)

8.51530

60120240

open: free-slip top

blue: free-slip top

red: laboratory

Figure A5. Characteristics of the stream function maximum as a function of the domain aspect ratio. In black, thenumerical experiments with a free surface, in blue the ones with the top open boundary, in red the laboratory experi-ments for temperature differences between 21 and 54�C. Symbols indicate the width of the models (indicated in cm inthe legend in frame a). a) non-dimensional streamfunction maximum, b) non-dimensional horizontal position of max-imum, c) non-dimensional height.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

176

Page 20: Dynamics of a laminar plume in a cavity: The in uence of

Cuvelier, C., A. Segal and A. A. van Steenhoven (1986), FiniteElement Methods and the Navier-Stokes Equations, Reidel,Dordrecht, The Netherlands.

Davaille A. (1999), Simultaneous generation of hotspots andsuperswells by convection in a heterogeneous planetarymantle, Nature, 402, 756–760.

Davaille A., F. Girard, and M. Le Bars (2002), How to anchorplumes in a convecting mantle? Earth Plan. Sci. Lett., 203,621–634.

Davaille, A., and A. Limare (2004), Laboratory studies ofmantle convection, in: Mantle Dynamics, edited by D. Ber-covici, pp. 89‐165, vol. 7, Treatise of Geophysics, edited byG. Schubert, Elsevier, The Netherlands.

Davaille, A., A. Limare, F. Touitou, I. Kumagai, and J. Vatteville(2011), Anatomy of a laminar starting thermal plume at highPrandtl number, Exp. Fluids, 285–300.

Farnetani, C. G., and H. Samuel (2005), Beyond the thermalplume paradigm, Geophys. Res. Lett., 32, 07311, doi:10.1029/2005GL022360.

Fujii, T. (1962), Theory of the steady laminar natural convec-tion above a horizontal line heat source and a point heatsource, Int. J. Heat Mass Transf., 6, 597–606.

Garnero, E., and A. K. McNamara (2008), Structure anddynamics of the Earth’s lower mantle, Science, 320, 626–628.

Gérault, M., T. W. Becker, B. J. P. Kaus, C. Faccenna,L. Moresi, and L. Husson (2012), The role of slabs andoceanic plate geometry in the net rotation of the lithosphere,trench motions, and slab return flow, Geochem. Geophys.Geosyst., Q04001.

Griffiths, R. W. (1986), Thermals in extremely viscous fluids,including the effects of temperature dependent viscosity,J. Fluid Mech., 166, 115–138.

Griffiths, G. W., and I. H. Campbell (1990), Stirring andstructure in mantle starting plumes, Earth Planet. Sci. Lett.,99, 66‐78.

Happel, J., and H. Brenner (1973), Low Reynolds NumberHydrodynamics, Noordhoff, Leiden, The Netherlands.

Hernlund, J., and P. J. Tackley (2008), Modeling mantle con-vection in the spherical annulus, Phys. Earth Planet. Inter.,171, 48–54.

Hwang, Y. K., J. Ritsema, P. E. van Keken, S. Goes, andE. Styles (2011), Wavefront healing renders deep plumesseismically invisible, Geophys. J. Inter., 187, 273–277,doi:10.1111/j.1365-246X.2011.05173.x.

Ito, G., and P. E. van Keken (2007), Hotspots and meltinganomalies, in: Mantle Dynamics, edited by D. Bercovici,pp. 371–345, Vol. 7, Treatise on Geophysics, edited byG. Schubert, Elsevier, The Netherlands.

Jadamec, M. A., and M. I. Billen (2010), Reconciling surfaceplate motions with rapid three-dimensional mantle flowaround a slab edge, Nature, 456, 338–339.

Jellinek, A. M., and M. Manga (2002), The influence of achemical boundary layer on the fixity, spacing and lifetime ofmantle plumes, Nature, 41, 760–763.

Kaminski, E., and C. Jaupart (2003), Laminar starting plumes inhigh-Prandtl-number fluids, J. Fluid Mech., 478, 287–298.

King, S. D., and J. Ritsema (2000), African hotspot volcanism:small-scale convection in the upper mantle beneath cratons,Science, 290, 1137–1140.

Kneller, E. A., and P. E. van Keken (2008), Effect of three-dimensional slab geometry on deformation in the mantlewedge: implications for shear wave anisotropy, Geochem.Geophys. Geosyst, 9, Q01003.

Korenaga, J. (2005), Firm mantle plumes and the nature of thecore-mantle boundary region. Earth Planet, Sci. Lett., 232,29–37.

Krishnamurthi, R. (1970a), On transition to turbulent convec-tion. 1. Transition from 2-dimensional to 3-dimensional flow,J. Fluid Mech., 42, 295–307.

Krishnamurthi, R. (1970b), On transition to turbulent convec-tion. 2. Transition to time-dependent flow, J. Fluid Mech.,42, 309–320.

Kumagai, I. (2002), On the anatomy of mantle plumes: effectof the viscosity ratio on entrainment and stirring. EarthPlanet. Sci. Lett., 198, 211–224.

Kumagai, I., A. Davaille, K. Kurita, and E. Stutzmann (2008),Mantle plumes: thin, fat, successful, or failing? Constraints toexplain hot spot volcanism through time and space, Geophys.Res. Lett., 35, L16301, doi:10.1029/2008GL035079.

Laudenbach, N., and U. R. Christensen (2001), An opticalmethod for measuring temperature in laboratory models ofmantle plumes. Geophys. J. Int., 145, 528–534.

Lin, S.-C., and P. E. van Keken (2005), Multiple volcanicepisodes of flood basalts caused by thermochemical plumes,Nature, 436, 250–252, doi:10.1038/nature03697.

McNamara, A. K., and S. Zhong (2005), Thermochemicalstructures beneath Africa and the Pacific Ocean, Nature, 437,1136–1139.

Montelli, R., G. Nolet, F. Dahlen, G. Masters, E. R. Engdahl,and S.-H. Hung (2004), Finite-frequency tomography revealsa variety of plumes in the mantle, Science, 303, 338–343.

Morgan, W. J. (1971), Convection plumes in the lower mantle,Nature, 230, 42–43.

Moses, E., G. Zocchi, and A. Libchaber (1993), An experimentalstudy of laminar plumes, J. Fluid Mech., 251, 581–601.

Namiki, A., and K. Kurita (1999), The influence of boundaryheterogeneity in experimental models of mantle convection,Geophys. Res. Lett., 26, 1929–1932.

Olson, P., and H. Singer (1985), Creeping plumes, J. FluidMech., 158, 511–531.

Olson, P., G. Schubert, C. Anderson, and P. Goldman (1988),Plume formation and lithosphere erosion: a comparison oflaboratory and numerical experiments, J. Geophys. Res., 93,15,065–15,084.

Olson, P., G. Schubert, and C. Anderson (1993), Structure ofaxisymmetric plumes, J. Geophys. Res., 98, 6829–6844.

Piromallo, C., T. W. Becker, F. Funiciello, and C. Faccenna(2006), Three-dimensional instantaneous mantle flow inducedby subduction, Geophys. Res. Lett., 33, L08304.

Ribe, N., A. Davaille, and U. Christensen (2006), Fluidmechanics of mantle plumes, in: Mantle Plumes: A Multi-discipinary Approach, edited by J. Ritter and U. Christensen,pp. 1–47, Springer, Heidelberg, Germany.

Rogers, M. C., and S. W. Morris (2009), Natural versus forcedconvection in laminar starting plumes, Phys. Fluids, 21,083601.

Schmalzl, J., M. Breuer and U. Hansen (2002), The influenceof the Prandtl number on the style of vigorous thermal con-vection, Geophys. Astrophys. Fluid Dyn., 96, 381–403.

Schmeling, H., A. Y. Babeyko, A. Enns, C. Faccenna,F. Funiciello, T. Gerya, G. J. Golabek, S. Grigull, B. J. P.Kaus, G.Morra, S.M. Schmalholz, and J. vanHunen (2008), Abenchmark comparison of spontaneous subduction models -towards a free surface,Phys. Earth Planet. Inter., 171, 198–223.

Shlien, D. J. (1976), Some laminar thermal and plume experi-ments, Phys. Fluids, 19, 1089‐1098.

Shlien, D. J. and R. L. Boxman (1979), Temperature-fieldmeasurement of an axisymmetric laminar plume, Phys.Fluids, 22, 631–634.

Shlien, D. J. and R. L. Boxman (1981), Laminar starting plumetemperature-field measurement, Int. J. Heat Mass Transf.,24, 919–931.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

177

Page 21: Dynamics of a laminar plume in a cavity: The in uence of

Solomatov, V. S., R. El-Khonzondar, and V. Tikare (2002),Grain size in the lower mantle: constraints from numericalmodeling of grain growth in two-phase systems, Phys. EarthPlanet. Inter., 129, 265‐282.

Stadler, G., M. Gurnis, C. Burstedde, L. C. Wilcox, L. Alisic,and O. Ghattas (2010), Dynamics of plate tectonics and mantleflow: from local to global scales, Science, 329, 1033–1038.

Tackley, P. J. (1998), Three-dimensional simulations of mantleconvection with a thermo-chemical basal boundary layer: D,in The Core-Mantle Boundary Region, Geodynamics Series,vol. 28, edited by M. Gurnis, M. E. Wyssession, E. Knittle,and B. A. Buffett, pp. 231‐253, American GeophysicalUnion, Washington, D.C.

Tan, E., M. Gurnis, and L. J. Han (2002), Slabs in the lowermantle and their modulation of plume formation, Geochem.Geophys. Geosyst., 3, 1067.

Tanny, J., and D. J. Shlien (1985), Velocity field measurementsof a laminar starting plume, Phys. Fluids, 28, 1027–1032.

Thompson, P. F., and P. J. Tackley (1998), Generation ofmega- plumes from the core-mantle boundary in a com-pressible mantle with temperature-dependent viscosity.Geophys. Res. Lett., 25, 1999–2002.

Turner, J. S. (1962), The starting plume in neutral surround-ings, J. Fluid Mech., 13, 356–368.

van Keken, P. E. (1997), Evolution of starting mantle plumes: acomparison between numerical and laboratory models, EarthPlanet. Sci. Lett., 1–14.

van Keken, P. E. (2001), Cylindrical scaling for dynamicalcooling models of the Earth, Phys. Earth Planet. Inter., 124,119–130.

Vasquez, P. A., A. T. Perez, and A. Castellanos (1996), Ther-mal and electrohydrodynamics plumes. A comparative study.Phys. Fluids, 8, 2091–2096.

Vatteville, J., P. E. van Keken, A. Limare, and A. Davaille (2009),Starting laminar plumes: comparison of laboratory and numeri-cal modeling, Geochem. Geophys. Geosys., 10, Q12013.

Vatteville, J. (2009), Dynamique des panaches thermiqueslaminaires: application aux panaches mantelliques, Ph.D.Thesis, Institut de Physique du Globe, Paris, France.

Whitehead, J. A., and D. S. Luther (1975), Dynamics of laboratorydiapir and plume models, J. Geophys. Res., 80, 705–717.

Whittaker, R. J., and J. R. Lister (2006a), Steady axisymmetriccreeping plumes above a planar boundary. Part 1. A pointsource, J. Fluid Mech., 567, 361–378.

Whittaker, R. J., and J. R. Lister (2006b), Steady axisymmetriccreeping plumes above a planar boundary. Part 2. A distrib-uted source, J. Fluid Mech., 567, 379–397.

Wolfe, C. J., S. C. Solomon, G. Laske, J. A. Collins, R. S.Detrick, J. A. Orcutt, D. Bercovici, and E. H. Hauri (2009),Mantle shear-wave velocity structure beneath the Hawaiianhot spot, Science, 326, 1388–1390.

Worster, M. G. (1986), The axisymmetric laminar plume:asymptotic solution for large Prandtl number, Stud. Appl.Math, 75, 139–152.

GeochemistryGeophysicsGeosystemsG3G3 VAN KEKEN ET AL.: STEADY STATE PLUME STRUCTURE 10.1029/2012GC004383

178