Sunteți pe pagina 1din 19

Physics of the Earth and Planetary Interiors 109 1998.

179197

The Gruneisen parameter for iron at outer core conditions and the resulting conductive heat and power in the core
Orson L. Anderson
) Department of Earth and Space Sciences and Institute of Geophysics and Planetary Physics, Uniersity of California, Los Angeles, Los Angeles, CA 90095-1567, USA Received 6 April 1998; accepted 11 August 1998

Abstract The aim of this paper is to find the conductive power in the outer core. Before the heat conduction equation can be usefully applied, however, a careful examination of the Gruneisen parameter, g , and the thermal conductivity, k c , of the core is required. The focus of this paper is on these two parameters at outer core conditions, using primarily experimental data and one theoretical evaluation of g . The melting gm of the core is found to be 1.3, the adiabatic gad , slightly less, and k c is found to be 44 W my1 Ky1 in the upper limit and 28 W my1 Ky1 in the lower limit. To obtain the cores temperature, Tm , from Tm of pure iron at core conditions., the freezing point depression is found. This requires assumptions about the impurities in the iron core. Recent experimental data greatly restrict the type and concentrations of these impurities. Several allowable combinations are found, however; all of these lead to a DT of ; 1000 K. At the core side of the core-mantle boundary CMB., the adiabatic temperature is 3900 K and dTrd r is 0.9 K my1, resulting in a conductive power of 4.4 " 1 TW 1 TW s 10 12 Watt. flowing out of the core along the adiabatic gradient. New experimental results on the thermal conductivity of mantle minerals and new theoretical insights into the heat flow in mantle plumes support the case that the conductive power along the cores adiabatic gradient passing to the upper mantle can be transported away by both conduction and convection in the upper mantle. q 1998 Elsevier Science B.V. All rights reserved.
Keywords: Gruneisen parameter; Iron; Core conduction; Mantle conduction

1. Brief description of the model This paper examines the conductive power flowing from the core and especially the parameters found in the thermal gradient at the core-mantle boundary CMB. and the thermal conductivity. Attention is focused on finding the values of the parameters involved in calculating the conductive heat,

) Corresponding author. Fax.: q1-310-206-3051; E-mail: olanderson@adam.igpp.ucla.edu

especially the Gruneisen parameter and the thermal conductivity. The values of these parameters are found to be similar to those by Braginsky and Roberts 1995., but different from those found in most recent reports. I disallow the cores thermal gradient from being subadiabatic because I find the conductive power from the core to be 3.4 to 5.4 TW 1 TW s 10 12 Watt., which is within the range of power leaving the CMB on the mantle side by conduction and convection. According to this model, any excess power, which will be convective, will be returned to

0031-9201r98r$ - see front matter q 1998 Elsevier Science B.V. All rights reserved. PII: S 0 0 3 1 - 9 2 0 1 9 8 . 0 0 1 2 3 - X

180

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

the core by the mechanism of compositional convection of Loper 1978a,b.. In this model, there is no attempt to explore the physics of the cores convective power or to find all the sources that comprise the cores conductive power. I do, however, estimate the power generated by the crystallization of the inner core boundary since the parameters for this power source have been determined.

2. Parameters needed for the heat flow analysis The heat conducted from the core to the mantle is given by: qc s yk c

/
Er

ET

,
S

1.
Fig. 1. Gruneisen parameter, g , vs. density for iron. The top scale, for P in GPa, shows various starting values of P s 0 for the bcc phase, the hcp phase, and liquid. The main focus is the gs e . vs. r curve starting at G, as deduced from Mao et al. 1990. EoS measurements, extending along the Hugoniot F. from Brown and McQueen 1986. and terminating at melting at the edge of the solid phase D. calculated by Stacey 1995.. Other lines represent g l vs. r at low pressure and gs a . E. at low pressure measured by Boehler and Ramakrishnan 1980.. Note that gm for the core is given as 1.3"0.3, which coincides with the solid phase D by X Stacey. The curve gm for the core, - - -, is coincident with D and Y D . The liquid for the isentrope, shown by P , is slightly lower.

where k c is the thermal conductivity of the core and r is its radiusthe radial distance from the center of Earth to the point considered. The outer core is known to be liquid and is ordinarily assumed to be under adiabatic compression, so one can obtain the slope of the temperature profile by means of the adiabatic equation:

/
Er

ET

s
S

ygg 1T

2.

where g 1 is the Gruneisen parameter of the liquid state, g is gravity, and f is the seismic parameter. The departure from adiabaticity is considered to be small. Although the values of g and f at any radius can be readily obtained from seismological tables, evaluation of the other parameters requires special effort. Values of k c and ETrE r . in Eq. 1. must be found. Before ETrE r . can be found, g l and T r . in Eq. 2. must be evaluated. T r . in turn depends on gm , the gamma of melting. Thus, to begin an evaluation of the power, Qc , one must first understand the Gruneisen parameter in the various states and phases of iron.

3. g in the solid, liquid, and melting states of iron Fig. 1 shows the theoretical and experimental values of g for pure iron in its various phases see

phase diagram of iron, Fig. 2.. We now review the literature data on g as plotted in Fig. 1. The line for values of g l has Points A, B, and C Stevenson, 1981; Verhoogen, 1980; Chen and Ahrens, 1997, respectively. marked as triangles. This line descends rapidly with pressure, P . Point E, showing g in the solid state gs ., represents the a-bcc. iron datum at P s 0. The line extending from Point E, called gs a ., represents measured values at higher densities see figure caption for data references.. The line emanating from E represents the measurements for bcc iron: g 0 s 1.66 and q s 0.6 in the equation:

g
s

g0

/
r

r0

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

181

X are listed K T 0 and V0 used in the calculation of K 0 in Table 1:

X K0 s4q

~P
V V0

/ /
3 K T0 V0 4 3
5 y1 3

y1

y1

/ /
V0

y1

.
4.

X Table 1 shows that the average value of K 0 for hcp iron is 5.31, as found from Eq. 4.. Using the Mao et

Table 1 X Values of K 0 for e iron from measured P , V dataa using the BirchMurnaghan EoS where K T 0 s 165 GPa and V0 s 6.73 cm3 moly1 P GPa. Fig. 2. The phase diagram of iron in V , T space. The shaded areas represent DV at phase boundaries. The beta phase b . is accepted as part of the phase diagram, making b the dominant phase at core conditions. The Hugoniot passes through the e and b phases well below melting, until it emerges into the liquid phase at 240 GPa. 34.8 35.6 56.1 80.0 101.6 112.4 125.4 132.1 141.2 142.2 156.7 170.8 197.3 208.7 210.1 212.1 227.2 235.7 244.8 247.3 256.2 258.2 264.0 270.1 270.4 277.8 292.8 293.5 294.7 304.1
a

V cm3 moly 1 . 5.8110 5.8120 5.4740 5.2760 5.0800 4.9300 4.9030 4.8090 4.7140 4.7010 4.6690 4.6130 4.4750 4.4580 4.4680 4.4170 4.3850 4.3260 4.3210 4.3100 4.2540 4.2870 4.2890 4.2430 4.2130 4.1700 4.1450 4.1370 4.1440 4.0990

V =10y 5 m3 kgy 1 . 10.408 10.410 9.805 9.450 9.099 8.831 8.872 8.614 8.443 8.420 8.362 8.263 8.015 7.985 8.383 7.912 7.854 7.748 7.740 7.720 7.620 7.679 7.682 7.600 7.546 7.469 7.424 7.410 7.423 7.342

K0 4.915 5.257 4.795 5.687 5.587 4.997 5.538 5.135 4.878 4.835 5.193 5.324 5.244 5.448 5.553 5.278 5.452 5.287 5.456 5.443 5.294 5.531 5.663 5.500 5.330 5.223 5.338 5.306 5.364 5.268

where the subscript 0 means at P s 0. The solid squares in Fig. 1 DX , D, and DY . represent gs of the core at the core-mantle boundary CMB., the inner core boundary ICB. and the middle of the outer core but in the solid state., as calculated by Stacey 1995.. To these literature data on g is now added the value of gs e ., Point G, corresponding to the e solid phase. We find gs e . at P s 0, Point G in Fig. 1, from X the relationship between K 0 and g 0 for hcp iron, as derived by Stacey 1995. see his Eqs. 34. and 35.., where:
X g 0 s 1r2 . K 0 y 0.95, X K0 s E K TrET .Ps 0 . X K0

3.

and where To obtain for e iron listed in Table 1., one uses P , V data on hcp iron measured to 330 GPa by Mao et al. 1990. at X is found from these data by solving T s 300 K. K 0 the BirchMurnaghan EoS, Eq. 4.. The values of

P and V data from Mao et al. 1990.. X Average value of K 0 is 5.31.

182

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

X al. P , V data with the Vinet EoS, one finds K 0 s 5.31 X and, using same data with the BornMie EoS, K 0 s X 5.11. For e iron, K 0 s 5.3 is selected from the Mao et al. 1990. data, yielding g 0 s 1.7 for e iron from Eq. 3.. The value of g 0 for hcp iron is represented by Point G in Fig. 1. We now seek gshcp. at high pressure. Stacey 1995. found theoretically the values of gs for close packed crystalline iron at core pressures shown by the squares D, DX , and DY . in Fig. 1, covering the pressure range of the outer core. They are g DX . s 1.27, g D. s 1.30, and g DY . s 1.33. Gilvarry 1956. argued that the melting of a solid occurs at the limit where T Tm , e.g., at the melt boundary of the solid and liquid, but on the solid side. Therefore the melting line shown as a dashed line in Fig. 1. is drawn adjacent to Staceys points solid squares.. Noting Staceys value of g at midcore pressure, e.g., near 240 GPa, we now connect the Hugoniot to Staceys Point D. I assume that for the hcp phase, gs is independent of T at constant V . This makes the Hugoniot value, g H P ., identical to the hcp value, gs P .. I now use the experimental observation that the curve representing the P , V data on the Hugoniot egresses from the solid hcp phase at P s 240 GPa, because at that pressure, the shear velocity of pure iron approaches zero, as seen in Fig. 3. Thus, the onset of melting for the Hugoniot, Tm 240., is at a pressure of 240 GPa. The g of the Hugoniot, g H , is connected to Staceys midpoint value, which is at 240 GPa, the same pressure at which the Hugoniot enters the liquid state. This gives g H 240. s gm s 1.30. Since gs e . is a function of P but insensitive to T at constant r , the Hugoniots g H is identical to gs e . or g of b iron above T s 1500 K.. Thus the terminus of the gs e . curve, as well as that of g H , is Staceys Point D. A line connects gs e . at P s 0 r s 8280 kg my3 . and gs e . at P s 240 GPa r s 12200 kg my3 .. From Eq. 3. it is found that q s 0.7, so that ge r . is:

Fig. 3. The Hugoniot experimental P , V data, plotting r vs. P . The sound velocity of iron measured along the Hugoniot path. The melting of the Hugoniot, placed at 240 GPa, occurs when ys plunges to zero. D r of melting is found from the separation between the two Hugoniot curves in the liquid state. Data represented by solid circles are from Brown and McQueen 1986.; data represented by diamonds are from Altshuler et al. 1962..

tion of state of hcp iron and the Hugoniot of iron to the melt at midcore pressure. We are now in a position to evaluate the melting temperature at 240 and 330 GPa. We follow the procedure shown by Anderson and Duba 1997. and also by Stixrude et al. 1997. evaluating TH at its melting pressure 240 GPa. and then finding Tm330. from Tm 240. by the Lindemann law. The temperature profile of the Hugoniot must now be evaluated.

4. Evaluation of the Hugoniot temperature The Hugoniot temperature, TH , at any PV point can be calculated McQueen, 1964. from the integration of: dT H s ygs TS V dV q

ge r . s 1.7

/
r

r0

0.7

5.

The g r diagram, Fig. 1, is now complete. What is significant for core physics from the diagram is that gm240. s 1.3, which is identical to the solid state value of Stacey 1995. and also connects the equa-

V0 y V . d P q P y P0 . dV
2 CV

6.

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197 Table 2 Calculation of TS and DTS the isentropic contribution to DT H . P J my3 . 0 40 60 80 100 120 140 160 180 200 220 240
a

183

r =10y 3 kg my 3 .
9.49 9.99 10.42 10.72 11.04 11.3 11.54 11.75 11.94 12.12 12.20

V =10y 6 m3 kgy 1 . 10.5 10.0 9.6 9.31 9.06 8.85 8.69 8.51 8.38 8.25 8.14

g
1.7 1.55 1.49 1.45 1.42 1.39 1.34 1.35 1.33 1.32 1.31 1.30

D rrr 0.0513 0.0490 0.0280 0.0215 0.0232 0.0210 0.018 0.016 0.0150 0.0130

gave
1.52 1.47 1.435 1.405 1.38 1.36 1.34 1.325 1.315 1.305

DTS K. 51 50.8 30.7 30.46 26.15 24.09 20.94 18.82 17.84 15.80

TS K. 650 a 701 752 783 814 840 864 885 904 922 937

P0 s P 40. s 40 = 10 8 J my 3 ; V0 s V 40. s 10.5 = 10y5 m3 kgy 1 . Andrews 1973..

or: d T H s d TS q dS CV ,

7.

where gs represents g of the solid state for hcp iron, dTS is the temperature increment due to isentropic

compression, and d SrC V is the temperature increment arising from P H y P . under compression dV . Thus d S can be considered as a type of work term, dW see also Anderson, 1995, p. 321.. The wellknown P and r data along the Hugoniot see Table 2 of Brown and McQueen, 1986. are reproduced as the first two columns in Table 2. The first term on

el Fig. 4. The electronic specific heat factor at pressures along the Hugoniot. The inset shows C V calculated by Stixrude et al. 1997. for el el isochores vs. T . The final C V rR curve was found by assuming a temperature distribution T vs. r so as to calculate C V vs. pressure and then recalculate from Eq. 6. the resulting T vs. P distribution and recycle until convergence has been achieved.

184

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

Table 3 Work contribution to TH , DTH P =10 9 J my 3 . 0 40 60 80 100 120 140 160 180 200 220 240
a b el CV in terms of R

C V Total =10 2 J kgy 1 Ky 1 . 4.461

Work a =10 2 J kgy 1 . 0 0 1054.8 1766.6 2084.2 2491.5 2729.9 2976.2 3110.9 3245.6 3406.0 3469.1

DT W : W %C V K.

D TS K.

DTH : DT W qDTS K.

Hugoniot T , TH K. 650 b 916 1311 1731 2205 2698 3210 3720 4229 4750 5271

0.3 0.45 0.60 0.77 0.94 1.10 1.28 1.45 1.55 1.62

4.907 5.130 5.353 5.606 5.859 6.097 6.364 6.617 6.766 6.870

215.0 344.4 389.4 444.3 465.9 488.1 488.8 490.5 503.4 505.0

51.0 50.8 30.7 30.5 26.2 24.1 20.9 18.9 17.8 15.8

266.0 395.2 420.1 474.3 492.1 512.1 509.7 509.4 521.2 520.8

P y P .dV q V0 y V .d P 4r2; P , P0 , V , V0 from Table 2. Boundary condition: Andrews 1973.. Also listed: DTS and T H .
el from theirs because the C V results of Stixrude et al. 1997. and a different and more exact solution of g V . for e iron were used. It is significant that the temperature rise due to isentropic compression is less than 10% of the total temperature rise. Therefore it would take an unreasonable value of gs to shift the Hugoniot temperature by a significant amount. We find that gs is exactly fixed by Eq. 5..

the right of Eq. 7. arises from the adiabatic temperature profile; step-by-step details of its evaluation along the Hugoniot are listed in Table 2. This term may be readily evaluated since g V . for e iron is now known from Eq. 5.. The numerator of the second term in Eq. 7. can be considered to be the work necessary to displace the isentrope to the Hugoniot: W s 1r2 V0 y V .d P q P y P0 .dV 4 . The chief problem in evaluating Eq. 6. consists of finding values of C V , composed of the classical lattice term 3 R , but with an additional term arising el from the free electrons, C V , which is linear in T and also depends on density. The calculation of the el fraction of dTH arising from C V must be done el cautiously because C V depends on T , and T depends el on C V . Stixrude et al. 1997. found and plotted C Ve of iron vs. T and r their Fig. 2.; their plot is shown el as an inset in Fig. 4. A solution was found for C V vs. P ; it is the prominent curve shown in Fig. 4. The solution for TH vs. P for the Hugoniot is listed in the last column of Table 3, the previous columns of which give the step-by-step evaluation of d S. The integrating constant, T0 , is the value of T at the phase boundary between the a and e phases, given as 650 K by Andrews 1973., Fig. 9. The solution for TH P . is given in Table 3, which lists T H 240. s Tm 240. s 5271 K 5300 K to two significant figures.. Boness and Brown 1990. found TH 240. s 5625 K and Brown and McQueen 1986. found T H 240. s 5510 K. Results found here are different

5. The value of Tm of iron at the ICB pressure Taking Tm 240. s 5300 K for iron, the value of Tm 330. for iron can be found from the Lindemann equation for melting. The Lindemann formulation in T , r space is Anderson, 1995, p. 281.:

E ln Tm E ln r

s 2 gm y 1r3 . .

8.

From Fig. 1, we see that gm changes very little between P s 135 and 330 GPa. Thus the above equation can be approximated by taking 2gm y 1r3. as a number independent of density, e.g., 2gm y 1r3. s 1.9 for gm s 1.3. across the outer cores density range. By integrating Eq. 8. for this special case, it is found that:

/ /
s Tm 1

Tm 2

r2 r1

1.9

9.

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197 Table 4 Values of parameters for pure iron and the core Radius and pressure km. r 135 GPa., CMB r 200 GPa. r 240 GPa. r 330 GPa., ICB
a b

185

r , pure iron:core conditionsa 10 3 kg my 3 .


11.28 12.12 12.47 13.38

r , core b
10 3 kg my 3 . 9.9035 11.29 12.166

A c surface area =10 13 m2 . 15.26 d

Mc mass of interior c =10 24 kg. 1.85

DVm pure iron m3 kgy 1 .

Tm pure iron K. 4380 5020 5300 6100

1.9

0.097

2 = 10y6 0.7 = 10y6

For e-iron solid. Mao et al. 1990. for room temperature data followed by T correction from a T . PREM. c Stacey 1992.. d Melchior 1986..

Values of r P . needed for the evaluation of Eq. 9. are listed in Table 4, from which we have r 330.rr 240. s 13.38r12.47. Thus, Tm 330.r Tm 240. s 1.15 by Eq. 9., and Tm 330. s 6095 K for pure iron, 6100 K to two significant figures. Using Eq. 9. and the densities listed in Table 4, we find the melting point of pure iron at mantle-core boundary MCB. pressure to be about 4380 K. At 200 GPa, Tm is 5020 K, considerably larger than the experimental value reported by Boehler 1993., Tm 200. f 3950 K. Boehler 1993. extrapolated his experimental Tm data and obtained Tm 330. s 4800 K, whereas 6100 K was obtained here. The value of Tm 240. obtained here involves no estimates. The calculations used in finding g for e iron were based on P , V measurements on hcp iron in the diamond cell Mao et al., 1990.. The determination that the Hugoniot crosses the melting curve at 240 GPa has been made experimentally see Fig. 3.. The Lindemann theory is used to find the value of Tm 330. from Tm 240.; gm is determined from Fig. 1. el For the calculation of TH along the Hugoniot, C V values at all pressures along the Hugoniot see Fig. 4. are needed. These were extracted from the paper el of Stixrude et al. 1997., who calculated C V vs. volume and temperature. My calculated value of Tm330. for pure iron, 6100 K, agrees reasonably well with the value calculated by Poirier and Shankland 1994. using dislocation theory, Tm 330. s 6150 K, and the value found by Stevenson 1981.. Many others have found Tm 330. to be near 6000 K, including Gilvarry 1956., Zharkov 1962., Birch 1972., Boschi 1975., Liu

1975., Brown and McQueen 1986., Anderson and Young 1988., Boness and Brown 1990., Kerley 1994., Wasserman et al. 1996. and Anderson and Duba 1997..

6. Models of core impurity composition Our value for Tm 330. is for pure iron. To find Tm 330. for the core, we must calculate the effect of the cores impurities on the melting temperature of pure iron. In looking for a plausible set of impurities for the core, one observes wide division of opinion on this matter among cosmochemists and geophysicists. Table 5 is a summary of various past models for core impurity composition. The uncertainty in the choice of light element impurities has been greatly reduced by several recent advances. First, Poirier 1994. provided equations and figures that help restrict the concentration of light elements. The concentration of all impurities must produce a total impurity density equal to 10% of the density of pure iron at core conditions. This is called the core mass deficit, e.g., the difference between the density of iron and PREM density at core conditions. The equations of Poirier 1994. allow one to derive the contribution of each impurity in the core to the 10% core mass deficit. The relation between the weight percent of an impurity and the core mass deficit is shown in Fig. 5. The three lower lines, for oxygen, sulfur, and silicon, are taken from the analysis of Poirier 1994. see his Fig. A1..

186

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

Table 5 Models of core composition impuritiesa Previous chemical models Sulfur only Murthy and Hall 1970. Ahrens 1982. Small amount of sulfur, large amount of silicon Allegre ` et al. 1995. Dreibus and Palme 1996. No sulfur, silicon only Saxena and Benimoff 1977. McDonough and Sun 1995. Silicon and sulfur MacDonald and Knopoff 1958. Sulfur and oxygen Boehler 1992. Features added Oxygen and silicon almost mutually exclusive Sulfur with large DV of mixing Sherman, 1997.
a

Impurities considered are the abundant light elements Si, S, and O.

Second, the paper by ONeill et al. 1998. provides experimental evidence on the solubility of Si, S, and O in an Fe-rich metal. Their results restrict the amount of S in an Fe-rich metal to less than 6% by weight because of high volatility, and they found

O to be even more cosmochemically volatile than S. The abundance of Si is tied to the abundance of SiO 2 , and there is no evidence that there was a large enough loss of O from the primitive Earth to free sufficient Si for the core. ONeill et al. concluded that . . . from the chemical view, none of the cosmochemically abundant light elements appears able to account for the mystery light component in the core, at least when considered individually. Third, Sherman 1995. pointed out that the presence of silicon eliminates oxygen as a core impurity candidate because oxygen demands oxidizing conditions and silicon demands reducing conditions. ONeill et al. 1998. came to the same conclusion about the incompatibility of silicon and oxygen. They noted, however, that a small amount of silicon and oxygen can coexist in iron, according to results known and used in the steelmaking industry, and they presented a curve taken from the steelmaking literature shown in Fig. 6. By this figure, 2-1r2 wt.% silicon could coexist with 2-1r2 wt.% oxygen. Fourth, Sherman 1997. discovered that there is a high excess volume of mixing density deficit. in FeFeS alloys at 250 GPa in the FeFeS phase diagram see Fig. 7.. This density deficit is sufficiently large that it shifts the sulfur curve towards smaller levels of weight impurity the S ) curve in

Fig. 5. Curves showing the contributions of a given weight percent of an impurity element towards the density deficit of the outer core. S represents sulfur assuming an FeS ideal solution. S ) results from the volume of mixing in the FeS solution according to Sherman 1997.. The curves for O, S , and Si are taken from Fig. A1 of Poirier 1994.. The curve for S ) is found from data presented by Sherman 1997..

Fig. 6. The mutual dependence of the solubilities of Si and O in liquid Fe at 1 bar, showing that Si and O are mutually exclusive except along the two axes. From the steelmaking literature and modified from ONeill et al. 1998. their Fig. 10..

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

187

Of the six proposed impurity compositional models listed in Table 6, only the first four satisfy the limits stated above. The last two are for the cases of S alone and Si alone. For all six cases considered, the value of the weight percent sum is constrained to account for the 10% core density deficit using the curves in Fig. 5.

7. Depression of Tm(ICB) by impurities Finding the temperature depression, DT , is straightforward if known impurities with known concentrations in iron form as a solid solution, but the problem is complex if the solute forms a eutectic with the solvent. The concentrations of the impurities in the first four cases listed in Table 6 are sufficiently small that I assume that the impurities do indeed form a solid solution. Case f probably does not form a solid solution, but it is listed in order to show, as I do below, that the melting point depression, DT , is unreasonably large. The equation used for calculating DT for outer core conditions is: DT s yT ) i ln 1 y x i ,l . ,

Fig. 7. Density vs. composition for the FeFeS phases at 250 GPa after Sherman, 1997.. The calculated nonideal mixing curve relative to Fe and FeS is compared with the ideal mixing curve. The first up arrow on the left corresponds to 2.3 wt.% sulfur projected to the calculated curve. The first down arrow on the left represents 4.3 wt.% from the ideal mixing curve. The second up arrow on the left represents 5.9 wt.% projected to the calculated curve. The second down arrow on the left represents 10 wt.% from the ideal mixing curve. A concentration of 5.9 wt.% sulfur satisfies the 10% density deficit of the core.

10 .

Fig. 5., so that only 5.9 wt.% of sulfur would offset all 10% of the outer cores density deficit. ONeill et al. 1998. conclude that the amount of S in the core should be less than 6 wt.% or about half the amount required for the density deficit, according to Fig. A1 of Poirier 1994.. However, that conclusion was based on the S curve in Fig. 5. Taking into account the discovery of Sherman 1997. of the high volume of mixing of sulfur in FeS, the S ) curve in Fig. 5 was constructed; it shows that sulfur could completely account for the core deficit and be within the limits stated by ONeill et al. 1998.. My approach is to find combinations of S, O, and Si that are within the limits described above: sulfur identified in this paper as S ) . to be held to less than 6 wt.%, and O and Si in combination to be taken from the 2750 K isotherm in Fig. 6.

where T ) is the pure iron melting point at the ICB pressure and x i ,l is the molar fraction of the i th impurity in the liquid outer core. Eq. 10. is derived in Appendix A. Evaluating Eq. 10. for case a in Table 6, where ) xS s 0.048, x Si s 0.074, and xo s 0.032: DT s yT ) ln 0.952 q ln 0.926 q ln 0.968 4 s y0.159T ) . Using T ) s 6100 K, DT s y970 K for case a. Results from the first four cases listed in Table 6 vary from y700 to y1070 K. It is interesting that Poirier 1994. stated that many authors assume that the depression of the melting point is 7001000 K, but they do not generally explain how this result was obtained. Following the comment of Poirier 1994. that the iron core would likely have more than one impurity, I average the results from cases a, b, and c, and choose DT s y1000 K. Accordingly, 6100 K y 1000 K s 5100 K is assigned as the value of Tm at the ICB Braginsky and Roberts, 1995,

188

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

Table 6 Calculation of the freezing temperature depression, DT , resulting from various models of the composition of the core Model Comment Impurities S Case a Wt.% Molar frac. Case b Wt.% Molar frac. Case c Wt.% Molar frac. Case d Wt.% Molar frac. Case e Wt.% Molar frac. Case f Wt.% Molar frac. Si-rich 3 0.048 O-rich 2 0.031 S ) -rich 4 .077 S ) alone 5.9 0.109 S alone 11 0.177 Si alone 18.5 0.311 y0.372 y2271 y0.194 y1183 y0.115 y700 1.2 0.022 1.2 0.038 y0.143 y871 0.5 0.009 4 0.125 y0.1755 1070 4 0.074 1 0.032 y0.159 y970

DTrT ) S
)

DT

Si

T ) represents melting at the ICB; S represents the case of sulfur with volume of mixing in FeS solution; S ) represents sulfur in an ideal solution of FeS. Only light impurities considered. Cases a, b, c, and d constrained by limits set by ONeill et al. 1998.. Cases e and f presented to illustrate extreme cases, though they exceeded limits set by ONeill et al. 1998..

estimated 5300 K, and Stacey and Stacey, 1998, assumed 5000 K.. The ratio of temperatures at the ICB core to pure iron., Tm crTm i , is therefore 0.936. Eq. 10. is based on the assumption of an ideal dilute solution. Since both silicon and sulfur have a volume of mixing, neither silicon nor sulfur forms an ideal solution with iron at high pressure. Thus there may be some error in the above DT calculations. Nevertheless, the results from Eq. 10. listed in Table 6 give a good idea of the DT effect since the magnitude of the errors arising from lack of ideality is probably small compared to the variability in DT values arising from the choice of impurities themselves. Except for case f, concentrations of impurities described in the six models satisfy the criteria for a dilute solution. I conclude that an FeSi core is improbable because if DT due to crystallization were y2270 K, then the adiabatic T of the core at the CMB would be about 2750 K, close to the adiabatic T of the mantle at the CMB. This would result in a negligible thermal gradient across DY , preventing DY from being a

useful thermal boundary as required by convection modelling. ONeill et al. 1998. also found an FeSi core to be improbable. They stated, . . . neither O nor a combination of O with Si can dissolve in liquid Fe-rich metal in sufficient amounts to account for the presumed density deficit in the Earths outer core.

8. The melting temperature and the thermal gradient Values of Tm and gm are needed for calculations of the thermal gradient. The Lindemann law in the form given by Eq. 8. is useful because it is expressed in terms of the density, values of which are listed in Table 4. Using Eq. 9. to find Tm CMB. from Tm ICB. s 5100 K requires the density ratio 9.904r12.166.. Thus, Tm CMB. s 3450 K see Table 7 for other values of Tm P ... The equation for finding the value of Tad along the isentrope anchored at 5100 K is found from Eq. 2.. Rewriting this

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197 Table 7 Details of calculating the thermal gradient and values of thermal conductivity Radius r 135 GPa., CMB r 240 GPa. r 330 GPa., ICB
a b

189

Tm , core K. 3450 4400 5100

g la
1.33 1.30 1.27

g b m sy 2 . 10.68 7.94 4.40

Tad , core K. 3900 4630 5100

f b km sy 1 . 2
67.33 89.95 105.38

dTrd r , core K kmy 1 . y0.82 y0.53 y0.27

c Qc TW.

3.4 d , 5.4 e

From Stacey 1995.. From PREM. c Dimensions, W my1 Ky1 . d k s 28.6 W my 1 kmy 1 from Stacey 1972.. e k s 43 W my1 kmy1 from this work.

equation in densitytemperature coordinates, we have:

E ln Tad E ln r

sgl.

11 .

Since the volume decreases at melting Fig. 2., it is expected that g l will be smaller than gm . The volume decrease at melting is on the order of 1%. Thus, we take g l s 1.29. Since g l changes very little at outer core conditions, Eq. 11. becomes:

= 10 6 J kgy1 from the above formula. To get power, we must know the age of the inner core t in billion years.. Labrosse et al. 1997. use t s 1.7, and Stacey and Stacey 1998. use t s 2.3. For t s 2, the power, Q m , is 1.5 TW. Glatzmaier and Roberts 1996. estimated Q m s 3 TW. Stacey and Stacey 1998. estimate a much smaller value of DV they find DVrV f 1.1%., which would depress the value of D Hm to 0.5 = 10 6 J kgy1 and that of Q m to 0.4 TW.

r2 Tad . 2 s r1 Tad . 1

1.29

12 .

9. The thermal conductivity in the core We have determined g and dTrd r across the core, but we also need the value of thermal conductivity to find the gradient power, Qc , across the core side of the CMB. According to the Fourier law, power is equal to Eq. 1. multiplied by area: dT Qc s A c k c , 13 . dz where A c is the surface area of the core, and k c is its thermal conductivity. There are two recent estimates of the value of k c . Braginsky and Roberts 1995. proposed 40 W my1 Ky 1 ; Labrosse et al. 1997. proposed 60 W my1 Ky 1. Other estimates of k c are 35 W my1 Ky 1 Buffett et al., 1996., 40 W my1 Ky 1 Stevenson, 1981.; and 28.6 W my1 Ky 1 Stacey, 1972; Stacey, 1992, p. 331.. In order to justify the magnitude of these values of k for the core, we need to explore the theoretical and experimental background by which k c is obtained. The chief problem for core physics is that values of k c are needed at temperatures very much larger than most measurements can be taken. Data on

Values of Tad for three radii in the outer core found from Eq. 12., including Tad ICB., are shown in Table 7. The core temperature at the CMB is 3900 K. These values are close to those presented by Braginsky and Roberts 1995. in their Table E2. With these data, dTrd r can be found using Eq. 1.; values of dTrd r . for the three selected radii are also shown in Table 7. Now that gm and Tm for the core at the ICB have been found, there is sufficient information to determine the heat of crystallization, D Hm . The formula for D Hm , derived from Anderson and Duba 1997., is: D Hm s DVm K S m 2 gm y 1r3 . 1 q agm Tm . ,

where DVm is the volume change at melting, K Sm at 330 GPa is 1370 GPa, according to PREM, a ICB. s 1.7 = 10y5 Ky 1 Stacey, 1995., and, from the discussion above, gm s 1.3 and Tm s 5100 K. Anderson and Duba 1997. found DV s 2 = 10y6 3 m kgy1 from Hugoniot data, so that D Hm s 1.38

190

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

k at P s 0 or at small pressure. even at high T are beset by several transitions, as can be seen by examining the P s 0 isobar in Fig. 2. In order to find k of the core, we need data at high P and T near the path shown by the Hugoniot in Fig. 2. A physical principle called the WiedemannFranz ratio, by which the value of thermal conductivity in metals is found from measurements of the electrical conductivity, is very helpful. Electrical conductivity, s , can be measured more easily at high pressure than thermal conductivity, k , because the latter is a transport property requiring accurate measurements of gradients, whereas the former requires the simpler measurements of current and field. The WiedemannFranz ratio, by which k is calculated from s , is founded in the classical free electron model of atoms. A summary of that theory following the derivation of Joos 1958. is presented in Appendix B, and the result is given by:
. 15 . s 3 e k e is the electronic contribution to the thermal conductivity, s is the electrical conductivity, k is the Boltzmann constant, and e is the electrons charge. The quantity on the right of Eq. 15. is often called the Lorenz number, K L : note that it is composed only of fundamental constants and numbers. K L can be simplified by including in the thermal conductivity equation the magnetic diffusivity Braginsky and Roberts, 1995., h s 8 = 10 5 sy 1 , where h is in units of m2 sy1 and s is in units of S my1 . For this case, k e s K L = Trh and K L s 0.02 W m sy1 Ky1 . The WiedemannFranz ratio is remarkable because, though it is derived from free electron theory, none of the parameters of that theory are retained. Many physicists early in the century did experiments to see whether the WiedemannFranz ratio is independent of T and of the metal tested. An example of such a test is given by Sprackling 1991., p. 305, who showed that srk T is the same for the metals copper, silver, lead, and zinc over a broad range of temperature. The WiedemannFranz ratio is also very useful for core physics because k e is proportional to s T at all pressures and temperatures. Thus if s is measured at a high T , then k is measured at the same high T . The WiedemannFranz ratio provides a very useful method for extrapolation s

of measurements made at accessible temperatures to core temperatures. Further, there is a suggestion that the measurement of the effect of an impurity on electrical conductivity may be independent of the kind of impurity, depending only on concentration. The electrical conductivity at pressures along the Hugoniot was measured by Matassov 1977. on samples of iron with various levels of silicon impurities. His shock wave results are reproduced in Fig. 8. The plot shows s vs. P along the Hugoniot up to 140 GPa, barely within outer core pressure. On the Hugoniot the temperature increases with P , but at maximum 140 GPa., TH ; 2100 K for an ironsilicon alloy, while the core temperature is about 3900 K see Table 7.. I assume that the effect of one mole of Si on electrical conductivity is the same as that of one mole of S or one mole of oxygen. Thus a mole fraction impurity level of roughly 0.150 Table 6. is

ke

p2 k

Fig. 8. Electrical conductivity of Fe with various levels of silicon impurities at high-pressure shock conditions measured by Matassov 1977. up to 140 GPa.. The FeSi curve for x Si s 0.181 is used to obtain the electrical conductivity, s s 8.5=10 3 mho cmy1 s 10 5 S my1 , at maximum P . Since T of the Hugoniot is about 2200 K, the value of s must be corrected downward. to obtain core temperatures ; 3900 K., and must be further corrected downward to go from the solid to the liquid state.

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

191

needed to achieve the 10% core deficit level. The FeSi curve in Fig. 8 for x Si s 0.181 is used to find s s 8.5 = 10 5 mho my1 s 8.5 = 10 5 S my1 . Matassov reports that in the high pressure range s scales as 1rT , so s at 140 GPa and the solidus T is 8.5 = 10 5 2100r3900. s 4.6 = 10 5 S my 1 following Braginsky and Roberts, 1995., but this is still in the solid state. Another reduction in s is required because melting always further reduces the conductivity. Assume a further 10% drop due to melting, obtaining s s 4 = 10 5 S my1 . Thus, for T s 3900 K and h s 2 m2 sy1 , k e s 0.02 = 3.9 = 10 3r2 s 39 W my1 Ky1 . In his Table 18.8, Matassov 1977. lists g for the Earths core as 7.304.87. = 10 5 S my 1, so apparently he did not reduce s due to melting as was done by Braginsky and Roberts 1995.. Consequently he finds k e for the core to be higher, 54 W my1 Ky1 . Gardiner and Stacey 1971., using data then available on ironsilicon alloys, extrapolated the resistivity of pure liquid iron at low pressure to core temperatures and found, after correcting for the expected impurity concentrations, that s s 3 = 10 5 S my1 , corresponding to k e s 28.6 W m Ky1 Stacey, 1992.. Secco and Schloessin 1989. made measurements of electrical conductivity on solid and molten pure iron up to 7 GPa in a large volume press and found s s 7.8 = 10 5 S my1 for pure iron at this pressure. They reasoned that this value could be assumed to be the same as the conductivity of the iron diluted by impurities at outer core P and T . Taken at face value, this leads to k e f 77 W my1 Ky1 , which is probably too large. References to the value of s at core conditions listed in the table of Matassov 1977., may not be corrected for temperature between Hugoniot T and core T and for the solidliquid state transition. This could lead to values on the order of k e s 6070 W my1 Ky1 . The value of k e s 39 W my1 Ky1 seems a reasonable upper limit for the core. To this must be added the lattice contribution, k l , so that:

10. Conductive heat from the core to DY and the mantle Choosing k c s 28.6 W my1 Ky1 as the lowest limit and k c s 43 W my1 Ky1 as the highest limit for the outer core, one calculates the limits in conductive heat flow from the core by Eq. 13.. A c is listed in Table 4. The two limiting values of the present-day power, Q c , leaving the core along the gradient are 3.4 and 5.4 TW, or Q c s 4.4 " 1 TW. According to the curve in Fig. 2 of Buffett et al. 1992., the power leaving the core is Qc s 5 TW corresponding to their t s 2 billion year lifetime. curve.. The power from the surface of the Earth is 44 = 10 12 W 44 TW., according to Pollack et al. 1993.. Therefore the conductive power flowing into the mantle is about 10% of the surface power leaving Earth. On the core side of the CMB, T c 2970 km. s 3900 K, as given by Tad in Table 7. Stacey and Loper 1983. found dTrd ZDY . s 9.6 K kmy1 . The change in temperature across DY is thus f 1440 K for a thickness of DY s 150 km. Therefore, on the mantle side of the CMB, T m 2820 km. f 2460 K, which agrees with the value Brown and Shankland 1981. found from their isentrope of the lower mantle. The Stacey and Loper 1983. thermal gradient in DY is 12 times the thermal gradient of the core at the CMB, and therefore DY would conduct the same power as the core if k DY is 1r12 that of k c , e.g., k DY f 3.7 W my1 Ky1 . This suggests that DY is composed primarily of mantle material with little core material contained in it. I therefore conclude that DY is a thermal boundary layer, not a compositional boundary layer.

11. Thermal conductivity and the power flowing into the lower mantle It is commonly accepted that the thermal conductivity in the deep lower mantle is about 45 W my1 Ky1 . Kieffer 1976. estimated that k lm s 4.2 W my1 Ky1 by calculating the heat transfer due to phonons in a dielectric solid. If the deep mantle has this value of k lm , then the lower mantle could conduct less than 4.2r43 s 11% of the cores con-

k c s ke q k l .

16 .

Taking k l to have the same value as the lattice k in the deep lower mantle, k l f 4 W my1 Ky 1 Kieffer, 1976.. Therefore, I take k c s 43 W my1 Ky 1.

192

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

ductive power, leaving one to find some other mechanism of heat transfer for the large amount of residual power 3.9 TW.. However, in the assumption that k lm is the same as the lattice thermal conductivity, there is the implicit assumption that conductivity is determined by phonon transfer alone. This implicit assumption will probably result in an underestimation of the capacity of the lower mantle to conduct power towards the surface. In addition to lattice thermal conductivity, k L ., there is also likely radiative transfer connected with infrared electromagnetic waves arising from the optical properties of a dielectric solid. The radiative thermal conductivity associated with infrared electromagnetic waves, k R , is given by Zharkov and Trubitsyn 1978., p. 57, according to the equation:

kR s

16 s) n2 T 3 " , 3 a

where s) is the StefanBoltzmann constant, a is the absorption coefficient, and n is the refractive index. Note the T 3 factor, which has a pronounced effect at lower mantle temperatures. In a recent report, Hofmeister 1998. showed that the optic modes are the primary mode of heat transport determining thermal conductivity at lower mantle conditions, contrary to the assumption made in the calculation of k lm where phonons are considered as the only transport mechanism. Taking into account recent infrared reflectivity measurements and correcting for the pressure dependence of the constants, she found that k R for the deep lower mantle is sufficiently large that k L q k R is about 2.3 times the value of k L . Thus we have sufficient evidence to suppose that k lm is roughly 10 W my1 Ky1 rather than about 4.2 W my1 Ky1 . Taking the thermal gradient of the lower mantle to be 0.5 K kmy1 Brown and Shankland, 1981., 0.6 TW and perhaps more is transported conductively towards the lithosphere. Could the residual of the cores conductive power, 3.8 TW, be transported upwards in the lower mantle by convection? Tackley et al. 1993, 1994. and Tackley 1995, 1996. have shown that the convective heat in the mantle is transported to the surface by a number of mantle plumes originating at the base of DY . When

these plumes interact with the lithosphere, they create hotspot intrusions that can be studied by geophysical methods. The number of plumes that are currently interacting with the Earths surface equals the number of known hotspot intrusions. G. Schubert private communication. reported that he, Don Turcotte and Peter Olson, who are writing a book on mantle geodynamics, have identified a total of 38 current hotspot intrusions. They have calculated that these intrusions collectively transport 2.3 TW of power to the Earths surface. Thus the lower mantles conductive heat, 0.6 TW, and the convective heat from the known hotspot intrusions, 2.3 TW, yield 2.9 TW, close to the lower limit of our suggested core conductive power, 3.4 TW. In addition to the plumes evidenced by hotspot intrusions, there are, however, plumes that have not yet been made evident by surface activity, as revealed by the snapshots of plume activity given by numerical simulation done by Tackley and his colleagues. Plumes grow from DY , slowly ascend to the surface, contact the surface, detach from DY , and eventually disappear, but, at any given time there are more active plumes than the number of plumes in contact with the surface giving hotspot intrusions.. Thus, a reasonable assumption is that the number of hot plumes that have not yet made contact with the surface is about half of the number that have been identified through surface activity. I assume that the total convective power arising from DY via the plumes is 1.5 times the power for known hotspot intrusions calculated by Schubert and his colleagues that is, a total of about 3.5 TW.. Under this assumption, the total estimated power coming from the CMB both convective and conductive. would be about 4.1 TW, close to the mean of the upper and lower limits of conductive core power 4.4 TW.. I therefore conjecture that a rough balance exists between the conductive power transmitted from the core to the mantle and the power transferred by conduction and convection from the upper mantle towards the lithosphere.

12. The convective heat in the outer core The outer core must be in a state of convection so that the Earths magnetic field can be maintained.

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

193

Maintenance of the magnetic field requires a small fraction of a TW 0.2 TW, according to Stacey and Loper, 1983. in mechanical energy. Loper 1984. and Loper and Roberts 1983. have proposed that the mechanical energy requirements may be largely supplied by vigorous convective motion driven by compositional buoyancy of the ejecta enriched with light elements. from the inner core to the outer core. The convective power arising from the core may be small compared to its conductive adiabatic power. Buffett et al. 1996. emphasized that a modest heat flux in excess of that conducted down the adiabatic gradient is sufficient to power the geodynamo, even in the absence of compositional convection and latent heat release. Gubbins 1977. stated that the gravitational energy released by rearrangement of matter in the core is completely converted to magnetic dissipation enabling a large magnetic field to be generated with a low heat flow from the core. Although convective heat may be transferred from the core to DY and then on to the lower mantle, there appears to be no convincing evidence that the cores convective power is large compared to that conducted down the adiabatic gradient. Any excess power may be returned to the core by compositional convection Loper, 1978a. or contained by a subadiabatic thermal gradient near the boundary of DY , as proposed by Labrosse et al. 1997.. In my proposed model, the mantle successfully transports away the conductive power arising from the cores adiabatic gradient. Therefore, I assume that a large subadiabatic gradient is not necessary and that any power from core convection alone in excess of the power conducted down the adiabatic gradient will be returned to the core by compositional convection.

quite helpful. Support from NSF grant EAR-9614654 acknowledged. Support by ONR acknowledged. IGPP contribution no. 5073.

Appendix A Calculation of the temperature depression requires the assumption of thermodynamic equilibrium between the solid and liquid phases. As a beginning, assume that all impurities reside in the liquid. Because the condition of thermal equilibrium requires that the Gibbs free energy be equal for both phases DG s 0., we must deal with the chemical potential of the pure liquid and the pure solid. Stevenson 1981. showed that if all the impurities reside in the liquid, the thermal equilibrium arising from DG s 0 is given by:

ml0 P ,T . q RT ln 1 y x i ,l . s ms 0 P ,T . ,

A1 .

Acknowledgements I acknowledge helpful comments from Frank Stacey, Stanislav Braginsky, Paul Roberts, Gary Glatzmaier, Jerry Schubert, and Paul Tackley on the geophysical aspects of an early version of this paper. I acknowledge helpful comments from Dave Sherois man, Giulio Ottonello, Surendra Saxena, Franc Robert, and John Wasson on its geochemical aspects. The critiques of two unknown reviewers were also

where m 0 is the chemical potential, and x i ,l is the molar fraction of the i th species in the liquid and where the super and subscripts l and s refer to the liquid and solid, respectively. Note that the weight percent of an impurity element was used in calculating the fraction of the cores density deficit that it can account for. But now, using thermodynamic functions to obtain freezing depression, we need to express concentrations in molar fraction because molar fraction identifies the number of atoms on a particular thermodynamic site. The molar fraction of the impurities in the six cases is listed in Table 6. Eq. A1. is oversimplified for use in core theory because we now are sure that light impurities exist in the solid inner core. As the core solidifies, a large fraction less than unity. of the impurities passes into the liquid, leaving a small fraction behind in the inner core see, for example, Jephcoat and Olson, 1987.. Impurities in the solid inner core are accounted for by adding a term to the right of Eq. A1., giving:

ml0 P ,T . q TR ln 1 y x i ,l .
s ms 0 P ,T . q TR ln 1 y x i ,s . , where x i ,s - x i ,l .

A2.

194

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

Assuming that v is the fraction of all impurities existing in the outer core, then 1 y v . is the fraction residing in the inner core. We make the approximation that the second term on the left of Eq. A2. and the second term on the right of Eq. A2. are multiplied by v and 1 y v ., respectively it would be more accurate to multiply all x i ,l by v and all x i ,s by 1 y v ., but for small concentrations the approximation used here is sufficiently accurate.. Thus, Eq. A2. can be replaced by the approximation:

the denominator of Eq. A4. that vrln 2 can be replaced by unity. Thus, Eq. A4. is simplified to: DT s yT ) ln 1 y x i ,l . .

A5.

Appendix B The free electron model of metals, which preceded wave mechanics, was very successful in the formulation of a number of properties of metals, including electrical conductivity. In this classical model, the valence electrons are able to move about the lattice freely having intermittent reactions with the lattice. Five parameters are of significance in the free electron theory of metals: the drift velocity of the electrons, y ; the relaxation time, t , measuring the time between collisions with the lattice; the number of electrons per unit volume, n; and the mass and electrical charge of the electron, m and e. The electrical conductivity, s , of a metal is the parameter relating the current j to the electrical field E, where: j s s E.

ml0 P ,T . q v RT ln 1 y x i ,l .
s s m0 P ,T . q 1 y v . RT ln 1 y xi ,s . .

Using the general relationship D m s D H y TD S, where H is enthalpy and S is entropy, and referencing Chapter 9 of Landau and Lifshitz 1958. as the authority, Stevenson 1981. replaced Eq. A1. eliminating ms, leaving the following for the freezing depression, DT , of the outer core: DT s yv

/
DS
)

RT )

ln 1 y x i ,l .

B1.

vT
sy

ln 2

ln 1 y x i ,l . ,

A3.

where D S is the entropy of melting and v has been incorporated. For the core, T ) is 6100 K. The parameter v takes into account the fact that the inner core contains some impurities. Thus: DT s y T )v ln 1 y x i ,l . .

Separate solutions of y with E giving y s errm. E and y with J J s ney . combined with Eq. B1. eliminate the parameters J, E, and . We then find the electrical conductivity in terms of the parameters of the free electron model:

s s ne 2trm .

B2.

0.693 .

A4.

We proceed to find the thermal conductivity of the free electron model. In solids, k is found from the Fourier law, Eq. 13., so that:

Anderson and Duba 1997. calculated that at the ICB, D r freezing. s 200 kg my3 . Masters and Shearer 1990. found from seismic data on the core that D r ICB. s 550 kg my3 , leaving D r chemical differentiation. s 350 kg my3 . The outer core has a density 1270 kg my3 lower than that of pure iron so that the concentration of impurities in the outer core accounts for 1270 y 350 or 920 kg my3 . Thus v s 920r1270 s 0.72. Within the approximations created by the uncertainty in exact chemical composition, the value of v is close enough to the value in

k solid. s

/
A

Q dz dT

Analogous to this is the thermal conductivity of gases, which, according to the kinetic theory of gases, is:

k gases . s 1r3 . Cu l ,
where u is the average particle velocity in the direction of heat flow, l is the mean free path, and C is the contribution of each gas molecule to the specific heat per unit volume. In the equation for k solid.,

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197

195

we note that k is inversely proportional to the temperature gradient, dTrd z. The thermal gradient is also found in k of gases, located in C, the specific heat: C s d ErdT . s d Erd z .d zrdT .. Thus k of gases is also inversely proportional to the thermal gradient.. The 1r3 comes from integration of the cosine angle, because the particles will be travelling in all directions with the velocity . The need to find the vector component in one selected direction leads, by integration, to the factor 1r3. The specific heat, C, is the product of n and k , where n is the number of molecules per unit volume, and k is the energy of one particle with one degree of freedom, the Boltzmann constant. The conduction electrons are in a state of chaotic thermal agitation, somewhat like the atoms in an ordinary gas, except that l of free electrons is the measure of the distance between collisions of the electron with the lattice. The equation for k e for free electrons is the same as for gases:

References
Ahrens, T.J., 1982. Constraints on core composition from shockwave data. Phil. Trans. R. Soc. London A 306, 3747. Allegre, C.J., Poirier, J.-P., Humler, E., Hofmann, A.W., 1995. ` The chemical composition of the Earth. Earth Planet. Sci. Lett. 134, 515526. Altshuler, L.V., Sinakov, G.V., Irunin, R.F., 1962. On the composition of the Earths core. Isv. Earth Phys. 1, 13. Anderson, O.L., 1995. Equations of State of Solids for Geophysics and Ceramic Science. Oxford Univ. Press, New York. Anderson, O.L., Young, D., 1988. Crystallization of the Earths inner core. In: Smylie, D.E., Hyde, R. Eds.., Structure Dynamics of Earths Deep Interior, Geophys. Monogr. Ser., 46. Am. Geophys. Union, pp. 8389. Anderson, O.L., Duba, A., 1997. Experimental melting curve of iron revisited. J. Geophys. Res. 102, 2265922669. Andrews, D.J., 1973. Equation of state of the alpha and epsilon phases of iron. J. Phys. Chem. Solids 34, 825840. Birch, F., 1972. The melting relations of iron and temperatures in the earths core. Geophys. J. R. Astr. Soc. 29, 373387. Boehler, R., 1992. Melting of the FeFeO and the FeFeS systems at high pressure: Constraints on core temperatures. Earth Planet. Sci. Lett. 111, 217227. Boehler, R., 1993. Temperatures in the Earths core from melting-point measurements of iron at high static pressures. Nature 363, 534536. Boehler, R., Ramakrishnan, J., 1980. Experimental results on the pressure dependence of the Gruneisen parameter. A review. J. Geophys. Res. 85, 69967002. Boness, D.A., Brown, J.M., 1990. The electronic band structures of iron, sulfur and oxygen at high pressures and the Earths core. J. Geophys. Res. 95, 2172121730. Boschi, E., 1975. The melting relations of iron and temperatures in the Earths core. Riv. Nuovo Cim. 5, 501531. Braginsky, S.I., Roberts, P.H., 1995. Equations governing the convection in Earths core and the geodynamo. Geophys. Astrophys. Fluid Dyn. 79, 197. Brown, J.M., Shankland, T.J., 1981. Thermodynamic parameters in the earth as obtained from seismic profiles. Geophys. J. R. Astr. Soc. 66, 579596. Brown, J.M., McQueen, R.G., 1986. Phase transitions, Gruneisen parameter and elasticity for shocked iron between 77 GPa and 400 GPa. J. Geophys. Res. 91, 74857494. Buffett, B.A., Huppert, H., Lister, J.R., Woods, A.W., 1992. Analytical model for solidification of the Earths core. Nature 35, 329331. Buffett, B.A., Huppert, H.E., Lister, J.R., Woods, A.W., 1996. On the thermal evolution of the Earths core. J. Geophys. Res. 101, 79898006. Chen, G.Q., Ahrens, T.J., 1997. Sound velocities of g- and liquid iron under dynamic compression, abstract.. EOS Trans. Am. Geophys. Union 78, F757, Suppl. Dreibus, G., Palme, H., 1996. Geochemical constraints on the sulfur content of the Earths core. Geochim. Cosmochim. Acta 60, 11251130.

k e free electrons. s

/
3

nku l ,

B3.

where k e is the electronic contribution to the total thermal conductivity, the subscript emphasizing that the lattice conductivity, k l , also has yet to be taken into account. Dividing Eq. B3. by Eq. B2.:

ke s

1 nku l m s 3 ne 2t

Using t s lru:

ke s

1 s

3 e2

u2 m .

B4.

Using the theorem of equal partition of thermal and kinetic energy: 1 2 mu 2 s kT .

Eq. B4. is transformed into the WiedemannFranz ratio, giving:

ke sT

2 k s 3 e

B5.

196

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197 McQueen, R.G., 1964. Laboratory techniques for very high pressures and the behavior of metals under dynamic loading. In: Schneider, K.A.G., Hepworth, M.T., Parlee, N.A.D. Eds.., Conference on Metallurgy at High Pressure, Vol. 22. Gordon & Breach, New York, pp. 44132. Melchior, P., 1986. The Physics of the Earths Core: an Introduction. Pergamon, Oxford. Murthy, V.R., Hall, H.T., 1970. The chemical composition of the Earth core: Possibility of sulfur in the core. Phys. Earth Planet. Inter. 2, 276282. ONeill, H.St.C., Canil, D., Rubie, D.C., 1998. Oxide-metal equilibria to 25008C and 25 GPa. Implications for core formation and the light component in the Earths core. J. Geophys. Res. 103, 1223912260. Poirier, J.-P., 1994. Light elements in the Earths outer core: a critical review. Phys. Earth Planet. Inter. 85, 319337. Poirier, J.-P., Shankland, T.J., 1994. Dislocation theory of melting for iron, revisited. In: Schmidt, S.C., Shaner, J.W., Samara, G.A., Ross, M. Eds.., High-Pressure Science and Technology. American Institute of Physics, College Park, MD, pp. 927930. Pollack, H.N., Hurter, S.J., Johnson, J.R., 1993. Heat flow from the Earths interior: analysis of the global data set. Rev. Geophys. 31, 267280. Saxena, S.K., Benimoff, A., 1977. Formation of FeNiSi planetary cores. Nature 270, 333334. Secco, R.A., Schloessin, H.H., 1989. The electrical resistivity of solid and liquid Fe at pressures up to 7 GPa. J. Geophys. Res. 94, 58875893. Sherman, D.M., 1995. Stability of possible FeFeS and FeFeO alloy phases at high temperature and the composition of the Earths core. Earth Planet. Sci. Lett. 132, 8789. Sherman, D., 1997. The composition of the Earths core: new constraints on S and Si versus temperature. Earth Planet. Sci. Lett. 153, 149155. Sprackling, M., 1991. Thermal Physics. American Institute of Physics, New York, p. 305. Stacey, F.D., 1972. Physical properties of the Earths core. Geophys. Surv. 1, 99119. Stacey, F.D., 1992. Physics of the Earth. 3rd edn. Brookfield Press, Kenmore, Queensland, Australia. Stacey, F.D., 1995. Theory of thermal and elastic properties of the lower mantle and core. Phys. Earth Planet. Inter. 89, 219246. Stacey, F.D., Loper, D.E., 1983. The thermal boundary layer Y interpretation of D and its role as a plume source. Phys. Earth Planet. Inter. 33, 4555. Stacey, F.D., Stacey, C.H.B., 1998. Gravitational energy of core evolution: implications for thermal history and geodynamo power. Submitted to Phys. Earth Planet. Inter. Stevenson, D.J., 1981. Models of the Earths core. Science 214, 611619. Stixrude, L., Wasserman, E., Cohen, R.E., 1997. Composition and temperature of Earths inner core. J. Geophys. Res. 102, 2472924739. Tackley, P.J., 1995. Mantle dynamics: Influence of the transition zone. U.S. National Report to International Union of Geodesy and Geophysics, 19911994 American Geophysical Union,

Gardiner, R.B., Stacey, F.D., 1971. Electrical resistivity of the core. Phys. Earth Planet. Sci. 4, 406410. Gilvarry, J.J., 1956. The Lindeman and Gruneisen laws. Phys. Rev. 102, 308316. Glatzmaier, G.A., Roberts, P.H., 1996. An anelastic evolutionary geodynamo stimulation driven by compositional and thermal convection. Physica D 97, 8194. Gubbins, D., 1977. Energetics of the Earths core. J. Geophys. 43, 453464. Hofmeister, A.M., 1998. Mantle values of thermal conductivity from infrared phonon lifetimes: relationships between transport and thermodynamic properties abstract S31A-08.. EOS Trans. AGU Suppl., p. S214. Jephcoat, A.P., Olson, P., 1987. Is the inner core of the Earth pure iron?. Nature 325, 325335. Joos, G., 1958. Theoretical Physics, 3rd edn., Chapter XXIV. Blackie, London. Kerley, G.I., 1994. A new multiphase equation of state of iron. In: Schmidt, S.C., Shaner, J.W., Samara, G.A., Ross, M. Eds.., High-Pressure Science and Technology-1993. American Institute of Physics, College Park, MD, pp. 903906. Kieffer, S.W., 1976. Lattice thermal conductivity within the Earth and considerations of a relationship between the pressure dependence of the thermal diffusivity and the volume dependence of the Guneisen parameter. J. Geophys. Res. 81, 3025 3030. Landau, L.D., Lifshitz, E.M., 1958. Statistical Physics. Pergamon, London. Labrosse, S., Poirier, J.-P., le Mouel, J.-L., 1997. On cooling of the Earths core. Phys. Earth Planet. Inter. 99, 117. Liu, L., 1975. On the g , e , l . triple point of iron and the Earths core. Geophys. J. R. Astr. Soc. 43, 697705. Loper, D.E., 1978a. The gravitationally powered dynamo. Geophys. J. R. Astr. Soc. 54, 389404. Loper, D.E., 1978b. Some thermal consequences of a gravitationally powered dynamo. J. Geophys. Res. 83, 59615970. Loper, D.E., 1984. Structure of the core and lower mantle. Adv. Geophys. 26, 134. Loper, D.E., Roberts, P.H., 1983. Compositional convection and the gravitationally powered dynamo. In: Soward, A.M. Ed.., Stellar and Planetary Magnetism. Gordon and Breech, New York, pp. 297327. MacDonald, G.F.K., Knopoff, L., 1958. On the chemical composition of the outer core. Geophys. J. R. Astr. Soc. 1, 284297. Mao, H.K., Wu, Y., Chen, L.C., Shu, J.F., Jephcoat, A.P., 1990. Static compression of iron to 300 GPa and Fe 0.8 Ni 0.2 alloy to 260 GPa: implications for composition of the core. J. Geophys. Res. 95, 2173721742. Masters, T.G., Shearer, P.M., 1990. Seismological constraints on the structure of the Earths core. J. Geophys. Res. 95, 21691 21695. Matassov, G., 1977. The electrical conductivity of ironsilicon alloys at high pressures and the Earths core. PhD thesis, Lawrence Livermore Laboratory, Rpt. UCRL-52322, Livermore, CA. McDonough, W.F., Sun, S.S., 1995. The composition of the earth. Chem. Geol. 120, 223253.

O.L. Anderson r Physics of the Earth and Planetary Interiors 109 (1998) 179197 Washington, DC, Rev. Geophys. Suppl., July 1995, pp. 275 282. Tackley, P.J., 1996. Effects of strongly variable viscosity on three-dimensional compressible convection in planetary mantles. J. Geophys. Res. 101, 33113332. Tackley, P.J., Stevenson, D.J., Glatzmaier, G.A., Schubert, G., 1993. Effects of an endothermic phase transition at 670 km depth in a spherical model of convection in the Earths mantle. Nature 361, 699704. Tackley, P.J., Stevenson, D.J., Glatzmaier, G.A., Schubert, G., 1994. Effects of multiple phase transitions in a 3-D spherical model of convection in the Earths mantle. J. Geophys. Res. 99, 1587715901.

197

Verhoogen, J., 1980. Energetics of the Earth. National Academy of Sciences, Washington, DC, pp. 5165. Wasserman, E., Stixrude, L., Cohen, R.E., 1996. Thermal properties of iron at high pressures and temperatures. Phys. Rev. B 53, 82968309. Zharkov, V.N., 1962. Physics of the Earths core. Inst. Earth Phys. Acad. Sci. USSR Engl. Transl. 20, 187. Zharkov, V.N., Trubitsyn, V.P., 1978. Physics of Planetary Interiors. In: Hubbard, W.B. Ed.., Astronomy and Astrophysics Series, Vol. 6. Pachart Publishing Housing, Tucson, AZ, p. 57.

S-ar putea să vă placă și