Vous êtes sur la page 1sur 95

8 July 2009 Physics of the Earth, F D Stacey and P M Davis Solutions to problems in Appendix J

1.1 (a) Consider a mass element dm at (x,y,z) in a thin spherical shell of radius r, centred at the coordinate origin, so that r2 = (x2 + y2 + z2). The moments of inertia of dm about the x, y and z axes are dI x = ( y 2 + z 2 )dm; dI y = ( x 2 + z 2 )dm; dI z = ( x 2 + y 2 )dm;
2 2 2 2 so that dI x + dI y + dI z = 2( x + y + z )dm = 2 r dm The same relationship applies to all mass elements and therefore to 2 their total, I x + I y + I z = 2mr But by symmetry I x = I y = I z so that for a spherical shell

I = (2 / 3)mr 2 (b) Now let this be an element of a uniform sphere, so that dm = 4 r 2dr and 2 8 R 8 I = r 2dm = r 4dr = R 5 . 3 3 0 15 3 Substituting M = ( 4 3) R we have the result for a uniform

sphere:

I = ( 2 5 ) MR 2

(c)

Let = 0 ( R / r ), where 0 is the surface density (at R)


M = 4 r dr = 4 r 20 ( R / r )dr
2 0 0 R R

= 2R 0
3

2 I = r 2 dm 3 0 where dm = 4r2 dr = 4R0rdr is the mass of a spherical shell of radius r.


R

8 2 I = R0 r 3dr = 0 R 5 3 3 0

I /MR2 = 1/3, compared with 0.3307 for the Earth.

1.2

(a) The total mass and moment of inertia are obtained as a sum of the values for a sphere of radius R and density and a superimposed sphere of radius R/2 and density (f - 1)
4 4 R M = R 3 + ( f 1) = R 3 ( 7 + f ) 3 3 2 6 3 R 2 5 2 4 2 4 R I = R 3 R 2 + ( f 1) = R ( 31 + f ) 5 3 5 3 2 2 60 ( 60 )( 31 + f ) = 31 + f I = 2 MR ( 6 )( 7 + f ) 10 ( 7 + f )
2 For I MR = 0.3307, f = 3.403 (Earth) 0.365 2.06 (Mars) 0.391 1.25 (Moon) (b) Let the core have radius kR; then 4 4 4 M = R 3 + (kR)3 2 = R 3 (1 + 2k 3 ) 3 3 3 8 5 8 8 5 5 I= R + ( kR ) 2 = R (1 + 2k 5 ) 15 15 15 5 2 (1 + 2k ) I = MR 2 5 (1 + 2k 3 )
3

(Earth) For I MR 2 =0.3307, k = 0.5480 0.365 0.385 (Mars) 0.391 0.230 (Moon) The algebra admits alternative, very high, values of k that are incompatible with plausible densities. 1.3 As a simple approximation, assume each layer to have uniform density, equal to the average of values at the top and bottom of the layer. Then, starting from the outside we can obtain the increments in mean density for successive layers. This allows us to use the method of solving Problem 1.2, by taking the sum of moments of inertia for a series of uniform spheres with densities , as listed r rS 0 0.04 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.95 1.0 160000 141000 89000 41000 13300 3600 1000 350 80 18 2 0.4 0 150500 115000 65000 27150 8450 2300 675 215 49 10 1.2 0.2 35500 50000 37850 18700 6150 1625 460 166 39 8.8 1.0 0.2

Thus I =

5 8 5 5 5 rS (1.0 ) 0.2 + ( 0.95) 1.0 + ( 0.9 ) 8.8 15 5 +.... + ( 0.04 ) 35500 5 8 = ( 6.96 108 ) 254 = 7.0 1046 kg m 2 15 The mass of the model is also an overestimate 4 3 3 3 M = rS3 (1.0 ) 0.2 + ( 0.95) 1.0 + .......... + ( 0.04 ) 35500 3 4 8 3 30 = ( 6.96 10 ) 1640 = 2.3 10 kg 3 Thus, if we reduce the moment of inertia estimate by the factor needed to bring the mass of the model into line with the observed mass, we have I sun 6.0 1046 kg m 2 For a uniform sphere of the same mass and radius I 0 = 0.4 MrS 2 = 3.85 1047 kg m 2

I 5.7 1046 = = 0.148 Using Allens value, I 0 3.85 1047 1.4 The Doppler shift is seen by a stationary observer. To an observer rotating with the Sun the radiation would not appear Doppler shifted. Thus the radiation exerts no retarding torque on the Sun. The angular momentum lost by radiation of energy dE is the angular momentum of a spherical shell of radius R and mass dE/c2 (c is the speed of light). Thus the rate of loss of angular momentum is 2 daS dt = R 2 dE dt . c 2 3 2 dE dt = 4rE S = 3.825 1026 W , where rE is the radius of the 2 Earths orbit and S = 1360W m is the solar constant. Thus daS / dt = 3.94 1021 kg m 2s 2 compared with the present angular momentum aS = I = 5.7 1046 2.87 106 = 1.6 1041 kg m 2s 1 1 daS = 2.4 1020 s 1 = 7.6 1013 /year aS dt This would not be a significant effect, even in the life-time of the Sun. Corresponding slowing of rotation depends upon the radial movement of mass to compensate for the mass lost from the surface. If the radius of gyration, k = I M , is maintained constant then the moment of inertia is proportional to the mass and its rate of change is given by 1 dI 1 dM 1 dE . = = = 2.14 1021 s 1 2 I dt M dt Mc dt 2 and since aS = Mk ,
3

1 d 1 das 1 dM = = 2.2 1020 s 1 = 6.9 1013 year 1 dt as dt M dt 1.5 The cross-sectional area intercepting radiation is r , so energy 2 2 absorbed is r S . Area radiating is 4r , so energy lost is 4r 2T 4 . Equating these 1 T = ( S 4 ) 4 = 278K (5o C)
2

1.6

In this case the same area (A) receives and radiates energy so 1 AS = AT 4 ; T = ( S ) 4 = 393K (120o C)
Power/unit area at the solar surface is P = S (rEarth orbit / RSun ) 2 = TS 4
2 TS = S ( rE / RS ) / = 5770K
4 1

1.7

1.8

Power/unit area received by planet ( rorbit ) Power radiated per unit area T 4 Therefore T 1 rorbit

1.9

Luminosities of the planets are proportional to the squares of their radii and inversely proportional to squares of distances from the Sun. Both quantities are listed, relative to Earth values, in Table 1.1. The order is Planet Jupiter Venus Earth Mercury Saturn Mars Moon Uranus Neptune Pluto Luminosity, relative to Earth, ( Rplanet Rorbit ) 4.45 1.726 1.000 0.979 0.918 0.122 0.074 0.043 0.016 2 105
2

1.10 If we assume the flux of dust particles towards the Sun to be in equilibrium with their production in the asteroidal belt , then they have the same size distribution. However, the radial speed of their motion towards the Sun, obtained by differentiating Eq. (1.25), is 1 1 proportional to d and therefore to m 3 , where d is particle diameter and m is mass. The number density in space is the ratio flux/speed, and is therefore 1 dN m n dm m 3 = m ( n 1 3)dm

The Earth samples this distribution. 1.11 On free fall from infinite distance the meteoroids gain kinetic energy equal to the gravitational potential energy lost, so the speed at the Earth is v, given by 1 2 GM Sun GM Earth v = + 2 rorbit rEarth
2 This is compounded with the Earths orbital speed, ( GM Sun rorbit ) which is perpendicular to the free fall into the Sun. Since the Earths gravity gives the minor component of v above and so deflection by the Earth is small, we can add the orbital velocity in quadrature also. Thus 1

vTotal

3GM Sun 2GM Earth 1 = + = 52787m s rEarth rorbit


2

2GM Earth 1 = 10883m s Escape speed from the Earth is vescape = rEarth which is about 20% of the speed of escape from the solar system when the Suns gravity is also included.
2

2.1

m + 0.1mchondrites Core mass = 0.326 = irons Earth mass mirons + mchondrites whence mirons = 0.251 mTotal Let r = core radius / total radius Then, = r 3core + (1 r 3 ) mantle = 5280kg m 3 Assuming uncompressed densities (low pressure phases), core = 7800kg m 3 , mantle = 3300kg m 3 , r = ( m ) / ( c m ) 3 = 0.76
1

2.2

2.3

The simple supposition that the isotopes with the smallest (or most negative) mass excesses would be the most abundant is not generally correct. The table gives some clues to the reason for this. First, we note that 1H and 3He are the only isotopes with fewer neutrons than protons. This is true for the whole periodic table and not just for the elements selected in Table 2.1. Generally the proportion of neutrons increases with atomic number and 40Ca is the heaviest isotope for which the number of protons is even equal to the number of neutrons. The reason is that the mutual electrostatic repulsion of the protons reduces the binding energy, so that the total binding energy is increased by diluting the protons with neutrons. This effect increases with atomic number and all the heavy elements are neutron-rich. As noted in Section 2.1, isotopes with atomic masses that are multiples of 4 are favoured and, of those in the table, all except 56Fe are multiples of 4He, reflecting the strong binding of the 4He structure. In comparing these nuclei with one another we have a valid indication of the general 5

variation of binding energy with nuclear size, but we need caution in comparing them with the less abundant isotopes of the same elements, all of which have extra neutrons. The neutron mass (1.008665u, where u is the atomic mass unit, as listed in Table A1) exceeds the sum of the proton and electron masses (1.007276u), so that the added neutrons increase the nuclear mass excess without any consideration of nuclear binding. In spite of this, the binding energies apparent from the mass excesses appear to favour many of the heavier, more neutron-rich isotopes, reinforcing the argument that dilution of the neutrons reduces the repulsive energy. This does not very effectively answer the question, why is there not a more obvious correlation between binding energy and abundance? There is no simple answer. The relative abundances are consequences of multiple processes of nuclear synthesis, as illustrated by the fact that isotopic ratios in some of the pre-solar system grains found in carbonaceous chondrites are quite different from the bulk of the solar system (Section 4.5). Abundances of isotopes reflect the processes that produced them, with binding energy introducing only a statistical bias. 3.1 (a) If the density is , then at radius r, accumulated mass m = ( 4 3) r 3 , the gravitational potential energy released by adding a shell of mass dm = 4r 2dr is Gmdm 16 2 2 4 dE = = G r dr r 3 Integrating from r = 0 to R and substituting M = ( 4 3) R 3 , E = ( 3 / 5 ) GM 2 / R (b) See Eq. (4.1). 3.2 If the cosmic ray flux is , the rates of production of S,R are dS dt = S P, dR dt = R P , where P is the concentration of parents, from which they are produced. For uniform over exposure time t, S = S Pt , but if t is long compared with the half-life of R, then an equilibrium concentration is established, such that dR dt = R P R = 0 so R = R P /

Thus S R = ( S R ) t Eq.(1.8) follows and Eq.(1.9) assumes an effective production rate of S proportional to ( S + R ) .
t Given the number remaining at time t, N = N 0 e , the number decaying in the interval t to (t + dt ) is dN * = dN = N 0 e t dt The average life is therefore

3.3

t =

1 N0

t =0

tdN * = te t dt =
0

3.4

(a) by Eq.(3.7),

d ( 40 Ar ) d ( K)
40 86

= 0.105 ( e K t 1) = 0.098 0.005

9 whence tAr = (1.20 0.05 ) 10 years.

By Eq.(3.12),

d ( 87 Rb d ( 207 Pb d ( 206 Pb

d ( 87 Sr

Sr )

86

Sr )

= e Rbt 1 = 0.0215 0.0007

9 whence tSr = (1.50 0.05 ) 10 years. 204 204

By Eq.(3.15),

Pb )

Pb )

U e235t 1 = 0.090 0.005 238 U e238t 1


235

tPb = (1.43 0.10 ) 109 years. whence (b) Disparities are well outside the formal errors of individual ages and the Ar age is lowest, implying an argon loss which (with the small error) is consistent for all of the minerals used. This suggests a 9 reasonably complete argon loss at about 1.2 10 years rather than partial loss later. The Sr and Pb ages are not clearly different, but consistent with a slight U-Pb diffusion. In this case it must be noted that the error of the Pb age is greatest, implying variability of diffusion. We must therefore take the Sr age as the most reliable estimate of the formation age and treat the Ar age as evidence of a 9 metamorphic event at about 1.2 10 years, although argon diffusion without appreciable reheating cannot be ruled out.

3.5

As noted in Section 4.3, the Pb-Pb isochron gives an age = ( 4.54 0.03) 109 years for the meteorites. The Rb-Sr isochron gives eRb 1 = 0.0664 (see Eq.(4.4). Thus Rb = ln(1.0664) year 1 = (1.416 0.009) 1011 year 1 9 (4.54 0.03) 10

4.1

The rate of production of 129 I during the interval was 127 I / where 127 I represents the total production of that isotope. The net rate of production of 129 I during this interval is therefore 127 d 129 I I) = 129 I ( dt so that I 129 I 129 I 1 which gives 127 = (1 e ) I
0 0

( I ) d ( 129 I )
129

127

= dt

The Xe eventually produced in the meteorite corresponds to the 129 I remaining at time t after cessation of synthesis, i.e. 129 Xe 129 I e t = 127 = (1 e ) 129 I I +t
1 For << ,

129

129

Xe e t = (1 1 + ..) = et 127 I

4.2

235 U 235 U e 235t 235 U t = 238 . 238t = 238 e( 238 235 ) 238 U now U 0 e U 0 9 235 For t = 4.67 10 years and ( U 238 U ) = 0.007254

235

238

U ) = 0.349
0

now

4.3 4.4

A formed first, by one half-life of

129

I, 1.6 107 years.

5 Putting R = S in Eq.(1.9), with S R = 4.52 10 and t1 2 = 12.25 years (Table H.2) t = 4.0 106 years.

5.1

Using numbers of atoms, the present production rates are d ( 40 Ar ) / dt = 0.105. 40 K. K d ( 4 He ) / dt = 8. 238 U. 238 + 7. 235 U. 235 + 6.Th. Th
=
238

whence

d ( 40 Ar dt ) d ( He dt )
4

U ( 8 238 + 0.05145 235 + 22.8 Th )


238

= 0.02406 ( 40 K

U)

Conversion to the (K/U) mass ratio by Eq.(5.1) gives Eq.(5.2). 9 Accumulations over = 4.5 10 years are 40 Ar=0.105. 40 K ( e K t 1)
4

He = 238 U 8 e238 1 + 0.05145 e235 1


Th
40

Total

Ar

( ) + 22.8 ( e 1) He ) = 0.0647 ( K
40 2

238

U)

which gives Eq.(5.3) when converted to the (K/U) mass ratio. 5.2 (a) Ar in the Earth is given by dAr = 0.105K 0e t .Ar dt t where K 0 = Ke is the initial amount of 40 K . t Let Ar = ue dAr du t = .e ue t Then dt dt So that, combining (1) and (3) with substitution for Ar by (2)

(1) (2) (3)

du = ( ) u + 0.105K 0 dt Integrating from u = 0 at t = 0 0.105K 0 ( )t e u= 1 ( ) So, in the Earth 0.105K 0 t t e e ArE = ( - )


0.105K ( )t 1 e ( - ) (b) The total Ar in the Earth and atmosphere is ArT = 0.105K ( et 1) =

and the atmospheric component is therefore ArA = ArT ArE 0.105K = ( ) et e( )t + - ) ( (c) (d)
t t ArA ( ) e + e = ArE ( e t e t )

In a form more convenient for numerical calculation t ArA ( ) ( e 1) = 1 ArE e( )t 1

9 Putting ArA ArE = 1 at t = 4.5 10 years gives = 3.187 1010 year 1 10 1 so = 2.357 10 year

5.3 The gravitational energy released by accretion of any spherically symmetrical body is calculated by considering the work done in peeling off layers and removing them to infinity. In this problem it is convenient to start with the following parameters: 4 3 Total mass of Earth, M = R 3 4 3 Mass of core, M C = (0.55R) .2 = 0.33275M 3 Mass of mantle, M M = M M C = 0.66725M Volume of mantle, 4 4 VM = R 3 (0.55 R)3 = R 3 (0.833625) 3 3 Density of mantle = M M / VM = 0.80042 The gravitational energy release of a uniform (undifferentiated) Earth is the standard result 3 GM 2 E0 = = 2.24 1032 J 5 R 9

Similarly the gravitational energy released by formation of the core alone is 2 3 GM C EC = = 0.201314 E0 5 0.55R In calculating the gravitational energy released by addition of the mantle to the core we need the mass inside radius r where r 3 (0.55 R )3 0.55R r R . This is m(r ) = M C +M M 3 3 R (0.55 R )
3 3 = 0.19958M + 0.80042 M r R The mass element of the mantle is dm = 4r 2 M dr = 0.80042 4 r 2dr so that the gravitational energy released by accumulation of the mantle on an already established core is

EM =

Gm( r )dm r4 = 4 GM 0.159748r + 0.640672 3 dr r R 0.55 R

= 4GM R 2 0.177398 2 = 0.532193 GM R = 0.886989 E0 The energy released by separation of the core from a homogeneous Earth is E = EM + EC E0 = 0.0883E0 = 1.98 1031 J. Using the heat capacity = 9.931027 J K-1, the average temperature rise is T = E / = 1994 K For the result of a more rigorous calculation, with allowance for self-compression, see Table 21.1. 6.1 With no change in size or shape the angular velocity of the Sun would be 0 = 3.15 1043 / 5.7 1046 = 5.5 104 rad s 1 Centrifugal force per unit mass at the equator would then be 2 0 R = 213m s 2 Equatorial gravity is (ignoring rotation and ellipticity) GM / R 2 = 274m s 2 which is slightly greater, so that the Sun would appear marginally stable, but the spherical approximation is a poor one. Even the calculation of ellipticity in Section 6.3 is inaccurate with large ellipticity. A calculation similar to that in Section 8.5, but with nonuniform density is needed for an accurate result. However, we can see what to expect from the equations for slight ellipticity in Section 6.3. Using Eq.(6.19), with f = (1 c a ) , a, c being the equatorial and polar radii, we can impose the following conditions. Conservation of volume: a 2 c = R 3 = ( 6.96 108 m ) ,
3

10

so that c a = (1 f ) = ( R a )

( C A) / Ma 2 = ( C A) / C . C / Ma 2
2 2 where C Ma = 0.4 0.148 = 5.92 10 by Problem 1.3

and C a 2 , A ( a 2 + c 2 ) / 2 , so that

( C A) / C = ( a 2 c 2 ) / 2a 2 = f (1 f 2 )
2

2 2 and C = 0C0 or a = 0 R , because the angular momentum is


3 fixed, and therefore = 0 (1 f ) . Thus, Eq.(6.19) becomes

2 f = 5.92 102 f 1/ (1 f ) + (1 f ) / 2 + 0.384 (1 f ) 3 which gives f=0.276. Although this is not an accurate result it demonstrates that by allowing the Suns ellipticity to be selfadjusted, its stability is increased.
4

6.2

gSun M R2 = S S 2 = 2.20 g Earth M E RE The Suns gravity is twice as strong, so the net acceleration of the Moon is towards the Sun, not the Earth. But the Earth is also accelerated towards the Sun, at almost the same rate and so the two bodies stay together in orbit about the Sun. Use Eq.(6.19), where, for a uniform ellipsoid C = 0.4Ma 2 , A = 0.2 M ( a 2 + c 2 ) , so that
J2 =
c2 CA = 0.2 1 2 Ma 2 a

6.3

also M =

4 2 c a c and = (1 f ) 3 a With these substitutions, Eq.(6.19) becomes 1 1 f 32 f = 0.2 f ( 2 f ) + + (1 f )2 2 8G For small f this reduces to 152 f = 1 for fixed 16G 2 For a uniform ellipsoid, C Ma = 0.4, so that Eq.(6.39) reduces to fH = (5 4) m

2 2 As in the substitutions above, m = a c GM = 32 4G and therefore f H is the same as f calculated above.

6.4

As in the solution to Problem 6.3, the moments of inertia of a uniform ellipsoid are

11

C = 0.4Ma 2 ;

A = 0.2 M ( a 2 + c 2 )

C A 1 c2 = 1 2 so H = 2 a C

c c = 1 f but f = 1 ; a a 1 2 H = 1 (1 f ) = f (1 f 2 ) 2 If the ellipsoid consists of an outer layer of density and an inner region of the same flattening with density , then we can consider moments of inertia of a large uniform ellipsoid of density , superimposed on a smaller one of density ( - ). Using primes for parameters of the smaller one C = 0.4Ma 2 + 0.4M a2 A = 0.2 M ( a 2 + c 2 ) + 0.2M ( a2 + c2 )
2 2 2 2 C A 0.2M ( a c ) + 0.2M ( a c ) = so that H = C 0.4Ma 2 + 0.4M a2 0.2Ma 2 (1 c 2 a 2 ) + 0.2 M a2 (1 c2 a2 ) = 0.4Ma 2 + 0.4M a2 But since the ellipticities are the same, (1 c2 a 2 ) = (1 c2 a2 ) = f ( 2 f ) is a common factor in the

numerator. Thus 1 H = f ( 2 f ) = f (1 f 2 ) 2 and the analysis can be extended to any number of superimposed homologous ellipsoids. 6.5 2 a 3 = 4.585 103 GM By Eq.(6.2) 3 1 f = J 2 + m = 5.02 103 2 2 2 Making ( C / Ma ) the subject of Eq.(6.39) m=
12 C 2 25 m = 1 1 = 0.365 Ma 2 3 5 2 f

6.6

Equating equatorial centrifugal force to gravity 4 2 c R = GM / R 2 where M = R 3 3 12 12 4 so that c = ( GM / R 3 ) = G 3 With uniform density, the critical angular momentum is

12

2 MR 2 c 5 13 Substituting R = ( 3M / 4 ) , this gives Lc = 0.32G1 2 M 5 31 6 Lc = I c = For the Earth, I = 5.861033 kgm2s-1, which means a numerical factor of 0.015 in the expression for Lc, a low value. This is reasonably explained by tidal friction. If the lunar orbital angular momentum were added, the numerical factor would be 8.8, still not anomalously high, although we argue in Section 1.15 that this is irrelevant. This is a contradiction of an argument, used to support the giant impact hypothesis for the origin of the Moon, that the angular momentum of the Earth-Moon system requires a special explanation, such as a giant impact on the Earth. For Jupiter, = 1.7710-4 rad s-1 and I 0.236MR2 = 2.831042 kg m2 (Ward, W. R. and Canup, R. M., 2006, The obliquity of Jupiter. Astrophys. J. 640: L91-L94), giving a numerical factor 0.070. Earlier estimates of I give a numerical factor ~ 0.075. 6.7 (a) At low latitudes, sin , sin 2 2, the latitude variation of gravity (Eq.6.38) can be simplified to allow the latitude (in radians) to be written in terms of g = ( g 9.780327 ) / 0.051628
2 12

Since it is that is calculated from g there is a sign ambiguity in , which could be in either hemisphere. Thus a graph of vs day number is a pair of lines which are straight at low latitudes and meet at the instant when the equator was crossed. This was 0.25 day before the noon measurement on Day 10, so the equator was crossed at 6 a.m. on Day 10. The gradient of the graph at low latitudes is 0.075 radian/day. Multiplying by the equatorial radius, a=6378km, we obtain the ships speed, 480 km/day (20km/hour or 5.56m s-1). (b) When the ship turns east, it adds its speed v= 5.56 m s-1 to its rotation with the Earth. At latitude the radius of its circular motion is acos , so that the centripetal acceleration increases from 2 2 0 a cos to ( 0 + v / a cos ) a cos where 0 = 7.292 105 rad s 1 is the Earths rotational speed. Neglecting the small second order term in v, the difference is 20 v This is an acceleration perpendicular to the rotational axis so the component normal to the surface gives the apparent change in g g = 20v cos = 8.1104 cos m s 2 7.1 To use Eq.(7.12), we need the orbital angular velocity of Mars. Eq.(B.23) relates this to the orbital radius, which by Table 1.1 is 13

7 1 1.524 times the Earths orbital radius. S = 1.99110 rad s for

We note also that H = ( C A ) / C = J 2 / ( C / Ma 2 ) by Eq.(7.3) so that p = J2 3 2 S . cos = 1.08 1012 rad s 1 2 ( C / Ma 2 ) = 7.06 arc sec/year 7.2 Assuming mass M = 1010kg, mean speed v = 15ms and lane separation W = 30m , the daytime angular momentum of North 12 2 1 American traffic is MvW = 4.5 10 kg m s . This is insignificant compared with the Earths angular momentum C = 5.86 1033 kg m 2s 1 . By giving the Earth an angular momentum opposite to that of the traffic, to conserve the total, and o assuming the traffic to be centred at colatitude = 50 , the instantaneous rotational pole is shifted away from North America by an angle , such that = , where tan = C sin / ( C cos MvW )
1

of this value for Mars. Thus the Earth and (1.524 ) 2 ( Mars ) = 1.12 1014 rad 2 s 2 . S
3 2

22 16 so that = ( MvW / C ) sin = 6 10 rad = 1.2 10 arc sec

7.3

(a) For the inner sphere rotational speed i is self-adjusted so that there is no torque about its axis. The vector difference in angular velocities is therefore perpendicular to i and this requires i = 0 cos (b) The differential rotational speed is = 0 sin At to the pole of this differential rotation its surface speed is v = .r sin and the consequent shear stress across the fluid is = v / d = ( r / d ) sin Taking the element of surface area to be dA = 2r sin d with the power dissipation per unit area, v the total power dissipation is
2

P = vdA = 2r 4 () 2 / d sin 3 d
0

= 8r ( ) / 3d 4 2 2 = 8r 0 sin / 3d
4 2

(c)

= 0 sin = 4.4 1010 rad s 1 P = K R ( ) = 6.1109W


2

14

(d) This would be an apparent contribution to tidal friction, but one which progressively reduces the obliquity of the ecliptic, that is the misalignment of rotation and orbital motion. However, it is small enough not to be observed, or to have a noticeable effect even over geological time. The thickness of a core layer according to part (b) would be about 0.4 mm. 7.4 If the elastic strain due to a wobble of amplitude is = 1.0 103 , then the strain energy is 1 1 ES = 2V = .106 2V 2 2 where is the mean rigidity of the Earth and V is its volume. The wobble kinetic energy is given by Eq.(7.19), so that ES 106 V = = 0.114 EK AH 2 Note that an estimate of is required. The value used here, 147GPa, is the volume average for the Earth obtained by numerically integrating the data in Appendix F. EK is designated EW in Eq.(7.19), but is seen here not to be the total energy. Thus ES / ( EK + ES ) = 0.102 . By Eq.(7.21)
W = H = CA C A Ma 2 = C Ma 2 C

7.5

2 where (C A) / Ma = J 2 and C / Ma 2 = 0.3307 is a constant specified by the radial density profile. By the equations in Section 2 6.3, equilibrium flattening and J 2 both vary as . Thus W 3 . To make the wobble period (now 430 days) coincide with the year (365 present days), would need to be increased by the factor (430 / 365)1 3 , making = 7.7 105 rad s 1 . This would have occurred more than 200 million years ago, as the present rate of slowing (Eq.8.31) is faster than in earlier geological periods.

7.6

(a) The change in moment of inertia caused by the removal of mass m from a spherical shell of radius R and deposition at Rcos from the axis is m(Rcos) 2 (2/3)mR2 For this to be zero, as required for no LOD change cos2 = 2/3; = 35 This neglects the displacement of the centre of mass and realignment of the axis.

15

(b)

The centre of mass is shifted towards m by a distance Rm/(M + m) and the new symmetry axis passes through this point, making an angle with the original axis, where is to be determined and M, R are the mass and radius of the Earth. The distance from the original m R cos( ) and the distance of m centre to this new axis is m+M M R cos( ) . Noting that the moments of from this axis is m+M inertia of the Earth, ignoring m, are C, A about the polar and equatorial axes, the total moment of inertia of the whole becomes 2 m cos( ) + I = C cos 2 + A sin 2 + M R m+M
M cos( ) m R m+M
2

The first two terms give the moment of inertia of the Earth (without m) about an axis through its centre but parallel to the new axis. The third term applies the parallel axis theorem to add the moment of inertia of M about the new axis and the fourth term is the moment of inertia of m about this axis. We require the new axis to be the axis of maximum moment of inertia, that is dI/d = 0.
dI / d = 2C cos sin + 2 A sin cos + MmR 2 2 cos( ) sin( ) M +m

16

= (C A) sin 2 +

MmR 2 sin [ 2( ) ] M +m

We assume to be small, so that sin2 is very small compared with cos2, which is close to unity, and MmR 2 dI / d (C A) sin 2 + sin 2 cos 2 = 0 M +m R 2 Mm sin 2 So that tan 2 M +m C A and, with m << M, and (C A) = J2MR2 1 R2m 1 m = sin 2 = sin 2 2CA 2J 2 M ` The original axis is stabilized by the equatorial bulge, as represented by J2, so that is small. It is a maximum if = 45. (c) With sin2 = 1, J2 = 10-3, M = 61024 kg, if we require > 10-2 arc sec = 4.810-8 radian, to be noticeable on a record of polar motion, the solution to part (b) means m > 2 J2M = 5.81014 kg, which corresponds to 580 km3 of water. This is much more than the capacity of any man-made lake and would represent a level change of about 1.5 m on the largest natural lake, the Caspian Sea, or 7 m on Lake Superior. 8.1 With respect to the Earths axis the second term in Eq.(8.1) has an asymmetry because r is measured from an axis through the combined centre of mass of the Earth and Moon. This asymmetry is cancelled 1 by that in the first term, as becomes apparent when ( R ) is substituted by Eq.(8.3), leaving the geocentric rotational term (the third term in Eq.8.8). Note that the cos term formally disappears by the substitution of Eq.(8.5). 8.2 The Earth is deformed tidally to a prolate ellipsoidal form in response to potential W2 (Eq.8.9), being the angle to the EarthMoon axis. The additional potential due to the Earth itself is then that arising from its ellipticity, as may be expressed by the second term of Eq.6.13, where we now put = ( / 2 ) , being the angle to the plane normal to the Earth-Moon axis, so that (C-A) is negative. Then k2 is defined as the ratio of the angular dependence (second term) in V to the excitation (W2), so that C A R M C A R k2 = = 2 ma r m Ma 2 r
3 3

17

The density is assumed to be uniform so that ( A C ) Ma 2 is very simply related to the semi-axes, a and c (noting that c > a in the present situation), because a2 + c2 C = 0.4 Ma 2 and A = 0.4M 2 Then
c2 AC c c = 0.2 2 1 0.4 1 , for 1 << 1 n 2 Ma a a a
3

c M R Therefore k2 = 0.4 1 a m r GM GMa 2 c 3 1 and V = 0.4 3 1 cos 2 r r a 2 2 c The elongation 1 is self-adjusted so that the total potential a (V + W2 ) is constant over the surface, and we can equate the values of (V + W2 ) at (r = c, = 0) and (r = a, = ) : 2

GM 0.4M Ga 2 3 c c

GM Ga 2 0.4M c m 1 + 3 = + a 2 a3 a R

c m 1 + 3 a R

which gives
a3 1 3 a3 m c a 1 0.4 3 + = 3 a c c 2 2 R M 2 c Since 1 is very small, the square bracketted factor is . 5 a c Thus substitution for 1 in the expression for k2 , with neglect a 3 of the distinction between r and a, gives k2 = . 2 8.3 Using Eq.(8.14) and allowing for the additional tidal potential due to the deformation itself, the peak-to-peak amplitude of the tide is a 3 ma = h(1 + k ). a 2 M R where M, a are now the mass and radius of Mercury, m, R are the mass and distance of the Sun and h, k are the Love numbers (see Section 8.2). Since Mercury is smaller than the Earth and rigidity increases with pressure, we expect Mercury to deform more readily than Earth. Thus we may take as extreme values of the numerical factor the values for the Earth and a uniform fluid planet 0.75 < h(1 + k ) < 2.5 giving 0.5 106 < a / a < 1.7 106
3

18

This strain amplitude is much less than any plausible yield point, but more than 10 times the tidal strain of the Earth, making tidal dissipation (per unit volume of solid) more than 100 times as great (for the same rotation rate). 8.4 Present angular momenta are 33 2 1 Earth rotation: C = 5.860 10 kg m s mM R 2 L = 4.872C (See Eq.8.38) Lunar orbit: aL = m+M 2 3 (a) For a circular orbit, L R = G (m + M ) , so aL = [G /(m + M ) ] mMR1 2
12

The original orbital angular momentum is assumed to have been aL0 = aL C = 3.872C so that aL0 aL = 3.872C = 0.7947 4.872C

and therefore R0 / R = aL0 / aL

= 0.632

8 making R0 = 2.43 10 m (38 Earth radii).

(b) On a parabolic orbit the total orbital energy is zero, that is 1 mM 2 2 GmM 0 R0 =0 2 m+M R0 where R0 is the distance of closest approach and 0 is the angular speed of the Earth and Moon about their common centre of mass. Thus in this case 2 3 0 R0 = 2G (m + M ) and the angular momentum is mM 2G 12 aL = 0 R02 = mM .R0 m+M m+M R0 is half of its value in the case of the circular orbit, i.e. so that R0 = 1.2 108 m (19 Earth radii). 8.5 The speed of the resonant tidal wave is ( L )a = gh R = L = 3.467 105 rad s 1 (1) Conservation of angular momentum gives Eq.(8.38) mM C + R 2 L = 5.872C 0 (2) M +m and the third equation relating , L and R is Keplers third law 2 R 3 = G ( M + m) (3) L Eliminating R from (2) and (3) and normalizing L , to 0
12

19

L G 2 M 3 m3 5.872 = = 4.247 4 0 0 ( M + m)C 30 Substituting for L by (1) we can solve for / 0 and then by (1) for L / 0 , from which Eq.(3) gives R = 0.50290 = 3.667 105 rad s 1 L = 2.002 10 6 rad s 1 R = 4.652 108 m = 1.21R0 8.6. Eq. (8.20) gives the torque exerted on the lunar orbital motion, and on the Earths rotation, arising from dissipation in the Earth of the energy of the tide raised by the Moon. Here we reverse the situation and consider the tide raised in the Moon by the Earths gravity. The same equation applies, but with m now the Earths mass instead of the Moons mass and a is the radius of the Moon, not the Earth. We can take k2 0.3 without serious error, but is more uncertain. It is much less than for Earth tides, for which most of the dissipation occurs in the oceans, giving = 2.9, and the dryness of the Moon suggests a high value for its anelastic Q, with = tan -1(1/Q). Taking = 0.00329 rad (0.2), and substituting other numerical values, including the present Earth-Moon distance for R, the torque is L=3.721016 Nm. The moment of inertia of the Moon is I = 0.393 Mr2 = 8.721034 kgm2 (see Table 2.1). The rate of slowing of the rotation is d/dt = L/I = 4.2710-19 rad s-2 and if we assume the Moons initial rotation rate to be i = 7.310 -5 rad s -1 (a 24 hour rotation period), the rotation will stop in i/(d/dt) = 1.71014 s = 5.4106 years. The slowing would have been much more rapid when the youthful Moon was closer. Note that the slowing is a linear process and not an exponential decay of , because the torque is independent of , so rotation reaches a sharp and unambiguous end point. From these numbers we can see how inevitable it is that the Moon presents a constant face to the Earth. 8.7

We apply the sine rule (e.g. Dwight, 1961, item 410.02) to the triangle formed by the vectors k2, kS, kO a b c = = sin A sin B sin C where a, b, c are the lengths of the sides opposite angles A, B, C respectively. Thus

20

ko kS kS k = 2 = = o o o sin 2.9 sin sin(180 2.9 ) sin( + 2.9o ) These are two equations with the two unknowns, kO, the potential Love number for the ocean tide, and the phase lag, (180 ), relative to the forcing potential, which is very close to kS. The solution is kO = 0.0547, = 13, that is a phase lag of 167. To see what this means in terms of the marine tide, we compare the tideraising potential of the Moon (Eq. 8.9) with the potential due to a surface layer of water deformed to the same shape. This can be obtained by considering a uniform ellipsoid with the density of sea water, W = 1025 kg m-3, and calculating the gravitational potential from its moments of inertia by Eq. (6.13). With equatorial and polar radii a and c the moments of inertia are C = 0.4MWa2 and A = 0.2MW (a2 + c2), that is (C A) = 0.2MW (a2 c2), where MW = 1.11024 kg is the mass of a volume of sea water equal to the volume of the Earth. Only the surface deformation contributes to (C A); there is no contribution from a sphere. Thus, we can equate the potential inequality of the water ellipsoid to kO times the deforming potential of the Moon. This means 0.2M W (a 2 c 2 ) ko mMoon a 2 = a3 R3 a and c differ so slightly that this reduces to k m a4 a c = o Moon 3 = 0.15m 0.4M W R This is the peak-to-peak amplitude of the marine tide, averaged over the surface of the Earth (including the land areas where it does not occur at all). It is much smaller than the observed tide because tidal phase is locally very variable and the shape of the ocean tide is not well represented by an ellipsoid. This calculation gives the averaged ellipsoidal component. 9.1 (a) Taking the density of ice to be 920 kg m-3, the effect of melting 14 1000km3 of Greenland ice is to remove a mass m = 9.2 10 kg from a point at 75N (=15) and to distribute it uniformly (as we assume) over the surface of the Earth. (i) Assuming the melting to occur in a few summer months, that is in a time short compared with a half wobble period (7 months), removal of the ice mass shifts the axis of maximum moment of inertia towards Greenland by a small angle, , and so excites a wobble of this amplitude. With progressive damping of the wobble, this becomes an annual shift of latitude, so that Greenland moves towards the pole by /year . Addition of the same mass uniformly over the Earth has no effect on the axis of symmetry and we assume that it spreads out in all directions from Greenland. To calculate the magnitude of , consider the addition of m at =15 to an otherwise rotationally symmetrical Earth, with moments of inertia C, A. The axis of maximum moment of inertia is shifted by an angle , and its magnitude is I = C cos 2 + A sin 2 + ma 2 sin 2 ( + ) 21

where dI / d = 0 , so that ( C A) sin 2 = ma 2 sin 2 ( + ) which is, for small and the assumed numerical values ma 2 sin 2 = = 3.55 108 rad = 7.3 10 3 arc sec 2 ( C A) (ii) Substitution for in I gives, with small I C + ma 2 sin 2 The axial moment of inertia with m distributed over the surface is 2 I = C + ma 2 3 The fractional increase in moment of inertia due to removal of the ice and replacement with a shell of water is 2 I / I 1 = ( ma 2 / C ) sin 2 3 so that the fractional change in rotation rate is 2 / = ma 2 / C sin 2 = 2.8 1010 3 (b) Isostatic rebound adjusts the ground level so that the mass in any vertical column remains the same. All changes in moments of inertia are therefore reduced by a factor of order (h/a) where h is the thickness of lost ice and a is the Earths radius. However, rebound is a slow process (Section 9.5) and reduces the effects considered in (a) only on a time scale of thousands of years.

22

9.2

(a)

Gravity due to element of disc between r and (r+dr) is Gdm G 2rdrzh dg = 2 2 cos = 3 (h + r ) ( h2 + r 2 ) 2 For the complete disc g = 2Ghz
R

rdr

(h

+ r2 )

2 g = 2Gz 1 h ( h 2 + R 2 ) 12 2 h h g = 2Gz 1 1 + (b) R R h 1 h 2 = 2Gz 1 1 ... R 2 R h 2Gz 1 R g = 2Gz at R (c) (d) For a spherical shell of radius R and thickness z , mass, m = 4R 2 z 2 surface g = Gm R = 4Gz This result is independent of h when R is assumed to be infinite, but is twice the infinite sheet result. For the infinite sheet geometry, 90o , cos 0 at r >> h. cos does not vanish in this way for the infinite sphere. Note also an explanation in terms of Gausss theorem, as in electrostatics. The gravitational field lines of a thin isolated plane sheet are normal to the sheet and are equally distributed on either side of it, whereas a spherical shell exerts no gravitational field inside itself and all of the field lines from the 1

23

surface distribution of mass emerge from one side of it (the outside) making the field there twice as strong. 9.3 (a) In the approximation of a spherically symmetric Earth, surface gravity is g = GM/R2, independently of the radial density distribution, and M = (4/3)R3 . g is well measured, so if is determined then G is fixed, or vice-versa. (b) At depth z (radius r = R z) gravity is due only to the material inside r, mass m(r), that is g = Gm(r)/r2 The material between r and R is a spherical shell outside the point of observation and does not contribute. If the density of any layer is then the gravity gradient within it is dg G dm 2Gm(r ) = dr r 2 dr r3 G 2g 2g = 3 4r 2 = 4G r r r This is zero if 2G = g/r = Gm(r)/r3 = G(4/3) r3/r3= (4 / 3)G
that is = (2/3) Note that the free air gradient just above the surface, where 0, is (8/3)G = 2g/R = 3.08610-6 s-1 = 0.3086 mGal m-1 taken as negative here because g decreases with r. (c) If = /2, then we can rewrite the gradient as dg/dr = 2G 2g/r = 2G (8/3)G = (2/3)G = 0.07715 mGal m-1, a quarter of the free air gradient. (d) If = 1025 kg m-3, the density term in the gradient is 4G = 8.6010-7 s-2 = 0.0860 mGal m-1 and the gradient becomes 0.2226 mGal m-1 (e) The density of air at sea level is 1.23 kg m-3, so 4G = 110-4 mGal m-1, 0.03% of the free air gradient. This is the difference between the free air and free vacuum gradients. 2(1 + ) 3(1 2)

10.1 From Appendix D

K=

so that 4 2(1 + ) 4 2 (1 + ) K+ = + = + 2 3 3(1 2) 3 3 (1 2) 1 = 2. 1 2 Thus K + 4 VP 2(1 ) 3 = = VS (1 2)


2

24

V 2 V 2 P P From which = 2 / 2 1 VS VS

10.2 The simpler case is the calculation of V , for which we need the modulus for the combined layers, for compression perpendicular to the layers. If stress is applied, with lateral constraint preventing dimensional changes in the plane of the layers, then the strains are 1 and 2 . The strain for the 1 combined layers is 2 (1 1 + 1 2 ) = so that
= 21 2 / ( 1 + 2 )

and therefore
2 2 2 21 2 2 ( 1 ) . ( 2 ) V = = = ( 1 + 2 ) 1 + 2 2V1V2 = 1 2 (V1 + V22 ) 2 1 1 1

To calculate V|| , consider Eqs.(10.4) for each material. (x-y) is the plane of layering, with propagation in the x direction and z perpendicular to the layers. There are six equations relating the strain components x1 , y1 , z1 , x 2 , y 2 , z 2 , to the normal stresses x1 , y1 , z1 , x 2 , y 2 , and z 2 . The following conditions simplify the equations: 1. Strains must be the same in both media in the plane of layering, so x1 = x 2 = x and y1 = y 2 . 2. The total strain in y and z directions is zero, so y1 = y 2 = 0 and z1 + z 2 = 0 . 3. In the z direction stresses are equal in the two media, z1 = z 2 = z With these constraints the six equations are x = x1 ( y1 + z ) / E1 z1 = z ( x1 + y1 ) / E1 0 = y1 ( x1 + z ) / E1

(1) (2) (3) (4) (5) (6)

z1 = z ( x 2 + y 2 ) / E2

x = x 2 ( y 2 + z ) / E2 0 = y 2 ( x 2 + z ) / E2

From these equations we must calculate the modulus for strain in the x direction, that is || = x x = ( x1 + x 2 ) / 2 x .

25

First, use (2) and (5) to substitute for y1 and y 2 in the other four equations. Then add (3) and (6) to eliminate z1 . This gives x = x1 (1 2 ) z ( 2 + ) / E1 2 2 x = x 2 (1 ) z ( + ) / E2 z = (1 ) .[ E2 x1 + E1 x 2 ] / ( E1 + E2 ) (1) (4) (3)+(6)

Substituting for z in (1) and (4), we have


2 (1 + ) E2 x1 + E1 x 2 x = x1 (1 2 ) . (1) / E1 E1 + E2 (1 ) 2 (1 + ) E2 x1 + E1 x 2 x = x 2 (1 2 ) . (4) / E2 E1 + E2 (1 ) These are simultaneous equations that can be solved for x1 and x2 .

At this point it is convenient to substitute 1 and 2 for E1 and E2, using the relation (Appendix D) E = (1 + )(1 2 ) (1 ) Then (1) and (4) become
2 x (1 2 )( 1 + 2 ) = x1 (1 ) + (1 2 ) 2 2 x 2 1 1 2 2 = x 2 (1 ) + (1 2 ) x1 2

These give
2 1 2 2 2 + 2 (1 ) + (1 2 ) + (1 2 )( 1 + 2 ) 1 2 1 = 2 1 2 4 2 2 (1 ) + (1 2 ) . (1 ) + (1 2 ) 2 1

|| =

x1 + x 2 2 x

which simplifies to
1 2 2 (1 2 ) 2 + + + 4 2 1 ( + 2 ) . || = 1 2 1 2 2 (1 ) 2+ + 2 1
1 2 2 2 2 1 1 + 2 1+ . 2 = 2 (1 ) 1 2 2+ + 2 1

26

+ 2 = 1 2

2 ( 2 ) 2 1 . 1 2 2 (1 ) ( 1 + 2 )

2 Substituting VP for each 2 2 2 (VP12 VP22 ) 2 VP1 + VP2 || = 1 . 2 2 2 2 (1 ) (V + V ) 2 P1 P2 For slight anisotropy we can write it as 12 VP|| 1 2 2 2 2 2 2 2 1 = (VP1 + VP2 ) (1 )2 (VP1 VP2 ) 1 2VP1VP2 VP For = 0.27, VP1 VP2 = 1.164 gives 1% anisotropy VP1 VP2 = 1.402 gives 5% anisotropy. 12

10.3 (a) By Eq.(10.23) ( / QP ) T = 6s where T = 20.25min = 1215s is the trans-Earth travel time for a PKP wave (see Fig. 17.8). Thus QP = 636 . (b) If the core Q is assumed to be infinite, Eq.(10.23) still applies but QP is the mantle value and T is the travel time in the mantle only. This is equal to the PcP travel time at = 0, T = 8.5min. = 510s. Thus QP (mantle) = 267. (c) Taking QS QP / 2 (Eq.10.24), QS (mantle) 133 . (d) S waves do not penetrate the core. If we consider SKS waves with no attenuation in the core then Eq.(10.23) applies with T = 15.6 min = 936s for ScS waves at = 0 and QS 133 as above, giving d ln A / df 22 10.4 If the total width of a hysteresis loop, measured on the strain axis, is and the peak-to-peak stress is 2 (amplitude of the stress cycle), then the area of an elliptical loop is E = ( 4 ) .2 = ( 2 ) . . The peak elastic energy, for strain is (1 2 ) . So that E ( ) . 2 = 21 = = E Q ( 2 ) .
8

Thus = 2 / Q = 1.0 10 To measure this to 10% requires a strain resolution of 10-9.

27

11.1 We determine the vertical strain to obtain displacement which will have a maximum value at the upper surface. Let z be measured downwards from the top of the block. Taking compression as negative, Eq. (11.34) gives the vertical stress as zz = gz
Because the forces at the sides are zero, xx = yy = 0 Hooke's law Eq.(11.1) gives zz = (exx + eyy + ezz ) + 2ezz The strains are related by Eq.(11.11) ezz = exx = eyy The vertical stress becomes zz = (1 2)ezz + 2ezz = ( 2 + 2)ezz gz = ( 2 + 2) Integrating g z2 w = A ( 2 + 2) 2 where A is the constant of integration. Applying boundary conditions w = 0, z = L; g L2 A= ( 2 + 2) 2 L2 z 2 g ( 2 + 2) 2 2 if = w= g L2 = 3 103 10 108 /(5 3 1010 ) (5) = 20 m Note that the shear stress implied by this deformation would exceed the yield stress of rock. A mountain with these dimensions would collapse. wmax = w z

11.2 For a block of material as described in problem 11.1 the vertical stress is zz = gz (now taken positive as we are considering friction). The normal stress, n , and shear stress, , on a plane dipping at angle

28

are obtained from Eq.(11.26, assuming zz = 22 ) as n = gz cos 2 =(1/2) 22 sin 2 = (1/ 2)gz sin 2 From Eq. (11.36) slipping occurs when S0 + f n
(1/ 2)gz sin 2 S0 + f gz cos 2 gz sin cos S0 + f gz cos 2

S0 (1) gz cos 2 The angle is steeper at low elevations. Note that if we consider a mass m of area A, on a dipping plane and resolve forces rather than stresses, we obtain tan f + S0 A / mg cos tan f +

Since m = Az cos , we also obtain Eq.(1) tan f + S0 /(gz cos 2 )

11.3 (a) To find the constant term we equate the radial stress to the pressure at r=R, A rr (r = R) = 3 = P; R 3 A = PR , so PR 3 PR 3 PR 3 , (r ) = 3 , (r ) = 3 r3 2r 2r To find the displacements we integrate the strain field, noting that rr (r ) =

29

rr + + = 0 e = err + e + e = 0 Considering the x-direction, Hooke's law, Eq.(11.6) gives xx = e + 2


u 1 P0 R 3 = x 2 x 3 ux = A ux = P0 R 3 , u x = 0, x , A=0, 4x 2

P R3 u u = 2 = rr ( = 0, = 0) = 0 3 x x x

P0 R 3 4x 2 From symmetry P0 R 3 , u = u = 0 4r 2

ur =

(b) A uniform pressure at infinity corresponds to rr = = = P In order to satisfy the boundary condition of zero radial stress at the cavity we superpose the negative solution from part (a) with that of a uniform P throughout the medium PR 3 r3 R3 = P(1 + 3 ) 2r R3 = P (1 + 3 ) 2r rr = P 11.4 Cracks in a hydrofracture experiment open against the least principal stress (compressions positive). The transition from horizontal to vertical fracture will occur when the least principal stress equals the overburden pressure. Above this depth we infer that the least principal stress is the overburden pressure because the cracks are horizontal.

30

1 is larger than overburden pressure at depths < 1 km 3 =gz and 2 >3 . We use the condition that at the transition depth the least and intermediate principal stresses are equal to one another and to the overburden pressure to find 2 2 = 3 =gz=3 103 9.8 103 = 29.5 MPa Then at 2 km if the only change is the overburden pressure it then becomes the intermediate principal stress ( 2 ). The principal stresses at 2km depth become, [1 , 2 , 3 ] = [100,59, 29.5] MPa

11.5 (a) From Eq.(11.53) Pb = 33 1 =40 MPa; 3 = Pc = 25 MPa


1 = 3Pc Pb = 35 MPa 3 = gh=3 103 10 103 = 30MPa =[35,30,25] MPa, [North-South, Vertical, East-West] (b) Strike slip earthquakes with maximum and minimum principal stresses horizontal. (c) The angle the slip plane makes with 1 is given by 1 1 = tan -1 (1/ f ) = tan -1 (1/ 0.6) = 29.50. 2 2 2 The dip angle will be 90 degrees and the strike angle 29.5 degrees Eq.(11.39) as = to north.

11.6

(a) The radial and tangential stresses for a cylindrical pressurized magma chamber are AP AP rr = 2 0 ; = 2 0 ; = 0 r r The quilibrium equations equations are given by Eq.(11.49) as

rr 1 r rr + + =0 r r r r 1 + + 2 r = 0 r r r since r = 0; rr rr + r r AP0 AP0 + 2 2 AP0 r2 r =0 = 3 + r r

31

(b) Since rr ( r = R ) = P0 = A = R2. AP0 R2

(c) To obtain displacement we integrate the strain field noting, P0 R 2 P R2 + ( 0 2 ) + 0 = 0 r2 r so e = err + e + e = 0 rr + + = rr = e + 2 u u = 2 r r

u P0 R 2 = 2 r r P R2 u = C 0 , where C is a constant. As r , u = 0, so C = 0. 2r 2 u= P0 R 2 2r for P0 negative, u decays as 1/ r.

11.7 (a) We substitute in the equilibrium eq.(11.58) derivatives of zz = 2 Xz 2 y 2 Xy 3 2 Xzy 2 ; yy = ; zy = r 4 r 4 r 4

i.e. zz zy 4 Xzy 8 Xz 3 y 4 Xzy 8 Xz 2 y 2 + = + z y r 4 r 6 r 4 r 6 = 8 Xzy 8 Xzy ( z 2 + y 2 ) =0 r 4 r 6 yy zy 6 Xy 2 8 Xy 4 2 Xy 2 8 Xz 2 y 2 + = + y z r 4 r 6 r 4 r 6

8 Xy 2 8 Xy 2 ( z 2 + y 2 ) = =0 r 4 r 6 (b) At the free surface tractions are generated by normal and shear stresses, 2 Xz 2 y 2 Xy 3 2 Xzy 2 zz = ; yy = ; zy = ; r 4 r 4 r 4 At z = 0, zz = zy = 0, i.e., surface tractions Eq(11.19) are zero.

32

(c) To find the net force we integrate the surface tractions around a cylindrical surface surrounding the line of forces. Consider cylindrical coordinates with measured from y. zz = 2 X cos sin 2 2 X cos3 2 X cos 2 sin ; yy = ; zy = r r r Traction on any area is (Eq.11.18) ijn j , where n are direction cosines. T2 = 23 n3 + 22 n2 = yy n y + zy nz For an area rd on the edge of a cylinder direction cosines are [n z ,n y ]= [sin, cos],
/2

F=Ty =
/2

2T rd
y 0 yy

= = = = =

2(
0 0

n y + zy nz ) rd
4

4X 4X 4X

/2

(cos (cos
0

+ cos2 sin 2 )d + cos2 (1 cos2 ))d

/2

/2

cos
0

4 X sin 2 [ + ] 2 2 4 =X So F=X 11.8 Stresses for a normal line load at the surface of a half space are given by 2 Xz 3 2 Xzy 2 2 Xz 2 y zz = ; yy = ; xy = r 4 r 4 r 4
As in Problem 11.8, consider cylindrical coordinates with measured from z. 2 X cos3 2 X cos2 sin 2 X cos2 sin ; yy = ; zy = r r r Traction on any area is ijn j Eq.(11.18), where n are direction cosines. zz = T3 = 33n3 + 32n2 = zz nz + zy n y For an area rd on the edge of a cylinder surrounding the line load, direction cosines are [n z ,n y ]= [cos, sin],

33

/ 2

F = Tz = = = = 4X 4X 4X
/2

/2

2Tz rd =
4 2

2(
0 2

zz z

n + zy n y )rd

(cos +cos sin )d


0

/2

(cos
0

+ cos 2 (1 cos 2 ))d 4 X sin 2 [ + ] 2 4

/2

cos
0

d =

=X So F=X

11.9 (a) Applying the first of Eqs. (10.4), we have x = 0 and z =0, so that x = y. Substituting this in the second of Eqs. (10.4), with z =0 as before, y = y(1 2)/E. The required modulus is E* = y / y = E/(1 2). (b) We first allow the strip to expand freely in width, with strain e0, and then apply line forces to its edges, such that the strain is reduced to the value, e, that would be obtained by heating it while welded to the half space. Then the stress in the strip is yy = E*(e0 e). This requires a force per unit length F = hE*(e0 e). Equal and opposite forces are applied to the surface of the half space and, using the equation from Problem 11.7, at distance W from each of these forces the stress is yy = 2F/W, so that at the mid-point of the strip (at distance W from each of the two line forces), the half space stress is 4F/W. The corresponding strain is e = (4F/W)/ E*. Equating the forces, F, in the strip and the half space, (W/4)e = h(e0 e), so that 4h e0/(W + 4h) (4h/W)e0 The possibility that uneven diurnal (or annual) heating of surface rocks generates stresses that penetrate to kilometre depths can be considered as an explanation of the paradoxical observation of a diurnal variation in earthquake frequency in the vicinity of Kilauea volcano, Hawaii (Rydelek, Davis and Koyanangi, J. Geophys.Res., 93,B5, 4401-4411, 1988). A variation at tidal frequency is not seen and this means that, to be an effective earthquake trigger, thermal stress would need to exceed the tidal stress, ~ 25 kPa. We can imagine a surface strip of rock exposed to strong sunlight, with surroundings insulated by soil cover, to have a free thermal strain e0 210-4 ( = 10-5 K-1 (linear coefficient), T = 20 K). Its effective thickness is h = (2/) 0.19 m, that is the skin depth for a thermal wave of diurnal frequency, 7.310-5 s-1, in rock of thermal diffusivity = 1.310-6 m2s-1. The necessary width of the strip, W, is

34

the depth of seismic activity, ~ 3000 m, so that the deep strain is e (4h/W)e0 = 1.610-8 and if E* 80 GPa, the thermal stress is 1.3 kPa, less than 10% of the tidal stress, making it an unpromising earthquake trigger. 12.1 Using subscripts I, P, H for the Indian plate, the Pacific plate and the Hawaii hot-spot, Eq.(12.1) is H I = H P + P I A straightforward way to add these vectors is to resolve them into components in directions x (through 0N, 0E), y (through 0N, 90E) and z (90N). Then for each vector with a pole at (, ) the components are x = cos cos ; y = cos sin ; z = sin Corresponding components are added so that xHI = xHP + xPI , etc The magnitude of H I is
H 2 2 2 I = ( xHI + + yHI + zHI ) 12

and its direction is given by sin HI = zHI / H I tan HI = yHI / xHI with the convention that -90<<90 when x is positive and 90<||<270(= 90o ) when x is negative. 12.2 If the cooled subducting lithosphere is 105 m thick, then the volume flux of cooled material is 3.41011 m3/year. Taking the average density to be 3300 kg m-3, the mass flow is dm/dt = 1.11015 kg/year = 3.6107 kg s-1. Much of the material is still at a high temperature when it subducts and its specific heat is about 1200 J kg-1 K-1, over the whole of its cooling range, but some has reached the surface temperature, at which the specific heat is 850 J kg-1 K-1.We assume the average effective specific heat to be CP = 1100 J kg-1 K-1. Then, the rate at which coolness (negative heat) is carried down is CPTdm/dt = 271012W, from which T = 682K. This would be consistent with the plate model of the oceanic lithosphere, according to which mature lithosphere becomes a passive conductor, with a linear temperature gradient from 280 K to a basal temperature Tmax, if Tmax = 1644 K. This means an average temperature of 962 K, that is, 280 K plus 682 K of average cooling. The basal temperature cannot reasonably be as high as this, but a modest increase in the assumed thickness, to 120 km, reduces the average cooling to 568 K and the basal temperature to 1416 K, which is more conventional. A more likely alternative is that the temperature gradient is not linear, but has strong hydrothermal cooling of the shallow layers with diffusive cooling only of the deeper part. With greater average cooling, a thinner lithosphere is required. 12.3 If the total length of plate boundaries, ignoring transforms, is taken as (67000 + 51000) km = 118000 km, drawn on the surface of the

35

Earth, 5.1108 km2 , then their average separation is (5.1108/118000) km = 4320 km. We can compare this with the plate speed required to produce 3.4 km2/year of ocean crust from 67000 km of ridge, an average spreading rate of 5 cm/year or 2.5 cm/year each way from the ridges. In 90 million years this motion produces 2250 km of ocean crust on each side of a ridge. On the other hand, at trenches only one plate subducts and with 51000 km of trench the speed required to consume 3.4 km2/year of crust is 6.6 cm/year, implying a reach of 5940 km. These numbers make the implicit assumption that ridges and trenches are geometrically fixed features, with the plates moving with respect to them. It is clear that this is not a valid model. Both ridges and trenches are themselves moving with respect to one another at speeds comparable to the spreading and subduction rates. The inference that subduction speeds substantially exceed spreading rates means that Birds numbers require trenches to be approaching the ridges that feed them faster, on average, than the ridges are producing new crust. Although observations of the plates give only relative motions, we are interested also in the motions relative to the deeper mantle. This is difficult to infer for spreading centres, but for subduction zones the inclinations of the Wadati-Benioff zones under the over-riding plates are consistent with motion relative to the mantle. If we suppose plate speed to be proportional to the force acting on it, and that this is due to its total cooling at the point of subduction, then the force and speed, v, would be proportional to 1/2, where is the plate lifetime. This is the correlation which Carlson et al (1983) suggested is observed. Alternatively, if plate motion is restricted by a viscous drag which increases with plate size, that is the product v, as well as directly on v, then v 1/4. Although this is not favoured, we can consider an arbitrary case, v f , that is v1/f. Consider plates 1 and 2 with speeds v1, v2, and lifetimes 1, 2, each produced by length L/2 of ridge. Their total area, which must be taken as the area of the ocean floor and therefore constant, is given by A = (L/2)v11 + (L/2)v22 where v1 = v /2, v2 = 3 v /2, so that A = L v Also 1 = Cv11/f = C( v /2) 1/f 2 = Cv21/f = C(3 v /2) 1/f with C as a proportionality constant. The average lifetime is = (1+ 2)/2 = C( v /2) 1/f (1 + 31/f )/2 so that 1 = 2 /(1 + 31/f ), 2 = 2 31/f /(1 + 31/f ) Substituting these relationships in the expression for A A = L v [(1 + 31+1/f )/2(1 + 31/f )] If f =1/2, A = 1.4 L v and if f = 1/4, A = 0.5 L v

36

The assumptions here set up a simple model, but show that the relationship between v and is very dependent on the v - correlation. 12.4 This problem involves calculation of small differences between much larger quantities and requires care not to obscure the answer with rounding errors. By the simple isostatic assumption, which we reconsider later, the mass of the continent, MC, is equal to the sum of the masses of the ocean, MO, and mantle, MM, that it displaces. We calculate the difference between their contributions to the moment of inertia of the Earth, using the depths of the centres of mass, relative to sea level, of the continent (zC = 19160 m), ocean (zO = 2250 m) and mantle (zM = 21830 m). We choose plausible densities for the ocean (1025 kg m-3), mantle (3300 kg m-3) and continent, subject to the condition MC = MO + MM, which gives a continent density of 2975 kg m-3, but this is a rounded number and we can avoid using it directly. A different choice of densities, within a reasonable range, makes little difference to the calculation if the isostatic condition is imposed. With these numbers the masses per unit area are: ocean, 4.6125107 kg m-2; mantle, 1.4378108 kg m-2; and the sum of these for the continent. These numbers are multiplied by the area. A = 1.251013 m2 to obtain the total masses. The moment of inertia of the continent about the centre of the Earth (radius R) is I C = I i + M C ( R zC ) 2 , where I i is the moment of inertia about its own centre of mass, but is neglected here, not primarily because it is small, relative to the second term, but because the moment of inertia of the displaced (ocean +mantle) about its centre is almost the same and we are interested only in the difference. In this approximation the moment of inertia contribution of the continent exceeds that of the (ocean +mantle) by I = I C ( I O + I M )
= M C ( R z C ) 2 M O ( R zO ) 2 M M ( R z M ) 2

Substituting ( M O + M M ) for M C , the R 2 terms cancel and this reduces to I = M M ( zM zC )(2 RM zM zC ) M O ( zO zC )(2 R zC zO ) It is the R terms that are important. Taking the Earth average value, R = 6.371 106 m, I = 3.63 1031 kg m 2 , which is 4.5 107 of the moment of inertia of the Earth. There is no correction here for surface curvature, which is the same for the continent as for the (ocean plus mantle) that it replaces. We use this result to infer the effect on rotation of the transfer of the continent from the pole to the equator. For a rough idea of the effect we can simply increase the axial moment of inertia, C, by 4.5 parts in 107 and take this as the fractional decrease in the rotation rate. That neglects the displacement of the centre of mass of the Earth. The centre of mass of the continent is elevated, relative to the material it displaces, by C = (MCzC MOzO MMzM)/MC = 1911 m The corresponding shift in the centre of mass of the whole Earth, mass, ME, is E = (MC/ME)C = 0.48 m

37

The total moment of inertia must be referred to an axis through this displaced centre, adding {ME E2 + MC [(R E)2 R2]} = MCE( 2R + C+ E)= 9.21027 kg m2 to the estimate of I above. This is not significant. The answer to part (a) of this question is, therefore that the rotation rate is reduced by 4.5 parts in 107 and the axis of rotation (and the pole) are shifted towards the continent by 0.48 m. The effect of shifting the continent to the equator is to add I to the axial moment of inertia, C, and now we consider the ellipticity to relax to restore the original value. No external torques are invoked, so the angular momentum is unchanged and after completion of the continental migration to the equator there is no asymmetry that would deflect the rotation axis. But the increase in C is matched by the decrease in A, so that (C A) is increased by 2I and the increase in J2 is J2 = 2I /Ma2. The rest of the Earth relaxes to cancel this excess, and the reduction in rotation rate, at least to first order in what is already a very small quantity, becomes negligible. We now reconsider briefly the appeal to isostasy. Strictly, we must consider isostatic compensation to mean that the pressure at the base of the continent is equal to that at the same level in the surrounding mantle, but that would be equivalent to equal overburden masses only if gravity is the same in both. With the different mass distributions that cannot be quite true. This is the density dipole problem discussed in Section 9.4, but not pursued further here. 12.5 Sea level reflects the average lithospheric shrinkage, z . Taking the local shrinkage to increase with age, t, as Ct1/2, where C is a constant, the average for lithosphere age is
z = 1 1 2 1/ 2 1/2 zdt = Ct dt = 3 C 0 0

the local shrinkage at that age being zmax = C1/2, so that z = (2/3)zmax. We take this simple relationship to apply directly to the increase in depth with age, by allowing C to include the factor for isostatic adjustment (Section 9.3 and Problem 20.1). With local deepening of 3000 m at 90 million years, the average deepening for ocean overlying lithosphere with an age range 0 to 90 million years is 2000 m. If this was once 200 m less then we can estimate the duration, , of lithosphere that gives an average deepening of 1800 m: 1800/2000 = (/90million years) 1/2, = 72.9 million years. This applies to both cases (i) and (ii), so that if the plates were the same size as at present (assuming an average to be appropriate), then their speed would have been greater by the factor 1.23. Alternatively, if the speed has not changed, then they were smaller by the factor 0.81. In both cases the heat transferred to the oceans in time was proportional to 1/2 and the average rate of heat transfer proportional to -1/2, being greater by the factor 1/0.9 1.1 than at present. This means an average ocean floor heat flux of 34.8 tW, compared with

38

the present 31.3 tW. If t1/3 replaces t1/2 in this calculation, then z = (3/4)zmax, with a present average deepening of 2250 m and an earlier deepening of 2050 m. This requires = (2050/2250)390 million years = 68.1 million years and an ocean floor heat flux of 34.4 tW. Neither of these simple models can be taken seriously, but they indicate the magnitude of changes to plate motion that must be contemplated if a 200 m sea level change is to be explained by temperature variations in the oceanic lithosphere. 13.1 Using Eq. (13.6), the limit is obtained by equating the force to zero, so that 0.915(mzm + czc) = (m c) zc We treat zm as unknown, with values of the other parameters as applied in Eq. (13.6), giving zm = 895 m. The corresponding increase in ocean depth is 1.44895 m = 1289 m. By Fig. 20.2, this corresponds to a lithospheric age of about 15 million years. 13.2 For any mass of material of density that is thermally contracted, relative to its surroundings, so that the total volume contraction is V, the negative buoyancy force is gV. The mechanical work done by this force as the material sinks through a depth range z is gVz and, if this occurs in time t, the power is gVz/t. As a first rough step we ignore the variations in and V with depth and introduce numerical values: t = 1 year = 3.156107 s, z = 2.8106 m (the mantle depth), V = 3.36 km2 2.1 km = 7.06109 m3, g=10 m s-2, = 3300 kg m-3, to obtain a power dissipation of 20.7 tW. Recognizing that increases with depth, but V decreases because decreases, we can use the simple approximation that the product () decreases by a factor 2 over the mantle depth range. On this basis its mean value over this range is of its surface value. Reducing the power estimate by this factor we have 15.5 tW, approximately twice the thermodynamically allowed maximum. Before discussing the implication, we apply a more rigorous method of calculation to the same problem. The assumed linear variation of () neglects the facts that there are discontinuities in properties in the transition zone and that the temperature difference between subducted material and its surroundings increases with adiabatic compression, partly compensating for the decrease in . By replacing V with VT and then V = m in the expression for power, we have Power = mgT/t (1) in which the only depth variation, apart from z itself, is in the product ( T ) . Its variation with compression on an adiabat is directly proportional to T, as in Eq.(19.56), and so is conveniently represented by a thermodynamic identity (Appendix E, Eq. E4)

39

ln( T / CP ) = S ln V S At the high temperatures of the mantle, and especially on an adiabat, the variation in CP is negligible, so we can use the tabulation of S (see Appendix G) to obtain the variation of (T) over any range V2 ( T ) 2 exp S d ln V (T )1 V1 In using this expression we must note that it applies only to material in constant phase and cannot be carried through the transition zone, but must be integrated in stages. Integrating over the whole mantle we find that (T) diminishes by a factor 2.1 and the average is 0.74 of the surface value. This is fortuitously close to the factor estimated above from the linear () approximation. It confirms the factor 2 discrepancy between the subduction energy estimated by assuming that all subducted heat reaches the base of the mantle and the available energy derived from thermodynamics. The unavoidable explanation is that the subducted coolness does not all reach the base of the mantle, but is distributed through it, in essentially the same way as heat is derived from the body of the mantle and not all from the base. This may seem obvious, but theories based on a supposition that all subducted material eventually reaches the bottom of the mantle, being recycled from there, are often presented.

13.3 Opinion is divided over whether the crust has been derived from the whole mantle or primarily from the upper mantle and we consider both alternatives. Either way, we assume that it was derived uniformly from the source material, that is in proportion to local density, and not by volume. We assume also that the compressibility-pressure relationship is the same for both mantle and crust materials, so that their density ratio is the same at all pressures. With numerical values, m/c = 3370 kg m-3/ 2870 kg m-3, so that (mc 1) = 0.174. In calculating gravitational energy, we consider a surface layer of density m to be replaced by material of density c, with extraction of the difference and its distribution in the mantle. This is energetically equivalent to light material rising. The total mass of the crust is Mc = 2.81022 kg, so the mass of material removed from it is (mc 1) Mc = 4.881021 kg. This mass is taken from the crust, average radius R = 6.355106 m, and deposited in a shell between radii R and Ri = 3.48106 m if we consider distribution through the whole mantle, or 5.711106 m for the upper mantle only. We apply the simplifying approximation that the density structure of the Earth is such that, over this range, gravity is constant, g = 10 m s2 , independent of depth. This means that the mass inside radius r is m(r) = (g/G)r2 and density is (r) = (g/2G)/r, where G is the gravitational constant (see also Problems 1.1(c) and 17.1). Then the mass in a range r to (r + dr) is a fraction 2rdr/(R2 Ri2) of the mass in the shell and the mass deposited in this range is dm = (mc 1) 40

Mc2rdr/(R2 Ri2). It moves through a gravitational potential difference (R r)g, so that the total energy released by the crustal separation is R R Mcg m ( R r ) gdm = 2( c 1) ( R 2 Ri2 ) R ( R r )rdr Ri i
M g m 1) 2 c 2 R 3 / 6 RRi2 / 2 + Ri3 / 3 c ( R Ri ) Substituting numerical values, E = 6.331028 J if the whole mantle is involved, or 1.441028 J for the upper mantle only. The corresponding item in Table 21.1 is a compromise value, consistent with stronger upper mantle depletion in the light elements but with some contribution by the lower mantle. For several reasons these estimates are approximate and for one particular reason they are underestimates. Crustal separation increases the inward concentration of mass and this increases the internal pressure, and hence compression, causing the whole Earth to contract and adding to the gravitational energy release. Some of this additional energy is released in the core and some is stored as compressional energy and is not available to support mantle convection. An iterative calculation is needed to address this issue, but from a similar calculation on inner core development (Stacey and Stacey, 1999) we conclude that it is only an incremental effect and does not alter the conclusion that, from the numbers in Table 21.1, the energy of crustal differentiation is small enough to be lost in the uncertainties of the other numbers. = 2(

13.4 Distributed through the mantle mass (4.01028 kg), the total heat & generation of 201012 W means an average generation q = 510-12 -1 W kg . In a layer of density and thickness t at the base of the mantle the heat generation is & & Q = (4/3)(r3 rc3) q , where rc is the core radius and r = (rc + t). If this is assumed to be conducted through radius r, then & Q = 4r2(dT/dr) This heat suffices to cause convection if (dT/dr) exceeds the adiabatic temperature gradient given by Eq. (19.55), so a condition for convection is & (4/3)(r3 rc3) q > 4r2(T/KS)g which can be rewritten to solve for r & (r3 rc3)/ r2 > 3Tg/(KS q ) = 152000 m giving r > 3532 km, that is t > 52 km. This thickness is well inside the boundary layer for transfer of core heat to the mantle and cannot imply a static boundary layer, but demonstrates the inevitability of mantle convection. 13.5 Eq. (13.13) gives plume radius as a power of viscosity. Thus, a factor 100 change in viscosity only changes the radius by a factor

41

(100)1/4 = 3.16. If the viscosity is 1020 Pa s, then the radius is 66 km or if viscosity is 1016 Pa s, the radius becomes 6.6 km.

13.6 We derive Eq.(13.20) zz xx = (1/ 2)s ge using Eq.(13.25)


( zz xx ) = =
e

1 z( z ) gdz D 0
zc

1 1 = z (0 S ) gdz + zc 0 zc = = =

z(
e

S ) gdz

1 1 g S e 2 + g ( f S )( zc 2 e 2 ) 2 zc 2 zc 1 1 1 1 1 g S e 2 + g f z c 2 g f e 2 g S z c 2 + g S e 2 2 zc 2 zc 2 zc 2 zc 2 zc

1 1 1 g f z c g f e 2 g S z c 2 2 zc 2 1 1 = g ( f S ) zc g f e 2 2 2 zc

(1)

Isostasy=> S zc = f ( zc e); f = S zc ( zc e ) (2)

Substitute (2) in (1) z 1 z 1 g S c e2 ( zz xx ) = g ( S c S ) zc 2 zc ( zc e ) 2 ( zc e ) z 2 c z 2 c + ezc e 2 1 z 2c 1 zc ( zc e ) e2 = g S = 2 g S ( z e ) ( z e ) ( z e ) 2 ( zc e ) c c c 1 = g S e 2

13.7 (a) Substitution in Eq.(13.43) with = atan(0.85), = 6o , = 2o , w = 0.7 , w = b = 0.85 ,


b = 1 {( + )[1 + (1 w ) / ( csc sec(2 + 2) 1)] }/ b ,

gives b = 0.88 . (b) The effective friction for the low Andes is b (1 b ) = b (1 0.88) = 0.12b . Double the friction for the high Andes means that b (1 b ) = 0.24 b , b = 0.76. Substitution back into Eq.(13.43) rearranging terms and solving for w = 0.5 .

42

These values are consistent with the basal shear zone of the high Andes being drier b = 0.70 compared with b = 0.88 for the low Andes and the overlying wedge material also being drier, w = 0.70 compared with w = 0.50 , by a similar amount. 14.1 At distance r from the axis of the dislocation the strain energy 1 density is 2 = S 2 / 82 r 2 . Thus the energy per unit length of an 2 annulus between r and (r+dr) is dE = (S 2 / 8 2 r 2 ) 2rdr and so the total energy between radii r1 and r2 is
E1,2 = S 2 4

r2

r1

dr S 2 r2 = ln r 4 r1

Thus we cannot consider either r1 0 or r2 . The difficulty with the lower limit arises from the fact that the equations refer to a simple dislocation, that is one in which there is a sudden change in displacement on the axis, and not a more realistic graded dislocation. r1 = 0 implies infinite stress and energy density on the axis. Thus r1 cannot be less than the distance at which the failure stress ()max is reached. The upper limit problem is typical of the difficulties that arise when two parameters are allowed to become infinite. Thus by assuming that the dislocation is infinitely long, we mean longer than any other dimension including r. With this assumption we cannot also allow r to approach infinity. The strain falls off more rapidly than by Eq.(14.1) at distances comparable to the dislocation length. An approximate solution is therefore to put r2 equal to the (finite) length. 14.2 Eliminating MS from Eqs. (14.34) and (14.36) a log10 E = 1.5log10 + 4.8 + f ( , h ) T Considering earthquakes all at the same depth and distance 1.5 E (a / T )
2 But strain, a / T and E where is the duration of a wave passing a stationary point and is proportional to the total length of the wave train. Thus 0.5 (a / T ) This appears to say that, as the amplitude (or magnitude) increases, so the wavetrain becomes shorter and is clearly not correct. To see 0.3 why, consider the body wave case, for which ( a / T ) . This is

reasonable because the surface observations of a, T are representative of the spherical spreading wave-front. But for surface waves the energy is spread over different depths, depending on T. Although a/ T measures the surface strain in a surface wave and (a/

43

T)2 gives the energy density, we really need the integration of the energy flux with depth. Thus the energy density argument cannot be applied directly to surface wave energies. 14.3 We can write the displacement y in a propagating wave t x as y = a sin 2 2 T = T where dy 2 t x Strain = = a. cos 2 2 dx VPT VPT T so that max = 2a VPT

Energy/unit vol =

1 2 = 2 2 a 2 / VP 2T 2 2 This is the total wave energy/unit vol which oscillates between kinetic energy and strain energy. The volume through which wave 2 energy is distributed is 2R R where R = (assuming a surface earthquake that radiates into a hemisphere). Thus total energy E = 2R 2VP .2 2a 2 / VP 2T 2 s But = VP 2 , so

E = 4 3VPR 2a 2 / T 2 14.4 Differentiating Eq. (14.53) and substituting dv/dk in Eq.(16.49), we have 1/ 2 k g g g u = v + tanh(kh) 2 tanh(kh) + h sec h 2 (kh) k 2 k k 1 = v + v 2 + gh(1 tanh 2 (kh) 2v v gh 1 tanh 2 (kh) = + 2 2v This general expression can be written in alternative ways by substituting for one of the parameters by Eq. (14.53), but is convenient for demonstrating the two limits. If (kh) is very large, that is << h, then tanh(kh) 1 and the second term vanishes, leaving u = v/2. This applies to normal ocean waves in deep water. If (kh) is small, >> h, as for long waves, such as tsunamis, then tanh2 (kh) (kh) 2, which is negligible compared with unity and the second term reduces to gh/2v. But, by Eq. (14.53), in this case v (gh)1/2, and the second term becomes v/2, so that u = v. 14.5 (a) Eq.(14.8) gives the displacement in the i-direction from a unit point force in the k-direction as

44

1 ik 1 2r 4 r 4(1 ) xi xk Expanding this expression 1 1 1 1 1 1 1 1 x2 x G11 = = + 4 r 4(1 ) xr 4 r 4(1 ) r 4(1 ) r 3 Gik =

1 1 (3 4) x 2 = + 3 4 4(1 ) r r 2 + 2 4 6 + 6 4 ( + 3) (3 4) = 3 = = 2 + 2 2 + 2 ( + ) Recalling (Table D1) that = 1 1 ( + 3) x 2 G11 = + 4 4(1 ) r ( + ) r 3 1 1 yx G12 = 4 4(1 ) r 3 1 1 xz G13 = 4 4(1 ) r 3 (b) The solution for displacement from an earthquake source is equivalent to that of a double couple as given by Eq.(14.10). For the first couple, note the derivative w.r.t. the source point, x , x ' , is negative with respect to a field point, x, i.e., x' = x . Using the results from part (a) ( + 3) x 2 y ( + 3) 3 yx 2 + 3 ; G11,2 = B 3 + 5 ; G11 = B r r ( + ) r r ( + )
yx x 3xy 2 G21 = B 3 ; G21,2 = B 3 + B 5 r r r xz 3 yxz G31 = B 3 ; G31,2 = B 5 r r For the second couple yx y 3 yx 2 ; G12,1 = B 3 + B 5 r3 r r 2 ( + 3) y x ( + 3) 3xy 2 G22 = B + 3 ; G22,1 = B 3 + 5 ; r r ( + ) r r ( + ) G12 = B G32 = B xz 3 yxz ; G31,2 = B 5 3 r r

45

Adding individual couples to make a double couple located at z = c


ux = G12,1 + G11,2 = B y ( + 3) 3 yx 2 y 3 yx 2 + B 5 + B 3 + 5 3 r r r r ( + )

y y ( + 3) 6 x 2 y = B 3 + 3 + 5 r ( + ) r r Let x = 0, = y y 1 y ux = B 3 = B 2 = 2 3/ 2 2 r (y + c ) 12 ( y + c 2 )3/ 2

(c) To find the displacement from an earthquake source (Eq. 14.10), multiply the double couple displacements given in part (b) by the total moment M 0 = bS to obtain bS y bS y ux = = 2 2 3/ 2 2 12 ( y + c ) 12 ( y + c 2 )3/ 2
(d ) The moment magnitude relation is given by Eq.(14.38) as M W = (2 / 3) log10 M 0 6.07 M 0 = 103/ 2( M W + 6.07) For an equidimensional fault, length is related to moment and stress drop though Eq.(15.23) 103/ 2( M W + 6.07) M l = 0 = = 7.515 km. b= = l
1/ 3 1/ 3

3 106 7515 = 0.751 m 3 1010

(e) x km -5 -4 -3 -2 -1 0 1 2 3 4 5 ux (metres) -0.0214 -0.0249 -0.0270 -0.0252 -0.0161 0 0.0161 0.0252 0.0270 0.0249 0.0214 46

14.6 It is convenient to break the integral in Eq.(14.27) into three regimes (Figures 1 and 2). We compare the pulse received from the end of the fault ( r = L / 2 ) with that received from the beginning of the fault ( r = L / 2 ). For rupture towards the end of the fault, the integral in Eq.(14.21) becomes L/2 M0 P(t ) = I [t r / V ( L / 2 + y ) / VR + ycos / V )]dy LT L / 2

L/VR+T VR

L/VR

t t l2 l1 VS L/VS+T

L/VS T

Figure 1. Time space development of the slipping Haskell patch. The thick lines denote the start (lower) and end (upper) of the region that is slipping. The thin lines describe the propagation of the radiation. Points l1 and l2 denote the beginning and end of the slipping patch. Radiation from all points along the diagonal line between l1 and l2 arrive simultaneously at time t. At the beginning of rupture l1 increases at the rupture speed while l2 remains at zero until the patch grows to its final size. Similarly at the end of the rupture, when l1 reaches the end, l2 continues until the patch size decreases to zero.

Referring to Figure (1) for a pulse received from the end of the fault,

M P(t ) = 0 LT P(t ) =
M0 LT

L/2

L/ 2

I [t r / V ( L / 2 + y ) / V

+ ycos / V )]dy

M0 (l1 (t ) l2 (t )) LT For the starting phase, L/V < t < L / V + T


L/2

L/ 2

I [t L / V

+ y (1/ V 1/ VR )]dy

47

t ' = l1 / VR = t - L / V + l1 / V ; l1 = P (t ) = = M 0 t L /V T L / VR L / V

t L /V 1/ VR 1/ V

M 0 t L /V T X For the intermediate phase, L / VS + T < t < L / VR

t ' = T + l2 / VR = t L / V + l2 / V ; l2 = l1 l2 = P (t )= T 1/ VR 1/ V

t L /V T 1/ VR 1/ V

M0 M T = 0 T L / VR L / V X For the ending phase, L / VR < t < T + L / VR l1 = L l1 l2 = = l2 = t L /V T 1/ VR 1/ V

L / VR L / V t + L / V + T 1/ VR 1/ V

L / VR t + T 1/ VR 1/ V M 0 L / VR t + T M 0 L / VR t + T = T L / VR L / V T X

P (t ) =

For a pulse received from the beginning of the fault For the starting phase, 0 < t < T t ' = t l1 / V = l1 / VR , l2 = 0 l1 = t 1/ VR + 1/ V
L L + VR V

For the intermediate phase, T < t < t ' = t l2 / V = T + l2 / VR l2 = t T 1/ VR + 1/ V T 1/ VR + 1/ V

l1 l2 =

48

For the ending phase, l1 l2 = L =

L L L L + <t<T + + VR V VR V

t T 1/ VR + 1/ V

L / VR + L / V t + T 1/ VR + 1/ V

T+L/VR+L/V t L/VR+L/V

L/VR+T L/VR

l2 T 0 l1 L

Figure 2. See caption for figure 1, but for a pulse received from the start of the fault, at y=0. Radiation for slip between points l1 and l2 is superposed in the integral Eq.(1) to form the pulse at time t. 14.7 We compare the Brune spectrum given by Eq.(14.32), with the Fourier transform of the pulse, i.e.,
P( ) = K t exp(it at )dt
0

Integrating by parts

49

t exp(it at ) exp(it at )dt P( ) = K i a i a 0 0 exp(it at ) exp(it at ) =K (i a ) 2 0 i a

1 1 =K =K 2 2 2 a 2ia (i a ) 1 a 2 2 + 2ia =K 2 2 2 2 ( a 2ia ) ( a + 2ia ) a 2 2 + 2ia =K 4 4 2 2 2 2 + a 2a + 4 a

a 2 2 + 2ia =K 4 2 2 4 + 2a + a a2 2 2a =K 2 +i 2 2 2 ( + a 2 ) 2 ( + a ) The amplitude spectrum is given by K P( ) = 2 ( a 2 2 ) 2 + 4a 2 2 2 2 ( + a ) K = 2 a 4 + 2a 2 2 + 4 ( + a 2 ) 2 K = 2 (a + 2 ) Then Brunes spectrum Eq.(14.32) may be rewritten u(0) u(0) 2 P( ) = u() = = 2 02 1 + (/0 ) 2 0 +

K = u(0)02 a = 0
2 P(t ) = u(0)0 t exp( 0t )

15.1 Except close to the surface, as considered below, the power -2 dissipation per unit area of fault face is v = 0.016Wm , where = 107 Pa is the shear stress and v = 1.6 109 m sec 1 is the speed of fault movement. It is here supposed that this power appears as heat in the fault plane. We may therefore consider a vertical plane heat source of this strength, effectively having infinite horizontal extent along the fault and extending vertically between limits which we examine as part of the problem. Intuitively we expect a heat flow anomaly of similar magnitude. The vertical planar heat source may be equated to a sum or integral of a series of horizontal line sources. We consider first one line source at depth z, having width dz and heat production per unit
50

length 0.016 dz Wm-1. In equilibrium in an infinite medium heat would diffuse radially outwards and at distance r the heat flux would 0.016dz Wm 2 . The effect of the free isothermal surface (z=0) is be 2 r to deflect the heat flow so that it flows normally through the surface and this may be simulated by adding to the source in the infinite medium a heat sink of strength -0.016dz Wm-1 at -z, i.e. above the surface. Then the heat flow in the semi-infinite medium is simply that obtained by superimposing radial flows outwards from the heat source, below the surface, and inwards towards the heat sink, above the surface. The heat flux through the surface at a horizontal distance y from the fault is the sum of resolved fluxes of the source and sink 0.016dz dQ = 2. sin 2 r 0.016 zdz . = where tan = z / y . z2 + y2 [As a check, note that the total heat flow through the surface is

y =

dQ.dy = 0.016dz which coincides with the heat input.]

The total heat flux due to a planar source extending between depths z1 and z2 is
2 0.016 z2 zdz 0.016 z2 + y 2 = ln 2 2 z = z1 z1 z 2 + y 2 2 z1 + y This is a double-infinity problem, similar to that in Problem 14.1. We cannot consider an infinitely long fault and integrate to infinite depth. In this case we are justified in considering an indefinitely long fault and equating z2 to the depth of the lithosphere (~100km). The logarithmic dependence on z2 makes the precise value unimportant. Equating z1 to the depth at which overburden pressure equals z1 = / g = 380m

Q1 =

z2

dQ =

For z < z1 we consider stress proportional to z. Then the heat flux due to this component of the fault is 0.016 z1 z 2dz 0.016 y 1 z1 = Q2 = 1 tan z1 0 z 2 + y 2 z1 y The total fault-generated heat flux is the sum of Q1 and Q2 . Putting y 0 and substituting numerical values
Q ( y = 0) = 0.016 z2 0.016 ln + = 0.0335Wm 2 z1

15.2 A displacement of 2105 m under a shear stress of 107 Pa dissipates 21012 Jm -2. If the rock was already at its melting point, then this much heat suffices to melt a total thickness of 1665 m, 830 m on each side. If, instead, we assume that the rock starts cold (300 K) and must first be raised to 1500 K, then the additional

51

heat required is about 1.3106 J kg-1 and the thickness of melt is 214 m on each side. The energy required for other processes, such as production of fault gouge, is much less than for melting and would require correspondingly larger volumes. This is not a realistic situation, but demonstrates the impossibility of storing locally the energy dissipated in fault zones.

15.3 As in the development of Eq.(15.33), let time to failure, t, be t = t0 n ,


t = ( ) 1/ n t0

(1)

1 t 1/ n 1 = ( ) t nt0 t0 The number of patches with shear stress between and + d , S ( )d , is equated to the number with failure times between t and t+dt, R(t )dt R(t )dt = S ( )d (2) d R ( t ) = S ( ) dt From Eq.(1)

S ( ) t 1+ 1 t0n( ) n t0

(3)

S ( ) for n large, and K is a constant t S ( ) R(t ) = , < 0 , which gives Omoris law provided S ( ) is a t weak function of compared with n . K

15.4 (a) The Brune spectrum Eq.(14.27) can be written M0 & . M0( f ) = 1 + ( f/ f 0 ) 2 Then

52

2 fM 0 && M0( f ) = 1 + ( f /f 0 ) 2
2 2 2 4 2 f 2 && 2 ( f )df = 4 f M 0 df =2 M 2 M0 df 0 (1 + ( f/f0 )2 )2 (1 + ( f/f 0 ) 2 ) 2 0

= 2 4 2 M 02 = 2 4 M f
2

f2 df ; Let y = f / f 0; dy = df / f 0 (1 + ( f/f 0 ) 2 ) 2 0

2 3 0 0

y2 (1 + y 2 )2 dy. See Dwight (1981, Item 122.2) for the 0

standard integral.
= 2 4 2 M 02 f 0 3
3 2 3 0 0

4 (1)

= 2 M f Inserting Eq.(1) into Eq.(15.47) && M 02 dt 2 M 02 f 03 S ER = = 10VS5 5VS5 0 From Eq.(15.39) E = M0 2


S ER 2 M 0 f 03 = M0 5VS5

(b) Combining Eqs.(15.40) (15.41) and (15.42) 3.15 VS M 0 = 2 f 0 S (c) Radiation efficiency Eq.(15.37) is R = ER / E
S 2 ER 2 2 M 0 f 03 2 2 M 0 f 03 R = = = M 0 5VS5 5VS3 3 3

2 2 M 0 f 03 = 5VS3 2 2 f 03 3.15 VS = 5VS3 2 f 0


3 3 3

2 2 3.15 = = 0.4975, (as obtained by Singh and Ordaz, 1994) 5 2 15.5 (a) We use relationships between magnitude, moment, rupture speed and fault dimension to find how corner frequency varies with magnitude.

53

0 = VR / l f0 = VR 2 l
1/ 3

(Eq. 15.41)

M l = 0 (Eq. 15.23) M W = (2 / 3) log10 M 0 6.07 (Eq. 14.38) 3/ 2 M W + 3/ 2 6.07 = log10 M 0 M 0 = 103/ 2M W +3/ 26.07 so f0 = VR M 2 0
1/ 3

V = R 2

M0
1/ 3

1/ 3

V = R 3/ 2M W +3/ 26.07 2 10 V = R 1/ 3101/ 2( M W +6.07) 2 (b) Magnitude Moment Nm l (Size, m)

T , Rupture time, sec 0.25 0.79 2.51 7.92 25.05

b (Slip, m)

f 0 (Corner frequency) Hz 2.22 0.70 0.22 0.07 0.02

4 5 6 7 8

1.27E+15 4.03E+16 1.27E+18 4.03E+19 1.27E+21

751.55 2376.62 7515.53 23766.19 75155.29

0.08 0.24 0.75 2.38 7.52

15.6 (a) From kinematics the accelerated and decelerated motion gives T 2 ye = 2 (1/ 2)a0 ( ) 2 4 2 ye y a0 = 32 e = 32 2 (2) 2 T (b) For the cosine case =2 / T 42 ac = 2 ye = 2 T 2 4 ac / a0 = = 1.234 32 (c) For the step function case, Eq.(15.46) gives the radiated energy as proportional to the integral of the acceleration squared.
54

2 2 a dt = 32 o o

T 2

T 2

2 2 2 4 ye 4 ye T 4 ye dt =322 = 32 2 (2 ) 4 (2 ) 4 2 (2) 4

64

y 3
3

2 e

(1) dX ,
2 ye 3 2

For the cosine case, X = t , dt =


T 2 T 2

2 4 2 2 3 2 2 a dt = ye cos tdt = ye cos XdX = o o o

(2)

(d) Taking the ratio of Eqs.(1) and (2) 2 3 ye 64 3 128 Radiation efficiency ratio = 2 = 4 = 1.314 3 ye 2 Efficiencies are 0.65% and 3.2% compared with 0.5% and 2.5% for the cosine case (Table 15.1). (e) The radiation efficiency of the stepped case is larger, even though the maximum accleration in the smoother cosine case is 23% greater. The overall integral of acceleration squared in the cosine case is less. In general, seismic and radiatation efficiencies are greater than the 0.5% and 2.5% of the cosine slider. This occurs because accelerations are more staccato than this smooth end-member model. A step function scenario requires the driving force minus friction to remain positive and constant during the first half of the cycle (y<ye ) and then to reverse, and remain constant and negative in the second half. For the stretched spring model, this means the friction must decrease in the same way as the spring force decreases and then, at y e , increase suddenly and decrease in parallel with the spring force decrease until the block stops at 2ye . This requires a change in properties that is unrealistic.

16.1 Since the wave motion at the boundary must be the same on both sides of it, the transmitted amplitude AT must be equal to the sum of incident and reflected amplitudes AI and AR. For this purpose we may assume either that the reflected wave has the same phase as the incident wave or opposite phase. In the latter case the sign convention is AT = AI AR (1) The strain amplitude of the incident wave, frequency f, is 2AI 2fAI AI I = = = I VP1 VP1

55

and the corresponding energy density in the wave is therefore V 2 A 2 E I = 1 ( I ) = P1 1 I = 1 2 AI2 2 2 VP1 2 2 where 1 = VP1 1 is the elastic modulus appropriate to the wave. The power per unit area of wavefront is therefore PI = E IVP1 = 1 1VP12 AI2 2 Similarly the powers per unit area of the reflected and transmitted waves are 2 PR = 1 1VP12 AR 2
2 PT = 1 2VP22 AT 2 But the power entering the surface must be equal to that leaving it, so that PI = PR + PT
2 2 2 or 1VP1 AI = 1VP1 AR + 2VP2 AT Substituting for AT from Eq.1 we relate AR, AI: 2 1VP1 ( AI2 AR ) = 2VP2 ( AI AR ) 2 2

(2)

1VP1 ( AI + AR ) = 2VP2 ( AI AR )

AR 2VP2 1VP1 = AI 2VP2 + 1VP1 If instead of Eq.1 we had put AT = AI + AR , then the result (3) would, have been reversed in sign. Thus the motion of the reflected wave is reversed if 2VP2 > 1VP1 and then Eq.3 applies with AR AI positive, but if 1VP1 > 2VP2 , the reflected wave motion is the same as that of the incident wave and Eq.3 gives negative AR.

(3)

16.2 Consider a ray refracted from a medium of velocity V1 as it crosses an interface with medium of velocity V2 . Let the i be the angle of incidence and f the angle refraction.
Y

V1

h i x f d

V2

With the geometry given in the figure the travel time is

56

T=

(Y x) 2 + d 2 x2 + h2 + V1 V2

dT x Y x =0= dx V1 x 2 + h 2 V2 (Y x) 2 + d 2 sin(i ) sin( f ) = V1 V2 sin(i ) V1 = sin( f ) V2

16.3 The displacement expressed as a sum of normal modes is given by u( x, t ) = an un ( x, t ) = an sin( kn x ) exp( in t )
n i

where kn = / c, Take the Fourier transform with respect to time u( x, ) = an sin( kn x )


n

1 i + in

(1)

Note that
2 nx m x sin( L ) sin( L )dx = mn L0
L

(2)

Suppose at time t = 0 , a force per unit volume is applied in the y( x x s ) direction at a location x = xs , given by f = f 0 H (t ) , a where f 0 is the force and a is the cross-sectional area. The equation of motion (Eq.16.70) becomes ( x xs ) H (t ) = 2u / t 2 2u / x 2 + f 0 a Taking the Fourier transform in time we obtain ( x xs ) 1 () ( + ) = 2u () 2u () / x 2 f 0 a i 2

and replacing u by Eq.(1) or (16.67) an kn 2 sin( kn x )


i

1 f ( x xs ) 1 () 2 + 0 ( + ) = an sin( kn x ) i + i n a i 2 i + in n (3)
L

2 m x If we apply the operator sin( ) f ( x )dx to Eq.(3), noting Eq.(2) L0 L

57

c 2 am km 2

2 f sin(km xs ) 1 () 1 1 + 0 ( + ) = am 2 i + im aL i 2 i + im (4)

am 2 f 0 sin(km xs ) 1 () m 2 2 ( + )= aL i i + im 2 that is am 2 f 0 sin(km xs ) 1 () 1 = ( + ) i + in aL ( 2 + 2 ) i 2 m

Substituting back into Eq.(1)


2 f sin(k x ) 1 () ) sin(ki x) u ( x, ) = 0 2 i s2 ( + 2 n M ( + i ) i

(5)

where M = aL is the mass of the string. Finally we invert the Fourier operator u ( x, t ) =
n

2 f0 1 cos i t sin(ki xs ) sin(ki x) M i2

(6)

16.4 Recalling Eqs.(16.35)-(16.37) for a P wave incident on a free surface u inc = [u x , uz ] = [cos j,sin j ]exp [i( px z t ) ] (16.35) The reflected P wave is u refl = [u x , uz ] = SP[sin i,cos i ]exp [i( px + z t ) ] and the reflected S wave u refl = [u x , uz ] = SS [cos j, sin j ]exp [i( px + z t ] (16.36) (16.37)

Equating shear tractions to zero on the surface: u u xz = x + z = 0 x z xz / / i = ( cos j + p sin j ) + SP( sin i + p cos i ) + SS ( cos j p sin j )
0 = ( 2VS + p 2VS ) + SP (pVP + pVP ) + SS (2VS p 2VS ) 0 = SP(2pVP ) + ( SS 1)(2VS p 2VS ) Now from Eq.(16.32) and Eq.(16.34) 2 + p 2 = 0 = SP(2pVPVS ) + (1 SS )(2V 2 S p 2V 2 S ) 0 = SP(2pVPVS ) + (1 SS )(( 1 p 2 )V 2 S p 2V 2 S ) 2 VS 1 , so VS2

0 = SP(2pVPVS ) + (1 SS )(1 2 p 2V 2 S )

58

Equating normal stress at the surface to zero: u u u u u zz = x + z + 2 z = x + VP2 z z z x z x u u zz = 0 = (VP2 2VS2 ) x + VP2 z x z u x u z 0 = (VP2 2VS2 ) + VP2 x z 2 2 0 = (VP 2VS ){ p cos j + SPp sin i + SSp cos j} +
VP2 { sin j + SP cos i SS sin j}

0 = (VP2 2VS2 ){ pVS + SPp 2VP + SSpVS } + VP2 {pVS + SP2VP SS pVS } 0 = SP( p 2VP3 2 p 2VPVS2 + 2VP3 ) + SS ( pVSVP2 2 pVS 3 pVSVP2 ) + pVSVP2 2 pVS3 pVSVP2
Note from Eq.(16.32) and Eq.(16.33), 2 = 0 = SP( p 2VP3 2 p 2VPVS2 + ( 1 p 2 , so 2 VP

1 p 2 )VP3 ) + SS (2 pVS 3 ) 2 pVS3 2 VP

0 = SP(VP 2 p 2VPVS2 ) + SS (2 pVS 3 ) 2 pVS3 0 = SP(1 2 p 2VS2 ) + ( SS + 1)(2 pVS 3 / VP )

16.5 It is possible to integrate, over the inner core, the force on a point mass , but there is an easier way. We can use the facts that the gravitational force between two uniform spheres is equal to the force that would be experienced if the masses were concentrated at their centres and that a spherical shell exerts no gravitational force on a body inside it. The force exerted on the inner core arises from its density excess, relative to the outer core, so the effective gravitational mass is m1 = (4/3)ri3(i o). When its centre is displaced from the centre of the Earth by a small distance , the restoring attractive force is effectively that arising from a sphere of density o and radius , that is m2 = (4/3)3o. The centres of the two spheres are separated by distance , so that the force between them is F = Gm1m2/2 = (162/9)Go(i o)ri3. Note that the centre of mass of m1 is inside all of the outer core except for the sphere of radius . When the inner core moves, its inertial mass is not m1 but (4/3)ri3i, that is its total mass. More than that, it cannot move alone but must cause outer core material to flow around it, so that the total inertial mass is M (4/3)ri3(i + o). Thus, the period of oscillatory motion is
T 2 M 3(i + o ) = ( F / ) G o (i o )

59

Taking o = 12163 kg m-3 and (i o) = 820 kg m-3, we have T = 18869 s = 5.24 hours, compared with the average period of the triplet of frequencies reported by Pagiatakis et al, ~ 4.5 hours. Apart from neglect of splitting of the mode frequency by rotation, assumptions and approximations primarily concern uncertainty in (i o), uniformity of the outer and inner cores and the pattern of outer core flow that accommodates inner core motion. 17.1 At radius r gravity is Gm(r ) g= r2 where m(r) is the mass inside r. If is the density at r, 2 then dm( r ) / dr = 4r dg G Gm( r ) 2g = 2 .4r 2 2 = 4 G so that 3 dr r r r The condition g = constant, dg/dr = 0, gives g = 2Gr = 2G0 R where 0 is the surface density, at radius R. Thus = 0 R r Total mass m = 4 r 2 ( R r ) dr = 2 R 3 0 0
0 R

4 3 = R 3 so 0 = ( 2 3) Thus = ( 2 3) R r 17.2 The geometry of this problem is given in Fig.17.1 and the algebra in Section 17.2. To find the series of slownesses, dT/dS, and travel time intercepts, take differences, as in the following table and fit equations with the form of Eq. (17.7) to linear segments. S S (km/s) + Tn S(km) T(s) Layer No. (n) T = VP n T 0 0 3.03 S T= 1 0.33 1 3 2.99 3 1.00 3.77 5 1.53 5.00 10 2.53 5.00 S T = + 0.53 20 4.53 2 5

60

5.00 30 50 60 80 6.53 5.00 10.53 5.41 12.38 6.51 15.45 6.49 100 120 140 160 200 250 300 18.53 6.49 21.61 6.64 24.62 8.23 27.05 8.20 31.93 8.20 38.03 8.20 44.13 4 T= S + 7.54 8.2 3 T= S + 3.15 6.5

The velocities of the layers, 3, 5, 6.5 and 8.2 km/s are seen immediately. The thickness of layer 1 is then given by Eq.(17.5). Using this and Eq.(17.7), the thickness of layer 2 is found, and so on. Layer 1 2 3 4 Velocity 3.0km/s 5.0km/s 6.5km/s 8.2km/s Thickness 1.0km 10.0km 20.0km -

17.3 (a) The travel time for the direct path through the sediment is T1 = S VP1 where VP1 = 2.5km/s and the travel time, T2, for rays penetrating the igneous rock is given by Eq.(17.5). Equating T1 and T2 , we find the distance at which the travel time break occurs
2 2 (VP2 VP1 ) S S = + 2 z1 VP1 VP2 VP1VP2 which gives S = 0.748km with the estimated z1 , VP1 VP2 . Thus, seismometers should be spread around this range, say 0.2 to 2.5km. 12

61

(b)

Paths to (1) and (2) are common up to point (A), so the time difference is z1 z tan z z T1 T2 = + 1 VP1 cos VP1 cos VP2
1 tan = z VP1 cos VP2 VP2 , so that
12

But sin = VP1

(V cos =

2 P2

2 VP1 )

VP2

; tan =

(V

VP1

2 P2

2 VP1 ) 12

12

and therefore T1 T2 =

2 2 z (VP2 VP1 )

VP1VP2

17.4

The difference between distances travelled by the two rays shown is S = S1 + S2 and the travel time difference is due to the path elements A-B-C S1 S2 T = + VP2 cos VP1 sin( + ) and x = S1 tan = S 2 / tan( + )
S = S1 + S 2 = S1 [1 + tan .tan( + ) ]

62

S1 = S / [1 + tan .tan( + ) ] S 2 = S .tan .tan( + ) / [1 + tan .tan( + ) ]

Thus
dT T 1 tan .tan( + ) = = + dS S VP2 cos [1 + tan .tan( + )] VP1 sin( + ) [1 + tan .tan( + ) ]

1 1 1 sin V + V . cos( + ) cos 1 + tan . tan ( + ) P2 P1

cos ( + ) sin 1 + cos .cos ( + ) + sin .sin ( + ) VP2 VP1 = 1 1 1 V ( cos cos sin sin ) + V sin cos P2 P1

cos sin tan sin + VP2 VP2 VP1 cos Substituting for by sin = VP1 / VP2 and therefore =
2 2 2 2 cos = VP2 VP1 / VP2 , tan = VP1 / VP2 VP1 2 2 dT cos (VP2 VP1 ) sin = + we obtain dS VP2 VP1VP2 The refraction back into the upper medium would give horizontal rays that could not be observed, even in principle if ( + ) = 90o . The limit is therefore max = 90o sin 1 (VP1 VP2 ) . 12

17.5 The triplication range is bounded by rays which (1) just graze the 5800km boundary and (2) are critically internally reflected at the same boundary.

(1) Geometry of the grazing ray gives 5800 cos ( 1 2 ) = 2 6400 o 2 = 50.02

63

(2) Geometry of the critically reflected ray gives 3 sin i1 = (refracted angle = 90o ) 4 i1 = 48.59o 5800 sin i0 = sin i1 (parameter p is constant) 6400 i0 = 42.82o
1 2

1 = ( i1 i0 ) (i1 is the external angle of the triangle)

1 = 11.54o Thus the triplication range is 11.54 to 50.02. The core shadow zone begin at the range 3 of the mantle ray which just grazes the core. Then 3 = 2 ( 1 + 2 ) , as in the diagram 3500 cos 2 = ; 2 = 52.88o 5800 3 3 3500 sin i1 = sin i2 = 4 4 5800 5800 3 3500 sin i0 = sin i1 = . 6400 4 6400

1 = i1 i0 = (26.91o 24.21o ) = 2.70o


3 = 2 ( 1 + 2 ) = 111.16o

The general ray penetrating the core travels an angular distance = 2 ( 1 + 2 + 3 ) , as represented in the diagram, where

64

1 = i1 i0 2 = i3 i2 3 = 90o i4 and by constancy of the ray parameter, p R 4 R sin i1 = sin i0 ; sin i2 = sin i0 ; R2 3 R2 4 R R sin i3 = sin i0 ; sin i4 = sin i0 ; 3 Rc Rc ( R = 6400km; R2 = 5800km; Rc = 3500km) so that we can write in terms of the single variable i0 : 4 R R sin i0 = 2 sin 1 sin i0 i0 + sin 1 R2 3 Rc R 4 R sin i0 + 90o sin 1 sin i0 sin 1 3 R2 Rc Examination of this function shows that at the critical angle i0 = 24.21o of the ray which just penetrates the core (i.e. at i3 = 90o ), = 193.98o which is 13.98 beyond 180. As i0 is decreased to 18.20, decreases to its minimum value 158.74; o this is more than 13.98 away from 180 . Beyond this 180

65

as i0 0. Thus the range of the core shadow zone is 111.16 to 158.74. 17.6 Speed VP varies with depth z as VP = VP0 + V z The angle of a ray to the vertical is i where sin i / VP = sin i0 / VP0 s
sin i = (1 + V z / VP0 ) sin i0

If x is the horizontal distance travelled then


2 dx dz = tan i = (1 + V z / V0 ) sin i0 / 1 (1 + V z / VP0 ) sin 2 i0 Integrate by putting u = (1 + V z / V0 ) sin i0 ; dz = V0du / V sin i0 12

With the condition x =0 at z = 0, this gives the equation for the ray path 12 2 V z VP0 V0 2 x= cos i0 1 1 + sin i0 + V sin i0 VP0 V sin i0 which can be rewritten VP0 VP0 VP0 x cot i0 + z + = V V V sin i0 This is the equation of a circle of radius (VP0 / V sin i0 ) , centred at x = (VP0 / V cot i0 ) , z = VP0 / V (above the surface). 17.7 p = r sin i / VP = constant along any ray path. The extreme range of angles in the crust for rays from the core is bounded by rays that graze the core surface at i = 90. Thus the angle i0 at the base of the crust is given by 0.55RE RE sin 90o = sin i0 p= 13.7km/s 6.4km/s i0 = 14.9o (for P waves) 18.1 (a) The mass inside radius r is m(r ) = M (r / R) and the gravity at Gm GMr that level is g = 2 = 3 . r R The pressure increment over the range dr at r is GM dP = g dr = 3 rdr R Integrating from R (where P = 0) to r
3 2 2 2

GM GM 2 2 (R r ) P ( r ) = 3 r dr = R R 2R3

66

Substituting for density, = 3 GM 2 P= 8 R 4

3M , 4R 3

(b) Since surface gravity is g = GM R we can substitute M R 2 = g G in the result for part (a) so that
2

3g 2 P= 8G
R R

18.2 (a) M = 4 r 2dr = 4 ( ar 2


0 0

b 3 r )dr R

4 = R 3 ( a b ) 3 = 5.96 1024 kg

(b) In this case the mass inside radius r is r b 4 r4 m( r ) = ( ar 2 r 3 )dr = ar 3 b 3 R R 0 and the pressure increment over the range dr is br 2 Gm(r ) 4 r dr = G ( ar )(a b )dr 2 r 3 R R 2 2 3 4 7 r br = G ( a 2 r ab + 2 )dr R R 3 3 Integrating from R to r dP =
2 7 b2 2 7 r 3 b2 r 4 P (r ) = G R 2 ( a 2 ab + ) a 2 r 2 + ab 3 9 4 3 9 R 4 R2

At r = 0, P = 311 GPa, compared with 364 GPa for the Earth (by the PREM model, see Appendix F) and 172 GPa for a uniform body with the size and mass of the Earth (Problem 18.1). The sharp density jump at the core-mantle boundary gives a much larger central pressure than the smoothly graded density profile. (c) The moment of inertia of a spherical shell of radius r and thickness dr and density is, with the graded density profile dI = (2/3)r2dm = (8/3)r4dr = (8/3)(ar br/R)dr Integrating over the whole body ( r = 0 to r = R )

67

I = (8/3)(a/5 b/6)R5 With the mass in the answer to part (a) I 8 8 4 = ( a b) /( a b) = 0.339 2 MR 15 18 3 compared with 0.3307 for the Earth.
2 4

a a 18.3 (a) = A + B r r 2 4 d a a = 2 A 3 4B 5 dr r r But d dr = 0 at r = a, so B = A 2


a2 a4 d = 2 A 3 5 dr r r

By Eq.(18.18) P =

a2 a4 f 1 .2 A 3 5 g r2 r r

a2 d 2 a4 = 2 A 3 4 + 5 6 dr 2 r r
f a2 a4 A 10 5 + 14 7 r r 3g f 4 4 f A A = At r = a, K = K 0 = 3g a 3 3 g a3 7 5 3 a a P = K 0 Thus 2 r r 5 2 5 2 3 3 3 3 a a = K 0 1 = K 0 1 2 r r 2 0 0

By Eq.(18.19) K =

(b) To find K = dK/dP we can use Eq.(18.20), but an alternative is fA a 2 a4 50 6 98 8 3g r r dK dK dr = = K = 2 dP dP dr a f a4 .2 A 5 6 + 7 8 g r r


a 2 49 25 1 r

3 a 2 7 5 r At r = a, K = K 0 = 4

68

18.4 (a)

= A(r r0 ) 2 d dr = 2 A( r r0 ) d 2 dr 2 = 2 A d 3 dr 3 = 0

f 2A . ( r r0 ) g r2 f A ( 4r0 2r ) By Eq.(18.19) K = 3g r 2 2f A . , so that At r = r0 , K = K 0 = 3g r0 By Eq.(18.18) P =


2 3 1 3 r0 2 r0 P = 3 K 0 = 3 K 0 0 r r 0 2 3 1 3 r0 2 r0 K = K0 2 = K0 2 0 r r 0

(b) By Eq.(18.20) K =

4r0 r 4( 0 )1 3 1 = 6r0 3r 6( 0 )1 3 3 At ( 0 ) = 1, K = K 0 = 1

(c) Substituting P, K and dK dP in Eq.(19.39)


0 ) For f = 1.44, as favoured in Section19.3, 0 = (f 1)/3 = 0.15. Negative values, and therefore the harmonic model, which necessarily gives negative , must be rejected. 1 = 3 0
1/ 3

(2 f

( f 1) 1) 2 ( f 1)(

1/ 3

19.1 (a) Given the ideal gas equation, PV = nRT, we can differentiate with respect to T at constant P: P(V/T)P = nR, so that = (1/V)( V/T)P = nR/PV = 1/T Differentiating with respect to V at constant T: V(P/V)T + P =0, so that KT = V(P/V)T = P The mass of n moles of material with atomic mass m is nm, so that = nm/V Substituting for , KT and in the thermodynamic formula for we have KT (1/ T ) P PV = = = CV (nm / V )CV TnmCV With the relationship CP CV = R/m, we substitute for m to obtain

69

PV CP CV PV CP CV CP = = 1 TnCV R nRT CV CV (b) With the part (a) result, the remaining parts of this question can be gleaned from elementary physics texts, but it is useful to see the adiabatic relationship between T and V rewritten in terms of the Grneisen parameter because, in the approximation that is constant, the relationship is quite general and is not restricted to ideal gases. To see a different perspective on this, use Table E2 (Appendix E) to find (U/T)S = PV/T, where U is internal energy and, if heat input dQ is applied at constant P, dU = dQ PdV. For an adiabatic change dQ = 0 and dU = PdV, so that (U/T)S = P(V/T)S, but the relationship extracted from Table E2 identifies this with PV/T. Thus, (V/T)S = V/T or (lnT/lnV)S = . For constant , this integrates to T1/T2 = (V2/V1). For ideal gases this is familiar, with = CP/CV 1. = (c) The result in part (b) can be rewritten TV = constant. Substituting for T from the ideal gas equation, we have PV +1 = constant, but this is specific to ideal gases and is not general. 19.2 For the lower mantle, both equations for can be used, but for the outer core the acoustic formula is not applicable. The results are obviously a bit rough because they ignore the difference between KS and KT in calculating . Such a calculation can be used iteratively, applying the resulting calculation to this difference and starting again. However, the figures here suffice for a good idea of the significance of the KS/KT problem. (a) (b) (10-6 K-1) Lower mantle (CV = 1134 JK-1kg-1) Bottom (Eq. 19.33) 1.0763 10.18 (Eq. 19.39) 1.1500 10.88 Top (Eq. 19.33) 1.2144 20.10 Top (Eq. 19.39) 1.3037 21.63 -1 -1 Outer core (CV = 840 JK kg ) Bottom (Eq. 19.39) 1.3894 10.91 Top (Eq. 19.39) 1.4324 18.45 (c) (1 + T) 1.041 1.047 1.047 1.054 1.076 1.132

19.3 (a) If we consider a layer in the Earth within which the temperature gradient differs from the adiabat by , then we can calculate the adiabatic density change, by the first term of Eq.(17.32), and then consider the additional temperature to be applied at constant pressure. Thus ( d dz )excess = ( T ) P . ( dT dz ) excess =

70

(b) We note that dP dP = = g dr dz dr 1 so that g = dP dr d d 1 d Therefore 1 g = 1 + . = 1 + dr dP dr dP K d 1 dK K d 1 dK = 2 = 1 But = so that dP dP dP dP dP where we have made use of the identity K = . d dK 1 d in a homogeneous layer. Thus 1 g = dr dP 19.4 (a) (i) Equating the heat generated inside radius r to the heat conducted through r 4 3 & 2 dT r q = ( 4r ) dr 3
T0

& R q dT = 3 r dr 0 Tc

& q 2 R 6 (ii) With conductivity = AT 3 , 4 3 & 3 2 dT r q = AT ( 4r ) dr 3 T0 R & q T 3 dT = r dr 3A 0 Tc

Tc = T0 +

& 2q 2 4 Tc = T04 + R 3A

(b) Heat generation in a shell of radius r and thickness dr is l & & dQ = q0 ( r / R ) .4r 2dr Integration from 0 to r gives the heat generation inside r & 4q & Q(r ) = l 0 r l +3 R (l + 3) & 4q0 3 4 3 & & But Q( R) = (l + 3) R = 3 R < q > l +3 & & <q> so q0 = 3 4 r l +3 & & Q(r ) = <q> l Therefore 3 R & ( r ) = 4r 2 dT dr Since Q 71

& dT < q > r l +1 = dr 3 R l Integrating over radius R as before & < q > R2 Tc T0 = 3(l + 2) which coincides with 19.4(a)(i) if l = 0.
19.5 With spherical symmetry the diffusion equation, with constant diffusivity, , is T 2 T = (r ) t r 2 r r We seek a solution of the form T = T0(r) exp(t/) where T0(r) is the initial temperature profile, the form of which is maintained as the temperature decays exponentially. Differentiating this expression, we have T 1 = T0 ( r ) exp( t / ) t T T = r 2 0 exp( t / ) r2 r r T 2 T 2T (r ) = 2r 0 + r 2 20 exp( t / ) r r r r Substituting the derivatives in the diffusion equation, we have the differential equation for T0(r) d 2T0 2 dT0 T0 + + =0 dr 2 r dr Now put x = r / d 2T0 2 dT0 + + T0 = 0 dx 2 x dx Then substitute T0 = xnu, where n is a constant adjustable for convenience in obtaining a solution, and represent derivatives with respect to x by primes, so that T '0 = nxn-1u + xnu' T "0 = n(n1) xn-2u + 2n xn-1u' + xnu" Substituting these derivatives in the equation for T0(x), the equation becomes

72

xnu" + (2n + 2) xn-1u' + [n(n + 1) xn-2 +xn]u = 0 Now we can choose a value of n to simplify the equation. There are at least two ways of doing this. The most straightforward is to put n = 1, which gives the differential equation for a sinusoidal oscillation u" + u = 0 (an alternative is to put n = , which leads to the solution via a half-order Bessel function). Solving for u u = Asinx + Bcosx sin x cos x T0(x) = A x + B x At r 0, x 0, (cosx)/x , so B =0, and (sinx)/x 1, and A = T0(0), that is the central temperature at t = 0. Thus, with substitution for x sin( r / ) T0 ( r ) = T0 (0) ( r / ) With the temperature at the surface (r = R) taken as zero, R/ = , that is, the relaxation time is 1R = and T (r ) = To
2

sin(r / R) exp(t / ) (r / R)

19.6 Taking the logarithm of the Lindemann expression, we have ln TM = (2/3) lnV +2lnD + constant Although, from observations of specific heat, D is found to be a slight function of temperature (as in the tabulations of Anderson and Isaak,1995), this is not a feature of the Debye theory itself, but an illustration of the fact that the theory is an approximation. So, in this approximation, we do not need Eq. (19.29) to be a partial derivative and can differentiate the Lindemann expression to obtain

73

d ln TM 2 2d ln D 2 = + = 2 D d ln V 3 d ln V 3 This avoids the logical difficulty of taking a derivative of TM at constant T, but we can write dP/dlnV = K, which is KT if this is a derivative at constant T. So d ln TM 2 K = 2 D dP 3 1 dTM 2( D 1/ 3) = TM dP KT 19.7 The significance of Eq, (17.32) to this problem is that we are generally interested in the difference between the actual temperature gradient and the adiabatic gradient. The first term in this equation gives the density variation on an adiabat and the second term accounts for the departure from an adiabatic temperature gradient. If we make the density gradient zero, then g g g = = K S CP with the adiabatic gradient given by Eq.(19.56). Tg g T = = T CP z Adiabatic CP which is small compared with . Alternatively, we obtain the total temperature gradient for constant density. g g T T = + = (1 + T ) = CV z Constant z Adiabatic CP = We can obtain the same result by a direct calculation of the actual temperature gradient by noting that, in a uniform layer, is a function of P and T only, so that d P dT = + =0 dz P T z T P dz
dT =0 g + ( ) dz KT dT g g = = dz KT CV

Noting that KS/KT = CP/CV = (1 + T) does not differ greatly from unity, we see that the temperature gradient required for constant density departs very strongly from the adiabat. It is interesting to note that neither of the expressions for or the constant density profile has an obvious dependence on absolute temperature, although

74

the difference is the adiabatic gradient, which is proportional to T. To verify this, write g g g 1 g T g T dT = = 1 = = dz Adiabatic CV CP CV 1 + T CV 1 + T CP as in Eq. (19.56). Thus, we can use Eq. (17.32) to calculate the temperature gradient for constant density by writing dT g T dT = += + dz CP dz Adiabatic g T g g g = + = ( T + 1) = CP CP CV CP as in the direct calculation. Introducing lithospheric values, the required temperature gradient is about 7.910-3 K m-1 = 7.9 K km-1. This is exceeded by the average lithospheric gradient and by a factor 3 or 4 in the crust. In this range density decreases with depth in material that is compositionally uniform.

19.8 It maybe useful to look first at the derivation of Eq. (19.51) and an alternative form of it. For any pressure change in uniform material P P dP = dV + dT V T T V so that the variation in pressure of the solid phase on the melting curve is P P P T P = + V T =TM V T T V P T =TM V T =TM P T P P 1 = V T =TM T V P T =TM V T dTM KT KM = 1 KT dP V V from which Eq. (19.51) follows immediately. Alternatively, dT K M = KT /(1 KT M ) dP and, with substitution for dTM/dP by Eq. (19.50) K M = KT (1 + 2TM ) or dT KT = K M /(1 + K M M ) dP Using this to substitute for KT in Eq. (19.50), the result simplifies to

75

1 dTM 2 = TM dP K M dP and, since K M = d T =TM d ln TM = 2 d ln 19.9 At first sight it would appear from Eqs. (19.19) and (19.52) that the gradients of TM and TS differ by a factor 2, in conflict with Eq. (19.50), but that is not so, because the variations of with P are different for the two temperature profiles. As discussed in Problem 19.8, the modulus describing compression on the melting curve, KM, is larger than KS and the increase in density with pressure is therefore smaller than on the adiabat at the same (P, T) conditions. This introduces a complication to the application of Eq. (19.52). The reason for using Eqs. (19.19) and (19.52) to calculate temperature profiles is that, assuming that the equations themselves are satisfactory, their use is both easier and more accurate than the use of equations such as (19.52), because the only variable, apart from , is and its variation with is slight. The complication that arises is that the density limits on the two integrals are different. Over any pressure range, the density ratio of the TM profile is less than for the TS profile, requiring an ancillary calculation to determine what these limits are. There are additional complications because the object of the exercise is to estimate the cooling required for a fully solidified core and there are further effects, neglected here. With the increasing concentration of light elements in the outer core, as they are rejected by the inner core, the process is not a simple freezing of a constant composition and, in any case it must eventually reach the point of a eutectic mixture, or a complicated series of eutectic points for different light ingredients. Another effect that must be allowed for in a more rigorous calculation is that the solid is denser than liquid of the same composition and the whole Earth contracts as the inner core freezes, increasing internal pressure and raising the melting point at every level and so reducing the cooling required for inner core growth. This is addressed in Section 21.3 for the present inner core (see Fig. 21.2), and here is qualitatively extrapolated to the whole core. There are several ways of tackling the problem, with varying degrees of sophistication, and a rigorous approach requires an iterative calculation because TM is related to , which is calculated from KM, but that requires TM. A straightforward, step by step integration of KM, TM and , starting at the present inner core boundary, is an obvious way to go. Here we short-circuit the process by noting that by integrating melting point and adiabatic temperature profiles with respect to density, rather than pressure, and making the assumption that is independent of T at constant , Eqs. (19.19) and (19.52) can be combined: (lnTM/lnTS) = 2, that is (TM2/TM1) = (TS2/TS1)2. Then, with TM1 = TS1 = 5000 K and TS2 = 3739 K from Table G1, TM2 = 2796 K. This is the resulting estimate of TM at the CMB density, 9902 kg m-3, not the CMB pressure. To correct it to CMB pressure (by compensating for the excess decompression), we take the thermal expansion coefficient at that pressure, 1810-6 K-1, with the

76

inferred difference between TM and TS, 943 K, to correct the decompression on the melting curve by 168 kg m-3. By Eq. (19.52), with = 1.443, this correction raises the melting point to 2935 K, compared with 3007 K in Table G1. The CMB cooling to reach this point is 806 K (732 K with the Table G1 numbers). Extrapolating to the whole core the argument in Section 21.3 concerning compression caused by contraction, the estimated cooling is reduced to 635 K. 20.1 Let the magnitudes of the two effects on ocean depth be x due to lithospheric shrinkage and y due to subsidence. Then the additional depth of water, density w , is (x + y). Due to the subsidence by y, a depth y of material with density m is displaced from the vertical column with the contracted lithosphere. Isostasy requires the mass in any vertical column to be conserved, so ( x + y )w ym = 0; y / x = w /(m w ) m x+ y y = 1+ = = (1 w m ) 1 From which x x m w 20.2 (a) Differentiation gives T = T0 exp ( z ) cos ( t z ) t 2T 2 = ( 2 2 ) T0 exp ( z ) sin ( t z ) z +.2 T0 exp ( z ) cos ( t z ) The diffusion equation (Eq.6.3) requires that these two derivatives be equal (both when the cosine terms vanish and when the sine term vanishes), so that, 2 2 = 0 and 2 = from which = = ( / 2) , as in Eq.(20.23).
12 12 (b) We require exp ( / 2) z = 1/ 30 000 7 1 6 2 1 With 1.99110 s , = 1.3 10 m s , z = 37m

20.3 Differentiation of Eq.(20.23) gives dT T0 exp z + cos t z = sin t dz 2 2 2 2 = T0 exp z . 2 cos t z 2 2 2 4

77

This function has a maximum value when the cosine term is unity. dT = T0 exp z Thus dz max 2 2 [Note that the maximum occurs when T zt = 0]. 20.4 Putting = 2
12

1 for convenience, the temperature z*

z difference becomes T = T ( z ) T 2 z = T0 e sin(t z ) T0 e z 2 sin(t z 2) But we can also write T = T0 f sin(t ) where f is the factor to be found. Simultaneous equations for f and are obtained by considering t = 0 and 2 , giving e z sin(z ) + e z 2 sin ( z 2 ) = f sin
e z cos(z ) e z 2 cos ( z 2 ) = f cos

Squaring and adding these equations we obtain f independently of f = e2 z + e z 2e


3 z 2

cos ( z 2 )

1 2

z = z. which is the required result, where z = x = z * 2 df = 0 for maximum f gives Putting dx x x 3cos + sin = 2e x 2 + e x 2 2 2 from which x = 1.335 and hence fmax = 0.347. The corresponding value of is 2 = 2 ( x 2) = 4.49 1012 s 1 which gives a period T = 2 = 1.40 1012 s = 44000 years [Note: For a surface temperature wave of peak-to-peak amplitude 2T0, the maximum error in temperature difference over the lower half of a bore hole is 0.347 T0 or 17% of the incident wave, in the most unfavourable case.]

12

& & 20.5 Heat generation, q = q0 (1 z z0 )


z0 & & & Heat flux at depth z, Q = Qmantle + qdz z

78

1 z 1 z 2 & & = Qmantle + q0 z0 + 2 z0 2 z 0 & =Q =Q & & & At z = 0, Q 0 mantle + q0 z0 / 2 & & = q z /6 Thus Q
mantle 0 0 2 & = q z 2 z + 1 z = dT Q &0 0 and dz 3 z0 2 z0 2 3 z & dz = q z 2 z 1 z + 1 z &0 0 So T = 0 Q 2 2 z0 6 z0 3 But at z = z0 , T = T , so

& T = q0 z0 2 / 3
z 3 z 2 1 z 3 T = T 2 + z0 2 z 0 2 z 0 T dT =2 z0 dz z = 0 T dT = dz Average z0 Ratio of surface to average gradients = 2:1.

21.1 Let = c/r, where c is a constant to be determined by the total mass, M. The mass inside radius r is m(r), where dm(r)/dr = 4r2 = 4r2 (c/r), so that m(r) = 4c rdr =2cr2 and the total mass is M = 2cR2, so that c = M/2R2 and m(r) = Mr2/ R2 If mass is progressively added, then when the body has reached radius r, is surface potential is V = Gm(r)/r, and the energy released by depositing an additional layer dr is dE = Vdm = [Gm(r)/r)]4r2(c/r)dr = (2GM2/R4)r2dr Integration from 0 to R gives E = (2/3)GM2/R. For this model, the factor f in Eq. (21.2) is 2/3, which is much nearer to the value for the Earth (0.6654) than either is to the value for a uniform sphere (3/5). 21.2 From values in Table 21.3 the radiogenic heat production of the mantle plus crust is 7.11012 W kg 1 . Multiplying by the Moons mass (Table A5) the total heat production is 5.2 1011 W . Distributed over a surface area of 4(1.74 106 ) 2 m 2 = 3.80 1013 m 2 , it gives a heat flux of 13.7mW m 2 21.3 The classical (high temperature) heat capacity of the Earth is

79

M mantle M core 27 1 + 3R = 5.84 10 JK mmantle mcore Note that if Mmantle, Mcore are in kg, the gas constant R must be taken in JK-1(kg mole-1) see Table A.1. It should be remarked that this estimate neglects the electron contribution to the heat capacity of the core, which adds about 50% to its value, raising the total to 6.4 1027 JK 1 27 1 Using = 5.84 10 JK , with a heat flux, dQ / dt = 4.42 1013 W for = 4.5 109 years = 1.42 1017 s , the total heat loss is Q = ( dQ dt ) . = 6.28 1030 J

and T = Q = 1075K

This is an entirely reasonable total cooling of the Earth in its life-time. Kelvins problem arose not from a lack of heat capacity of the Earth, but because he assumed diffusive cooling which is ineffective. 22.1 As in Problem 18.1, the central pressure is 3 GM 2 2 2 2 = P= G R 8 R 4 3 The corresponding compression is given by Eq.(18.37) c P = 1 + K 0 K0 0 and the adiabatic temperature ratio over this compression range is given approximately (assuming constant ) by Eq.(19.22) Tc c = TS 0 so that
1/ K 0

2 K Tc = TS 1 + G2 0 R 2 3 K0

K0

22.2 The fraction of the mantle between r and (r+dr) is 4 3 3 3 3 dV V = 4r 2dr / ( RM RC ) = 3r 2dr / ( RM RC ) 3 The thermodynamic efficiency for convective transport of heat from it is = max ( RM r ) / ( RM RC ) The average efficiency of mantle convection is therefore RM 1 3max 2 = dV = 3 ( RM r ) r dr 3 V ( RM RC ) ( RM RC ) RC 80

4 3 max RM RC ( 4 RM 3RC ) . 3 = 0.4max 3 4 ( RM RC )( RM RC )

11 11 22.3 With K 0 = 2 10 Pa, K 0 = 4 and P = 1.3 10 Pa , Eq. (18.32)gives

P = 1 + K 0 = 1.377 0 K0 Then Eq. (19.22) gives (with constant )

1 K 0

T = = 1.468 T0 0 Thermodynamic efficiency = (T T0 ) / T = 0.32 22.4 The effect of self-compression in Mercury, a small planet, is modest and, assuming similar bulk moduli for inner and outer core materials, has little effect on the energy of compositional separation. We neglect it by making the simplifying assumption that the densities of the components are independent of pressure. Then the size of the core does not change and a straightforward way of calculating the energy of inner core formation is to calculate the energy of accretion with and without the inner core: E = Ei + Eo Eu where Ei is the energy that would be released in forming the inner core in isolation, Eo is the energy released by adding the outer core to it and Eu is the formation energy of the uniform core, before separation. Since it is assumed that the core size does not change, the mantle has no effect. In the real situation, the increased inward concentration of mass increases the internal pressure, causing contraction and increased gravitational energy release. This is partly offset by the irretrievably stored energy of elastic compression. It is significant for the Earth, but is not a major contribution in Mercury. E is a small difference between much larger quantities, so that care is required not to compromise it with rounding errors. Eu and Ei are given directly by the gravitational energies of uniform spheres: 3 GM 2 3 GM i 2 Eu = ; Ei = 5 R 5 0.8R where the core radius is R, M is its total mass and Mi is the inner core mass. Calculation of Eo requires an integral. The mass inside radius r within the outer core range, 0.8R r R, is
m( r ) = M i + M o r 3 (0.8R )3 0.512 M o r3 = (Mi M o )+ 0.488 0.488 R 3 R 3 (0.8 R )3

81

where M o = (MMi) is the outer core mass. The gravitational potential at r is V = Gm(r)/r and the energy released by deposition of an incremental layer is dE = Vdm where dm = 4r2dro = 4r2dr M o /[(4/3)R3 (4/3) (0.8R)3] = [3 M o /(0.488R3)]r2dr and o is the outer core density. Thus
dE 0 = 3GM o 0.512 M o r 4 dr (M i M o ) r dr + 0.488R 3 0.488 0.488 R 3

which is integrated over the range 0.8R to R, giving


E0 =

3GM o 0.512 Mo (M i M o )0.18 R 2 + 0.134464 R 2 3 0.488 R 0.488 0.488 3GM o = [(0.08784 M i + 0.042304 M o ] (0.488) 2 R

Writing Mo = fM, Mi = (1 f)M, we can collect the three terms in the expression for E
GM 2 E = R

0.2352 0.126912 2 2 0.75(1 f ) + (0.488) 2 f (1 f ) + (0.488) 2 f 0.6

0.26352 GM 2 2 2 0.15 + f (1.5 (0.488) 2 ) + f (0.75 0.136608 /(0.488) ) R with no rounding of the numerical values. Now Mo (1 (0.8)3V o 0.512 i 1 f = = = (1 + ) 3 3 M (0.8) V i + (1 (0.8) )V o 0.488 o =

where V is the core volume. Taking i/o = 1.05, E = 0.00272GM2/R, but if the 5% density contrast is interpreted as i/o = 1/0.95, then E = 0.00286GM2/R. With numerical values from Table 1.2, M = 2.2421023 kg, R = 1.913106 m, E = 4.771027 J or 5.021027 J, according to interpretation of i/o. In Section 24.8 the heat released by inner core formation in Mercury is estimated to be 9.41028 J, slightly less than the conducted heat. On this basis there would be no dynamo, but, with the compositional energy estimated above, it becomes viable. The efficiency of thermal convection would be about 7%, so refrigerator action would be very efficient, requiring power only 7% of the heat carried down, and leaving sufficient convective energy to maintain a weak dynamo for the entire life of the planet. Although these numbers depend on guesses, such as the size of Mercurys inner core, it is hard to avoid

82

the inference that compositional convection is necessary for dynamo action in Mercury. 23.1 The cooling of the mantle is estimated from calculations in Chapter 23 to be about 5.510-8 K/year. The volume expansion coefficient varies between 910-6K-1 and 4010-6 K-1 according to the data in Appendix G. If we assume an average value of 2110-6K-1 then the 1 12 1 volume contraction of the mantle is V dV / dt = 1.15 10 year . The core cools more slowly than the mantle, so we estimate the contraction of the whole Earth generously by assuming the mantle value throughout. Since mass is constant and C Ma 2 V 2 3 , C 1dC / dt = (2 3)1.15 1012 year 1 = 7.7 1013 year 1 . The corresponding change in rotation is 1 13 1 therefore d / dt = +7.7 10 year . The tidal friction effect 10 1 (Eq.8.31) is 2 10 year and the thermal contraction is not of observable magnitude. 23.2
& &2 Convective strain rate Q & & Mechanical power W Q (constant efficiency) & 1 where is stress, so Q & & 1 & 2 & 3 Viscosity = Q / Q = Q

23.3 Substituting the numerical values and converting to more convenient units, the heat balance equation can be rewritten
& dT/dt = 1.313107 exp(31250/T) Q R04.2645910-3 exp(3.465710-4 t) & where t is time measured backwards in units of 106 years and Q R0 is the present radiogenic heat in terawatts. The object is to find the & & value of Q R0 that gives T = 2500 at t = 4500. The result is Q R0 = & & 25.7 terawatts. The corresponding present cooling rate is ( Q 0 Q R0 )/m = 27.1 K/109 years. This is the rate of fall in potential temperature, Tp, that is the mantle temperature adiabatically extrapolated to P = 0. The cooling rates for the higher temperatures in the deep mantle are higher by the factor T/Tp, which is about 1.64 near the base of the mantle. Although this is a simplified model, it is realistic and conveys the essential physics of mantle cooling.

23.4 The standard formula for calculation of an adiabatic temperature profile is


P 1

ln(T1/T2) =

P2

( / K

)dP

83

but it is simpler to use


1

ln(T1/T2) =

d( ln )

which is identical if the temperature gradient is precisely adiabatic (and if the P, K and tabulations are compatible). In the case of the core, extrapolation from 5000 K at the inner core boundary gives T = 3739 K at the core-mantle boundary, as in Table G1. The outer core is believed to be very close indeed to adiabatic so there is no difficulty in the interpretation. In the mantle, adiabatic extrapolation from 1950 K at r = 5701 km yields 2616 K at the core-mantle boundary, implying a 1122 K temperature increment across the thermal boundary layer at the bottom of the mantle. This is somewhat larger than the increment modelled in Appendix G, because seismological evidence, discussed in Section 17.5, indicates that (apart from the thermal boundary layer) the lower mantle is slightly superadiabatic. Using the increment calculated here, the gradient in a 200 km thick boundary layer is 5.610-3 K m-1. With a thermal conductivity = 5 Wm-1K-1, and a surface area A = 4R2 = & 1.521014 m2, the conducted heat flux would be Q = AdT/dz = 12 4.2710 W. In section 21.4 conducted heat from the core is estimated to be 3.79 1012 watts. 23.5 The mantle potential temperature falls at a rate & & dTP / dt = (Q0 QR 0 ) / m and the upper bound on the core cooling rate (at the core-mantle boundary, dTCMB /dt ) is 1.64 times this. This imposes an upper bound on the rate of heat loss from the core, & max Q core = c dTCMB/dt if there is no radiogenic heat. Any radiogenic heat is added to this maximum. A table summarizes the results
& Q R0 (tW) d Tp/dt (K/109 years) d TCMB/dt (K/109 years) & max Q core (tW) (no radioactivity) & max Q core (2 tW of 40K)

12 87 143 19 21

20 53 87 11.7 13.7

24 36 59 7.9 9.9

29.5 13 21 2.8 4.8

Without implausibly strong radioactive heating of the core, the high & Q R0 model does not allow sufficient core heat loss for dynamo action. This model must be disallowed. The other models are not affected by this argument.

84

24.1 In the absence of magnetizable materials, the field energy per


2 unit volume is B 2 0 . If the internal field Bi = 2 B0 , then the

internal field energy is Bi2 4 3 8 B02 3 . R = Ei = R 2 0 3 3 0 Here B0 is the equatorial field just outside the sphere, which has half the field strength at the poles, which is equal to Bi . The external field intensity at (r , ) is given by Eq. (24.10) Be = B0 (1 + 3cos 2 ) R 3 / r 3
12

Thus for a volume element at (r , ) , which is a ring of radius


r sin , width rd and thickness dr, the field energy is

dEe = ( Be2 / 20 ) .2r 2 sin d dr

2 B02 R 6 (1 + 3cos ) sin d dr = r4 0

Integrating over ranges 0 < < and R < r < 4 B2 Ee = 0 R 3 = Ei / 2 3 0

24.2

(a) Consider the elementary loop, shown shaded in the crosssection. The cross-sectional area is dS = 2 R cos dr where
r = R sin so that dr = R cos d ; dS = 2 R 2 cos2 d

85

If the current density is i, then the magnetic moment of the element is dm = i dS . A = i.2 R 2 cos2 d.( R sin ) 2 The total magnetic moment is m = 2R 4i
2
0

cos2 sin 2 d =

4 Ri 8

4 3 2 2 The dissipation is i edV = i e . R 3 256 m 2e = 3 3 R 5 (b) With i = i0 r / R , dm = i0 ( r R ) 2 R 2 cos2 d.( R sin ) 2 (where r = R sin ) = 2R 4i0 cos2 sin 3 d
m = ( 4 15 ) i0 R 4
2 dE = i 2edV = (i0 sin )2 .e .2 R 2 cos2 d.2R sin 0 dt
2 = 4 R 3i0 e

cos2 sin 3 d

8 3 2 15 m 2e R i0 e = 15 2 R 5

24.3 Using the result of Problem 24.1 B02 3 3 E = Ei = 4 R 2 0 m 6 where B0 = 0 3 ; R = core radius = 3.48 10 m 4 R m2 E= 0 3 4 R From Problem 24.2(a) dE 256 m 2e = dt 33 R 5 The time constant for decay of current is twice the time constant for decay of magnetic energy, because energy varies as the square of current.

86

Thus =

2E 32 0 R 2 = = 2.20 1011 s=6970 years dE dt 512 e

With the same numerical values Eq.(24.43) gives 12215 years.

24.4 By Eq.(24.35), skin depth is z0 = ( 2 0 )


12

With = 2.76 105 Sm -1 at the top of the core, this gives z0 = 170 km at = (2 /1000) year 1 or 17.0 km at = (2 /10) year 1 .

24.5 Using the result of Problem 7.3 i = 0 cos


2 we have = 0 i = 0 (1 cos ) 0 / 2

Taking the westward drift rate as 0.2/year, = (2 0 )1 2 = 1.7 103 rad

24.6 We use the same core gravity and density profiles as in Eqs. (22.31) and (22.32), so that the fraction of core mass in the radius range dr is given by Eq. (22.33). Now we require the gravitational potential difference between arbitrary radius r and the core-mantle boundary at R V = gdr =
r R

g R x +1 x +1 ( R r ) /( x + 1) Rx

so that the equation corresponding to (22.35) is R dm( r ) g R ( x + 2) x +1 x +1 2 x +2 E = m V =m x +2 ( R r r )dr Mc R ( x + 1) ri R ri x + 2 ri 2 x +3 ( x + 2) x +1 + 2 x +2 =mg R 2 x + 3 R ( x + 1) R ( x + 1)(2 x + 3) With numerical values, this is 3.011028 J, that is of the energy released by an equivalent light mass arriving from the inner core. The possibility of a continuing chemical exchange across the coremantle boundary has been suggested from time to time, but this calculation assumes that the deposited material has the same density as the inner core. If the deposited material were less dense than the 87

outer core, it would remain at the top as a gravitationally stable layer and not contribute to core energy. If the deposited material were more dense then it might either mix into the outer core or sink to add to the inner core. Only in the latter case would the convective energy be competitive with that derived from conventional inner core growth. 24.7 For both parts of this problem we start with B = B (1 + 3cos2) 1/2 and integrate either B2 or B over the surface, with the geometry indicated in the figure. In both cases it suffices to integrate over a hemisphere because of symmetry about the equator.

B2 (a) < B >= 0 2 2a


2

/2

(1 + 3cos
0

)2a 2 sin d = 2 B02

Brms = 2 B0

(b)
/2

< B >= B0

(1 + 3cos
0

1/ 2

sin d = B0 (1 + 3x ) dx = 3B0 ( x 2 + 1/ 3)1/ 2dx


2 1/ 2 1 1

Use a standard integral (e.g. Dwight, 1961, item 230.01):

88

2 2 1/ 2 ( x + a ) dx =

x 2 a2 ( x + a 2 )1/ 2 + ln x + ( x 2 + a 2 )1/ 2 , with a 2 = 1/ 3 2 2


1

1 x < B >= 3B0 ( x 2 + 1/ 3)1/ 2 + ln x + ( x 2 + 1/ 3)1/ 2 6 2 0 = 1.38 B0

25.1 The probability that both intrusive and intruded rocks self-reverse is f1 f 2 and the probability that neither reverses is (1 f1 )(1 f 2 ) . Thus the probability of agreement between them is P = f1 f 2 + (1 f1 )(1 f 2 ) = 1 f1 f 2 + 2 f1 f 2 Therefore the probability of disagreement is F = 1 P = f1 + f 2 2 f1 f 2 For F= 0.02 and f1 = f 2 = f , there are two solutions f = 0.01 or 0.99. 25.2 With 0 = 109 s-1, when increases from 100 s to 108 s, E/kTB increases from 25.32 to 39.14. This arises from changes in both E, which increases, and TB, which decreases, and we need to relate these changes. We write ln(E/kTB) = 0.436 = lnE lnTB (Eq. 1)

By Fig. 25.2, at T/C = 0.9, ms/m0 = 0.44 and d(ms/m0)/d(T/C) = 2.2, so that dln(ms/m0)/dln(T/C) = 4.5. But with E ms2 (assuming the barrier to moment reversal to be magnetostatic energy of grain shape), dlnE/dlnms = 2 and therefore dlnE/dlnT = 9.0. Identifying T with TB, this result combines with Eq. (1) above to give the changes in TB and E for a 106-fold increase in cooling rate: lnTB = 0.0436, lnE = 0.392. This is taken at TB/C = 0.9, with C = 853 K, that is TB = 768 K and TB = 33.5 K. Thermoremanence in small fields is a fractional alignment of grain moments represented by Eq. (25.4), with B/kT << 1, so that, at the blocking temperature, TB, m/ms = B/kTB. With cooling to a low temperature, at which m is measured, it increases by a factor close to m0/ms, so that the measured moment is mcold = m0(B/kTB). In this approximation, the value of ms at T = TB is irrelevant to the strength of mcold because the higher moment of the slowly cooled sample at its blocking temperature is compensated by the greater increase in ms from the higher blocking temperature of the rapidly cooled sample. Thus, the difference in their moments arises only from the difference 89

in 1/ TB. The ratio B / kTB << 1 , so that tanh(B / kT ) in Eq. (25.4) and the error in the paleointensity estimate is linear in the difference in 1/ TB for the two cooling rates. The slowly cooled sample would have remanence stronger by the factor (1 + TB/TB), that is 4.2% by these numbers. Paleointensities are sufficiently uncertain for this not to be a significant problem. 25.3 Although paleomagnetic pole positions are distributed over a sphere, the angular spread is small enough to allow a statistical treatment of the two dimensional spread as being on a plane surface, with displacements from the centre in degrees. Another standard assumption, which we follow, is that paleomagnetic poles from rocks with an age range short enough for polar wander to be negligible are randomly distributed about their mean, with a normal (Gaussian) distribution. We assume also that the scatter is symmetrical, with equal standard deviations in x and y directions and no correlation between x and y displacements. Although we are assuming a circular distribution, we note that distributions may be elliptical because the non-linearity of the relationship between magnetic dip and angle to the pole (Eq. 24.11) means that a circular distribution of magnetization directions gives an elliptical distribution of poles and vice versa. 30 million years of polar wander at 0.3 per year means a 9 displacement of the pole, so that adequacy of the data to discern this is represented by 95% confidence that the mean of the observations falls within a 9 circle. The 15 rms scatter refers to the scatter in r = (x2 + y2) 1/2. The usual probability integral, applicable to scatter in one dimension, must be modified to recognize that in two dimensions the areas of successive rings are proportional to their radii. In this case the probability that any point will lie within radius R is
P=

1 exp( r 2 /(22 )) rdr 2 0

and, for P = 0.95, R = 2.448, where is the rms radial displacement from the mean. [Tables of this integral can be found, for example, in R. S. Burington and D. C. May, 1970, Handbook of probability and statistics, 2nd ed. New York: McGraw-Hill, but it is not a difficult numerical problem]. With = 15, R = 36.72. The displacement of the mean position of N data points will, on average, differ from the true mean by /(N + 1) 1/2 and R scales accordingly, so, to bring the circle of 95% confidence within 9, we require N (36.72/9) 2 = 16.6, that is a minimum of 17 data points. 25.4 The logarithmic time variation of viscous remanent magnetization (VRM) is very generally observed in a wide range of materials and it is not surprising that it can be explained by more than one theory. This question is concerned with the single domain theory, emphasised by Dunlop and zdemir (1997), which relates directly to 90

the theory of thermoremanence in Section 25.2. At the end of this answer we summarize also an alternative theory that is relevant to larger grains. A basic difficulty is that, although VRM is very well observed, the microscopic processes that give rise to it are not and there is, therefore, little effective direct check on the theories. The attempt frequency, 0, in Eqs. (25.1) and (25.2), is very high and, for viscous magnetization to occur on laboratory or geological time scales, E/kT >> 1. This is why blocking temperatures appear sharp. The transition from the superparamagnetic state, in which spontaneous changes in magnetization occur more rapidly than observations, to a stable state, in which there are no changes on the time scale of observation, occurs over a very limited range of E/kT. The postulate that, over the range of interest, values of E are uniformly distributed means that, by Eq. (25.2), at a fixed temperature the values of ln are uniformly distributed. Although the thermally activated viscous changes in magnetization occur statistically and not at specific times, t, on a lnt scale the changes in moment are concentrated in a limited range in lnt around ln. With a uniform distribution of ln, the magnetization varies linearly with lnt. This single domain theory of VRM assumes that the grains, and their magnetic moments are independent, but the lnt law is not restricted to this situation. It is evident that, for the purposes of paleomagnetism, the most important grains are those of the pseudosingle domain type and that they may have volumes up to 104 times the single domain size. It is improbable that they have independent psd moments and a compromise with multidomain theory is needed. The essential feature of multidomain theory is that magnetostatic energy is associated with every grain moment and can be represented by a self-demagnetizing field proportional to the moment. This field arises from the magnetic polarity at grain surfaces (and at some internal domain boundaries). In zero external field it acts in a direction that opposes the magnetization, but in an external field the lowest energy state is one with a magnetic moment parallel to the field. In an external field B, there is an internal field (B Nm), where N is a demagnetizing coefficient that depends on grain and domain geometry. The lowest energy state is one for which this internal field is zero and the effect of the field is to drive the magnetization towards this state. The internal (self-demagnetizing) field biases the potential barriers opposing magnetization changes. For the case of zero external field, a barrier of energy E0 in the absence of magnetization becomes (E0 + cm) for changes that would increase m and (E0 cm) for changes that reduce m, where c is a constant. The sharpness of the dependence on E of the probability of a change is so strong that the first of these can be neglected (there is a negligible probability of a change that increases m). Thus, the rate of change of viscous magnetization is dm/dt dP/dt = 0exp [(E0 cm)/kT] = {0exp [E0/kT]}exp(cm/kT)

91

The quantity in curly bracket is a constant that we can couple with the proportionality constant and write as A, so that dm/dt = Aexp(cm/kT) Integrating from m0 at t = t0, we have the required lnt variation m = m0 (kT/c)ln[(cA/kT)(t t0)] A more detailed discussion of the multidomain approach to VRM is given by Stacey F. D., 1963, The physical theory of rock magnetism. Advances in Physics 12: 45-133; and by Stacey, F. D. and Banerjee, S. K., 1974, The physical principles of rock magnetism. Amsterdam: Elsevier. 25.5 Analysis of the statistics of geomagnetic reversals is very dependent on the completeness of the reversal record. Although periods of constant polarity are generally much longer than the time taken for a reversal, there is no lower bound and brief polarity intervals merge with events and excursions that defy objective judgement over whether they are properly counted as reversals. Nevertheless, all of them potentially provide information on the reversal mechanism and the state of the core during a reversal. The interest in very short polarity intervals concerns the doubt over whether reversals are strictly independent events. Immediately after a reversal are further reversals either more or less likely to follow? Without a perfect record, these alternatives may be hard to distinguish. Suppose that, superimposed on a random pattern, there is an increased probability p that any reversal will be followed immediately by another. The probability that the first reversal will stick is (1 p). But the probability of the second one sticking, once it has occurred, is also (1 p). The chance that the final state is opposite to the initial one is an infinite sum of probabilities 1 p + p2 p3 = 1/(1 + p) In the limit p 1, this sum becomes . If fine details of the record are unresolved, this could appear as a reduced probability of very short polarity intervals, when in fact it is an enhanced probability. 26.1 Let the orbital angular momentum per unit mass of the Earth be L m = l = r 2 d / dt where r is the instantaneous distance from the Sun. Then dt = ( r 2 l )d and the solar energy received in time dt is dE = SA( r0 r ) 2 dt = SA( r02 l )d

92

where S is the solar constant, A is the Earths cross-sectional area and r0 is a reference distance (1 A.U.). Thus the average rate at which solar energy is received is dE SAr02 d SAr02 2 < >= < >= . dt l dt l T where T is the orbital period. But, by Keplers third law 2 3 2 p a , where a = (Eq.B.23), T = by Eq.(B.15) and GM (1 e 2 ) p = l 2 GM by Eq.(B.11). Thus dE < >= (GM ) 2 SAr02 / l 4 (1 e 2 )3 2 dt where factors in the square bracket are constant (assuming constant angular momentum). 26.2 Solar tidal friction transfers rotational angular momentum (C) to the Earths orbital motion. Assuming linearity of tidal friction, so that solar and lunar components can be separated, 21% of the total is caused by the solar tide, so, with C from Appendix A and d/dt by Eq. (8.31), the rate of change of orbital angular momentum is da/dt = 0.21C(d/dt) = 1.0971016 kg m2 s-1 = d(MEr2)/dt where ME is the Earths mass, r is the orbital radius and = 2/(1 year) = 1.9910-7 rad s-1 is the orbital angular velocity. By Keplers third law (Eq. B23, Appendix B), 2r3 = GMS, where Ms is the mass of the Sun, so we can write the variations in and r separately dr 2 r 1/ 2 da ( ) = = 1.233 1013 ms-1 = 3.9 104 cm/yr dt M E GM S dt d 3 da = = 2.46 1031 rad s-1 2 dt M E r dt which corresponds to 1.210-9 s annual increase in the length of the year. Both numbers are completely insignificant and have no climatic implication. To emphasize the point, if the present angular momentum, 5.861033 kg m2 s-1, were transferred to the orbit, the orbital angular momentum would increase by only 0.22 parts per million. 26.3 (a) If sea level rises by transfer of mass m from polar regions, where it contributes negligibly to the axial moment of inertia, to the oceans 93

which are assumed to be spread without a latitude bias (and if the transfer occurs too rapidly for restoration of isostatic balance), then the axial moment of inertia of the Earth is increased by (2/3)mR2. With sea covering 3.621014 m2, 2 mm of sea water (density 1025 kg m-3) has a mass of 7.421014 kg and its moment of inertia is 2.01028 kg m2. With conservation of angular momentum, the annual change in rotation rate is / = C/C = 2.510-10, which means an annual increase in the length of day of 0.022 ms. This is comparable to the long term annual increase caused by tidal friction, 0.024 ms, but much less than the short term fluctuations seen in Fig. 7.4. If a surface layer of the ocean, depth z, expands to depth (z + z) then the mean radius of the layer increases by z /2. The resulting change in moment of inertia is C = (2/3)(zA)[(R + z/2) 2 R2] = (2/3)ARzz = 3.151021 z (kg m2 for z in metres) where (zA) is the mass of the layer. If z is assumed to be 1000 m, then C/C = 3.910-14, almost a factor 104 less than for the ice melting case. 26.4 (a) The heat required (annually) for melting is Q = ML = 2.51020 J, where L = 3.35105 J kg-1 is the latent heat of melting. This amounts to 7.9 Twatts. (b) The heat required to raise the temperature of a volume V by T is Q = VCP T and the expansion that it causes is V = VT, so that Q = CPV/ = CPAz/ Notice that we do not need to know V or T. CP = 3990 J K-1 kg-1 and = 1025 kg m-3 are well known, but the choice of a value for is problematic because it is very temperature sensitive. The range of sea surface temperatures is 0 C to 30 C and if we take the average, 15 C, then = 1.510-4 K-1 and Q = 2.01022 J, almost 100 times the heat required for the ice melting case. Since we are considering an annual sea level rise of 2 mm, these estimates of the heat required can be converted to watts of average heating rate, 7.91012 W or 6.31014 W for cases (a) and (b). Comparison with the energies in Tables 26.1 to 26.3 shows that these are big numbers. In case (b) the absorbed heat is 0.6% of the solar energy reaching the surface. It is obviously easier to envisage ice melting than ocean heating as the cause of sea level rise. 26.5 The rate of tidal dissipation is the product of tidal torque and angular speed. For the purpose of this problem, the change in angular speed is not significant and the effect of harnessing tides is to increase the

94

phase lag, , in Eq. (8.20). There may also be an effect on the Love number k2, but this is almost certainly secondary, and the other parameters in the equation can be treated as constants. Thus, tidal energy dissipation and rotational slowing are each proportional to , so rotational slowing is directly proportional to the extracted power. Adding 14.4 terawatts to the 3.7 terawatts of natural dissipation (Section 8.3), increases the rate of slowing by a factor 4.9, to 3.21021 rad s-2, corresponding to a length of day increase by 12 ms/century.

95

Vous aimerez peut-être aussi