Stress, Balance Equations, and Isostasy

Chapter 3 Stress, Balance Equations, and Isostasy We define stress to consider forces acting at depths, and derive the equations of mass, linear mome...
Author: Shana Robertson
4 downloads 0 Views 1MB Size
Chapter 3

Stress, Balance Equations, and Isostasy We define stress to consider forces acting at depths, and derive the equations of mass, linear momentum, angular momentum, and energy conservation. Very slow tectonic movements allow us to neglect inertia force and simplify the equation of motion to a balance equation. Isostasy is the most basic but important application of force balance equations to tectonics.

3.1

Definition of stress

Suppose that a square pole is composed of five rectangler solids (Fig. 3.1(a)). If the pole bears a load at the top and base, the parts may push each other across the bounding planes. Parts A and B may push each other with that load. However, parts C and D may push each other with much less force. Force across a plane in a solid depends on the direction of the plane. In order to consider a force exerted on a rock mass, suppose there is a surface element δS on the mass, and a unit vector N normal to the surface and pointing outward indicates the attitude of the surface element (Fig. 3.1(b)). In addition, let the vector δF be the traction force acting on the element. The force may increase with the area δS so that we define a vector quantity t(N ) as the force per unit area exerted there. N is in parentheses because the force depends on the attitude of the surface element. The quantity is called a stress vector. Mathematically, the stress vector is defined by δF . (3.1) t(N ) = lim δS→0 δS The stress vector and its components are expressed in the units Nm−2 = pascals (Pa)1 . 1 The same units are used to count pressure. 1 bar = 105 Pa = 1000 hPa = 0.1 MPa. Engineering literature often uses the units kgf /cm2 which are approximately equal to 0.98 MPa. The unit dyn cm−2 is used in older literature. It is converted by the factor 1 Pa = 10 dyn cm−2 .

59

60

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.1: (a) A square pole loaded at the top and base is composed of five rectangular solids, A through E. (b) A rock mass is pulled from the adjacent mass with the force δF on an element of the surface δS. A unit vector N indicates the attitude of the surface element. (c) Stress vector t(N ) and Cartesian coordinates O-123 whose third axis is parallel to the unit normal N. Consider a surface element δS with the unit, outward normal N, to which the third axis of Cartesian coordinates O-123 is adjusted (Fig. 3.1(b) and (c)). Let e(i) be the ith base vector of the coordinates, then we write t(N ) = S31 e(1) + S32 e(2) + S33 e(3) , (3.2) where S31 , S32 and S33 are the components of the stress vector t(N ), that is, t(N ) = (S31 , S32 , S33 )T . The stress vector depends on the direction of the surface on which it acts. The first subscript ‘3’ indicates accordingly that the surface is normal to the third coordinate axis and its outward normal points positive direction of the axis. The second subscript distinguishes the components of the stress vector. A stress component normal to the surface on which it acts is known as the normal stress on the surface, whereas the component tangent to the surface is known as shear stress. Generalizing Eq. (3.2), the components of stress vectors acting on a surface elements normal to the coordinate axes, we have   Sij = tj e(i) . (3.3) We define the stress tensor from the components as ⎞ ⎛   S11 S12 S13 · · · components of te(1)  S = ⎝ S21 S22 S23 ⎠ · · · components of te(2)  · · · components of t e(3) . S31 S32 S33

(3.4)

The limit in Eq. (3.1) indicates that stress vector is defined at each point on an infinitesimally small surface. The diagonal components are the normal stresses acting on a surface element normal to the coordinate axes, while the other components represents shear stresses acting on the surface elements. Sign conventions Figure 3.2(a) shows the positive direction of those components. Since we have defined the unit vector N as being the outward normal to the surface on which the stress vector is

3.1. DEFINITION OF STRESS

61

Figure 3.2: Sign convention in continuum mechanics (a, b) and solid earth science (c, d). Arrows indicate the direction of the positive stress components acting on a cube whose surfaces are parallel to the coordinate planes. (a) A stress component is positive when it acts in the positive direction of the coordinate axis, and on a surface of the cube whose outward normal points in the one of the positive coordinate directions. (b) A stress component is positive when its direction is opposite to that of (a).  and ⊗ indicate vectors pointing to this and opposite sides of the page, respectively.

defined, the opposite faces of a cube have the unit vectors in the opposite directions. Accordingly, the normal stress on the faces has the opposite positive directions, that is, the normal stresses are positive in sign if the cube is under a tensional stress (Fig. 3.2(b)). Compressive stresses have negative normal stresses. We have seen that the strain tensor E has positive diagonal components if the length parallel to coordinate axes become larger. Therefore, it is natural that positive stress components correspond to positive strain components. However, compressive stresses are common in the crust due to overburden pressure. It is the sign convention in solid-earth science to define the sign of stress as being positive for compression, opposite to the sign convention of continuum mechanics (Table 3.1). Accordingly, we will use the symbol r for the stress tensor with positive components for compression to describe tectonic models. Positive directions of the components of r are shown in Fig. 3.2(c) and

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

62

Table 3.1: Symbols and signs used in this book.

Continuum Mechanics Solid-Earth Science

Stress

Unit Normal Vector

Strain

S (tension = +) r (compression = +)

N (outward) n (inward)

E (elongation = +) e (shortening = + )

Figure 3.3: Direction of the unit normal vectors n and N at the surface of a rock mass. (d). The stress tensor r should be used with the inward normal n (Fig. 3.3), where S = −r,

N = −n.

(3.5)

Stress components in cylindrical coordinates Cylindrical coordinates O-rθz are sometimes convenient (Fig. 3.4). Transformation between the cylindrical and rectangular Cartesian coordinates O-xyz is given by the equations, x = r cos θ,

y = r sin θ,

r =x +y , 2

2

2

z=z

−1

θ = tan (y/x).

Stress components in the coordinate systems are transformed by the two sets of equations σrr = σxx cos2 θ + σyy sin2 θ + 2σxy sin θ cos θ,

(3.6)

σθθ = σxx sin θ + σyy cos θ − 2σxy sin θ cos θ, 2

2

(3.7)

σrθ = (σyy − σxx ) sin θ cos θ + σxy (cos θ − sin θ). 2

3.2

2

(3.8)

Cauchy’s stress formula

  % (i) It is seen from Eq. (3.3) that ti e(j) = Sji = k Sjk ek . Therefore, we have   t e(i) = ST · e(i) .

(3.9)

3.2. CAUCHY’S STRESS FORMULA

63

Figure 3.4: Stress components in the cylindrical coordinates O-rθ. Note that the choice of frame of reference, the base vectors of which are e(i) , is arbitrary. Equation (3.9) is generalized as follows: the stress vector acting on a surface element with the unit outward normal N is (3.10) t(N ) = ST · N, where N may be oblique to the coordinate axes2 . This is called Cauchy’s stress formula. For the sign convention of solid-earth science we have the the equivalent formula t(n) = rT · n.

(3.11)

If t(N ) is the traction on a body at its surface element with the normal N, the adjacent body has the outward normal −N at the surface element. The action-reaction law requires that t(N ) = −t(−N) = −t(n),

(3.12)

which is shown in Fig. 3.3. Normal and shear stresses Normal and shear stresses are the components of t(n) parallel and normal to n, respectively. Thus, Normal stress: Shear stress:

σN = t(n) · n,  σS = |t(n)|2 − σN2 .

(3.13) (3.14)

Since the vector n points inward at the surface of a rock mass (Fig. 3.3), σN is positive if the mass is pushed inward at the surface. By contrast, the normal stress defined by the equation SN = t(N ) · N

(3.15)

is positive for tension. The shear stress for this sign convention is  SS = |t(N )|2 − SN2 . 2 Equation (3.10) is demonstrated via the force balance equation [134], which is introduced in §3.3. However, axiomatic continuum mechanics defines a stress tensor as the linear transformation from t(N ) to t(N ), rather than the force per unit area on the surface elements parallel to coordinate planes. The interested reader will find more information in [243].

64

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

The vector σN (n) = σN n has a magnitude obviously equal to the normal stress σN and is parallel to n. If σN is positive, σN (n) points inward at the surface of the rock in question. The vector σN (n) is obtained by the orthogonal projection of t(n) onto the line parallel to n, so that the projector P (n) = nn (Eq. (C.37)) is used to describe the vector:   σN (n) = P (n) · t(n) = nn · r · n = n · (r · n) n. (3.16) On the other hand, σS (n) is defined as the vector that has a magnitude equal to the shear traction and indicates the direction of the shear traction at the surface element of the mass. The elementary orthogonal projector P⊥ (n) = I − nn (Eq. (C.38)) is useful:   σS (n) = P⊥ (n) · t(n) = I − nn · t(n) = r · n − n · (r · n) n = t(n) − tN (n). (3.17) Hydrostatic pressure From Eq. (3.11), let us derive a useful equation for tectonics: the tensorial formulation of hydrostatic state of stress. Consider the stress tensor defined by r = p I,

(3.18)

where I and p are the unit tensor and a scalar quantity. Coordinate rotation by Q transforms the stress as r = Q · r · QT = Q · (p I) · QT = p Q · QT = p I = r, indicating that the stress defined by Eq. (3.18) is isotropic, that is, independent from orientations. The state of stress is anisotropic if there is the dependence. In this case, the force acting on a surface element dS is t(n)dS = (r · n)dS = (pdS)n, which is perpendicular to the surface and pointing inward with a constant magnitude of pdS. The stress vector is parallel to n for hydrostatic stress state: shear stress always vanishes. Therefore, Eq. (3.18) is known as the hydrostatic state of stress, and p is hydrostatic pressure. The equation S = −p I expresses the same stress for the sign convention of continuum mechanics. Hydrostatic pressure constricts a body: it shrinks with a similar figure with the original one if the body is homogeneous inside. If the body is subject to an anisotropic stress, the shape may change.

3.3

Force balance

Tectonic motion to create a large-scale geologic structure is very slow. Forces acting on a rock mass at depth may be thought of being balanced at every moment. Accordingly, we shall derive balance equations for forces, and derive vertical stress at depth. There are two categories of forces: surface force per unit area of the surface t and body force per unit mass X. The former was discussed in the last section. The latter acts on all the internal elements of a body from its exterior. Examples are gravity and electromagnetic forces. As an example, gravity is written as the vector X = (0, 0, g)T , (3.19) where g is gravitational accerelation and the third coordinate axis points vertically downward. Body force per unit volume is ρX.

3.3. FORCE BALANCE

3.3.1

65

Equation of motion

Consider a rock mass defined by a volume V and a closed surface S. Body and surface forces, X and t(N ) act on every portion in the body and on the surface, respectively. If the forces are balanced, we have   (3.20) 0 = t(N ) dS + ρX dV. Since

S

& V

V

ρv dV is the total linear momentum of the rock mass, the material derivative  D ρv dV Dt V

represents the inertia force. Therefore, the equation of motion of the body is    D ρv dV = t(N ) dS + ρX dV. Dt V S V

(3.21)

The surface integral in the left-hand side of Eq. (3.21) is transformed into a volume integral with Gauss’s divergence theorem (Eq. (C.63)). In addition, combining Eq. (3.10), we obtain    D ρv dV = ∇ · ST dS + ρX dV. (3.22) Dt V V V In this case the order of D/Dt and integral is exchangable3 so that we have    ρv˙ − ∇ · ST − ρX dV = 0. V

Since the choice of the volume is arbitrary, the integrand must vanish to satisfy this equation. Consequently, we obtain the differential form of ρv˙ = ∇ · ST + ρX

or

ρv˙ = −∇ · rT + ρX.

(3.23)

or

0 = −∇ · rT + ρX.

(3.24)

or

0=−

If the forces are balanced, we have 0 = ∇ · ST + ρX The indicial notation of Eq. (3.24) is 0=

 ∂Sji j

∂xj

+ ρXi

 ∂σji j

∂xj

+ ρXi .

(3.25)

3 If V and V are the initial and instantaneous volumes that the rock mass occupies, respectively, at time t = 0 and t, 0 the exchangeability is demonstrated as follows. The initial volume V0 does not depend on time, so that the operator D/Dt can be included in the integral over V0 . In addition, because of V = J V0 where J is the Jacobian, the initial and temporal densities, ρ0 and ρ, are related to each other as ρ = ρ0 /J . The initial one does not depend on time, either. That is, Dρ0 /Dt = D(ρJ )/Dt = 0. Therefore,        D Dv D Dv Dv D(ρJ ) ρ ρv dV = ρvJ dV0 = ρ + ρJ dV0 = J dV0 = dV. v· Dt V Dt V0 Dt Dt Dt Dt V0 V0 V

66

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Based on Eq. (3.24), we shall derive a simple but important quantity for tectonics. Suppose that the ground surface is flat and there is no horizontal density variation. Rocks are subject to the gravitational force shown in Eq. (3.19). Defining z-axis vertically downward from sea level, the force balance in this direction is obtained from Eq. (3.24) as ∂σxz ∂σyz ∂σzz + + = ρ(z)g, ∂x ∂y ∂z where density, ρ(z), is the function of depth, z. Since there is no horizontal variation at all, the first and second terms in the above equation vanish. We therefore have ∂σzz = ρ(z)g. ∂z Integrating both sides, we obtain

z σzz =

ρ(ζ)g dζ,

(3.26)

h

where h is altitude. This equation enables us to calculate the vertical stress σzz from the density &z profile ρ(z). If g is constant over the depth range in question, Eq. (3.26) becomes σzz = g h ρ(ζ) dζ. The integral in the right-hand side of this equation is the mass of the vertical column with the unit cross-section of the height from the surface to the depth z. Therefore, σzz is the stress at the base of the column due to its own weight4 . The vertical stress by the weight is called overburden stress. If density can be assumed to be constant, the overburden at the depth z is simply pL = ρgz.

(3.27)

We have found how the vertical stress σzz is determined. What are the other stress components? Areas of compressional and extensional tectonics may have different stress states. Stress may not only have spatial but also tempral variations, which are represented by components other than the vertical one. However, it is convenient to define a reference stress state for later discussions. Although choice of the reference is arbitrary, the isotropic stress r = pL I is often used for the reference in the models of tectonics, where z pL = ρ(ζ)g dζ

(3.28)

(3.29)

h

is the overburden at depth z. This is the overburden with the same dimension with pressure. Equation (3.28) has the same form as Eq. (3.18), so that the stress indicated by Eq. (3.28) is known as lithostatic stress, and pL is called lithostatic pressure.

3.4. SYMMETRY OF STRESS TENSOR

67

Figure 3.5: Force F at the point P with a position vector x yields a moment of that force M about the origin O. The magnitude of M equals |x||F | sin θ, where θ is the angle between the vectors.

3.4

Symmetry of stress tensor

Both surface and body forces acting on a rock mass exert moments (torques) on the mass. If a force F acts at a point P whose position vector is x (Fig. 3.5), the moment of the force about the origin is & M = x × F . Likewise, the moment about the origin by surface force is S x × t(N ) dS, where x is the position vector of the point where the surface force t(N )dS acts on the rock mass. The moment & by body force is V x × ρX dV , where x is the position vector of a particle in the mass where the body force acts. Therefore, in case that the total moment is in equilibrium to conserve the angular momentum, we have the balance equation of momentum as   0 = x × t(N ) dS + x × ρX dV. (3.30) S

V

If they are imbalanced, the residual moment causes the acceleration in the angular velocity of the body:    D (x × ρv) dV = (x × ρX) dV + (x × t) dS. (3.31) Dt V V S The choice of position of the origin is arbitrary. Hhowever, Eq. (3.31) holds wherever the origin is defined5 . 4 In-situ stress measurements have shown that this picture gives a close approximation to observed σ zz [3]. However, deviations from predicted σzz from the density profile are rarely observed [30, 75]. 5 Let us calculate the total moment about a position with the position vector o, from which position vector x is defined. That is, x = x + o. By this additive decomposition of x, Eq. (3.31) becomes     Left-hand side = (x + o) × ρv˙ dV = x × ρv˙ dV + o × ρv˙ dV, V V V   Right-hand side = (x + o) × ρX dV + (x + o) × t dS V S       = x × ρX dV + x × t dS + o × ρX dV + o × t dS. V

S

V

S

Combining the terms in and outside the square brackets in the above equations, we have the equations    x × ρv˙ dV = x × ρX dV + x × t dS, V

V

S

which is identical to Eq. (3.31) provided that x is replaced by x and      o× ρv˙ dV = o × ρX dV + t dS . V

V

S

()

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

68

It is important for discussions in the rest of this book that the stress tensor is symmetric: S = ST ,

r = rT .

(3.32)

These formulas hold for both static and dynamical problems unless there is neither “couple stress” nor “body torque”. The former is the surface force acting on a body to rotate it about the axis perpendicular to the object surface, and the latter is the distributed moment in the body6 . Equation (3.32) is derived from the conservation of angular momentum7 . Plane Stress It is a basic theorem in linear algebra that all the eigenvalues are of real for a real, symmetric matrix. In addition, the eigenvectors are perpendicular to each other. If one and only one of the eigenvalues is zero, a state of plane stress is said to exist. In that case, taking coordinate axes parallel to the eigenvectors, the stress tensor is written as ⎛ ⎞ S1 0 0 S = ⎝ 0 S2 0⎠ , 0 0 0 where S1 and S2 are non-zero eigenvalues of S. The physical interpretation of the stress state is that traction always vanishes on a plane perpendicular to the eigenvector corresponding to the zero eigenvalue. This is demonstrated by substituting N = (0, 0, 1)T into t(N ) = ST · N. It should be noted that plane strain and plane stress do not go together for most cases (§7.1.3).

3.5

Conservation of energy

Let us derive an equation governing temperature changes from the conservation of energy. Multiplying both sides of Eq. (3.23) by v, we have      ρv · a dV = v · ∇ · ST dV + ρv · X dV, (3.33) V

V

V

where V stands for the volume that the rock mass in question occupies. The left-hand side of this equation is rewritten as     ρ|v|2 v·v D D ˙ ρ ρv · a dV = ρv · v˙ dV = dV = dV = K. Dt V 2 Dt V 2 V V The symbol K stands for the total kinetic energy, so that K˙ is the rate of its increase. On the other hand, the integrand v ·(∇·ST ) in the first term in the right-hand side of Eq. (3.33) has the components vi

∂(vi Sji ) ∂vi ∂(vi Sji ) ∂Sji = − Sji = − (Dij + ωij )Sji , ∂xj ∂xj ∂xj ∂xj

This is identical to the cross-product of the constant vector o and Eq. (3.22), so that Eq. () is valid for any o. Therefore, Eq. (3.31) holds wherever the origin is chosen. 6 See advance continuum mechanics books such as [132]. 7 The derivation needs a tricky calculation. See, for example, [134].

3.5. CONSERVATION OF ENERGY

69

where Lij and ωij are stretching and spin tensors, respectively. Therefore, Eq. (3.33) is rewritten as  K˙ +





D : S dV = V

∇(v · S) dV − V

 W : S dV +

V

ρv · X dV.

(3.34)

V

Since W = (ωij ) is antisymmetric we have W : S = 0. Combining this and Eq. (3.34), we obtain  K˙ +

 D : S dV =

V

 ∇(v · S) dV +

V

ρv · X dV.

(3.35)

V

Converting the first term on the right-hand side of this equation into a surface integral by Gauss’s divergence theorem, we obtain the energy conservation equation    (3.36) K˙ + D : S dV = [v · t(N )] dS + ρv · X dV. V

S

V

Energy equals the product of force and distance, so that the rate of energy change equals the product of force and velocity. The right-hand side of Eq. (3.36) indicates the energy that the rock mass accepts from outside. The second term on the left-hand side is the energy dissipation due to the deformation of the mass against the stress S. The dissipated energy is transformed to thermal or internal energy. Let e be the internal energy per unit mass, then the internal energy per unit volume is equal to ρe, and   ρe˙ dV = D : S dV. (3.37) V

V

Therefore, the energy conservation equation is rewritten as D Dt



 ρ V

   v2 + e dV = v · t(N ) dS + ρv · X dV. 2 S V

(3.38)

The left-hand side of this equation represents the energy increase of the rock mass, and the right-hand side indicates the work done by its outside. Equation (3.38) was derived from the equation of motion multiplied by velocity. That is, only the balance of kinetic energy was taken into account. However, vertical motion of a rock mass causes cooling or heating of the mass. Temperature is an important factor in controlling the mechanical properties of rocks. Accordingly, heat transfer is important in understanding the mechanical aspect of tectonics. Heat flux is defined by the amount of heat that passes through a unit area in a unit of time. The heat flux is a vector quantity, so let us use the symbol q for the quantity. The heat energy passing through the surface dS of a rock mass in a unit time equals N ·q. Substituting this into the right-hand side of Eq. (3.38), we have D Dt



 ρ V

    v2 + e dV = v · t(N ) dS + ρv · X dV − N · q dS. 2 S V S

(3.39)

70

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Using Gauss’s divergence theorem again, the last term in this equation is converted to a volume & integral, and we combine all terms into the form V (· · · ) dV = 0. Since the volume V is arbitrary, this equation is satisfied only if     D v2 + e = ∇ · S · v − q + ρX · v. (3.40) Dt 2 This is called the energy equation. Combining Eqs. (3.40) and (3.33), we have e˙ = −∇ · q + ∇ · (S · v) − v · (∇ · S). Components of the first and second terms in the right-hand side are     ∂(Sij vj ) ∂Sij ∂vi − vi Sij = S : (∇v). = ∂x ∂x ∂x i j j i,j i,j Therefore, we finally obtain the equation ρe˙ = S : (∇v) − ∇ · q.

3.6

(3.41)

Self gravitational, spherically symmetric body

The direct objects of geological observations occupy shallow levels in the Earth. However, those objects sometimes provide constraints on ancient global changes including planetary differentiation. The differentiation occurred at the beginning of planetary evolution, so that later geological processes have erased all traces of the event in the Earth. However, satellites and terrestrial planets smaller than the Earth have geological structures that evidence the early history of the planetary bodies. For example, wide rift zones on the Jovian satellite Ganymede are thought to be evidence of planetary differentiation (Fig. 3.6). An icy mantle ocupies one third of the radius, and 60% the mass of the satellite. The surface of Ganymede has high and low albedo regions which appear as light gray and dark terrains, respectively, in Fig. 3.6. Number density of impact craters indicates that the dark terrains are older than the lighter ones. White spots in this photograph are ejecta blankets from young impact craters. The light, young terrains with stripes are called sulci. These are thought to be the surface manifestation of normal faults, and the sulci are wide rift zones [171]. The formation of those normal normal faults is modeled in Section 12.4. Ganymede was initially wrapped up in a dark icy layer, but global expansion broke the layer into dark terrains. Young, light gray ice spread between the terrains. The global expansion was due to core formation in the satellite; initially deep-seated ice was raised to shallow levels in the moon by sinking rocks and metalic materials. The phase transformation of the ice by decompression caused the global expansion. We have seen in the previous section (§3.3.1) that the vertical stress σzz is determined by gravity, where the gravitational acceleration g is assumed to be constant. However, g may vary with depth

3.6. SELF GRAVITATIONAL, SPHERICALLY SYMMETRIC BODY

71

Figure 3.6: Surface of Ganymede, the largest Jovian satellite with a radius of 2640 km. The circular region labeled “E” is the palimpsest Epigeous, an old and degraded large impact basin 350 km in diameter. Voyager image, 0370J2-001. Courtesy of NASA.

in the real Earth. Accordingly, in this section we study the depth-dependence of g and lithostatic pressure. Firstly, let us assume a density distribution with spherical symmetry, and M (r) be the mass

72

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

within the distance r from the center. The gravitational acceleration at the radius is g(r) =

GM (r) , r2

(3.42)

where G = 6.67 × 10−11 Nm2 kg−2 is the gravitational constant. If density at the radius is ρ(r), we have  r

M (r) =

4πr  2 ρ(r ) dr .

(3.43)

0

Combining Eqs. (3.42) and (3.43), we obtain the gravitational acceleration as a function of r as  G r (3.44) g(r) = 2 4πr  2 ρ(r ) dr . r 0 Scondly, pressure p at r is obtained from the force balance equation (Eq. (3.25)). As the force in this case has merely originated from the self-gravity of the spherically symmetric body, there is no horizontal variation in the stress field. Equation (3.25) is rewritten for the spherical coordinates O-rθφ as ∂σrr 1 ∂σθr 1 ∂σφr 0=− − − − ρg, (3.45) ∂r r ∂θ r sin θ ∂φ where both ρ and g are the functions r and the latter is given by Eq. (3.44). The gravity term ρg in Eq. (3.45) is negative in sign, because gravity points downward, i.e., in the direction of −r. If the spherical body is composed of viscous fluid at rest, the state of stress in the body is hydrostatic. In this case, shear stresses σθr and σφr vanish, and in Eq. (3.45) σrr should be replaced by the hydrostatic pressure p. Consequently, we have dp = −ρg. dr

(3.46)

As ρ and g are always positive in sign, Eq. (3.46) indicates that p should decrease with r. Let us calculate g and p as the functions of r for the two cases in Fig. 3.7. Density in the layers is assumed to be constant. The former is a simple model for the Moon. Our Moon had its own magnetic field a few billion years ago, suggesting that there was a melted metalic core at the center. However, the core is so small that it is difficult to estimate its dimension. Accordingly, we assume the Moon to be a rocky spherical body with a constant density ρ = 3.3 × 103 kg m−3 to roughly estimate the gravitational acceleration and pressure at depths8 . In the case of a constant density ρ, the integral in Eq. (3.43) results in M (r) = 4πρr 3 /3, and we obtain g(r) =

4π Gρr. 3

(3.47)

8 The Moon has a less dense crust than the mantle. However, the mean thickness of the crust is less than 100 km, two orders of magnitude smaller than the radius of the Moon, which is about 1740 km. Accordingly, the crust is neglected in this estimation.

3.6. SELF GRAVITATIONAL, SPHERICALLY SYMMETRIC BODY

73

Figure 3.7: Pressure p and gravitational acceleration g versus distance r from the center of a spherically symmetric body. (a) Single-layer model Moon with a constant density with the parameters ρ = 3.3×103 kg m−3 and R = 1740 km. (b) Two-layer model Earth with the parameters ρm = 4×103 kg m−3 , ρc = 10 × 103 kg m−3 , Rc = 3490 km, and R = 6380 km. It follows that gravitational acceleration for a constant density, spherically symmetric body is proportional to the distance from the center (Fig. 3.7(a)). Substituting Eq. (3.47) into (3.46), we have 4πGρ2 dp =− r. dr 3 Using the boundary condition p = 0 at r = R, where R stands for the radius of the Moon, we obtain the pressure distribution 2π 2 (3.48) ρ G(R2 − r2 ). p(r) = 3 The density distribution in the Earth is roughly simulated by a two-layer model, as about half the radius of the Earth is occupied by a high density, metalic core at the center and the crust is negligibly thin compared to the core and mantle. We do not take into account the tiny density difference between the inner and outer cores. Let R and RC be the radius of the Earth and the core, respectively, then the density is ' ρ=

ρC ρm

(r ≤ RC ) (RC < r ≤ R).

Gravitational acceleration in the core is the same as Eq. (3.47) and we have g(r) =

4π GρC r 3

(r ≤ RC ),

(3.49)

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

74

where ρC is the density of the core. If that of the mantle is ρm , gravitational acceleration in the mantle is obtained by using Eqs. (3.44) and (3.49) as 4πG g(r) = 3

ρm r + (ρC − ρm )

RC3 r2

! (RC < r ≤ R).

Figure 3.7(b) shows g(r) for a double-layered model Earth. It is seen in this figure that g is roughly constant in the upper mantle. The models of tectonics in the following sections of this book deal with the crust and the upper mantle, so that we will always assume that gravitational acceleration is constant. Pressure distribution in the model Earth is given by integrating Eq. (3.46) as   1 1 4π Gρm (ρC − ρm ) RC3 − 3 r R    2π  2π + Gρ2m R2 − RC2 + Gρ2C RC2 − r2 3  3    4π 1 2π 1 3 p(r) = − + Gρm Rc (ρC − ρm ) Gρ2m R2 − r2 3 Rc R 3

p(r) =

(r < RC ) , (RC < r < R) .

The gravitational acceleration and pressure in the model Earth are calculated with the values, ρm = 4.0 × 103 kg m−3 , ρcore = 10 × 103 kg m−3 , Rc = 3490 km, R = 6380 km, and G = 6.7 × 10−11 Nm2 kg−2 , and are shown in Fig. 3.7. It is seen that gravitational acceleration is more or less constant in the upper mantle. Therefore, the acceleration is always assumed to be constant in the following sections of this book.

3.7

Isostasy

Buoyancy is an important force to drive tectonics. The crust has smaller density than the underlying mantle so that the crust is subject to buoyancy forces from the mantle and is floating on it. The tectonic thinning and thickening of the crust cause the subsidence and uplift of the surface. We will derive Archimedes’ principle from the force balance equation. Suppose a solid body is immersed in a fluid with a density of ρ (Fig. 3.8). The body and fluid are assumed to stand still so that all forces acting on the body are balanced. The force balance equation (Eq. (3.20)) in this case is   (3.50) 0 = t(n) dS + X dV, S

V

where V and S are the volume and surface of the body. The first and second terms on the right-hand side of this equation represent surface force from the fluid to the body and gravitational force (Eq. (3.19)), respectively. The buoyancy force exerting on the body is the former. The traction at the surface is pressure, p. That is, t(n) = pn. Using Gauss’s divergence theorem, the surface integral is

3.7. ISOSTASY

75

Figure 3.8: Schematic illustration to explain Archimedes’ principle. converted to a volume integral. Therefore, the buoyancy term is rewritten as   pn dS = − ∇p dV. S

(3.51)

V

A minus sign is placed on the right-hand side of this equation because the unit normal n is defined inward, unlike the case of Eq. (C.63). It should be noted that the integration on the right-hand side of Eq. (3.51) is taken inside the body, although p is not the pressure in the body but fluid pressure at the surface. Since the fluid is at rest, the pressure depends only on the depth z. The pressure gradient, ∇p, and the gravity term in Eq. (3.50) have no horizontal component. Therefore, an equation of scalar quantities is enough to consider the vertical force balance of this system. Let Fb be buoyancy, and ρ be constant, then Eq. (3.50) is converted to the equation  Fb = −ρ dV = −ρV, V

where the negative signs came from the definition of the force Fb that is positive upward. This equation stands for Archimedes’ principle: an object completely immersed in a fluid experiences an upward buoyant force equal in magnitude to the weight of the fluid displaced by the object. The concept of isostasy is derived from that principle. Suppose the structure shown in Fig. 3.9, where a continental crust with a thickness of tc is floating on a mantle. According to Archimedes’ principle, the buoyancy force exerted at the base of the crust is the weight of the mantle displaced by the part of the crust that is immersed in the mantle. The weight is ρm ghc S, where ρm the mantle density and S the area of the base. Other symbols are shown in Fig. 3.9. The gravitational force that the crustal block experiences is ρc gtc S, which should be balanced with the buoyancy force. Therefore, we have ρc gtc = ρm ghc .

(3.52)

76

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.9: A simple model for a continental crust floating on a mantle. This equation says that the lithostatic pressure at the level of the base is spatially constant. If density variations in the upper mantle are not negligible, this level should be taken to be deeper. The lithospheric plate that is composed of a crust and upper mantle is often envisaged as floating on a fluid asthenosphere. Rocks behave as fluid in a geological timescale, if they have high temperature. On the other hand, if the lower mantle is heated to become fluid, upper crustal blocks with smaller densities float on it. The level is placed in the middle of the crust. The level under which rocks behave as fluid and they cannot support a lithostatic pressure difference is called the depth of compensation. A simple model of isostasy asserts that overburden is constant at a depth of compensation. The following equation represents this concept:  dc

ρg dz = constant,

(3.53)

surface

where dc is the depth of compensation, and the z-axis is defined downward with the sea level at z = 0. If an offshore region is considered, the lower bound of this integral is taken to be at sea level to take into account the load of the ocean water. The integral is spatially or temporarily constant. When it is thought to be temporarily constant, we can calculate subsidence and uplift accompanied by the temporal variation of density distribution at depths, i.e., vertical movements that are recognized from stratigraphy give constraints to the variation. Equation (3.53) represents a one-dimensional model of isostasy, where density variations only along the z direction are taken into account. This idea is sometimes referred to as local isostasy compared to regional isostasy that is introduced in Section 8.1. In the latter model, horizontal density variations are explicitly included in its equation. In the rest of this chapter we will use a simple isostatic model to calculate topography and vertical tectonic movements.

3.8

Balance between ocean and continent

There are two tectonic classes of regions, continents and oceans. This division is reflected in the frequency distribution of the height or water depth of the surface of the solid Earth (Fig. 3.10).

3.8. BALANCE BETWEEN OCEAN AND CONTINENT

77

Figure 3.10: Cumulative frequency distribution of the altitude of the surface of the solid Earth. There are apparently two modes at about 1 km above and 3–4 km below sea level. They correspond to the average altitude of the continents at 0.84 km and the average depth of oceans at 3.4 km. Let us see that the two levels are in isostatic equilibrium [78]. Assuming g = 10 ms−1 and the depths and densities shown in Fig. 3.11(a) for representative continent and ocean, we have the overburden at the Moho depth under the continent pL = ρc gto = 0.98 GPa.

(3.54)

The overburden under the ocean basin at the same level is pL = ρw gb + ρo gto + ρm g (tc − h − b − to ) ≈ 1.006 GPa,

(3.55)

which is about the same as that under the continent (Eq. (3.54)). Therefore, continents and oceans are roughly in isostatic equilibrium. Seawater pushes the oceanic crust downward by its weight, and pushes the continents upward. What if the Earth loses oceans by extreme global warming? The average continental altitude from the average oceanic surface, x in Fig. 3.11(b), would have been decreased. Neglecting other factors including thermal expansion of rocks, the overburden at the level of continental Moho is pL = ρc gto + ρm g (tc − to − x) . Equating this pressure and that in Eq. (3.54), we obtain x = (tc − to ) (ρm − ρc )/ρm ≈ 3.3 km. The frequency distribution of global topographic height, hypsometry, is used to investigate the tectonics of terrestrial planets and moons. If impact cratering was the most significant process for shaping the surface of a planetary body, the resultant frequency distribution would be something like a normal distribution. In spite of the existence of numberless impact craters, the Moon has two modes in its distribution. They correspond to highlands and maria, suggesting that this distinction was formed through some global processes.

78

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.11: (a) Representative large-scale structure of the shallow part of the Earth. (b) A structure without seawater. The bimodal distribution for the case of the Earth reflects the density contrast of the continental crust to the oceanic crust and to the mantle. The density distribution with spherical symmetry is obviously the most stable structure in consideration of potential energy. However, the Earth has the dichotomy of continents and ocean basins. Plate tectonics is an important action to keep the dichotomy. Surface processes, including erosion and sedimentation, are also important for redistributing rocks and sediments. The strength of the lithosphere also affects the distribution. The continental lithosphere has mechanical strength so that plate convergence causes continental thickening with a limiting average altitude. As the average altitude exceeds it, increased gravitational force flattens the crust to lower the surface (§3.11).

3.9

Sediment load

A thick sedimentary pile pushes its basement downward by its weight. Let us consider the effect of sediment loading, using the thick sedimentary sequence resting on the Atlantic passive continental margin (Fig. 3.12) as an example. Sedimentary layers that lie beneath the continental shelf accumulated in sublittoral paleoenvironments9 . The total thickness of the sediments reaches several kilometers. The paleo water depth, which is called paleobathymetry, and the thickness indicate that the basement has subsided by the same amount as the thickness. Some old literature explained the subsidence by the increasing sediment load. Sediments have greater densities than water, and sedimentation replaces water by sediments. Therefore, sedimentation increases loading to the basement. Namely, sedimentation created the sedimentary basin. Now we are ready to deny this idea for the case of passive margin subsidence. Consider, for 9 Sublittoral zone lies immediately below the intertidal zone and extending to a depth of about 200 m or to the edge of the contiental shelf [2].

3.9. SEDIMENT LOAD

79

Figure 3.12: Crustal structure of the southern Baltimore Canyon Trough, Atlantic margin of North America. LJ, Lower Jurassic; MJ, Middle Jurassic; UJ, Upper Jurassic; LK, Lower Cretaceous; UK, Upper Cretaceous; T, Tertiary. Simplified from [101].

Figure 3.13: Sedimentary basin subsidence by sediment load. simplicity, a constant density for sediments ρs . If an ocean basin with an initial depth of b is filled up with sediments (Fig. 3.13), the final thickness of the sedimentary pile x is given by the isostasy equation ρw gb + ρc gtc + ρm g (x − b) = ρs gx + ρc gtc . The left- and right-hand sides of this equation correspond to the left and right columns, respectively, in Fig. 3.13. Therefore, we obtain   ρm − ρw x= b. (3.56) ρm − ρ s The density of sediment depends on the lithology. Using a representative value ρs = 2.3 × 103 kg m−3 , we have x = 2.3b. If excessive sediments are supplied from the hinterland, the sediments would be exported to the lower reaches. For the case of the thick shallow marine sequence on the passive continental margin, b ≤ 0.2 km results in x ≈ 0.5 km. Therefore, the basement subsided actively, not passively, because of the sediment load (§3.14). It is necessary to grasp subsidence history, and not only the initial water depth and final sediment

80

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

thickness, to understand the mechanism of subsidence. Quantitative stratigraphy provides a way to reconstuct the history.

3.10

Quantitative stratigraphy

The subsidence history of a sedimentary basin is an observable function of time. Quantitative stratigraphy or geohistory analysis is the technique used to reconstruct the history from formations, where isostatic balance plays an important role. The result of the analysis designates the vertical component of the tectonic movements. For this purpose, the necessary data are the present depth, lithology, porosity, sedimentation age, the paleobathymetry of formations, and the eustatic sea level curve for the period covering the ages of the formations. The paleobathymetry is estimated from sedimentary structures characteristic for a certain depth such as hammocky cross-stratification, which indicates the depths were so shallow that the sedimentation was affected by surface waves. Some kinds of molluscs and benthic foraminifers live on the ocean bottom at certain water depths so that benthic fossils are another indicator of paleobathymetry. Figure 3.14 shows the paleo topography inferred from those kinds of data around Japan. The Japanese island arcs experienced vertical movements on the order of kilometers in the Miocene [263, 272]. The basic idea of geohistory analysis is that the sum of the formation thickness and paleobathymetry gives the depth to the basement from the present sea level, and applies several corrections. There are three key factors for the accumulation of thick sedimentary pile. Firstly, tectonic subsidence is necessary to form a container of sediments. Secondly, sediment supply is important. If the supply is scarce, a sediment-starved basin is the result. Thirdly, eustasy affects not only paleobathymetry but also sedimentation. A eustatic rise is equivalent to tectonic subsidence for sedimentation. We can abstract vertical tectonic movements from a sedimentary basin by taking into account these and additional factors. The mathematical procedure employed for this purpose is comparable to the stripping of formations from the top to lowest stratigraphic horizons (Fig. 3.15), so that it is known as the backstripping technique. The density of rocks is important for isostasy. Sediments lose their porosity and get a larger density while they are buried. Pores between sedimentary particles are usually filled with formation fluids that are mostly water. For example, mudstone has a porosity of some 50% shortly after deposition, However, increasing compactness of the particles expels the fluids during burial, i.e., compaction occurs. The porosity φ is the volume fraction of a pore. The porosity of mudstone is less than 10% at depths of a few kilometers. The total volume of sedimentary particles in a unit volume is 1 − φ. It is known that porosity of sedimentary rocks decreases roughly exponentially with depth, φ = φ0 e−Cζ , (3.57) where ζ is depth from the ocean bottom. The parameters φ0 and C depend on lithology. Accordingly, it is assumed that the parameters are determined for rock types at depths, e.g., in boreholes.

3.10. QUANTITATIVE STRATIGRAPHY

81

Figure 3.14: Paleo topography around Japanese islands from 23 to 15 Ma inferred from sedimentary environments. Benthic fossils and depositional facies are the keys for estimating the environment. The Japan arc drifted to form the Japan Sea backarc basin in the Early Miocene. The paleo position of the arc before 15 Ma, when the opening was completed, is controversial so that the paleo environments are indicated on the present map of this region. If a formation has a horizonally infinite extension, the porosity reduction results in the thinning of the formation. Consider a formation with an initial thickness of δd0 decreasing to δdN by increasing burial depth from d0 to dN . A column of sedimentary rocks with a unit sectional area has the total & d +δd volume of sedimentary particles d00 0 (1 − φ) dζ between the depths d0 and d0 + δd0 . Pore fluid is expelled from this part of the column by burial, but the particles are not. The conservation of

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

82

Figure 3.15: Schematic illustration to explain the backstripping technique. The left column shows the present state. The central and right columns indicate the situation just after the ith and (i + 1) layers deposited, respectively. particles is indicated by the equation  d0 +δd0 d0

(1 − φ) dζ =

 dN +δdN

(1 − φ) dζ.

dN

Substituting Eq. (3.57), we have δd0 +

  φ0 Cd0  −Cδd0 φ0 CdN  −CδdN e e e e − 1 = δdN + −1 . C C

(3.58)

The parameters C and φ0 are known for specific rock types. Therefore, if the thickness of a formation δdN when the depth of the formation was dN is given, we can calculate the thickness δd0 at a given depth of d0 through Eq. (3.58). That is, this equation corrects the effect of compaction on the thickness of a formation. In addition, it is easy for us to grasp the present depth and thickness of a formation. The ancient depth of the formation is obtained as the total thicknesses of the overlying formations. Therefore, the right-hand side of this equation is known for the youngest formation that occupies the top of the sedimentary pile. Accordingly, Eq. (3.58) works as a recursive function to calculate ancient thicknesses which are determined successively from the top to the base of the layers. This procedure is called decompaction, which increases layer thickness. The next task is the correction of sediment loading. Let us assign a number for each layer from the top (Fig. 3.15). The mass of the ith layer includes those of pore water φi ρw and sedimentary grains (1 − φi )ρg , where φi is the porosity of the layer and ρg is the average density of the grains. If

3.11. TECTONIC FORCE CAUSED BY HORIZONTAL DENSITY VARIATIONS

83

δdi is the thickness of this layer just after it was deposited, we obtain δdi using the recursive equation Eq. (3.58). The mass of the layer per unit basal area is   (3.59) mi = φi ρw + (1 − φi ) ρg δdi . If the lowermost layer has the ordinal number n, the sedimentary column with a unit sectional area has the mass of Mi = mi + mi+1 + · · · + mn−1 + mn (3.60) when the ith layer was deposited. Consider that the Moho is displaced downward by the weight of the ocean water and the ith layer. If the displacement is δhi , we have the equation ρw (bi − ei ) + Mi = ρw (bi+1 − ei+1 ) + Mi+1 + ρm δhc

(3.61)

from isostasy10 , where bi is the water depth and ei is the level of the ocean surface relative to the present one when the ith layer was deposited. That is, the sequence {en , en−1 , . . . , e0 } represents the eustatic sea level changes through time. They are estimated from ancient marine transgression and regression observed around stable continents11 . Combining Eqs. (3.58)–(3.60), we obtain Mi . Therefore, δhc is calculated through Eq. (3.61). We are able to estimate the ancient depth of burial for every formation from a sedimentary record using this backstripping technique, and to quantify the history of tectonic subsidence. Quantitative stratigraphy has a precision of ∼100 m for a sedimentary sequence that was deposited near sea level. Not only subsidence, but also uplift can be evaluated with this method. A sedimentary pile records the uplift as a decrease of paleo water depth or regression. We are able to quantify these phenomena, provided that sedimentation lasted during the uplift. If vertical movement occurred above sea level, fission track thermochronology is useful to evaluate uplifts of the order of kilo meters. Recently, a geochemical signature of continental carbonates was used to estimate paleo altitudes [66]. Figure 3.16 shows the subsidence history of the continental shelf off the Grand Banks of Canada determined by the backstripping technique. The vertical movements of the basement is illustrated by a subsidence curve. The subsidence of passive continental margins is modeled as the surface manifestation of post-rift lithospheric cooling (§3.14). Kominz et al. [103] exhibited a eustatic sea level curve from the subsidence curves of passive margins combined with the cooling model.

3.11

Tectonic force caused by horizontal density variations

Assuming the lithostatic state of stress, horizontal forces in the lithosphere are discussed in this section. Under this condition, lithostatic pressure is proportional to depth z, and the constant of 10 We asssume that the crustal thickness except for the sedimentary pile does not change so that the thickness does not appear in Eq. (3.61). Local isostasy is assumed here, but a technique called flexural backstripping has been developed to account for the strength of the lithosphere [74]. 11 Epeirogeny (§9.5) is an important but serious problem for this purpose.

84

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.16: Subsidence history of the continental shelf off the Grand Banks, Canada [1]. The present sea level is the origin of the vertical axis. Ages on the right side of this graph depict the depositional age of selected stratigraphic horizons.

Figure 3.17: Diagram showing the lithostatic pressure versus depth for high and low density rocks.

proportionality depends linearly on density (Fig. 3.17). We have an interesting consequence from this simple model if there is a horizontal density variation. Continents and oceans are isostatically compensated. If a depth of compensation is placed at the continental Moho, the depth dependence of the lithostatic pressures under a continent and an ocean basin are illustrated in Fig. 3.18(a), which is drawn with the same structure as shown in Fig. 3.11(a). Namely, the pressure is greater under the continent than under the ocean around the sea level because of the topographic bulge of the continent and of the density difference of the continental crust and seawater. However, the lithostatic pressure under the ocean catches up with that under the continent at the compensation depth. Pressure difference generally drives deformation. Accordingly, the difference between the lithostatic pressures produces tectonic force, which horizontally extends continents and constricts ocean basins (Fig. 3.18(c)).

3.11. TECTONIC FORCE CAUSED BY HORIZONTAL DENSITY VARIATIONS

85

Figure 3.18: (a) Lithostatic pressure under a continent and an ocean basin that are isostatically compensated. It is assumed that there is no density variation in the crust and in the mantle. (b) Vertical profile of density contrast between the regions. (c) Tectonic flow driven by the pressure difference.

A continent pushes an ocean basin with a force that is designated by the gap between the line graphs of the lithostatic pressures under the regions. The gap depends on the profile of the density difference δρ, which is schematically shown in Fig. 3.18(b). The density profile exhibits deviations with the opposite signs at two levels. The dipole moment is the magnitude of positive and negative anomalies multiplied by the distance between the anomalies. Accordingly, the gray area in Fig. 3.18(a) increases with the increasing dipole moment of the density profile. As the continental crust is thickened, both the average altitude and the Moho depth increase. They expand the gap, and further strengthen the horizontal force (Fig. 3.18(a)). Consequently, a continent pushes oceans with a force of 1.6 × 1012 Nm−1 = 1.6 TNm−1 per unit length of the continent-ocean boundary, if the structure shown in Fig. 3.11(a) is assumed. This value is obtained by the downward integration of the gap between the pressures. The force is as great as other tectonic forces. That is, large-scale variation of topography can drive tectonic deformation (§12.2). The same argument also applies to the horizontal force accompanied by a large-scale variation in altitude. That is, highlands tend to extend horizontally and pushes the neighboring lowland. The extensional force by this effect builds up with an uplift of the highland. The result is that the altitude of an extensive highland cannot exceed a certain limit. Gravitational collapse occurs when

86

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.19: Viscous fluid has a level surface at rest. However, a moving belt installed on the bottom of a container produces a slope at the surface. the highland is uplifted above the limiting altitude. The Tibetan Plateau has been uplifted to an altitude of ∼5 km by N–S cotinental convergence, and the crustal thickness reaches ∼70 km. E–W trending thrust faults are dominant along the Himalayas, but there are N–S or NNE–SSW trending normal faults and strike-slip faults oblique to the normal faults in the plateau [12, 240]. The crustal thickness is attenuated rather than increased in the plateau, and the normal and strike-slip faulting transfers crustal blocks eastward, suggesting that the plateau has reached the limit12 . The limiting altitude depends on the strength of the continental lithosphere, which has temporal and spatial variations13 . To see this, consider viscous fluid like honey in a container that has a moving belt at its bottom. This thought experiment models a convergent orogen, where the fluid simulates the continental crust. The belt drags the fluid to form a slope at the surface of the fluid (Fig. 3.19). The slope is dynamically maintained by the competing factors, the shear traction and gravitational spreading caused by the slope. The driving force of the latter results from the horizontal density difference of the fluid and air across the slope. The inclination of the surface is enhanced by the shear stress caused by the belt, and the shear stress increases with increasing viscosity. The ratio of the gravitational and viscous forces is called the Argand number [56] and indicates the tendency of gravitational spreading of large-scale positive undulations. Several factors including lithology, thermal regime and pore pressure, control the effective viscosity of rocks (§10.8). Since these factors may have temporal and spatial variations, the limiting altitude has these variations, also. If the lithosphere is soft or softened, topographic bulges on the lithosphere collapse.

3.12

Thermal isostasy

In this section we consider ocean-floor subsidence (Fig. 3.20) by combining the isostasy and thermal evolution of the oceanic lithosphere. The solid Earth discharges energy mostly through the cooling oceanic plates. Thermal conduction is the primary mechanism of heat transfer in the lithosphere14 . As rocks have low thermal 12 The Himalayas have higher altitudes than the plateau. The peaks are supported not only by the buoyancy force associated with isostatic compensation but also by the flexural strength of the lithosphere. 13 See Chapter 12. 14 The Galilean Satelite, Io, is a remarkable extra terrestrial body. The moon is as the size of the Moon, but discharges with

3.12. THERMAL ISOSTASY

87

Figure 3.20: Depth–age relationship of ocean basins [228]. The vertical axis shows the moving average with a window of 5 m.y. of the depth that is corrected for the load and thickness of sedimentary cover. GDH1, PSM and HS are synthesized depth-age curves based on different models. PMS assumes the cooling oceanic plate with the constant thickness and constant basal temperature [174]. HS is the curve calculated by the cooling half-space with the thermal properties derived from PMS. GDH1 is the compromised model that assumes a cooling half-space for the lithosphere younger than 20 m.y. and the plate model for the older one [228]. conductivity, the conduction is associated with steeper temperature gradients in the lithosphere than in the asthenosphere where heat is transferred by convection. That is, the lithosphere is the thermal boundary layer of the convecting mantle. Rocks lose their ductile strength with increasing temperature. Therefore, the lithosphere is the cold and hard lid of the convecting cells. The thickness of the thermal boundary layer changes with cooling and heating. The density of rocks depends on temperature so that a change in the temperature field affects isostatic balance. This is known as thermal isostasy. Therefore, the thermal regime of the lithosphere affects the vertical movement of ocean basins. The cooling lithosphere causes subsidence. To understand this phenomenon, we firstly consider the heat conduction equation. Conduction of heat A large rock mass has a large thermal capacity so that temperature change takes a long time in the mass. Consider a mass with the diameter L at a standstill, The temperature change is described by the same great energy as the Earth and most of the energy is emitted from its hotspot volcanoes [135].

88

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.21: Diagram showing the model of cooling oceanic lithosphere. the heat conduction equation ∂T (3.62) = κ∇2 T, ∂t where T is the temperature that is a function of time t and position, and κ is the thermal diffusivity of the rock mass. If heat is conducted only in the x direction, this equation reduces to ∂T ∂2 T =κ 2. ∂t ∂x

(3.63)

The representative value for the thermal diffusivity of rocks is 1 m2 s−1 . This unit designates that κ/L2 has the dimension of time. Accordingly, we define the dimensionless length x¯ and time t¯ by dividing L and L2 /κ and the dimensionless temperature T by dividing by a reference temperature T0 . Namely, T = T/T0 , x¯ = x/L, t¯ = κt/L2 , (3.64) 2 ¯ dx = Ldx, ¯ dt = L dt/κ, dT = T0 dT . Using these dx, dt dT , we rewrite Eq. (3.62) in the form ∂2 T ∂T = ∂t¯ ∂x¯ 2

(t¯ > 0, 0 < x¯ < 1) .

This equation does not include material constants such as κ, so that its solution depends only on the initial and boundary conditions. The solution can be applied to specific problems using Eq. (3.64). The time representing the cooling speed of the rock mass is evaluated by the ratio, τ = L2 /κ. This is the time constant for thermal conduction. This equation demonstrates that the cooling of a large body takes a long time. Subsidence of the oceanic basins Consider an oceanic lithosphere that is moving with a constant velocity of v away from the oceanic ridge. We define the Eulerian coordinates O-xz with the origin at the spreading center and the Laglangian coordinate ξ that travels with the oceanic plate (Fig. 3.21). The origin of ξ-coordinate

3.12. THERMAL ISOSTASY

89

is placed at z = 0, i.e., z = ξ. For simplicity, we assume that heat is transferred only by conduction. To calculate temperature change, the thermal conduction equation, ∂2 T ∂2 T = κ ∂t2 ∂ξ 2 is solved with the initial and boundary conditions T = Ta T = T0 T → Ta

at at as

ξ ≥ 0, t = 0, ξ = 0, t > 0, ξ → ∞, t > 0.

That is, we assume a cooling half-space with the constant temperatures T0 and Ta at the ocean floor and at the deep mantle, respectively. As the lithosphere ages, the ocean floor subsides and departs from z = 0 by a few kilometers. Therefore, ξ = 0 is not the level of the floor, but this departure is neglected in the above conditions because it is negligible compared to the dimension of the mantle. After all, we have the solution of this problem,   ξ T − T0 = Erf √ , (3.65) Ta − T0 2 κt where Erf( ) is the error function (Fig. 3.22) defined by the equation  2 x −η2 Erf(x) = √ e dη (x ≥ 0). π 0

(3.66)

The graph of the error function designates that the shallow and deep parts of the suboceanic mantle have steep and gentle geothermal gradients (Fig. 3.22). The turning point gradually goes down √ in the mantle with the denominator, κt, in the operand of the error function in Eq. (3.65). The √ denominator has the dimensions of length so that κt is called the diffusion length. Obviously, an √ isotherm descends with age as ξ ∝ κt. The error function has an inverse function. Therefore, the depth of the isotherm with a specific temperature T is given by Eq. (3.65), where the left-hand side is known. Substituting t = x/v and ξ = x into the equation, we have  (  T − T0 z v = Erf . (3.67) Ta − T0 2 κx The thermal evolution has been investigated in the above argument. Now we can calculate the subsidence. It is know that the density of rock at the temperature ΔT , which is measured from the reference temperature T0 , changes as ρ = ρ0 (1 − αΔT ),

(3.68)

where ρ0 is the density at the reference temperature and α is the thermal expansion coefficient. Rocks have a representative value for the coefficient at about 3 × 10−5 K−1 . We use ρa = ρm0 (1 − αTa ) for

90

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

Figure 3.22: Graphs of error function Erf(x) and complementary error function Erfc(x). the density of the asthenosphere. If b is the depth of the sea floor from the level ξ = 0, the isostasy equation gives  dc ρw b + ρ dξ = const. (3.69) 0

The value of this equation is evaluated at the spreading center, where the asthenosphere comes immediately below the ridge axis. Therefore, we obtain the value of the above equation there, ρa dc + ρw bR , where bR is the water depth of the axis. Accordingly, Eq. (3.69) is rewritten as  dc ρw b + ρ dξ = ρa dc + ρw bR . (3.70) 0

We assume the half-space so that the compensation depth can be treated as dc = ∞. Combining Eqs. (3.70), (3.67) and (3.68), we obtain  ( ! ∞ v z dz 1 − Erf b(ρa − ρw ) = ρm0 α(Ta − T0 ) 2 κx 0  (  ∞ v z Erfc = ρm0 α(Ta − T0 ) dz 2 κx 0 Erfc(x) = 1 − Erf(x) is the complementary error function (Fig. 3.22). This function satisfies ∞ 1 Erfc(x) dx = √ π, 0 therefore, the depth of the sea floor at a distance of x from the spreading axis is given by the equation ( 2ρa α(Ta − T0 ) κx b= (3.71) + bR . ρa − ρ w πv

3.13. HORIZONAL STRESS RESULTING FROM TOPOGRAPHY

91

The depth increases with the square root of the distance, as we have assumed a constant spreading rate at the ridge center. We are able to estimate the effective values for the thermal diffusivity and thermal expansion coefficient of the oceanic mantle by comparing the above model with observations. However, it is important that old sea floors are shallower than the depth predicted by the model (Fig. 3.20) and the ocean basins have different deviations. Some factors other than the cooling are responsible for this flattening. It is pointed out that reheating by hotspots cannot account for the deviations [198].

3.13

Horizonal stress resulting from topography

Large-scale topographic undulation is one of the sources of tectonic stress. Ocean floors show the difference in their depth up to a few kilometers (Fig. 3.20) so that great stress may be generated. Following Dahlen [47], let us consider how large is the stress level associated with the variation of the depth of the ocean floor. In conclusion, the ridge push is the force resulting from the topography. The force is important among other origins of tectonic stresses including the colliding force of plates. The asthenosphere comes immediately below the spreading axis, therefore we assume a constant density ρ0 under the axis. The coordinates O-xz shown in Fig. 3.21 are used in this section. Consider the density distribution ρ(x, z) = ρ0 + ρ$(x, z), where the last term is the density anomaly from the reference density ρ0 . According to Eq. (3.65), the temperature and the resultant density anomalies were derived in the previous section as   z , T (z, t) = Ta Erf √ 2 κt   z ∴ ρ(z, ˜ t) = ρm 0 αTa Erfc . (3.72) √ 2 κt The force FR that a unit length of the ridge axis is determined by the dipole moment of vertical density distribution (§3.11). The moment is obtained from Eq. (3.72) so that ∞   ρ$gz dz = ρm 0 gαTa κ t. FR = 0

Consequently, the force is proportional to the age of the oceanic lithosphere15 . The constant of proportionality is about 0.035 TNm−1 /m.y. For example, the ocean floor with an age of 80 m.y. has FR = 2.8 TNm−1 . The force resulting from the ocean-continent topographic contrast has a magnitude of about 1.6 TNm−1 (§3.11). Therefore, the ridge push is as great as the force by the contrast. 15 The integral is taken over the range [0, ∞]. However, the contribution from the asthenosphere is negligible because of Erfc(x)  1 for x > 2 (Fig. 3.22).

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

92

In the above argument, we used the density profile under the spreading axis as the reference. It is the next problem where the proper area that provides the reference is sought. If the density distribution was under the summit of Mount Everest, the entire world would have been under a compressional stress field. However, this is not the case. Accordingly, Coblentz and others excluded the arbitrariness in the choice of reference by taking global topography into account [42]. Let the levels of the top and base of the lithosphere be z = h and z = L, respectively. Note that the vertial force per unit area is σzz × (1 m2 ) and that the product of force and length has the dimensions of energy. Then the quantity  L h

L σzz dz =

U= h

ρ(ξ)g dξdz h

ξ

is the potential energy of the lithospheric column with the unit sectional area. A region with a large U tends to be subject to extensional tectonics. That with a small U is ready to be horizontally contracted. The global average is estimated at 2.4 TNm−1 [42]. The threshold altitude for the continent is ±200 m, and the threshold depth of the ocean is about 4.3 km. Spreading centers have depths actually shallower than this threshold. However, the topography is not the only cause of lithospheric stress. Therefore, an extensional stress field rifts continents and eventually breaks up the continental lithosphere.

3.14

Thermal subsidence of continental rifts

It was shown in the previous chapter that the subsidence history of a passive continental margin is revealed from stratigraphic records. Here we consider Mckenzie’s simple stretching model to link the subsidence and rifting together. Post-rift subsidence Passive margin subsidence after continental breakup is generally associated with very weak or no fault activity. Extensional tectonics does not account for the subsidence. In contrast, the continental lithosphere is thinned and horizontally extended during the rift phase. The fact is that the rifting accounts for the post-rift subsidence. Consider the thermally equilibrium state of the lithosphere illustrated in Figs. 3.23(a) and (c), i.e., the temperature at the base of the lithosphere is Ta at a depth of z = tL . The temperature distribution before and long after rifting is assumed to be in equilibrium. The continental crust is thinned with the lithospheric mantle. The thinning of the crust results in subsidence. Let tc and tL be the initial thickness of the crust and the lithosphere, respectively (Fig. 3.23(a)). It is also assumed that the temperature at z = tL is kept at Ta . The top of the lithosphere was at sea level before rifting. Here, the lithosphere is thought to be subject to homogeneous deformation in the rift phase. If the lithosphere is horizontally extended by a factor of β, the thickness of the lithosphere and crust become tc /β and tL /β, respectively. The rift-stage subsidence is affected not only by the thinning

3.14. THERMAL SUBSIDENCE OF CONTINENTAL RIFTS

93

Figure 3.23: Diagrams showing the simple stretching model [138]. (a) Structure and temperature distribution in the lithosphere in the pre-rift stage, (b) at the end of rifting, and (c) in post-rift stage. but also by the temporal variation of the thermal regime, because rock density changes with the variation as ρ = ρ0 (1 − αT ),

(3.73)

where ρ0 is the reference density at T = 0◦ C. If the thinning of the lithosphere is quick compared to the time constant of heat conduction, isotherms in the lithosphere are uplifted adiabatically16 (Fig. 3.23(b)). If we assume isostatic compensation throughout the rifting process, we obtain the syn-rift subsidence   1 tL ρa − tc ρc − (tL − tc )ρm 1− , (3.74) Si = ρa − ρ w β where ρc , ρm and ρa are the representative densities of the crust, mantle lithosphere, and asthenosphere, respectively, and are given by ρa = ρm0 (1 − αTa ), 16 This

  αTa tc ρc = ρc0 1 − , 2 tL

  αTa αTa tc ρm = ρm0 1 − . − 2 2 tL

assumption is valid for the rifting with a duration shorter than ∼25 m.y. [92].

94

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY Table 3.2: Parameters for the simple stretching model. ρm0 = 3.3 × 103 kg m−3 ρc0 = 2.8 × 103 kg m−3 ρw = 1.0 × 103 kg m−3

α = 3.5×10−5 K−1 κ = 1.0×10−6 m2 s−1 Ta = 1333◦ C

tL = 100 km tc = 35 km

The coefficients, ρc0 and ρm0 are the reference densities of the crust and mantle. Si is called the initial subsidence in the passive margin formation. Substituting the representative densities into Eq. (3.74), we obtain the magnitude of the initial subsidence   1 , (3.75) Si = Si0 1 − β     αTa αTa tc tc − ρm0 tL (ρm0 − ρc0 ) 1 − 2 tL tL 2 Si0 = . (3.76) ρm0 (1 − αTa ) − ρw The latter quantity, (Si0 ) can have positive and negative signs depending on the initial thickness of the crust. Since the denominator in the right-hand side of Eq. (3.76) is positive, the sign depends on the numerator. Solving the equation, (numerator) = 0, we obtain the solution, 4(ρm0 − ρc0 )2 + 4α 2 (ρm0 − ρc0 )ρm0 Ta2 ρm0 − ρc0 tc = ± . tL α(ρm0 − ρc0 )Ta α(ρm0 − ρc0 )Ta Using the parameters shown in Table 3.2, we obtain the ratio tc /tL = 0.154 and 42.7. However, this ratio should be less than unity so that the latter is a meaningless solution. The former indicates a threshold initial thickness of tc ≈ 15 km. The lithosphere with a continental crust initially thicker than ∼15 km subsides through the lithospheric thinning, but the surface is uplifted by rifting if the crust is initially thinner than this value. The uplift is due to a steepening of the geothermal gradient. That is, the attenuation of the lithospheric mantle plays an opposite role to that of the continental crust. Island arcs have volcanism so that their lithosphere is thinner than the ordinary continental lithosphere. If an island arc has a lithosphere with a pre-rift thickness of tL = 50 km, then the arc has the threshold at about 7 km. The geohistory analysis of syn-rift sediments can constrain β with Eq. (3.75). This parameter is called the stretching factor. The thermal regime is altered by the rifting. Isotherms between 0◦ and Ta are raised by the rifting (Fig. 3.23(b)). After the rifting, the isostherms descend to the equilibrium level in the long run (Fig. 3.23(c)). This post-rift cooling causes a gradual subsidence of the passive margin. This is called the thermal subsidence of the lithosphere. The simple stretching model (Fig. 3.23) allows us to estimate the rate and amount of this subsidence [138]. In order to calculate the heat conduction after the termination of rifting, the initial subsidence is neglected because Si  tL . Therefore, the thermal conduction equation ∂T/∂t = κ∂2 T/∂z2 is solved with the boundary condition T = 0 at z = 0. In addition, the temperature at z = tL is assumed to be

3.15. KINEMATIC MODEL OF SYN-RIFT SUBSIDENCE

95

kept at T = Ta . The time t is measured from the termination. The initial condition is shown by the graph in Fig. 3.23(b), which is the temperature distribution at the end of rifting. Solving this system, we obtain the thermal evolution   ∞ z T 2  (−1)n+1 β nπ nπz −n2 t/τ = 1− e , (3.77) + sin sin Ta tL π n nπ β tL n=1

τ=

π 2 t2L /κ.

(3.78)

The parameters shown in Table 3.2 give the time constant τ = 3.2 m.y. Combining Eqs. (3.73) and (3.77), we obtain the evolution of the density structure ρ(z, t) . Isostatic compensation through time is expressed as  tL  tL ρ(z, t) dz + ρw St = ρ(z, 0) dz, 0

0

where St is the thermal subsidence,    4αρm0 Ta tL β π St ≈ sin 1 − e−t/τ . 2 β π (ρm0 − ρc0 ) π

(3.79)

Higher-order terms in Eq. (3.77) are neglected in deriving Eq. (3.79). Equation (3.79) shows that the rate of thermal subsidence decays as the exponential function. The subsidence curve shown in Fig. 3.16 has this shape. Fitting the curve given by Eq. (3.79), we obtain the stretching factor for the area the curve was determined.

3.15

Kinematic model of syn-rift subsidence

Post-rift subsidence results from the cooling of the lithosphere, whereas syn-rift subsidence is controlled by not only the thermal but also the mechanical properties of the lithosphere. The latter affects the characteristics of rift zones such as their widths and depths. Decompression melting of the asthenosphere under rift zones causes uplift through the buoyancy of magma. Therefore, the observation of syn-rift subsidence sheds light on the deep seated factors and on the secular changes of these factors. The simple stretching model (§3.14) assumed an instantaneous thinning of the lithosphere and neglected the details of the rifting processes. In this section, we consider a mathematical inverse method to constrain the rift-phase strain rate of the lithosphere from an observed rift-phase subsidence curve (Fig. 3.24) [259]. Consider a pure shear rift with horizontal and vertical strain rates of E˙ xx and E˙ zz , respectively. The lithosphere is extending horizontally so that we have E˙ zz < 0 < E˙ xx . The pre-rift levels of the top and base of the lithosphere is z = 0 and z = a. The lithosphere is subject to a pure shear, so that we have −ε˙zz = ε˙xx = G(t),

96

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

where the function G(t) is defined to describe the temporal variation of the strain rates and is linked to the rate of subsidence in the following discussion. Consider that the horizontal velocity field is symmetric with respect to the plane x = 0. The lithosphere is assumed to undergo homogeneous strain so that the horizontal velocity u(x, t) is proportional to x. That is, we have u(x, t) = G(t) x. (3.80) Syn-rift subsidence reaches several kilometers, but is negligible compared to the representative thickness of continental lithosphere. Therefore, we assume the top of the lithosphere is kept at z = 0 during the calculation of the lithospheric strain. This assumption allows us to write the descending velocity of rocks at depth z as v(x, t) = −G(t) z. (3.81) Note that this is the velocity in the +z direction. Subsurface temperature distribution affects the subsidence of this rift. The thermal conduction equation to be solved should include velocity, because heat is transported not only by conduction but also by travelling rocks. Hence, the left-hand side of the thermal conduction equation, ∂T/∂t = κ∇2 T , should be replaced by the material derivative (Eq. (2.20)). That is, the equation      2 ∂T ∂2 T ∂ T u (3.82) + + ∇· T =κ v ∂t ∂x2 ∂z2 governs the thermal evolution. From Eqs. (3.81) and (3.80), we have    ∂T ∂T u ∇· T = G(t)x − G(t)z . v ∂x ∂z

(3.83)

If there was no horizontal variation in the initial temperature distribution, the pure shear would not raise horizontal temperature variation. Therefore, we have ∂T ∂2 T = 0. = ∂x ∂x2 Combining this and Eqs. (3.82) and (3.83), we obtain ∂T ∂2 T ∂T + Gz =κ 2 . ∂t ∂z ∂z

(3.84)

The boundary conditions that we assume for the temperature evolution are T = 0 at z = 0 and T = Ta at z = a. The initial condition and the equilibrium temperature distribution is the constant geothermal gradient to a depth of z = a. The subsidence of this rift is the sum of the subsidence due to the thinning of the lithosphere and to the thermal perturbation. The former is described by Eq. (3.75). The equation describing the latter factor should have the form  a  (3.85) Q(t) = B T (z, t) − T (z, ∞) dz, 0

3.15. KINEMATIC MODEL OF SYN-RIFT SUBSIDENCE

97

Figure 3.24: (a) Strain rate history of a Triassic rift in northern Italy determined from a subsidence curve of the rift (b). The 95% confidence region is shown in (a) by white [259]. where B = αρm /(ρa − ρw ) is the factor to account for the buoyancy of the crust. Therefore, we have the total subsidence   1 − BQ(t), (3.86) S(t) = A 1 − β where A = tc (ρm − ρc )/(ρa − ρw ). Crustal thinning has a positive effect for the total subsidence, but the thinning of the lithosphere increases the geothermal gradient which has a negative effect for the subsidence. Rearranging Eq. (3.86), we have β=

A . A − S − BQ(t)

Equation (2.39) links the stretching factor β and G(t) as  tR  β = exp G(t) dt ,

(3.87)

(3.88)

0

where tR is the time ellapsed since the rifting began. Accordingly, combining Eqs. (3.87) and (3.88), we obtain the strain-rate history of the rift )  * A d ln . (3.89) G(t) = dt A − S(t) − BQ(t) The rift-phase strain-rate history can be evaluated from observations for a pure shear rift. The right-hand side of Eq. (3.89) has S(t) that is determined from the quantitative stratigraphy of the rift. However, Q(t) depends on the unknown G(t) through Eq. (3.84). Therefore, G(t) is determined by a mathematical inversion17 from the observed subsidence history S(t). Figure 3.24(a) shows the 17 See

[259] for detail.

CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY

98

optimal strain-rate history for the subsidence history of a Triassic rift zone in northern Italy (Fig. 3.24(b)). The subsidence curve of the Triassic rift has an inflection point at about 220 Ma (Fig. 3.24(b)). After the point, the subsidence curve is similar to an exponential curve, suggesting thermal subsidence. The point designates a rejuvenated rifting, and corresponding to the second peak in the strain-rate history. A sharp decline in the strain rate by several orders of magnitude indicates the termination of rifting; the rift was aborted or the area became a passive margin. Timing of the termination determined by the inverse method for the Triassic rift has a large uncertainty which is illustrated by the 90% confidence region between 100 and 220 Ma in Fig. 3.24(a). The umbiguity stems from the difficulty to detect very slow rifting if it overlaps simultaneous thermal subsidence. More definite constrains come from geological structures. The termination of rifting is marked by the cessation or distinct decline of fault activity. For example, the cross-section in Fig. 3.12 shows that the subsidence of wedge shaped grabens had ceased by the early Jurassic. If rifting was successful to form a passive continental margin, the magnetic anomaly near the continent-ocean boundary provides a constraint for the timing. Some passive margins were remarkably uplifted when continents were broken off the margins. Such a unconformity is known as a breakup unconformity [59].

3.16

Exercises

3.1 Assume a state of stress with the principal stresses σ1 = 4 MPa, σ2 = 5 MPa and σ3 = 8 MPa, and vertical σ1 -axis and horizontal σ3 -axis oriented due east. Determine the magnitude of the normal and shear stresses acting on the block beneath a fracture plane of which trend and dip are N030◦ E and 60◦ SE, respectively. How much is the rake of the maximum shear stress on the plane? 3.2 Explain qualitatively why the function g(r) shown in Fig. 3.7(b) is convex downward and why the function decreases with increasing r in the lower mantle. 3.3 What is the magnitude of the lithostatic pressure at the Moho interface. Assume the thickness and density of the crust at 35 km and 2.8 × 103 kg m−3 , respectively. 3.4

Derive Eq. (3.46).

3.5 Estimate how much is the force per unit boundary length with which the Tibetan Plateau pushes the Hindustan Plain due solely to the structure illustrated in Fig. 3.25.

3.16. EXERCISES

Figure 3.25: Lithostatic pressure under the neighboring plateau and lowland.

99

Suggest Documents