Вы находитесь на странице: 1из 41

Chapter 3

Stress, Balance Equations, and


Isostasy
We dene 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 Denition of stress
Suppose that a square pole is composed of ve 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 S 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 F be the traction force acting on the
element. The force may increase with the area S so that we dene 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 dened
by
t(N) = lim
S0
F
S
. (3.1)
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 = 10
5
Pa = 1000 hPa = 0.1 MPa. Engineering literature often uses the
units kg
f
/cm
2
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 ve rectangular solids, A
through E. (b) A rock mass is pulled from the adjacent mass with the force F on an element of the
surface S. 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 S 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) = S
31
e
(1)
+ S
32
e
(2)
+ S
33
e
(3)
, (3.2)
where S
31
, S
32
and S
33
are the components of the stress vector t(N), that is, t(N) = (S
31
, S
32
, S
33
)
T
.
The stress vector depends on the direction of the surface on which it acts. The rst subscript 3 in-
dicates 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
S
ij
= t
j
_
e
(i)
_
. (3.3)
We dene the stress tensor from the components as
S =
_
_
S
11
S
12
S
13
S
21
S
22
S
23
S
31
S
32
S
33
_
_
components of t
_
e
(1)
_
components of t
_
e
(2)
_
components of t
_
e
(3)
_
.
(3.4)
The limit in Eq. (3.1) indicates that stress vector is dened at each point on an innitesimally 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
dened 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.
dened, 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 over-
burden pressure. It is the sign convention in solid-earth science to dene the sign of stress as being
positive for compression, opposite to the sign convention of continuum mechanics (Table 3.1). Ac-
cordingly, 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
62 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
Table 3.1: Symbols and signs used in this book.
Stress Unit Normal Vector Strain
Continuum Mechanics S (tension = +) N (outward) E (elongation = +)
Solid-Earth Science r (compression = +) n (inward) 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-rz are sometimes con-
venient (Fig. 3.4). Transformation between the cylindrical and rectangular Cartesian coordinates
O-xyz is given by the equations,
x = r cos , y = r sin , z = z
r
2
= x
2
+ y
2
, = tan
1
(y/x).
Stress components in the coordinate systems are transformed by the two sets of equations

rr
=
xx
cos
2
+
yy
sin
2
+ 2
xy
sin cos , (3.6)

=
xx
sin
2
+
yy
cos
2
2
xy
sin cos , (3.7)

r
= (
yy

xx
) sin cos +
xy
(cos
2
sin
2
). (3.8)
3.2 Cauchys stress formula
It is seen from Eq. (3.3) that t
i
_
e
(j)
_
= S
ji
=

k
S
jk
e
(i)
k
. Therefore, we have
t
_
e
(i)
_
= S
T
e
(i)
. (3.9)
3.2. CAUCHYS STRESS FORMULA 63
Figure 3.4: Stress components in the cylindrical coordinates O-r.
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
t(N) = S
T
N, (3.10)
where N may be oblique to the coordinate axes
2
. This is called Cauchys stress formula. For the
sign convention of solid-earth science we have the the equivalent formula
t(n) = r
T
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:
N
= t(n) n, (3.13)
Shear stress:
S
=
_
t(n)
2

2
N
. (3.14)
Since the vector n points inward at the surface of a rock mass (Fig. 3.3),
N
is positive if the mass is
pushed inward at the surface. By contrast, the normal stress dened by the equation
S
N
= t(N) N (3.15)
is positive for tension. The shear stress for this sign convention is
S
S
=
_
t(N)
2
S
N
2 .
2
Equation (3.10) is demonstrated via the force balance equation [134], which is introduced in 3.3. However, axiomatic
continuum mechanics denes 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 nd more information in [243].
64 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
The vector
N
(n) =
N
n has a magnitude obviously equal to the normal stress
N
and is parallel
to n. If
N
is positive,
N
(n) points inward at the surface of the rock in question. The vector
N
(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:

N
(n) = P
[
(n) t(n) =
_
nn
_
r n =
_
n (r n)
_
n. (3.16)
On the other hand,
S
(n) is dened 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:

S
(n) = P

(n) t(n) =
_
I nn
_
t(n) = r n
_
n (r n)
_
n = t(n) t
N
(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 dened by
r = p I, (3.18)
where I and p are the unit tensor and a scalar quantity. Coordinate rotation by Qtransforms the stress
as r

= Q r Q
T
= Q (p I) Q
T
= p Q Q
T
= p I = r, indicating that the stress dened 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 gure 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 X.
3.3. FORCE BALANCE 65
3.3.1 Equation of motion
Consider a rock mass dened 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
0 =
_
S
t(N) dS +
_
V
XdV. (3.20)
Since
_
V
v dV is the total linear momentum of the rock mass, the material derivative
D
Dt
_
V
v dV
represents the inertia force. Therefore, the equation of motion of the body is
D
Dt
_
V
v dV =
_
S
t(N) dS +
_
V
XdV. (3.21)
The surface integral in the left-hand side of Eq. (3.21) is transformed into a volume integral with
Gausss divergence theorem (Eq. (C.63)). In addition, combining Eq. (3.10), we obtain
D
Dt
_
V
v dV =
_
V
S
T
dS +
_
V
XdV. (3.22)
In this case the order of D/Dt and integral is exchangable
3
so that we have
_
V
_
v S
T
X
_
dV = 0.
Since the choice of the volume is arbitrary, the integrand must vanish to satisfy this equation. Con-
sequently, we obtain the dierential form of
v = S
T
+ X or v = r
T
+ X. (3.23)
If the forces are balanced, we have
0 = S
T
+ X or 0 = r
T
+ X. (3.24)
The indicial notation of Eq. (3.24) is
0 =

j
S
ji
x
j
+ X
i
or 0 =

ji
x
j
+ X
i
. (3.25)
3
If V
0
and V are the initial and instantaneous volumes that the rock mass occupies, respectively, at time t = 0 and t,
the exchangeability is demonstrated as follows. The initial volume V
0
does not depend on time, so that the operator D/Dt
can be included in the integral over V
0
. In addition, because of V = JV
0
where J is the Jacobian, the initial and temporal
densities,
0
and , are related to each other as =
0
/J. The initial one does not depend on time, either. That is,
D
0
/Dt = D(J)/Dt = 0. Therefore,
D
Dt
_
V
v dV =
D
Dt
_
V
0
vJ dV
0
=
_
V
0
_
v
D(J)
Dt
+ J
Dv
Dt
_
dV
0
=
_
V
0

Dv
Dt
J dV
0
=
_
V

Dv
Dt
dV.
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 at and there is no horizontal density variation. Rocks are subject to the
gravitational force shown in Eq. (3.19). Dening z-axis vertically downward from sea level, the
force balance in this direction is obtained from Eq. (3.24) as

xz
x
+

yz
y
+

zz
z
= (z)g,
where density, (z), is the function of depth, z. Since there is no horizontal variation at all, the rst
and second terms in the above equation vanish. We therefore have

zz
z
= (z)g.
Integrating both sides, we obtain

zz
=
_
z
h
()g d, (3.26)
where h is altitude. This equation enables us to calculate the vertical stress
zz
from the density
prole (z). If g is constant over the depth range in question, Eq. (3.26) becomes
zz
= g
_
z
h
() d.
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,
zz
is the stress at the base of
the column due to its own weight
4
. 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
p
L
= gz. (3.27)
We have found how the vertical stress
zz
is determined. What are the other stress components?
Areas of compressional and extensional tectonics may have dierent 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 dene a reference stress state for later discussions. Al-
though choice of the reference is arbitrary, the isotropic stress
r = p
L
I (3.28)
is often used for the reference in the models of tectonics, where
p
L
=
_
z
h
()g d (3.29)
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 p
L
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 xF sin , where 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 XdV , 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 =
_
S
x t(N) dS +
_
V
x XdV. (3.30)
If they are imbalanced, the residual moment causes the acceleration in the angular velocity of the
body:
D
Dt
_
V
(x v) dV =
_
V
(x X) dV +
_
S
(x t) dS. (3.31)
The choice of position of the origin is arbitrary. Hhowever, Eq. (3.31) holds wherever the origin is
dened
5
.
4
In-situ stress measurements have shown that this picture gives a close approximation to observed
zz
[3]. However,
deviations from predicted
zz
from the density prole 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 dened.
That is, x = x

+ o. By this additive decomposition of x, Eq. (3.31) becomes


Left-hand side =
_
V
(x

+ o) v dV =
__
V
x

v dV
_
+
_
V
o v dV,
Right-hand side =
_
V
(x

+ o) X dV +
_
S
(x

+ o) t dS
=
__
V
x

X dV +
_
S
x

t dS
_
+
_
V
o X dV +
_
S
o t dS.
Combining the terms in and outside the square brackets in the above equations, we have the equations
_
V
x

v dV =
_
V
x

X dV +
_
S
x

t dS,
which is identical to Eq. (3.31) provided that x is replaced by x

and
o
__
V
v dV
_
= o
__
V
X dV +
_
S
t dS
_
. ()
68 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
It is important for discussions in the rest of this book that the stress tensor is symmetric:
S = S
T
, r = r
T
. (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 body
6
. Equation
(3.32) is derived from the conservation of angular momentum
7
.
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
S =
_
_
S
1
0 0
0 S
2
0
0 0 0
_
_
,
where S
1
and S
2
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) = S
T
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. Multi-
plying both sides of Eq. (3.23) by v, we have
_
V
v a dV =
_
V
v
_
S
T
_
dV +
_
V
v X dV, (3.33)
where V stands for the volume that the rock mass in question occupies. The left-hand side of this
equation is rewritten as
_
V
v a dV =
_
V
v v dV =
D
Dt
_
V

v v
2
dV =
D
Dt
_
V
v
2
2
dV =

K.
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(S
T
) in the rst term in the right-hand side of Eq. (3.33) has the components
v
i
S
ji
x
j
=
(v
i
S
ji
)
x
j

v
i
x
j
S
ji
=
(v
i
S
ji
)
x
j
(D
ij
+
ij
)S
ji
,
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 L
ij
and
ij
are stretching and spin tensors, respectively. Therefore, Eq. (3.33) is rewritten as

K +
_
V
D : S dV =
_
V
(v S) dV
_
V
W : S dV +
_
V
v X dV. (3.34)
Since W = (
ij
) is antisymmetric we have W : S = 0. Combining this and Eq. (3.34), we obtain

K +
_
V
D : S dV =
_
V
(v S) dV +
_
V
v X dV. (3.35)
Converting the rst term on the right-hand side of this equation into a surface integral by Gausss
divergence theorem, we obtain the energy conservation equation

K +
_
V
D : S dV =
_
S
[v t(N)] dS +
_
V
v X dV. (3.36)
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 e, and
_
V
e dV =
_
V
D : S dV. (3.37)
Therefore, the energy conservation equation is rewritten as
D
Dt
_
V

_
v
2
2
+ e
_
dV =
_
S
v t(N) dS +
_
V
v X dV. (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 ux is dened by the amount of heat that passes through a unit area in a unit of time. The
heat ux 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 Nq. Substituting this into the right-hand
side of Eq. (3.38), we have
D
Dt
_
V

_
v
2
2
+ e
_
dV =
_
S
v t(N) dS +
_
V
v X dV
_
S
N q dS. (3.39)
70 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
Using Gausss 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 satised only if
D
Dt
_
v
2
2
+ e
_
=
_
S v q
_
+ X v. (3.40)
This is called the energy equation. Combining Eqs. (3.40) and (3.33), we have
e = q + (S v) v ( S).
Components of the rst and second terms in the right-hand side are

i,j
_
(S
ij
v
j
)
x
i
v
i
S
ij
x
j
_
=

i,j
S
ij
v
i
x
j
= S : (v).
Therefore, we nally obtain the equation
e = S : (v) q. (3.41)
3.6 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 dierentiation.
The dierentiation 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 dierentiation (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
zz
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)
r
2
, (3.42)
where G = 6.67 10
11
Nm
2
kg
2
is the gravitational constant. If density at the radius is (r), we
have
M(r) =
_
r
0
4r
2
(r

) dr

. (3.43)
Combining Eqs. (3.42) and (3.43), we obtain the gravitational acceleration as a function of r as
g(r) =
G
r
2
_
r
0
4r
2
(r

) dr

. (3.44)
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 eld. Equation (3.25) is rewritten for the spherical coordinates
O-r as
0 =

rr
r

1
r


1
r sin

g, (3.45)
where both and g are the functions r and the latter is given by Eq. (3.44). The gravity term g
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 uid at rest, the state of stress in the body is hydrostatic.
In this case, shear stresses
r
and
r
vanish, and in Eq. (3.45)
rr
should be replaced by the
hydrostatic pressure p. Consequently, we have
dp
dr
= g. (3.46)
As 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 eld 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 dicult to estimate its dimension. Accordingly, we assume
the Moon to be a rocky spherical body with a constant density = 3.3 10
3
kg m
3
to roughly
estimate the gravitational acceleration and pressure at depths
8
. In the case of a constant density ,
the integral in Eq. (3.43) results in M(r) = 4r
3
/3, and we obtain
g(r) =
4
3
Gr. (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 spher-
ically symmetric body. (a) Single-layer model Moon with a constant density with the parameters
= 3.310
3
kg m
3
and R = 1740 km. (b) Two-layer model Earth with the parameters
m
= 410
3
kg m
3
,
c
= 10 10
3
kg m
3
, R
c
= 3490 km, and R = 6380 km.
It follows that gravitational acceleration for a constant density, spherically symmetric body is pro-
portional to the distance from the center (Fig. 3.7(a)). Substituting Eq. (3.47) into (3.46), we
have
dp
dr
=
4G
2
3
r.
Using the boundary condition p = 0 at r = R, where R stands for the radius of the Moon, we obtain
the pressure distribution
p(r) =
2
3

2
G(R
2
r
2
). (3.48)
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
dierence between the inner and outer cores. Let R and R
C
be the radius of the Earth and the core,
respectively, then the density is
=
_

C
(r R
C
)

m
(R
C
< r R).
(3.49)
Gravitational acceleration in the core is the same as Eq. (3.47) and we have
g(r) =
4
3
G
C
r (r R
C
),
74 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
where
C
is the density of the core. If that of the mantle is
m
, gravitational acceleration in the mantle
is obtained by using Eqs. (3.44) and (3.49) as
g(r) =
4G
3
_

m
r + (
C

m
)
R
3
C
r
2
_
(R
C
< r R).
Figure 3.7(b) shows g(r) for a double-layered model Earth. It is seen in this gure 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
p(r) =
4
3
G
m
(
C

m
) R
3
C
_
1
r

1
R
_
+
2
3
G
2
m
_
R
2
R
2
C
_
+
2
3
G
2
C
_
R
2
C
r
2
_
(r < R
C
) ,
p(r) =
4
3
G
m
R
3
c
(
C

m
)
_
1
R
c

1
R
_
+
2
3
G
2
m
_
R
2
r
2
_
(R
C
< r < R) .
The gravitational acceleration and pressure in the model Earth are calculated with the values,
m
=
4.0 10
3
kg m
3
,
core
= 10 10
3
kg m
3
, R
c
= 3490 km, R = 6380 km, and G = 6.7 10
11
Nm
2
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 oating 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 uid with a density of (Fig. 3.8). The body and uid 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
0 =
_
S
t(n) dS +
_
V
X dV, (3.50)
where V and S are the volume and surface of the body. The rst and second terms on the right-hand
side of this equation represent surface force from the uid 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 Gausss 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
_
S
pndS =
_
V
p dV. (3.51)
A minus sign is placed on the right-hand side of this equation because the unit normal n is dened
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 uid pressure at the
surface. Since the uid 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 F
b
be buoyancy,
and be constant, then Eq. (3.50) is converted to the equation
F
b
=
_
V
dV = V,
where the negative signs came from the denition of the force F
b
that is positive upward. This
equation stands for Archimedes principle: an object completely immersed in a uid experiences an
upward buoyant force equal in magnitude to the weight of the uid 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 t
c
is oating 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
m
gh
c
S, where
m
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
c
gt
c
S, which should be balanced with the buoyancy force.
Therefore, we have

c
gt
c
=
m
gh
c
. (3.52)
76 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
Figure 3.9: A simple model for a continental crust oating 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 litho-
spheric plate that is composed of a crust and upper mantle is often envisaged as oating on a uid
asthenosphere. Rocks behave as uid in a geological timescale, if they have high temperature. On
the other hand, if the lower mantle is heated to become uid, upper crustal blocks with smaller den-
sities oat on it. The level is placed in the middle of the crust. The level under which rocks behave
as uid and they cannot support a lithostatic pressure dierence 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:
_
d
c
surface
g dz = constant, (3.53)
where d
c
is the depth of compensation, and the z-axis is dened downward with the sea level at z = 0.
If an oshore 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 reected 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 34 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
p
L
=
c
gt
o
= 0.98 GPa. (3.54)
The overburden under the ocean basin at the same level is
p
L
=
w
gb +
o
gt
o
+
m
g (t
c
h b t
o
) 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
p
L
=
c
gt
o
+
m
g (t
c
t
o
x) .
Equating this pressure and that in Eq. (3.54), we obtain x = (t
c
t
o
) (
m

c
)/
m
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 signicant 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 reects the density contrast of the continen-
tal 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 redis-
tributing rocks and sediments. The strength of the lithosphere also aects the distribution. The
continental lithosphere has mechanical strength so that plate convergence causes continental thick-
ening with a limiting average altitude. As the average altitude exceeds it, increased gravitational
force attens 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 eect 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 ac-
cumulated in sublittoral paleoenvironments
9
. 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 sedi-
mentation 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. Simplied from [101].
Figure 3.13: Sedimentary basin subsidence by sediment load.
simplicity, a constant density for sediments
s
. If an ocean basin with an initial depth of b is lled
up with sediments (Fig. 3.13), the nal thickness of the sedimentary pile x is given by the isostasy
equation

w
gb +
c
gt
c
+
m
g (x b) =
s
gx +
c
gt
c
.
The left- and right-hand sides of this equation correspond to the left and right columns, respectively,
in Fig. 3.13. Therefore, we obtain
x =
_

m

w

m

s
_
b. (3.56)
The density of sediment depends on the lithology. Using a representative value
s
= 2.3 10
3
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 nal 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 stratig-
raphy 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-stratication, which indicates the depths were so shallow that
the sedimentation was aected 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 paleo-
bathymetry gives the depth to the basement from the present sea level, and applies several cor-
rections. 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 impor-
tant. If the supply is scarce, a sediment-starved basin is the result. Thirdly, eustasy aects 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 lled with forma-
tion uids that are mostly water. For example, mudstone has a porosity of some 50% shortly after
deposition, However, increasing compactness of the particles expels the uids during burial, i.e.,
compaction occurs. The porosity 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 . It is known that porosity of sedimentary rocks decreases roughly exponentially
with depth,
=
0
e
C
, (3.57)
where is depth fromthe ocean bottom. The parameters
0
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 posi-
tion 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 innite extension, the porosity reduction results in the thinning of
the formation. Consider a formation with an initial thickness of d
0
decreasing to d
N
by increasing
burial depth from d
0
to d
N
. A column of sedimentary rocks with a unit sectional area has the total
volume of sedimentary particles
_
d
0
+d
0
d
0
(1 ) d between the depths d
0
and d
0
+ d
0
. Pore uid
is expelled from this part of the column by burial, but the particles are not. The conservation of
82 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
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
_
d
0
+d
0
d
0
(1 ) d =
_
d
N
+d
N
d
N
(1 ) d.
Substituting Eq. (3.57), we have
d
0
+

0
C
e
Cd
0
_
e
Cd
0
1
_
= d
N
+

0
C
e
Cd
N
_
e
Cd
N
1
_
. (3.58)
The parameters C and
0
are known for specic rock types. Therefore, if the thickness of a formation
d
N
when the depth of the formation was d
N
is given, we can calculate the thickness d
0
at a given
depth of d
0
through Eq. (3.58). That is, this equation corrects the eect 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
i

w
and sedimentary
grains (1
i
)
g
, where
i
is the porosity of the layer and
g
is the average density of the grains. If
3.11. TECTONIC FORCE CAUSED BY HORIZONTAL DENSITY VARIATIONS 83
d
i
is the thickness of this layer just after it was deposited, we obtain d
i
using the recursive equation
Eq. (3.58). The mass of the layer per unit basal area is
m
i
=
_

w
+ (1
i
)
g
_
d
i
. (3.59)
If the lowermost layer has the ordinal number n, the sedimentary column with a unit sectional area
has the mass of
M
i
= m
i
+ m
i+1
+ + m
n1
+ m
n
(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 h
i
, we have the equation

w
(b
i
e
i
) + M
i
=
w
(b
i+1
e
i+1
) + M
i+1
+
m
h
c
(3.61)
from isostasy
10
, where b
i
is the water depth and e
i
is the level of the ocean surface relative to the
present one when the ith layer was deposited. That is, the sequence e
n
, e
n1
, . . . , e
0
represents
the eustatic sea level changes through time. They are estimated from ancient marine transgression
and regression observed around stable continents
11
. Combining Eqs. (3.58)(3.60), we obtain M
i
.
Therefore, h
c
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, ssion 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 o 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 exural 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 o 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 dierence 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 dierence generally drives deformation. Accordingly,
the dierence 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 prole of density contrast between the regions. (c) Tectonic ow driven by the pressure
dierence.
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 prole of the density
dierence , which is schematically shown in Fig. 3.18(b). The density prole 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 prole. 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 10
12
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 eect 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 uid 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 kmby NS cotinental convergence, and
the crustal thickness reaches 70 km. EW trending thrust faults are dominant along the Himalayas,
but there are NS or NNESSW 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 limit
12
.
The limiting altitude depends on the strength of the continental lithosphere, which has temporal
and spatial variations
13
. To see this, consider viscous uid like honey in a container that has a
moving belt at its bottom. This thought experiment models a convergent orogen, where the uid
simulates the continental crust. The belt drags the uid to form a slope at the surface of the uid
(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 dierence of the uid 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 eective 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-oor 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 conduc-
tion is the primary mechanism of heat transfer in the lithosphere
14
. 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 exural 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: Depthage 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 sedimen-
tary cover. GDH1, PSM and HS are synthesized depth-age curves based on dierent 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 tempera-
ture. 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 eld aects isostatic balance.
This is known as thermal isostasy. Therefore, the thermal regime of the lithosphere aects the
vertical movement of ocean basins. The cooling lithosphere causes subsidence. To understand this
phenomenon, we rstly 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
t
=
2
T, (3.62)
where T is the temperature that is a function of time t and position, and is the thermal diusivity
of the rock mass. If heat is conducted only in the x direction, this equation reduces to
T
t
=

2
T
x
2
. (3.63)
The representative value for the thermal diusivity of rocks is 1 m
2
s
1
. This unit designates that
/L
2
has the dimension of time. Accordingly, we dene the dimensionless length x and time

t by
dividing L and L
2
/ and the dimensionless temperature T by dividing by a reference temperature
T
0
. Namely,
x = x/L,

t = t/L
2
, T = T/T
0
,
dx = Ld x, dt = L
2
d

t/, dT = T
0
dT.
(3.64)
Using these dx, dt dT, we rewrite Eq. (3.62) in the form
T

t
=

2
T
x
2
(

t > 0, 0 < x < 1) .


This equation does not include material constants such as , so that its solution depends only on the
initial and boundary conditions. The solution can be applied to specic problems using Eq. (3.64).
The time representing the cooling speed of the rock mass is evaluated by the ratio, = L
2
/.
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 dene the Eulerian coordinates O-xz with the origin at the spreading center and the
Laglangian coordinate that travels with the oceanic plate (Fig. 3.21). The origin of -coordinate
3.12. THERMAL ISOSTASY 89
is placed at z = 0, i.e., z = . For simplicity, we assume that heat is transferred only by conduction.
To calculate temperature change, the thermal conduction equation,

2
T
t
2
=

2
T

2
is solved with the initial and boundary conditions
T = T
a
at 0, t = 0,
T = T
0
at = 0, t > 0,
T T
a
as , t > 0.
That is, we assume a cooling half-space with the constant temperatures T
0
and T
a
at the ocean oor
and at the deep mantle, respectively. As the lithosphere ages, the ocean oor subsides and departs
from z = 0 by a few kilometers. Therefore, = 0 is not the level of the oor, 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,
T T
0
T
a
T
0
= Erf
_

2

t
_
, (3.65)
where Erf( ) is the error function (Fig. 3.22) dened by the equation
Erf(x) =
2

_
x
0
e

2
d (x 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,

t, in the operand of the error function in Eq. (3.65). The
denominator has the dimensions of length so that

t is called the diusion length. Obviously, an
isotherm descends with age as

t.
The error function has an inverse function. Therefore, the depth of the isotherm with a specic
temperature T is given by Eq. (3.65), where the left-hand side is known. Substituting t = x/v and
= x into the equation, we have
T T
0
T
a
T
0
= Erf
_
z
2
_
v
x
_
. (3.67)
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 T, which is measured from the
reference temperature T
0
, changes as
=
0
(1 T), (3.68)
where
0
is the density at the reference temperature and is the thermal expansion coecient. Rocks
have a representative value for the coecient at about 3 10
5
K
1
. We use
a
=
m0
(1 T
a
) 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 oor from the level = 0, the isostasy
equation gives

w
b +
_
d
c
0
d = const. (3.69)
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,

a
d
c
+
w
b
R
, where b
R
is the water depth of the axis. Accordingly, Eq. (3.69) is rewritten as

w
b +
_
d
c
0
d =
a
d
c
+
w
b
R
. (3.70)
We assume the half-space so that the compensation depth can be treated as d
c
= . Combining Eqs.
(3.70), (3.67) and (3.68), we obtain
b(
a

w
) =
m0
(T
a
T
0
)
_

0
_
1 Erf
_
z
2
_
v
x
__
dz
=
m0
(T
a
T
0
)
_

0
Erfc
_
z
2
_
v
x
_
dz
Erfc(x) = 1 Erf(x) is the complementary error function (Fig. 3.22). This function satises
_

0
Erfc(x) dx =
1

,
therefore, the depth of the sea oor at a distance of x from the spreading axis is given by the equation
b =
2
a
(T
a
T
0
)

a

w
_
x
v
+ b
R
. (3.71)
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 eective values for the thermal diusivity and thermal expansion
coecient of the oceanic mantle by comparing the above model with observations. However, it is
important that old sea oors are shallower than the depth predicted by the model (Fig. 3.20) and the
ocean basins have dierent deviations. Some factors other than the cooling are responsible for this
attening. 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 oors show the
dierence 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 oor. 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
0
under the axis. The coordinates O-xz shown in Fig. 3.21 are used in this section. Consider
the density distribution
(x, z) =
0
+ (x, z),
where the last term is the density anomaly from the reference density
0
. According to Eq. (3.65),
the temperature and the resultant density anomalies were derived in the previous section as
T(z, t) = T
a
Erf
_
z
2

t
_
,
(z, t) =
m0
T
a
Erfc
_
z
2

t
_
. (3.72)
The force F
R
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
F
R
=
_

0
gz dz =
_

m0
gT
a

_
t.
Consequently, the force is proportional to the age of the oceanic lithosphere
15
. The constant of
proportionality is about 0.035 TNm
1
/m.y. For example, the ocean oor with an age of 80 m.y.
has F
R
= 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).
92 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
In the above argument, we used the density prole 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 eld. 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
zz
(1 m
2
) and that the product of force and length has the
dimensions of energy. Then the quantity
U =
_
L
h

zz
dz =
_
L
h
_
h

()g ddz
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 eld 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 Mckenzies 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 T
a
at a depth of z = t
L
. 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 t
c
and t
L
be
the initial thickness of the crust and the lithosphere, respectively (Fig. 3.23(a)). It is also assumed
that the temperature at z = t
L
is kept at T
a
. 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 , the thickness of the lithosphere and crust
become t
c
/ and t
L
/, respectively. The rift-stage subsidence is aected 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
=
0
(1 T), (3.73)
where
0
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 adiabatically
16
(Fig.
3.23(b)). If we assume isostatic compensation throughout the rifting process, we obtain the syn-rift
subsidence
S
i
=
t
L

a
t
c

c
(t
L
t
c
)
m

a

w
_
1
1

_
, (3.74)
where
c
,
m
and
a
are the representative densities of the crust, mantle lithosphere, and astheno-
sphere, respectively, and are given by

a
=
m0
(1 T
a
),
c
=
c0
_
1
T
a
2
t
c
t
L
_
,
m
=
m0
_
1
T
a
2

T
a
2
t
c
t
L
_
.
16
This 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.

m0
= 3.3 10
3
kg m
3
= 3.510
5
K
1
t
L
= 100 km

c0
= 2.8 10
3
kg m
3
= 1.010
6
m
2
s
1
t
c
= 35 km

w
= 1.0 10
3
kg m
3
T
a
= 1333

C
The coecients,
c0
and
m0
are the reference densities of the crust and mantle. S
i
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
S
i
= S
i0
_
1
1

_
, (3.75)
S
i0
=
_
(
m0

c0
)
_
1
T
a
2
t
c
t
L
_
t
c
t
L

m0
T
a
2
_
t
L

m0
(1 T
a
)
w
. (3.76)
The latter quantity, (S
i0
) 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,
t
c
t
L
=

m0

c0
(
m0

c0
)T
a

_
4(
m0

c0
)
2
+ 4
2
(
m0

c0
)
m0
T
2
a
_
(
m0

c0
)T
a
.
Using the parameters shown in Table 3.2, we obtain the ratio t
c
/t
L
= 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 t
c
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 t
L
= 50 km, then the arc
has the threshold at about 7 km. The geohistory analysis of syn-rift sediments can constrain with
Eq. (3.75). This parameter is called the stretching factor.
The thermal regime is altered by the rifting. Isotherms between 0

and T
a
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 S
i
< t
L
. Therefore, the thermal conduction equation T/t =
2
T/z
2
is solved
with the boundary condition T = 0 at z = 0. In addition, the temperature at z = t
L
is assumed to be
3.15. KINEMATIC MODEL OF SYN-RIFT SUBSIDENCE 95
kept at T = T
a
. 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
T
T
a
=
_
1
z
t
L
_
+
2

n=1
(1)
n+1
n

n
sin
n

sin
nz
t
L
e
n
2
t/
, (3.77)
=
2
t
2
L
/. (3.78)
The parameters shown in Table 3.2 give the time constant = 3.2 m.y.
Combining Eqs. (3.73) and (3.77), we obtain the evolution of the density structure (z, t) .
Isostatic compensation through time is expressed as
_
t
L
0
(z, t) dz +
w
S
t
=
_
t
L
0
(z, 0) dz,
where S
t
is the thermal subsidence,
S
t

_
4
m0
T
a
t
L

2
(
m0

c0
)
_

sin

_
1 e
t/
_
. (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 con-
trolled by not only the thermal but also the mechanical properties of the lithosphere. The latter
aects 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 subsi-
dence 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

zz
=
xx
= G(t),
96 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
where the function G(t) is dened 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 eld 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 pro-
portional 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 aects 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 =

2
T, should be replaced by the material derivative (Eq. (2.20)). That is, the equation
T
t
+
_

_
u
v
__
T =
_

2
T
x
2
+

2
T
z
2
_
(3.82)
governs the thermal evolution. From Eqs. (3.81) and (3.80), we have
_

_
u
v
__
T = G(t)x
T
x
G(t)z
T
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
x
=

2
T
x
2
= 0.
Combining this and Eqs. (3.82) and (3.83), we obtain
T
t
+ Gz
T
z
=

2
T
z
2
. (3.84)
The boundary conditions that we assume for the temperature evolution are T = 0 at z = 0 and
T = T
a
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
Q(t) = B
_
a
0
_
T(z, t) T(z, )
_
dz, (3.85)
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% condence region is shown in (a) by white [259].
where B =
m
/(
a

w
) is the factor to account for the buoyancy of the crust. Therefore, we have
the total subsidence
S(t) = A
_
1
1

_
BQ(t), (3.86)
where A = t
c
(
m

c
)/(
a

w
). Crustal thinning has a positive eect for the total subsidence, but
the thinning of the lithosphere increases the geothermal gradient which has a negative eect for the
subsidence. Rearranging Eq. (3.86), we have
=
A
A S BQ(t)
. (3.87)
Equation (2.39) links the stretching factor and G(t) as
= exp
__
t
R
0
G(t) dt
_
, (3.88)
where t
R
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
G(t) =
d
dt
_
ln
_
A
A S(t) BQ(t)
__
. (3.89)
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 inversion
17
from the observed subsidence history S(t). Figure 3.24(a) shows the
17
See [259] for detail.
98 CHAPTER 3. STRESS, BALANCE EQUATIONS, AND ISOSTASY
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 inection 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 rift-
ing; 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%
condence region between 100 and 220 Ma in Fig. 3.24(a). The umbiguity stems from the diculty
to detect very slow rifting if it overlaps simultaneous thermal subsidence. More denite 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 o 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
1
= 4 MPa,
2
= 5 MPa and
3
= 8 MPa,
and vertical
1
-axis and horizontal
3
-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 10
3
kg m
3
, respectively.
3.4 Derive Eq. (3.46).
3.5 Estimate howmuch 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 99
Figure 3.25: Lithostatic pressure under the neighboring plateau and lowland.

Вам также может понравиться