0% found this document useful (0 votes)
32 views30 pages

Elasticity, Flexure - 21

This document discusses elasticity and the flexure of the lithosphere. It introduces linear elasticity and how stress is proportional to strain for elastic materials. It provides evidence that the lithosphere behaves elastically on geological timescales from observations of structures, earthquakes, and flexural features like bending of islands and sedimentary basins.

Uploaded by

nitish3542
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
32 views30 pages

Elasticity, Flexure - 21

This document discusses elasticity and the flexure of the lithosphere. It introduces linear elasticity and how stress is proportional to strain for elastic materials. It provides evidence that the lithosphere behaves elastically on geological timescales from observations of structures, earthquakes, and flexural features like bending of islands and sedimentary basins.

Uploaded by

nitish3542
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 30

i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 133 — #147

3
Elasticity and Flexure
In this Chapter
In this chapter we introduce the fundamentals of elasticity. Elasticity is the principal deformation
mechanism applicable to the lithosphere. In linear elasticity strain is proportional to stress.
Elastic deformation is reversible; when the applied stress is removed, the strain goes to zero.
Deformation of the lithosphere, in a number of applications, can be approximated as the
bending (flexure) of a thin elastic plate. Examples include bending under volcanic loads, bending
at subduction zones, and bending that creates sedimentary basins.

3.1 Introduction to increasingly large confining pressures due to the


increasing weight of the overburden. When the con-
In the previous chapter we introduced the concepts fining pressure on the rock approaches its brittle fail-
of stress and strain. For many solids it is appro- ure strength, it deforms plastically. Plastic deforma-
priate to relate stress to strain through the laws of tion is a continuous, irreversible deformation without
elasticity. Elastic materials deform when a force is fracture. If the applied force causing plastic defor-
applied and return to their original shape when the mation is removed, some fraction of the deforma-
force is removed. Almost all solid materials, includ- tion remains. We consider plastic deformation in
ing essentially all rocks at relatively low temperatures Section 7.11. As discussed in Chapter 1, hot man-
and pressures, behave elastically when the applied tle rocks behave as a fluid on geological time scales;
forces are not too large. In addition, the elastic that is, they continuously deform under an applied
strain of many rocks is linearly proportional to the force.
applied stress. The equations of linear elasticity are Given that rocks behave quite differently in response
greatly simplified if the material is isotropic, that is, to applied forces, depending on conditions of temper-
if its elastic properties are independent of direction. ature and pressure, it is important to determine what
Although some metamorphic rocks with strong foli- fraction of the rocks of the crust and upper mantle
ations are not strictly isotropic, the isotropic approxi- behave elastically on geological time scales. One of
mation is usually satisfactory for the Earth’s crust and the fundamental postulates of plate tectonics is that
mantle. the surface plates constituting the lithosphere do not
At high stress levels, or at temperatures that are deform significantly on geological time scales. Several
a significant fraction of the rock solidus, deviations observations directly confirm this postulate. We know
from elastic behavior occur. At low temperatures that the transform faults connecting offset segments of
and confining pressures, rocks are brittle solids, and the oceanic ridge system are responsible for the major
large deviatoric stresses cause fracture. As rocks are linear fracture zones in the ocean. That these frac-
buried more deeply in the Earth, they are subjected ture zones remain linear and at constant separation is

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 134 — #148

134 Elasticity and Flexure

direct evidence that the oceanic lithosphere does not other island chains, individual islands, and seamounts.
deform on a time scale of 108 years. Similar evidence There is also observational evidence of the elastic
comes from the linearity of the magnetic lineaments of bending of the oceanic lithosphere at ocean trenches
the seafloor (see Section 1.8). and of the continental lithosphere at sedimentary
There is yet other direct evidence of the elastic basins – the Michigan basin, for example. We will
behavior of the lithosphere on geological time scales. make quantitative comparisons of the theoretically
Although erosion destroys mountain ranges on a time predicted elastic deformations of these structures with
scale of 106 to 107 years, many geological structures the observational data in later sections of this chapter.
in the continental crust have ages greater than 109 One important reason for studying the elastic behav-
years. The very existence of these structures is evidence ior of the lithosphere is to determine the state of stress
of the elastic behavior of the lithosphere. If the rocks of in the lithosphere. This stress distribution is respon-
the crust behaved as a fluid on geological time scales, sible for the occurrence of earthquakes. Earthquakes
the gravitational body force would have erased these are direct evidence of high stress levels in the litho-
structures. As an example, pour a very viscous sub- sphere. An earthquake relieves accumulated strain in
stance such as molasses onto the bottom of a flat pan. the lithosphere. The presence of mountains is also evi-
If the fluid is sufficiently viscous and is poured quickly dence of high stress levels. Elastic stresses must balance
enough, a structure resembling a mountain forms (see the gravitational body forces on mountains. Because
Figure 3.1a). However, over time, the fluid will even- of their elastic behavior, surface plates can transmit
tually cover the bottom of the pan to a uniform depth stresses over large horizontal distances.
(see Figure 3.1b). The gravitational body force causes
the fluid to flow so as to minimize the gravitational 3.2 Linear Elasticity
potential energy.
A number of geological phenomena allow the long- A linear, isotropic, elastic solid is one in which stresses
term elastic behavior of the lithosphere to be stud- are linearly proportional to strains and mechanical
ied quantitatively. In several instances the lithosphere properties have no preferred orientations. The princi-
bends under surface loads. Direct evidence of this pal axes of stress and strain coincide in such a medium,
bending comes from the Hawaiian Islands and many and the connection between stress and strain can be
conveniently written in this coordinate system as

σ1 = (λ + 2G)ε1 + λε2 + λε3 (3.1)


σ2 = λε1 + (λ + 2G)ε2 + λε3 (3.2)
σ3 = λε1 + λε2 + (λ + 2G)ε3 , (3.3)

where the material properties λ and G are known as


Lamé parameters; G is also known as the modulus of
rigidity. The material properties are such that a princi-
pal strain component ε produces a stress (λ + 2G)ε
in the same direction and stresses λε in mutually
perpendicular directions.
Equations (3.1) to (3.3) can be written in the inverse
form as
1 ν ν
ε1 = σ1 − σ2 − σ3 (3.4)
E E E
Figure 3.1 (a) Structure formed immediately after rapidly pouring
a very viscous fluid into a container. (b) Final shape of the fluid ν 1 ν
ε2 = − σ1 + σ2 − σ3 (3.5)
after a long time has elapsed. E E E

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 135 — #149

3.3 Uniaxial Stress 135

ν ν 1 3.3 Uniaxial Stress


ε3 = − σ1 − σ2 + σ3 , (3.6)
E E E
and E and ν are material properties known as In a state of uniaxial stress only one of the principal
Young’s modulus and Poisson’s ratio, respectively. A stresses, σ1 say, is nonzero. Under this circumstance
principal stress component σ produces a strain σ/E Equations (3.2) and (3.3), with σ2 = σ3 = 0, give
in the same direction and strains (−νσ/E) in mutually −λ
ε2 = ε3 = ε1 . (3.7)
orthogonal directions. 2(λ + G)
The elastic behavior of a material can be charac- Not only does the stress σ1 produce a strain ε1 , but it
terized by specifying either λ and G or E and ν; the changes the linear dimensions of elements aligned per-
sets of parameters are not independent. Analytic for- pendicular to the axis of stress. If σ1 is a compression,
mulas expressing λ and G in terms of E and ν, and then ε1 is a decrease in length, and both ε2 and ε3 are
vice versa, are obtained in the following sections. Val- increases in length. The element in Figure 3.2 has been
ues of E, G, and ν for various rocks are given in shortened in the y direction, but its cross section in the
Section B.5 of Appendix B. Young’s modulus of rocks xz plane has expanded.
varies from about 10 to 100 GPa, and Poisson’s ratio Using Equations (3.4) to (3.6), we can also write
varies between 0.1 and 0.4. The elastic properties of ν
ε2 = ε3 = − σ1 = −νε1 . (3.8)
the Earth’s mantle and core can be obtained from seis- E
mic velocities and the density distribution. The elastic By comparing Equations (3.7) and (3.8), we see that
properties E, G, and ν inferred from a typical seis-
λ
mically derived Earth model are given in Section B.6 ν= . (3.9)
2(λ + G)
of Appendix B. The absence of shear waves in the
outer core (G = 0) is taken as conclusive evidence From Equations (3.1) and (3.7), we find
that the outer core is a liquid. In the outer core ν has G(3λ + 2G)
σ1 = ε1 , (3.10)
the value 0.5, which we will see is appropriate to an (λ + G)
incompressible fluid.
which, with the help of Equation (3.8), identifies
The behavior of linear solids is more readily illus-
Young’s modulus as
trated if we consider idealized situations where several
of the stress and strain components vanish. These can G(3λ + 2G)
E= . (3.11)
then be applied to important geological problems. (λ + G)

Figure 3.2 Deformation under uniaxial stress.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 136 — #150

136 Elasticity and Flexure

Figure 3.3 Stress–strain curves for quartzite in uniaxial compression (Bieniawski, 1967).

Equations (3.9) and (3.11) can be inverted to yield The dilatation or fractional volume change in
the following formulas for G and λ in terms of E uniaxial compression is, according to Equation (3.8),
and ν
= ε1 + ε2 + ε3 = ε1 (1 − 2ν). (3.15)
E
G= (3.12)
2(1 + ν) The decrease in volume due to contraction in the direc-
Eν tion of compressive stress is offset by an increase in
λ= . (3.13) volume due to expansion in the orthogonal directions.
(1 + ν)(1 − 2ν)
Equation (3.15) allows us to determine Poisson’s ratio
The relation between stress and strain in uniaxial for an incompressible material, which cannot undergo
compression or tension from Equation (3.8), a net change in volume. In order for to equal zero
σ1 = Eε1 , (3.14) in uniaxial compression, ν must equal 1/2. Under
uniaxial compression, an incompressible material con-
is also known as Hooke’s law. A linear elastic solid tracts in the direction of applied stress but expands
is said to exhibit Hookean behavior. Uniaxial com- exactly half as much in each of the perpendicular
pression testing in the laboratory is one of the sim- directions.
plest methods of determining the elastic properties There are some circumstances in which the formulas
of rocks. Figure 3.3 shows the data from such a of uniaxial compression can be applied to calculate the
test on a cylindrical sample of quartzite. The rock strains in rocks. Consider, for example, a rectangular
deforms approximately elastically until the applied column of height h that is free to expand or contract
stress exceeds the compressive strength of the rock, in the horizontal; that is, it is laterally unconstrained.
at which point failure occurs. Compressive strengths By this we mean that the horizontal stresses are zero
of rocks are hundreds to thousands of megapascals. (σ2 = σ3 = 0). Then the vertical stress σ1 at a distance
As we discussed in the previous chapter, a typical y from the top of the column of rock is given by the
tectonic stress is 10 MPa. With E = 70 GPa, this weight of the column,
yields a typical tectonic strain in uniaxial stress of
1.4 × 10−4 . σ1 = ρgy. (3.16)

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 137 — #151

3.4 Uniaxial Strain 137

The vertical strain as a function of the distance y from


the top is
ρgy
ε1 = . (3.17)
E
The slab contracts in the vertical by an amount
 h 
ρg h ρgh2
δh = ε1 dy = y dy = . (3.18)
0 E 0 2E

3.4 Uniaxial Strain


Figure 3.4 Stresses on a surface covered by sediments of
The state of uniaxial strain corresponds to only one
thickness h.
nonzero component of principal strain, ε1 say. With
ε2 = ε3 = 0, Equations (3.1) to (3.3) give
It is of interest to determine the deviatoric stresses
σ1 = (λ + 2G)ε1 (3.19)
after sedimentation. The pressure at depth h as defined
λ by Equation (2.61) is
σ2 = σ3 = λε1 = σ1 . (3.20)
(λ + 2G)
1 (1 + ν)
Equations (3.4) to (3.6) simplify to p= (σ1 + σ2 + σ3 ) = ρgh. (3.25)
3 3(1 − ν)
ν
σ2 = σ3 = σ1 (3.21)
(1 − ν) The deviatoric stresses are then determined from
(1 − ν)Eε1 Equations (2.63) with the result
σ1 = . (3.22)
(1 + ν)(1 − 2ν) 2(1 − 2ν)
σ1 = σ1 − p = ρgh (3.26)
By comparing Equations (3.19) to (3.22), one can also 3(1 − ν)
derive the relations already given between λ, G and (1 − 2ν)
σ2 = σ2 − p = σ3 = σ3 − p = − ρgh. (3.27)
ν, E. 3(1 − ν)
The equations of uniaxial strain can be used to
determine the change in stress due to sedimentation The horizontal deviatoric stress is tensional. For ν =
or erosion. We first consider sedimentation and assume 0.25 the horizontal deviatoric stress is 2/9 of the
that an initial surface is covered by h km of sediments lithostatic stress. With ρ = 3000 kg m−3 and h =
of density ρ, as shown in Figure 3.4. We also assume 2 km the horizontal deviatoric stress is −13.3 MPa.
that the base of the new sedimentary basin is laterally This stress is of the same order as measured surface
confined so that the equations of uniaxial strain are stresses.
applicable. The two horizontal components of strain We next consider erosion. If the initial state of stress
are zero, ε2 = ε3 = 0. The vertical principal stress before erosion is that given above, erosion will result
on the initial surface σ1 is given by the weight of the in the state of stress that existed before sedimentation
overburden occurred. The processes of sedimentation and erosion
are reversible. However, in many cases the initial state
σ1 = ρgh. (3.23)
of stress prior to erosion is lithostatic. Therefore, at a
From Equation (3.21), the horizontal normal stresses depth h the principal stresses are
are given by
ν σ1 = σ2 = σ3 = ρgh. (3.28)
σ2 = σ3 = ρgh. (3.24)
(1 − ν) After the erosion of h km of overburden the vertical
The horizontal stresses are also compressive, but they stress at the surface is σ̄1 = 0 (an overbar denotes
are smaller than the vertical stress. a stress after erosion). The change in vertical stress

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 138 — #152

138 Elasticity and Flexure

σ1 = σ̄1 − σ1 is −ρgh. If only ε1 is nonzero, other (ε1 = 0). Determine ε2 and σ1 when y kilo-
Equation (3.21), gives meters of rock of density ρ are eroded away. Assume
 
ν that the initial state of stress was lithostatic.
σ2 = σ3 = σ1 . (3.29)
1−ν
The horizontal surface stresses after erosion σ̄2 and σ̄3
are consequently given by 3.5 Plane Stress
ν
σ̄2 = σ̄3 = σ2 + σ2 = ρgh − ρgh
(1 − ν) The state of plane stress exists when there is only one
  zero component of principal stress; that is, σ3 = 0,
1 − 2ν
= ρgh. (3.30) σ1 = 0, σ2 = 0. The situation is sketched in Figure 3.5,
1−ν
which shows a thin plate loaded on its edges. The
If h = 5 km, ν = 0.25, and ρ = 3000 kg m−3 , we strain components according to Equations (3.4) to
find from Equation (3.30) that σ̄2 = σ̄3 = 100 MPa. (3.6) are
Erosion can result in large surface compressive stresses
1
due simply to the elastic behavior of the rock. This ε1 = (σ1 − νσ2 ) (3.31)
mechanism is one explanation for the widespread E
occurrence of near-surface compressive stresses in the 1
ε2 = (σ2 − νσ1 ) (3.32)
continents. E
−ν
ε3 = (σ1 + σ2 ). (3.33)
Problem 3.1 E
Determine the surface stress after the erosion of
10 km of granite. Assume that the initial state of The geometry of Figure 3.5 suggests that the plane
stress is lithostatic and that ρ = 2700 kg m−3 and stress formulas may be applicable to horizontal tec-
ν = 0.25. tonic stresses in the lithosphere. Let us assume that
in addition to the lithostatic stresses there are equal
horizontal components of principal stress σ1 = σ2 .
Problem 3.2
An unstressed surface is covered with sediments with
a density of 2500 kg m−3 to a depth of 5 km. If the
surface is laterally constrained and has a Poisson’s
ratio of 0.25, what are the three components of stress
at the original surface?

Problem 3.3
A horizontal stress σ1 may be accompanied by stress
in other directions. If it is assumed that there is no
displacement in the other horizontal direction and
zero stress in the vertical, find the stress σ2 in the
other horizontal direction and the strain ε3 in the
vertical direction.

Problem 3.4
Assume that the Earth is unconstrained in one lat-
eral direction (σ2 = σ3 ) and is constrained in the
Figure 3.5 Plane stress.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 139 — #153

3.6 Plane Strain 139

According to Equations (3.31) to (3.33), the horizontal


tectonic stresses produce the strains
(1 − ν)
ε1 = ε2 = σ1 (3.34)
E
−2ν
ε3 = σ1 . (3.35)
E
If the horizontal tectonic stresses are compressive,
vertical columns of lithosphere of initial thickness
hL , horizontal area A, and density ρ will undergo a
decrease in area and an increase in thickness. The mass
in a column will remain constant, however. Therefore
we can write
δ(ρA hL ) = 0. (3.36)
Figure 3.6 Sketch of a triaxial compression test on a cylindrical
The weight per unit area at the base of the column rock sample.
ρghL will increase, as can be seen from
  small compared with the applied horizontal principal
1
δ(ρghL ) = δ ρghL A · stresses, we conclude that the plane stress assumption
A
  is valid for the Earth’s lithosphere.
1 1
= δ(ρghL A) + ρghL A δ
A A
    Problem 3.5
1 δA Triaxial compression tests are a common labora-
= ρghL A − 2 δA = ρghL − .
A A tory technique for determining elastic properties and
(3.37) strengths of rocks at various pressures p and tem-
The term δ(ρghL A)/A is zero from Equation (3.36); peratures. Figure 3.6 is a schematic of the exper-
δ(ρghL ) is positive because −δA/A is a positive quan- imental method. A cylindrical rock specimen is
tity given by loaded axially by a compressive stress σ1 . The sam-
ple is also uniformly compressed laterally by stresses
δA 2(1 − ν)
− = ε1 + ε2 = σ1 . (3.38) σ2 = σ3 < σ1 .
A E Show that
The increase in the weight per unit area at the base
ε2 = ε3
of the lithospheric column gives the increase in the
vertical principal stress σ3 . By combining Equa- and
tions (3.37) and (3.38), we get
σ1 − σ2 = 2G(ε1 − ε2 ).
2(1 − ν)ρghL
σ3 = σ1 (3.39) Thus if the measured stress difference σ1 − σ2 is plot-
E
ted against the measured strain difference ε1 −ε2 , the
or
slope of the line determines 2G.
σ3 2(1 − ν)ρghL
= . (3.40)
σ1 E
3.6 Plane Strain
Taking ν = 0.25, E = 100 GPa, ρ = 3000 kg m−3 ,
g = 10 m s−2 , and hL = 100 km as typical values In the case of plane strain, ε3 = 0, for example, and
for the lithosphere, we find that σ3 / σ1 = 0.045. ε1 and ε2 are nonzero. Figure 3.7 illustrates a plane
Because the change in the vertical principal stress is strain situation. A long bar is rigidly confined between

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 140 — #154

140 Elasticity and Flexure

Figure 3.7 An example of plane strain.

supports so that it cannot expand or contract parallel


to its length. In addition, the stresses σ1 and σ2 are Figure 3.8 Principal stresses and shear stresses in the case of
applied uniformly along the length of the bar. pure shear.
Equations (3.1) to (3.3) reduce to
σ1 = (λ + 2G)ε1 + λε2 (3.41) and from Equations (2.130) and (2.131) with θ = −45◦
we get εxx = εyy = 0 and εxy = ε1 . Equation (3.47)
σ2 = λε1 + (λ + 2G)ε2 (3.42)
then gives
σ3 = λ(ε1 + ε2 ). (3.43)
E
σxy = εxy . (3.48)
From Equation (3.6), it is obvious that 1+ν
σ3 = ν(σ1 + σ2 ). (3.44) By introducing the modulus of rigidity from Equa-
tion (3.12), we can write the shear stress as
This can be used together with Equations (3.4) and
(3.5) to find σxy = 2Gεxy , (3.49)
(1 + ν) which explains why the modulus of rigidity is also
ε1 = {σ1 (1 − ν) − νσ2 } (3.45)
E known as the shear modulus. (Note: In terms of γxy ≡
(1 + ν) 2εxy , σxy = Gγxy .) These results are valid for both pure
ε2 = {σ2 (1 − ν) − νσ1 }. (3.46) shear and simple shear because the two states differ by
E
a solid-body rotation that does not affect the state of
stress.
3.7 Pure Shear and Simple Shear Simple shear is generally associated with displace-
ments on a strike–slip fault such as the San Andreas
The state of stress associated with pure shear is illus- in California. In Equation (2.134) we concluded that
trated in Figure 3.8. Pure shear is a special case of the shear strain associated with the 1906 San Fran-
plane stress. One example of pure shear is σ3 = 0 cisco earthquake was 2.5 × 10−5 . With G = 30
and σ1 = −σ2 . From Equations (2.56) to (2.58) with GPa, Equation (3.49) gives the related shear stress as
θ = −45◦ (compare Figures 2.14 and 3.8), we find that 1.5 MPa. This is a very small stress drop to be associ-
σxx = σyy = 0 and σxy = σ1 . In this coordinate system ated with a great earthquake. However, for the stress
only the shear stress is nonzero. From Equations (3.31) drop to have been larger, the width of the zone of
and (3.32), we find that strain accumulation would have had to have been
(1 + ν) (1 + ν) even smaller. If the stress had been 15 MPa, the width
ε1 = σ1 = σxy = −ε2 , (3.47)
E E of the zone of strain accumulation would have had

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 141 — #155

3.9 Two-Dimensional Bending or Flexure of Plates 141

to have been 4 km on each side of the fault. We will Using Equations (3.11) to (3.13), we can rewrite the
return to this problem in Chapter 8. formula for K given in Equation (3.50) as

Problem 3.6 1 E
K= = . (3.55)
Show that Equation (3.49) can also be derived by β 3(1 − 2ν)
assuming plane strain. Thus as ν tends toward 1/2, that is, as a material
becomes more and more incompressible, its bulk mod-
ulus tends to infinity.
3.8 Isotropic Stress

If all the principal stresses are equal σ1 = σ2 = σ3 ≡ p, 3.9 Two-Dimensional Bending or Flexure of
then the state of stress is isotropic, and the princi- Plates
pal stresses are equal to the pressure. The principal
We have already discussed how plate tectonics implies
strains in a solid subjected to isotropic stresses are
that the near-surface rocks are rigid and therefore
also equal ε1 = ε2 = ε3 = 13 ; each component of
behave elastically on geological time scales. The thin
strain is equal to one-third of the dilatation. By adding
elastic surface plates constitute the lithosphere, which
Equations (3.1) to (3.3), we find
  floats on the relatively fluid mantle beneath. The
3λ + 2G 1 plates are subject to a variety of loads – volcanoes,
p= ≡K ≡ . (3.50)
3 β seamounts, for example – that force the lithosphere
The quantity K is the bulk modulus, and its recip- to bend under their weights. By relating the observed
rocal is β, the compressibility. The ratio of p to the flexure or bending of the lithosphere to known surface
bulk modulus gives the fractional volume change that loads, we can deduce the elastic properties and thick-
occurs under isotropic compression. nesses of the plates. In what follows, we first develop
Because the mass of a solid element with volume V the theory of plate bending in response to applied
and density ρ must be conserved, any change in vol- forces and torques. The theory can also be used to
ume δV of the element must be accompanied by a understand fold trains in mountain belts by model-
change in its density δρ. The fractional change in den- ing the folds as deformations of elastic plates subject
sity can be related to the fractional change in volume, to horizontal compressive forces. Other geologic appli-
the dilatation, by rearranging the equation of mass cations also can be made. For example, we will apply
conservation the theory to model the upwarping of strata overlying
igneous intrusions (Section 3.12).
δ(ρV ) = 0, (3.51)
A simple example of plate bending is shown in
which gives Figure 3.9. A plate of thickness h and width L is
pinned at its ends and bends under the load of a line
ρδV + V δρ = 0 (3.52)
or
−δV δρ
= = . (3.53)
V ρ
Equation (3.53) of course assumes to be small. The
combination of Equations (3.50) and (3.53) gives

δρ = ρβp. (3.54)
This relationship can be used to determine the increase Figure 3.9 A thin plate of length L and thickness h pinned at its
in density with depth in the Earth. ends and bending under and applied load Va .

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 142 — #156

142 Elasticity and Flexure

Figure 3.10 Forces and torques on a small section of a deflecting plate.

force Va (N m−1 ) applied at its center. The plate is quantities in Figure 3.10 are considered positive when
infinitely long in the z direction. A vertical, static force they have the sense shown in the figure. At location
balance and the symmetry of the situation require x along the plate the shear force is V , the bending
that equal vertical line forces Va /2 be applied at the moment is M, and the deflection is w; at x + dx,
supports. The plate is assumed to be thin compared the shear force is V + dV , the bending moment is
with its width, h L, and the vertical deflection of M + dM, and the deflection is w + dw. It is to be
the plate w is taken to be small, w L. The lat- emphasized that V , M, and P are per unit length in the
ter assumption is necessary to justify the use of linear z direction.
elastic theory. The two-dimensional bending of plates A force balance in the vertical direction on the
is also referred to as cylindrical bending because the element between x and x + dx yields
plate takes the form of a segment of a cylinder.
q(x) dx + dV = 0 (3.56)
The deflection of a plate can be determined by
requiring it to be in equilibrium under the action of or
all the forces and torques exerted on it. The forces and
dV
torques on a small section of the plate between hori- = −q. (3.57)
zontal locations x and x+dx are shown in Figure 3.10. dx
A downward force per unit area q(x) is exerted on the The moments M and M + dM combine to give a
plate by whatever distributed load the plate is required net counterclockwise torque dM on the element. The
to support. Thus, the downward load, per unit length forces V and V +dV are separated by a distance dx (an
in the z direction, between x and x + dx is q(x) dx. A infinitesimal moment arm) and exert a net torque Vdx
net shear force V , per unit length in the z direction, on the element in a clockwise sense. (The change in V
acts on the cross section of the plate normal to the in going from x to x + dx can be ignored in calculating
plane of the figure; it is the resultant of all the shear the moment due to the shear forces.) The horizontal
stresses integrated over that cross-sectional area of the forces P exert a net counterclockwise torque −P dw
plate. A horizontal force P, per unit length in the z on the element through their associated moment arm
direction, is applied to the plate; it is assumed that P −dw. (Note that dw is negative in going from x to
is independent of x. The net bending moment M, per x + dx.) A balance of all the torques gives
unit length in the z direction, acting on a cross section dM − P dw = V dx (3.58)
of the plate is the integrated effect of the moments
exerted by the normal stresses σxx , also known as the or
fiber stresses, on the cross section. We relate M to the dM dw
=V +P . (3.59)
fiber stresses in the plate later in the discussion. All dx dx

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 143 — #157

3.9 Two-Dimensional Bending or Flexure of Plates 143

The bending stress σxx is accompanied by longitudi-


nal strain εxx that is positive (contraction) in the upper
half of the plate and negative (extension) in the lower
half. There is no strain in the direction perpendicular
to the xy plane because the plate is infinite in this direc-
tion and the bending is two-dimensional or cylindrical;
that is, εzz = 0. There is also zero stress normal to
the surface of the plate; that is, σyy = 0. Because the
plate is thin, we can take σyy = 0 throughout. Thus
plate bending is an example of plane stress, and we can
Figure 3.11 The normal stresses on a cross section of a thin
curved elastic plate. use Equations (3.31) and (3.32) to relate the stresses
and strains; that is,

We can eliminate the shear force on a vertical cross 1


εxx = (σxx − νσzz ) (3.62)
section of the plate V from Equation (3.59) by differ- E
entiating the equation with respect to x and substitut- 1
εzz = (σzz − νσxx ). (3.63)
ing from Equation (3.57). One obtains E
d 2M d 2w In writing these equations, we have identified the prin-
2
= −q + P 2 . (3.60)
dx dx cipal strains ε1 , ε2 with εxx , εzz and the principal
stresses σ1 , σ2 with σxx , σzz . With εzz = 0, Equa-
Equation (3.60) can be converted into a differential
tions (3.62) and (3.63) give
equation for the deflection w if the bending moment
M can be related to the deflection; we will see that M E
σxx = εxx . (3.64)
is inversely proportional to the local radius of curvature (1 − ν 2 )
of the plate R and that R −1 is −d 2 w/dx2 .
Equation (3.61) for the bending moment can be rewrit-
To relate M to the curvature of the plate, we pro-
ten, using Equation (3.64), as
ceed as follows. If the plate is deflected downward, as  h/2
in Figure 3.11, the upper half of the plate is contracted, E
M= εxx y dy. (3.65)
and the longitudinal stress σxx is positive; the lower (1 − ν 2 ) −h/2
part of the plate is extended; and σxx is negative. The The longitudinal strain εxx depends on the dis-
fiber stress σxx is zero on the midplane y = 0, which tance from the midplane of the plate y and the local
is a neutral unstrained surface. The net effect of these radius of curvature of the plate R. Figure 3.12 shows
stresses is to exert a counterclockwise bending moment a bent section of the plate originally of length l
on the cross section of the plate. The curvature of the (l is infinitesimal). The length of the section measured
plate has, of course, been exaggerated in Figure 3.11 along the midplane remains l. The small angle φ is l/R
so that x is essentially horizontal. The force on an in radians. The geometry of Figure 3.12 shows that the
element of the plate’s cross section of thickness dy change in length of the section l at a distance y from
is σxx dy. This force exerts a torque about the mid- the midplane is
point of the plate given by σxx y dy. If we integrate this
l
torque over the cross section of the plate, we obtain the l = −yφ = −y , (3.66)
bending moment R
 h/2 where the minus sign is included because there is
M= σxx y dy, (3.61) contraction when y is positive. Thus the strain is
−h/2 l y
εxx = − = . (3.67)
where h is the thickness of the plate. l R

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 144 — #158

144 Elasticity and Flexure

change in α, that is, dα, in the small distance l or dx.


Thus
 
dα d dw d2w
φ = dα = dx = − dx = − 2 dx,
dx dx dx dx
(3.68)

and we find
1 φ φ d2w
= ≈ =− 2. (3.69)
R l dx dx
Finally, the strain is given by

d 2w
εxx = −y , (3.70)
Figure 3.12 Longitudinal extension and contraction at a distance dx2
y from the midplane of the plate. and the bending moment can be written

−E d 2 w h/2 2
M= y dy
(1 − ν 2 ) dx2 −h/2
 
−E d 2 w y3 h/2
=
(1 − ν 2 ) dx2 3 −h/2

−Eh3 d 2 w
= . (3.71)
12(1 − ν 2 ) dx2
The coefficient of −d 2 w/dx2 on the right side of
Equation (3.71) is called the flexural rigidity D of the
plate
Eh3
D≡ . (3.72)
12(1 − ν 2 )
According to Equations (3.69), (3.71), and (3.72), the
bending moment is the flexural rigidity of the plate
divided by its curvature
Figure 3.13 Sketch illustrating the geometrical relations in plate d 2w D
bending. M = −D = . (3.73)
dx2 R
Upon substituting the second derivative of Equa-
tion (3.73) into Equation (3.60), we obtain the general
Implicit in this relation is the assumption that plane equation for the deflection of the plate
sections of the plate remain plane.
The local radius of curvature R is determined by the d4w d 2w
D = q(x) − P . (3.74)
change in slope of the plate midplane with horizon- dx4 dx2
tal distance. The geometry is shown in Figure 3.13. If We next solve Equation (3.74) for plate deflection in
w is small, −dw/dx, the slope of the midplane, is also a number of simple cases and apply the results to the
the angular deflection of the plate from the horizon- deformation of crustal strata and to the bending of the
tal α. The small angle φ in Figure 3.13 is simply the lithosphere.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 145 — #159

3.10 Bending of Plates under Applied Moments and Vertical Loads 145

Figure 3.14 An embedded plate subject to an applied torque.


Figure 3.16 Bending of a plate pinned at both ends.

Figure 3.15 Force and torque balance on a section of the plate in


Figure 3.14.

3.10 Bending of Plates under Applied


Figure 3.17 An embedded plate subjected to a concentrated
Moments and Vertical Loads load.

Consider a plate embedded at one end and subject


to an applied torque Ma at the other, as shown in The bent plate has the shape of a parabola. w is nega-
Figure 3.14. Assume for simplicity that the plate is tive according to the convention we established if M is
weightless. With q = 0, Equation (3.57) shows that positive; that is, the plate is deflected upward.
the shear stress on a section of the plate V must be
a constant. In fact, V = 0, since there is no applied Problem 3.7
force acting on the plate. This can easily be seen by What is the displacement of a plate pinned at both
considering a force balance on a section of the plate, ends (w = 0 at x = 0, L) with equal and opposite
as shown in Figure 3.15. Since P = 0 and since we bending moments applied at the ends? The problem
have established V = 0, Equation (3.59) requires that is illustrated in Figure 3.16.
M = constant. The constant must be Ma , the applied
torque, as shown by a moment balance on an arbitrary As a second example we consider the bending of
section of the plate (Figure 3.15). a plate embedded at its left end and subjected to a
To determine the deflection of the plate, we could concentrated force Va at its right end, as illustrated in
integrate Equation (3.74) with q = P = 0. How- Figure 3.17. In this situation, q = 0, except at the point
ever, since we already know M ≡ Ma , it is simpler x = L, and Equation (3.57) gives V = constant. The
to integrate Equation (3.73), the twice integrated form constant must be Va , as shown by the vertical force
of the fourth-order differential equation. The bound- balance on the plate sketched in Figure 3.18. With P
ary conditions are w = 0 at x = 0 and dw/dx = 0 at also equal to zero, Equation (3.59) for the bending
x = 0. These boundary conditions at the left end of the moment simplifies to
plate clarify what is meant by an embedded plate; the
dM
embedded end of the plate cannot be displaced, and = Va . (3.76)
its slope must be zero. The integral of Equation (3.73) dx
subject to these boundary conditions is
This equation can be integrated to yield
−Ma x2
w= . (3.75)
2D M = Va x + constant, (3.77)

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 146 — #160

146 Elasticity and Flexure

Figure 3.19 A uniformly loaded plate embedded at one end.


Figure 3.18 Forces and torques on a section of a plate loaded at
its right end by a force Va .
This equation may be integrated twice more subject to
the standard boundary conditions w = dw/dx = 0 at
and the constant can be evaluated by noting that there x = 0. One finds
is no applied torque at the end x = L; that is, M = 0  
V a x2 x
at x = L. Thus we obtain w= L− . (3.83)
2D 3
M = Va (x − L). (3.78)
Problem 3.8
The bending moment changes linearly from −Va L at Determine the displacement of a plate of length
the embedded end to zero at the free end. A simple L pinned at its ends with a concentrated load Va
torque balance on the section of the plate shown in applied at its center. This problem is illustrated in
Figure 3.18 leads to Equation (3.78), since M must Figure 3.9.
balance the torque of the applied force Va acting with
moment arm L − x. As a third and final example, we consider the bend-
The displacement can be determined by integrating ing of a plate embedded at one end and subjected to
Equation (3.74), which simplifies to a uniform loading q(x) = constant, as illustrated in
Figure 3.19. Equation (3.74), with P = 0, becomes
d4w
= 0, (3.79) d4w q
dx4 = . (3.84)
dx4 D
when q = P = 0. The integral of Equation (3.79) is
We need four boundary conditions to integrate Equa-
d3w tion (3.84). Two of them are the standard conditions
= constant. (3.80)
dx3 w = dw/dx = 0 at the left end x = 0. A third
The constant can be evaluated by differentiating Equa- boundary condition is the same as the one used in the
tion (3.73) with respect to x and substituting for previous example, namely, d 2 w/dx2 = 0 at x = L,
dM/dx from Equation (3.76). The result is because there is no external torque applied at the right
end of the plate – see Equation (3.73). The fourth
d 3w Va
=− . (3.81) boundary condition follows from Equation (3.59) with
dx3 D
P = 0. Because there is no applied concentrated load
A second-order differential equation for w can be at x = L, V must vanish there, as must dM/dx and
obtained by integrating Equation (3.81) and evalu- from Equation (3.73), d 3 w/dx3 . After some algebra,
ating the constant of integration with the boundary one finds the solution
condition d 2 w/dx2 = 0 at x = L. Alternatively, the  
same equation can be arrived at by substituting for M qx2 x2 Lx L2
w= − + . (3.85)
from Equation (3.78) into Equation (3.73) D 24 6 4
d2w Va The shear force at x = 0 is −D(d 3 w/dx3 )x=0 .
= − (x − L). (3.82)
dx2 D From Equation (3.85), this is qL, a result that

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 147 — #161

3.11 Buckling of a Plate under a Horizontal Load 147

Figure 3.20 Section of a uniformly loaded plate.

also follows from a consideration of the over- Figure 3.21 A freely supported plate loaded sinusoidally.
all vertical equilibrium of the plate because qL is
the total loading. The shear stress on the section
Problem 3.12
x = 0 is qL/h. The bending moment on the section
Determine the deflection of a plate of length L that
x = 0 is −D(d 2 w/dx2 )x=0 or −qL2 /2. The maxi-
max = σ
is embedded at x = 0 and has equal loads Va applied
mum bending or fiber stress, σxx xx at y = at x = L/2 and at x = L.
−h/2, is given, from Equations (3.85), (3.64), and
(3.70), by
Problem 3.13
max E h d 2w 6 d 2w 6M Find the deflection of a uniformly loaded beam
σxx = = D 2 =− 2 .
(1 − ν ) 2 dx
2 2 h2 dx h pinned at the ends, x = 0, L. Where is the maximum
(3.86) bending moment? What is the maximum bending
At x = 0, σxx
max is 3qL2 /h2 . The ratio of the shear stress stress?
to the maximum bending stress at x = 0 is h/3L, a
rather small quantity for a thin plate. It is implicit in Problem 3.14
the analysis of the bending of thin plates that shear A granite plate freely supported at its ends spans a
stresses in the plates are small compared with the gorge 20 m wide. How thick does the plate have to
bending stresses. be if granite fails in tension at 20 MPa? Assume ρ =
2700 kg m−3 .
Problem 3.9
Calculate V and M by carrying out force and torque
balances on the section of the uniformly loaded plate
Problem 3.15
Determine the deflection of a freely supported plate,
shown in Figure 3.20.
that is, a plate pinned at its ends, of length L and
flexural rigidity D subject to a sinusoidal load qa =
M Problem 3.10 q0 sin π x/L, as shown in Figure 3.21.
A granite plate with ρ = 2700 kg m−3 is embedded
at one end. If L = 10 m and h = 1/4 m, what is
the maximum bending stress and the shear stress at 3.11 Buckling of a Plate under a Horizontal
the base? A MATLAB solution to this problem is Load
provided in Appendix D.
When an elastic plate is subjected to a horizontal force
P, as shown in Figure 3.22a, the plate can buckle, as
Problem 3.11
illustrated in Figure 3.22b, if the applied force is suffi-
Determine the displacement of a plate that is embed-
ciently large. Fold trains in mountain belts are believed
ded at the end x = 0 and has a uniform loading q
to result from the warping of strata under horizontal
from x = L/2 to x = L.
compression. We will therefore consider the simplest

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 148 — #162

148 Elasticity and Flexure

But w must also vanish at x = L, which implies that if


c1 = 0, then
  
P 1/2
sin L = 0. (3.92)
D

Thus (P/D)1/2 L must be an integer multiple of π ,


 1/2
P
L = nπ n = 1, 2, 3, . . . (3.93)
Figure 3.22 Plate buckling under a horizontal force.
D

Solving this equation for P, we get


example of plate buckling under horizontal compres-
sion to determine the minimum force required for n2 π 2
P= D. (3.94)
buckling to occur and the form, that is, the wavelength, L2
of the resulting deflection. In a subsequent section we
Equation (3.94) defines a series of values of P for
will carry out a similar calculation to determine if the
which nonzero solutions for w exist. The smallest such
lithosphere can be expected to buckle under horizontal
value is for n = 1 when P is given by
tectonic compression.
We consider a plate pinned at both ends and sub- π2
P = Pc = D. (3.95)
jected to a horizontal force P, as shown in Figure 3.22. L2
The deflection of the plate is governed by Equa- This is the minimum buckling load for the plate. If P is
tion (3.74) with q = 0: smaller than this critical value, known as an eigenvalue,
d4w d 2w the plate will not deflect under the applied load; that
D + P = 0. (3.87) is, c1 = 0 or w = 0. When P has the value given by
dx4 dx2
Equation (3.95), the plate buckles or deflects under the
This can be integrated twice to give
horizontal load. At the onset of deflection the plate
d 2w assumes the shape of a half sine curve
D + Pw = c1 x + c2 . (3.88)
dx2  1/2
P
However, we require that w is zero at x = 0, L and w = c1 sin x
D
that d 2 w/dx2 = 0 at x = 0, L, since there are no
πx
applied torques at the ends. These boundary condi- = c1 sin . (3.96)
L
tions require that c1 = c2 = 0, and Equation (3.88)
reduces to The amplitude of the deflection cannot be determined
by the linear analysis carried out here. Nonlinear
d 2w
D + Pw = 0. (3.89) effects fix the magnitude of the deformation.
dx2 The application of plate flexure theory to fold trains
Equation (3.89) has the general solution in mountain belts requires somewhat more complex
   1/2 
P 1/2 P models than considered here. Although a number of
w = c1 sin x + c2 cos x , (3.90) effects must be incorporated to approximate reality
D D
more closely, one of the most important is the influ-
where c1 and c2 are constants of integration. Because ence of the medium surrounding a folded stratum.
w is equal to zero at x = 0, c2 must be zero, and The rocks above and below a folded layer exert forces
   on the layer that influence the form (wavelength) of
P 1/2
w = c1 sin x . (3.91) the folds and the critical horizontal force necessary to
D
initiate buckling.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 149 — #163

3.12 Deformation of Strata Overlying an Igneous Intrusion 149

3.12 Deformation of Strata Overlying an in excess of the lithostatic pressure ρgh:


Igneous Intrusion
q = −p + ρgh. (3.97)
A laccolith is a sill-like igneous intrusion in the form of
This problem is very similar to the one illustrated in
a round lens-shaped body much wider than it is thick.
Figure 3.19. In both cases the loading is uniform so
Laccoliths are formed by magma that is intruded
that Equation (3.84) is applicable. We take x = 0
along bedding planes of flat, layered rocks at pres-
at the center of the laccolith. The required boundary
sures so high that the magma raises the overburden
conditions are w = dw/dx = 0 at x = ±L/2. The solu-
and deforms it into a domelike shape. If the flow of
tion of Equation (3.84) that satisfies these boundary
magma is along a crack, a two-dimensional laccolith
conditions is obtained after some algebra in the form
can be formed. Our analysis is restricted to this case.
A photograph of a laccolithic mountain is given in  
(p − ρgh) 4 L2 x2 L4
Figure 3.23 along with a sketch of our model. w=− x − + . (3.98)
24D 2 16
The overburden or elastic plate of thickness h is bent
upward by the pressure p of the magma that will form Note that because of the symmetry of the problem the
the laccolith upon solidification. The loading of the coefficients of x and x3 must be zero. The maximum
plate q(x) is the part of the upward pressure force p deflection at the center of the laccolith, x = 0, is

(a)

Figure 3.23 (a) A laccolith in Red and White Mountain, Colorado. The overlying sedimentary rocks have been eroded (University of
Colorado, Boulder). (b) A two-dimensional model for a laccolith.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 150 — #164

150 Elasticity and Flexure

(p − ρgh)L4 3.13 Application to the Earth’s Lithosphere


w0 = − . (3.99)
384D
In terms of its maximum value, the deflection is given When applying Equation (3.74) to determine the
by downward deflection of the Earth’s lithosphere due
  to an applied load, we must be careful to include
x2 x4
w = w0 1 − 8 2 + 16 4 . (3.100) in q(x) the hydrostatic restoring force caused by the
L L effective replacement of mantle rocks in a vertical
Problem 3.16 column by material of smaller density. In the case
Show that the cross-sectional area of a two- of the oceanic lithosphere, water fills in “the space
dimensional laccolith is given by (p − ρgh)L5 /720D. vacated” by mantle rocks moved out of the way by the
deflected lithosphere. In the case of the continental
lithosphere, the rocks of the thick continental crust
Problem 3.17 serve as the fill. Figure 3.24a illustrates the oceanic
Determine the bending moment in the overburden case. The upper part of the figure shows a lithospheric
above the idealized two-dimensional laccolith as a plate of thickness h and density ρm floating on a
function of x. Where is M a maximum? What is the “fluid” mantle also of density ρm . Water of density
value of Mmax ? ρw and thickness hw overlies the oceanic lithosphere.
Suppose that an applied load deflects the lithosphere
Problem 3.18 downward a distance w and that water fills in the
Calculate the fiber stress in the stratum overlying the space above the plate, as shown in the bottom part of
two-dimensional laccolith as a function of y (dis- Figure 3.24a. The weight per unit area of a vertical
tance from the centerline of the layer) and x. If column extending from the base of the deflected
dikes tend to form where tension is greatest in the lithosphere to the surface is
base of the stratum forming the roof of a laccol- ρw g(hw + w) + ρm gh.
ith, where would you expect dikes to occur for the The pressure at a depth hw + h + w in the surrounding
two-dimensional laccolith? mantle where there is no plate deflection is

Figure 3.24 Models for calculating the hydrostatic restoring force on lithospheric plates deflected by an applied load qa . (a) Oceanic case.
(b) Continental case.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 151 — #165

3.14 Periodic Loading 151

ρw ghw + ρm g(h + w). ρc ghc + ρm g(h + w) − {ρc g(hc + w) + ρm gh}


Thus there is an upward hydrostatic force per unit area = (ρm − ρc )gw. (3.104)
equal to The restoring force is equivalent to the force that
ρw ghw + ρm g(h + w) − {ρw g(hw + w) + ρm gh} results from replacing mantle rock by crustal rock in a
layer of thickness w. The net force per unit area acting
= (ρm − ρw )gw (3.101) on the elastic continental lithosphere is therefore
tending to restore the deflected lithosphere to its orig- q = qa − (ρm − ρc )gw. (3.105)
inal configuration. The hydrostatic restoring force per
unit area is equivalent to the force that results from Equation (3.74) for the deflection of the plate becomes
replacing mantle rock of thickness w and density ρm
d4w d 2w
by water of thickness w and density ρw . The net D + P + (ρm − ρc )gw = qa (x). (3.106)
dx4 dx2
force per unit area acting on the lithospheric plate is
therefore We are now in a position to determine the elastic
deflection of the lithosphere and the accompanying
q = qa − (ρm − ρw )gw, (3.102) internal stresses (shear and bending) for different
where qa is the applied load at the upper surface of the loading situations.
lithosphere. Equation (3.74) for the deflection of the
elastic oceanic lithosphere becomes 3.14 Periodic Loading
d4w d 2w How does the positive load of a mountain or the nega-
D + P + (ρm − ρw )gw = qa (x). (3.103)
dx4 dx2 tive load of a valley deflect the lithosphere? To answer
Figure 3.24b illustrates the continental case. The this question, we determine the response of the litho-
upper part of the figure shows the continental crust sphere to a periodic load. We assume that the elevation
of thickness hc and density ρc separated by the Moho of the topography is given by
from the rest of the lithosphere of thickness h and x
density ρm . The entire continental lithosphere lies on h = h0 sin 2π , (3.107)
λ
top of a fluid mantle of density ρm . The lower part where h is the topographic height and λ is its wave-
of Figure 3.24b shows the plate deflected downward a length. Positive h corresponds to ridges and negative
distance w by an applied load such as excess topogra- h to valleys. Since the amplitude of the topography
phy. The Moho, being a part of the lithosphere, is also is small compared with the thickness of the elastic
deflected downward a distance w. The space vacated lithosphere, the influence of the topography on this
by the deflected lithosphere is filled in by crustal rocks. thickness can be neglected. The load on the litho-
The crust beneath the load is effectively thickened by sphere corresponding to the topography given by
the amount w by which the Moho is depressed. The Equation (3.107) is
weight per unit area of a vertical column extending
x
from the base of the deflected plate to the surface is qa (x) = ρc gh0 sin 2π (3.108)
λ
ρc g(hc + w) + ρm gh. where ρc is the density of the crustal rocks associated
The pressure at a depth hc + h + w in the surrounding with the height variation. The equation for the deflec-
mantle far from the deflected plate is tion of the lithosphere is obtained by substituting this
expression for qa (x) into Equation (3.106) and setting
ρc ghc + ρm g(h + w). P = 0 to obtain
The difference between these two quantities is the d4w x
D + (ρm − ρc )gw = ρc gh0 sin 2π . (3.109)
upward hydrostatic restoring force per unit area dx4 λ

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 152 — #166

152 Elasticity and Flexure

Because the loading is periodic in x, the response or


deflection of the lithosphere will also vary sinusoidally
in x with the same wavelength as the topography. Thus
we assume a solution of the form
x
w = w0 sin 2π . (3.110)
λ
By substituting Equation (3.110) into Equation
(3.109), we determine the amplitude of the deflection
of the lithosphere to be
h0
w0 =   . (3.111)
ρm D 2π 4
−1+
ρc ρc g λ
Figure 3.25 Deflection of the lithosphere under a periodic load.
The quantity (D/ρc g)1/4 has the dimensions of a (a) Short-wavelength loading with no deflection of the lithosphere.
length. It is proportional to the natural wavelength for (b) Long-wavelength loading with isostatic deflection of the
the flexure of the lithosphere. lithosphere.
If the wavelength of the topography is sufficiently
short, that is, if
 
D 1/4
λ 2π , (3.112)
ρc g
then the denominator of Equation (3.111) is much
larger than unity, and
w0 h0 . (3.113)

Short-wavelength topography causes virtually no


deformation of the lithosphere. The lithosphere is
infinitely rigid for loads of this scale. This case is
Figure 3.26 Dependence of the degree of compensation on the
illustrated in Figure 3.25a. If the wavelength of the
nondimensional wavelength of periodic topography.
topography is sufficiently long, that is, if
 
D 1/4 w0
λ 2π , (3.114) C= . (3.116)
ρc g w0∞
then Equation (3.111) gives Upon substituting Equations (3.111) and (3.115) into
the equation for C, we obtain
ρc h0
w = w0∞ = . (3.115) (ρm − ρc )
(ρm − ρc ) C=   . (3.117)
D 2π 4
This is the isostatic result obtained in Equation (2.3). ρm − ρc +
For topography of sufficiently long wavelength, the g λ
lithosphere has no rigidity and the topography is fully This dependence is illustrated in Figure 3.26. For a
compensated; that is, it is in hydrostatic equilibrium. lithosphere with elastic thickness 25 km, E = 70 GPa,
The degree of compensation C of the topographic ν = 0.25, ρm = 3300 kg m−3 , and ρc = 2800 kg m−3
load is the ratio of the deflection of the lithosphere to we find that topography is 50% compensated (C = 0.5)
its maximum or hydrostatic deflection if its wavelength is λ = 420 km. Topography with a

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 153 — #167

3.15 Stability of the Earth’s Lithosphere Under an End Load 153

shorter wavelength is substantially supported by the Pc is the minimum value for P for which the ini-
rigidity of the lithosphere; topography with a longer tially horizontal lithosphere will become unstable and
wavelength is only weakly supported. acquire the sinusoidal shape. If P < Pc , the horizontal
lithosphere is stable and will not buckle under the end
load.
3.15 Stability of the Earth’s Lithosphere The eigenvalue Pc can also be written
under an End Load  3 
Eh (ρm − ρw )g 1/2
Pc = = σc h, (3.122)
We have already seen how a plate pinned at its ends can 3(1 − ν 2 )
buckle if an applied horizontal load exceeds the criti- where σc is the critical stress associated with the force
cal value given by Equation (3.95). Let us investigate Pc . Solving Equation (3.122) for the critical stress we
the stability of the lithosphere when it is subjected to find
a horizontal force P. We will see that when P exceeds  
Eh(ρm − ρw )g 1/2
a critical value, an infinitely long plate (L → ∞) will σc = . (3.123)
3(1 − ν 2 )
become unstable and deflect into the sinusoidal shape
shown in Figure 3.27. The wavelength of the instability that occurs when P =
The equation for the deflection of the plate is Pc is given by Equation (3.120):
   1/4
obtained by setting qa = 0 in Equation (3.103): 2D 1/2 D
λc = 2π = 2π
Pc g(ρm − ρw )
d4w d 2w  1/4
D + P + (ρm − ρw )gw = 0. (3.118) Eh 3
dx4 dx2 = 2π . (3.124)
12(1 − ν 2 )(ρm − ρw )g
This equation can be satisfied by a sinusoidal deflec-
tion of the plate as given in Equation (3.110) if We wish to determine whether buckling of the litho-
 4  2 sphere can lead to the formation of a series of synclines
2π 2π and anticlines.
D −P + (ρm − ρw )g = 0, (3.119)
λ λ We consider an elastic lithosphere with a thickness
of 50 km. Taking E = 100 GPa, ν = 0.25, ρm =
a result of directly substituting Equation (3.110) into
3300 kg m−3 , and ρw = 1000 kg m−3 , we find from
Equation (3.118). Equation (3.119) is a quadratic
Equation (3.123) that σc = 6.4 GPa. A 50-km-thick
equation for the square of the wavelength of the sinu-
elastic lithosphere can support a horizontal compres-
soid λ. Its solution is
 2 sive stress of 6.4 GPa without buckling. Because of the
2π P ± [P2 − 4(ρm − ρw )gD]1/2 very large stress required, we conclude that such buck-
= . (3.120)
λ 2D ling does not occur. The lithosphere fails, presumably
Because the wavelength of the deformed lithosphere by the development of a fault, before buckling can take
must be real, there can only be a solution if P exceeds place. In general, horizontal forces have a small influ-
the critical value ence on the bending behavior of the lithosphere. For
this reason we neglect them in the lithosphere bending
Pc = {4Dg(ρm − ρw )}1/2 . (3.121) studies to follow.

Figure 3.27 Buckling of an infinitely long plate under an applied horizontal load with a hydrostatic restoring force.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 154 — #168

154 Elasticity and Flexure

Horizontal forces are generally inadequate to buckle The general solution of Equation (3.125) is
the lithosphere because of its large elastic thickness.  
x/α x x
However, the same conclusion may not apply to much w=e c1 cos + c2 sin
α α
thinner elastic layers, such as elastic sedimentary strata  
x x
embedded between strata that behave as fluids and + e−x/α c3 cos + c4 sin , (3.126)
highly thinned lithosphere in regions of high heat α α
flow. To evaluate the influence of horizontal forces on where the constants c1 , c2 , c3 , and c4 are determined
the bending of such thin layers, we take h = 1 km by the boundary conditions and
and the other parameters as before and find from  1/4
Equation (3.123) that σc = 900 MPa. From Equa- 4D
α= . (3.127)
tion (3.124), we obtain λc = 28 km. We conclude that (ρm − ρw )g
the buckling of thin elastic layers may contribute to the The parameter α is known as the flexural parameter.
formation of folded structures in the Earth’s crust. Because there is symmetry about x = 0, we need
only determine w for x ≥ 0. We require that w → 0 as
x → ∞ and that dw/dx = 0 at x = 0. Clearly, c1 and
3.16 Bending of the Elastic Lithosphere
c2 must be zero and c3 = c4 . Equation (3.126) becomes
under the Loads of Island Chains  
−x/α x x
w = c3 e cos + sin x ≥ 0. (3.128)
Volcanic islands provide loads that cause the litho- α α
sphere to bend. The Hawaiian ridge is a line of vol- The constant c3 is proportional to the magnitude of
canic islands and seamounts that extends thousands the applied line load V0 . From Equation (3.81), we
of kilometers across the Pacific. These volcanic rocks have
provide a linear load that has a width of about 150 km
1 d3w 4Dc3
and an average amplitude of about 100 MPa. The V0 = D 3 (x = 0) = . (3.129)
bathymetric profile across the Hawaiian archipelago 2 dx α3
shown in Figure 3.28 reveals a depression, the Hawai- (Half the plate supports half the load applied at
ian Deep, immediately adjacent to the ridge and an x = 0. Note also that a downward force on the left
outer peripheral bulge or upwarp. end of the plate is negative according to the sign con-
To model the deflection of the lithosphere under lin- vention illustrated in Figure 3.10.) Substituting for
ear loading, let us consider the behavior of a plate c3 from Equation (3.129) into Equation (3.128), we
under a line load V0 applied at x = 0, as shown in obtain
 
Figure 3.29. Since the applied load is zero except at V0 α 3 −x/α x x
x = 0, we take qa (x) = 0 and P = 0 in Equation w= e cos + sin x ≥ 0. (3.130)
8D α α
(3.103) and solve
The maximum amplitude of the deflection at x = 0 is
d4w given by
D 4 + (ρm − ρw )gw = 0. (3.125)
dx V0 α 3
w0 = . (3.131)
8D

Figure 3.28 A bathymetric profile across the Hawaiian


archipelago. Figure 3.29 Deflection of the elastic lithosphere under a line load.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 155 — #169

3.16 Bending of the Elastic Lithosphere under the Loads of Island Chains 155

Figure 3.31 Deflection of a broken elastic lithosphere under a line


load.

Figure 3.30 Half of the theoretical deflection profile for a floating of the elastic lithosphere if we assume that it is equal
elastic plate supporting a line load. to xb . A representative value of xb for the Hawaiian
archipelago is 250 km; with xb = 250 km, Equa-
tion (3.135) gives a flexural parameter α = 80 km. For
In terms of w0 , the deflection of the plate is
  ρm − ρw = 2300 kg m−3 and g = 10 m s−2 Equa-
x x tion (3.127) gives D = 2.4 × 1023 N m. Taking E = 70
w = w0 e−x/α cos + sin . (3.132)
α α GPa and ν = 0.25, we find from Equation (3.72) that
the thickness of the elastic lithosphere is h = 34 km.
This profile is given in Figure 3.30.
The deflection of the lithosphere under a line load is Problem 3.19
characterized by a well-defined arch or forebulge. The (a) Consider a lithospheric plate under a line load.
half-width of the depression, x0 , is given by Show that the absolute value of the bending
3π moment is a maximum at
x0 = α tan−1 (−1) = α. (3.133) π
4 xm = α cos−1 0 = α (3.137)
The distance from the line load to the maximum 2
amplitude of the forebulge, xb , is obtained by deter- and that its value is
mining where the slope of the profile is zero. Upon 2Dw0 Dw0
Mm = − 2 e−π/2 = −0.416 2 . (3.138)
differentiating Equation (3.132) and setting the result α α
to zero (b) Refraction studes show that the Moho is
dw 2w0 −x/α x depressed about 10 km beneath the center of
=− e sin = 0, (3.134)
dx α α the Hawaiian Islands. Assuming that this is the
we find value of w0 and that h = 34 km, E = 70
GPa, ν = 0.25, ρm − ρw = 2300 kg m−3 , and
xb = α sin−1 0 = π α. (3.135)
g = 10 m s−2 , determine the maximum bending
The height of the forebulge wb is obtained by substi- stress in the lithosphere.
tuting this value of xb into Equation (3.132):
Since volcanism along the Hawaiian ridge has weak-
wb = −w0 e−π = −0.0432w0 . (3.136)
ened the lithosphere, it may not be able to sustain large
The amplitude of the forebulge is quite small com- bending moments beneath the load. In this case, we
pared with the depression of the lithosphere under the should consider a model in which the lithosphere is
line load. fractured along the line of the ridge. Let us accord-
This analysis for the line load is only approximately ingly determine the deflection of a semi-infinite plate
valid for the Hawaiian Islands, since the island load floating on a fluid half-space and subjected to a line
is distributed over a width of about 150 km. How- load V0 /2 at its end, as sketched in Figure 3.31. The
ever, the distance from the center of the load to the deflection is given by Equation (3.126), with the con-
crest of the arch can be used to estimate the thickness stants of integration yet to be determined. Since the

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 156 — #170

156 Elasticity and Flexure

plate extends from x = 0 to x = ∞ and we require


w → 0 as x → ∞, c1 and c2 must again be zero. We
have assumed that no external torque is applied to the
end x = 0. From Equation (3.73), we can conclude
that d 2 w/dx2 = 0 at x = 0. This boundary condition
requires that c4 = 0. Finally, by equating the shear on
the end x = 0 to the applied line load, we find
1 d3w 2Dc3
V0 = D 3 (x = 0) = . (3.139)
2 dx α3 Figure 3.32 The deflection of the elastic lithosphere under an end
With the value of c3 from Equation (3.139) and c1 = load.
c2 = c4 = 0, Equation (3.126) gives
D = 7.26 × 1023 Nm and h = 49 km. The thickness of
V0 α 3 −x/α x
w= e cos . (3.140) a broken lithosphere turns out to be about 50% greater
4D α
than the thickness of an unbroken lithosphere.
The maximum amplitude of the deflection at
x = 0 is Problem 3.20
V0 α 3 (a) Consider a lithospheric plate under an end load.
w0 = . (3.141)
4D Show that the absolute value of the bending
For the same load, the deflection amplitude of a bro- moment is a maximum at
ken lithosphere is twice as great as it is for a lithosphere π
xm = α tan−1 1 = α, (3.146)
without a break. By substituting Equation (3.141) into 4
Equation (3.140), we can write and that its value is
x
w = w0 e−x/α cos . (3.142) Mm = −
2Dw0 −π/4
e
π Dw0
sin = −0.644 2 .
α α2 4 α
This profile is given in Figure 3.32. (3.147)
The half-width of the depression and the position (b) Refraction studies show that the Moho is
and amplitude of the forebulge are given by depressed about 10 km beneath the center of
π the Hawaiian Islands. Assuming that this is the
x0 = α (3.143)
2 value of w0 and that h = 49 km, E = 70 GPa,
3π ν = 0.25, ρm − ρw = 2300 kg m−3 , and g = 10
xb = α (3.144)
4 m s−2 , determine the maximum bending stress in
3π the lithosphere.
wb = w0 e−3π/4 cos = −0.0670w0 . (3.145)
4
Bending of the lithosphere under a triangular load
The amplitude of the forebulge for the broken litho-
and under axisymmetric loads is discussed in Sections
sphere model, although still small compared with the
12.1 and 12.2, respectively.
deflection of the lithosphere under the load, is con-
siderably larger than the forebulge amplitude of an
unbroken lithosphere supporting the same load. 3.17 Bending of the Elastic Lithosphere at an
We again evaluate the model results for the deflec- Ocean Trench
tion of the lithosphere caused by the Hawaiian Islands.
With xb = 250 km, we find from Equation (3.144) that Another example of the flexure of the oceanic elas-
α = 106 km. This result, together with ρm −ρw = 2300 tic lithosphere is to be found at ocean trenches. Prior
kg m−3 , g = 10 m s−2 , E = 70 GPa, and ν = 0.25, gives to subduction, considerable bending of the elastic

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 157 — #171

3.17 Bending of the Elastic Lithosphere at an Ocean Trench 157

directly. Quantities that can be measured directly are


the height of the forebulge wb and the half-width of
the forebulge xb – x0 , as illustrated in Figure 3.33. We
therefore express the trench profile in terms of these
parameters. We can determine x0 by setting w = 0:

x0 αV0
tan =1+ . (3.152)
Figure 3.33 Bending of the lithosphere at an ocean trench due to α M0
an applied vertical load and bending moment.
Similarly, we can determine xb by setting dw/dx = 0:
lithosphere occurs. The bent lithosphere defines the
oceanward side of the trench. To model this behavior, xb 2M0
tan = −1 − . (3.153)
we will consider an elastic plate acted upon by an end α αV0
load V0 and a bending moment M0 , as illustrated in
Figure 3.33. The height of the forebulge is obtained by substituting
The deflection of the plate is governed by Equa- this value of xb into Equation (3.151):
tion (3.125), and once again the general solution is
 
given by Equation (3.126). We require w → 0 as α 2 −xb /α xb xb
wb = e −M0 sin + (M0 + V0 α) cos .
x → ∞ so that c1 = c2 = 0 and 2D α α
  (3.154)
x x
w = e−x/α c3 cos + c4 sin . (3.148)
α α
From Equations (3.152) and (3.154), we find
At x = 0, the bending moment is −M0 ; from Equa-
tion (3.73) we obtain
  xb x0 
xb − x0 sin −
−M0 α 2 α α
c4 = . (3.149) tan
α
= xb x0 
2D cos −
α α
Also, at x = 0, the shear force is −V0 ; from Equa-
xb x0 xb x0
tions (3.59) and (3.73) we find sin cos + − cos sin
= α α α α
α2 xb x0 xb x0
c3 = (V0 α + M0 ). (3.150) cos cos + sin sin
2D α α α α
The equation for w can now be written as xb x0
tan − tan
= α α =1 (3.155)
α 2 e−x/α x x xb x0
w= −M0 sin + (V0 α + M0 ) cos . 1 + tan tan
2D α α α α
(3.151)
and
Equation (3.151) reduces to Equation (3.140) in the
case M0 = 0. Note that the line load here is V0 ; it π
xb − x0 = α. (3.156)
was V0 /2 in Equation (3.140). 4
The elastic deflection of the oceanic lithosphere in
terms of the vertical force and bending moment at the This half-width is a direct measure of the flexural
ocean trench axis is given by Equation (3.151). The parameter and, therefore, of the flexural rigidity and
vertical force and bending moment are the result of thickness of the elastic lithosphere.
the gravitational body force acting on the descending By using Equation (3.152), we can rewrite Equa-
plate. Unfortunately, V0 and M0 cannot be determined tion (3.151) for the deflection of the lithosphere as

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 158 — #172

158 Elasticity and Flexure

 
α 2 M0 −x/α x x0 x The dependence of M(xb − x0 )2 /Dwb on (x − x0 )/
w= e −sin + tan cos
2D α α α (xb − x0 ) is shown in Figure 3.34b. The bending
moment is a maximum at (x−x0 )/(xb −x0 ) = −1. The
α 2 M0 −[(x−x0 )/α]−x0 /α shear force can be determined from Equations (3.59)
= e
2D and (3.160) to be
⎧ ⎫ √ 3 π/4
⎪ x0 x x0 x⎪ 
⎨ sin cos − cos sin ⎬ 2π e Dwb π(x − x0 )
× α α α α V =− cos
⎪ x0 ⎪ 32 (xb − x0 )3 4(xb − x0 )
⎩ cos ⎭
α   
  π(x − x0 ) π(x − x0 )
x − x0 + sin exp − .
sin 4(xb − x0 ) 4(xb − x0 )
α 2 M0 −[(x−x0 )/α] −x0 /α α
=−
2D
e e x0  . (3.161)
cos
α The dimensionless shear force V (xb − x0 )3 /Dwb is
(3.157) plotted vs. (x−x0 )/(xb −x0 ) in Figure 3.34c. The shear
The height of the forebulge is thus given by force is zero at (x − x0 )/(xb − x0 ) = −1.
  The universal flexure profile is compared with an
xb − x0 observed bathymetric profile across the Mariana trench
sin
α 2 M0 −[(xb −x0 )/α] −x0 /α α in Figure 3.35. In making the comparison, we take
wb = − e e x  .
2D cos
0
xb = 55 km and wb = 500 m (x0 = 0). From Equa-
α tion (3.156), we find that α = 70 km. With ρm − ρw =
(3.158)
2300 kg m−3 and g = 10 m s−2 , Equation (3.127)
Upon dividing Equation (3.157) by Equation (3.158) gives D = 1.4 × 1023 N m. From Equation (3.72) with
and eliminating α using Equation (3.156), we E = 70 GPa and ν = 0.25, we find that the thickness
obtain of the elastic lithosphere is 28 km. This value is in quite
      good agreement with the thickness of the oceanic elastic
π x − x0 x − x0
exp − sin π4 lithosphere obtained by considering island loads. The
w 4 xb − x 0 xb − x0
wb
= π sin π largest bending stress is 900 MPa, and it occurs 20 km
exp − 4
seaward of the trench axis. This is a very large deviatoric
4
      stress, and it is doubtful that the near-surface rocks have
√ π/4 π x − x0 π x − x0
= 2e exp − sin . sufficient strength in tension. However, the yield stress
4 xb − x 0 4 xb − x 0
of the mantle is likely to approach this value at depth
(3.159) where the lithostatic pressure is high.
Although the trench bathymetric profile given in
The plot of w/wb vs. (x − x0 )/(xb /x0 ) shown in
Figure 3.35 appears to exhibit elastic flexure, other
Figure 3.34a defines a universal flexure profile. The pro-
trench profiles exhibit an excessively large curvature
file is valid for any two-dimensional elastic flexure of
near the point of the predicted maximum bending
the lithosphere under end loading.
moment. This is discussed in Chapter 7, where we
We can solve for the bending moment in terms of
associate this excess curvature with the plastic failure
(x − x0 )/(xb − x0 ) by substituting Equation (3.159)
of the lithosphere.
into Equation (3.73)
√ 2 π/4  
2π e Dwb π(x − x0 ) 3.18 Flexure and the Structure of
M= cos
8 (xb − x0 )2 4(xb − x0 ) Sedimentary Basins
 
π(x − x0 ) Lithospheric flexure is also associated with the struc-
× exp − . (3.160)
4(xb − x0 ) ture of many sedimentary basins. A sedimentary

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 159 — #173

3.18 Flexure and the Structure of Sedimentary Basins 159

Figure 3.34 Universal solution for the deflection of an elastic lithosphere under a vertical end load and bending moment. (a) Dependence
of the nondimensional displacement w/wb on the nondimensional position (x − x0 )/(xb − x0 ). The profile is also shown at an
amplification of 10 to 1 to more clearly show the structure of the forebulge. (b) The dimensionless bending moment versus
(x − x0 )/(xb − x0 ). (c) The dimensionless vertical shear force as a function (x − x0 )/(xb − x0 ).

Figure 3.35 Comparison of a bathymetric profile across the Mariana trench (solid line) with the universal lithospheric deflection profile
given by Equation (3.159) (dashed line); xb = 55 km and wb = 0.5 km.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 160 — #174

160 Elasticity and Flexure

basin is a region where the Earth’s surface has been


depressed and the resulting depression has been filled
by sediments. Typical sedimentary basins have depths
up to 5 km, although some are as deep as 15 km.
Because sedimentary basins contain reservoirs of
petroleum, their structures have been studied in detail
using seismic reflection profiling and well logs.
Some sedimentary basins are bounded by near-
vertical faults along which the subsidence has
occurred. Others, however, have a smooth basement,
and the subsidence is associated with the flexure of
the elastic lithosphere. The horizontal dimensions of
these sedimentary basins, about 400 to 1000 km, reflect
the magnitude of the flexural parameter based on sed-
iments of density ρs replacing mantle rock of density
ρm , α = [4D/(ρm − ρs )g]1/4 .
Some sedimentary basins have a nearly two-
dimensional structure. They are caused by the loading
of a linear mountain belt and are known as foreland
basins. Examples are the series of sedimentary basins
lying east of the Andes in South America and the
Appalachian basin in the eastern United States lying
west of the Appalachian Mountains. Depth contours
of the basement beneath the Appalachian basin are
given in Figure 3.36a. A basement profile is shown in
Figure 3.36b. The depth w is the depth below sea level,
and the coordinate −x is measured from the point
where basement rocks are exposed at the surface.
Figure 3.36 (a) Contours of basement (in km) in the Appalachian
It is appropriate to model the structure of the
basin of the eastern United States. Data are from well logs and
Appalachian basin as a two-dimensional lithospheric seismic reflection studies. (b) The data points are the depths of
plate under a linear end load. Thus the universal basement below sea level as a function of the distance from the
flexure profile given in Equation (3.159) is directly point where basement rocks are exposed at the surface along the
applicable. In order to fit the basement profile given profile given by the heavy line in (a). The solid line is the universal
in Figure 3.36b we take xb = 122 km and wb = 290 m flexure profile from Equation (3.159) with xb = 122 km and
(x0 = 0). Since the forebulge has been destroyed by wb = 290 m (x0 = 0).
erosion this choice of parameters is not unique. They
that the thickness of the elastic continental lithosphere
can be varied somewhat, and a reasonable fit can still
is h = 54 km. This is somewhat larger than the val-
be obtained. However, these values are near the cen-
ues we obtained for the thickness of the elastic oceanic
ter of the acceptable range. From Equation (3.156),
lithosphere. Flexure studies of other sedimentary
we find that they correspond to α = 155 km. As we
basins give similar values of elastic thickness.
have already noted, the flexural rigidity must be based
on the density difference between the mantle and the
Problem 3.21
sediments ρm − ρs . With ρm − ρs = 700 kg m−3 and
An ocean basin has a depth of 5.5 km. If it is filled
g = 10 m s−2 we find D = 1024 N m. From Equa-
to sea level with sediments of density 2600 kg m−3 ,
tion (3.72) with E = 70 GPa and ν = 0.25, we find

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 161 — #175

Further Reading 161

what is the maximum depth of the resulting sedimen- at its center and that the elastic lithosphere is not
tary basin? Assume ρm = 3300 kg m−3 . broken, determine the corresponding thickness of
the elastic lithosphere. Assume E = 70 GPa,
ν = 0.25, and ρm − ρs = 700 kg m−3 . A
M Problem 3.22
MATLAB solution to this problem is provided in
The Amazon River basin in Brazil has a width of 400
Appendix D.
km. Assuming that the basin is caused by a line load

Summary
Deformation processes are essential to studies of geodynamics. In this chapter we have con-
sidered linear elasticity. The deformation of a linear elastic solid (strain) is proportional to the
applied force (stress). Elastic deformation is reversible; when an applied force is removed, the
associated deformation returns to zero. Many geodynamic processes result in permanent defor-
mation. When large forces (stresses) are applied to an elastic material, it may fracture or deform
ductilely (plastically). These processes will be considered in later chapters.
We have shown that the morphology of some ocean trenches can be explained by the elastic
bending of the lithosphere treated as a thin plate. At other ocean trenches it will be necessary to
include plastic deformation. We have shown that elastic plate bending can explain the morphol-
ogy of sedimentary basins. The near circular structure of the Michigan Basin is an example. An
essential distinction in geodynamics is the difference in thickness of the “elastic” lithosphere rel-
ative to the “thermal” lithosphere. The thickness of the elastic lithosphere is the thickness of the
thin elastic plate obtained from flexure applications. The formal definition of the thermal litho-
sphere will be given in Chapter 4. Stress relaxation of elastic stresses in the hotter, lower thermal
lithosphere is responsible for the thinner elastic lithosphere. This difference will be quantified in
Chapter 6.

FURTHER READING
Eringen, A. C., Mechanics of Continua (John Wiley, New viscoelasticity, plasticity, and finite deformation theory.
York, NY, 1967), 502 pages. The book begins with an introductory chapter on elas-
A comprehensive treatment of the mechanics of con- tic and viscoelastic behavior. Cartesian tensors are then
tinua at a relatively sophisticated level. The basic introduced and used in the discussions of stress, strain,
concepts of strain, stress, flow, thermodynamics, and and the conservation laws. Subsequent chapters deal with
constitutive equations are introduced. Applications are linear elasticity, solutions of elastic problems by poten-
made to elasticity, fluid dynamics, thermoplasticity, and tials, two-dimensional problems, energy theorems, Saint-
viscoelasticity. Venant’s principle, Hamilton’s principle, wave propagation,
Fung, Y. C., Foundations of Solid Mechanics (Prentice-Hall, elasticity and thermodynamics, thermoelasticity, viscoelas-
Englewood Cliffs, NJ, 1965), 525 pages. ticity, and finite strain theory. Problems for the student are
A graduate-level textbook on the mechanics of solids. included.
The text is mainly concerned with the classical theory Jaeger, J. C., Elasticity, Fracture, and Flow (Methuen,
of elasticity, thermodynamics of solids, thermoelasticity, London, UK, 1969), 3rd edn., 268 pages.

i i
i i

“9781107006539AR” — 2013/12/3 — 19:53 — page 162 — #176

162 Elasticity and Flexure

A monograph on the mathematical foundations of elastic- value problems of static elasticity. Part 2 treats planar
ity, plasticity, viscosity, and rheology. Chapter 1 develops problems whose solutions are obtained with the aid of
the analysis of stress and strain with emphasis on Mohr’s the stress function and its complex representation. The
representations. Chapter 2 discusses stress–strain relations technique of conformal mapping is introduced. Part 3 devel-
for elasticity, viscosity, and plasticity, and criteria for frac- ops the Fourier series approach to the solution of planar
ture and yield. Chapter 3 derives the equations of motion problems, while Parts 4 and 5 make use of Cauchy integrals.
and equilibrium. Chapters 4 and 5 deal with stresses in Part 6 presents solutions for special planar geometries and
the Earth’s crust, rock mechanics, and applications to Part 7 deals with the extension, torsion and bending of bars.
structural geology. Novozhilov, V. V., Thin Shell Theory (P. Noordhoff,
Jaeger, J. C. and N. G. W. Cook, Fundamentals of Rock Groningen, The Netherlands, 1964), 377 pages.
Mechanics (Chapman and Hall, London, UK, 1979), 3rd A mathematical analysis of stresses and strains in thin
edn., 593 pages. shells using linear elasticity theory. There are four chap-
See further reading list Chapter 2. ters on the general theory of thin elastic shells, the
Kraus, H., Thin Elastic Shells (John Wiley, New York, NY, membrane theory of shells, cylindrical shells, and shells of
1967), 476 pages. revolution.
An extensive mathematical treatment of the deformation Timoshenko, S. and J. N. Goodier, Theory of Elasticity
of thin elastic shells. It includes three chapters on the the- (McGraw-Hill, New York, NY, 1970), 567 pages.
ory of thin elastic shells, four chapters on static analysis, See further reading list for Chapter 2.
two chapters on dynamic analysis, and two chapters on Timoshenko, S. and D. H. Young, Elements of Strength of
numerical methods. Materials (Van Nostrand, Princeton, NJ, 1968), 5th edn.,
Muskhelishvili, N. I., Some Basic Problems of the Mathe- 377 pages.
matical Theory of Elasticity (P. Noordhoff, Groningen, The An undergraduate engineering textbook with an extensive
Netherlands, 1963), 718 pages. treatment of the bending of beams and elastic stability.
This treatise on the mathematical theory of elasticity is Problems with solutions are included.
divided into seven major parts. Part 1 deals with the fun- Ugural, A. C., Mechanics of Materials (McGraw-Hill, New
damental equations of the mechanics of an elastic body. It York, NY, 1991), 441 pages.
includes separate chapters on analyses of stress and strain, This book covers the fundamentals of elasticity. It includes
relation between stress and strain, the equilibrium equa- a comprehensive treatment of beams as well as discussions
tions of an elastic body, and the fundamental boundary of inelastic behavior. It is written at an introductory level.

i i

You might also like