You are on page 1of 19

5

Stress

In all quantitative studies on the We are all familiar with daily effects of forces and stresses. The most perva-
sive force affecting us is gravity: it holds the atmosphere, the oceans, and us
relationship between original forces
to the Earth, and keeps the Earth and other planets from fragmenting. The
and the resulting deformations, an Earth is a large enough body to be deformed into a nearly spherical shape
because of gravity, whereas small objects in space, like asteroids and smaller
intermediate field of investigation enters,
bodies, are affected by gravity, but their mass and the influence of gravity
the condition of stress in the earth’s crust. are not great enough to transform them into spherical shapes.
Forces change the velocity or direction of bodies in motion. We use
W. HAFNER, 1951, Geological Society forces to open and close doors, ride bicycles and exercise machines, turn
of America Bulletin on lamps, push a gas or brake pedal on a car, and perform other everyday
activities. In the Earth, forces act on rock bodies and drive certain kinds
of deformation, but force is not as meaningful as stress in the study of
rock deformation, because we must consider the area affected by the force.
Stress (σ) is a force applied to an area—force per unit area. Hafner’s chapter
opening statement above about the direct connection between stresses and
deformation can be demonstrated only for elastic deformation. There, a
direct proportion exists between the amount of stress and the amount of
deformation that results; elastic deformation (Chapter 6) is, by definition,
recoverable. Other forms of deformation, or strain, are not as easily related
to forces, or stresses, in the Earth (Figure 5–1).
We commonly think of tectonic structures as products of stress
(Figure 5–1), but ironically, we observe most of the effects of stress in tec-
tonic structures without being able to measure the stress that produced
the structure. A tectonic structure is a manifestation of deformation that
is assumed to result from stress on a rock mass. Stress originates in pro-
cesses that generate, move, and consume lithospheric plates, with the aid
of gravity. Gravity alone induces the stresses primarily responsible for de-
formation in salt and impact structures (Chapter 2). Thus, understanding
the nature of stress is important to understanding tectonic structures. One
of the goals of structural geology is to reconstruct the orientations and
magnitudes of stresses that produced the structures we encounter in the
field. ­Unfortunately, these stresses are usually no longer present, and thus
cannot be measured directly. Experiments that duplicate natural struc-
tures in rocks also help us to estimate the orientations and magnitudes of
stresses needed to produce deformation in rocks.

112
Stress | 113

FIGURE 5–1  The strongly


folded Mesoproterozoic Moine
metasandstone, pegmatite, and schist
exposed at Cluanie Lake in the Scottish
Highlands, like most other structures,
are assumed to be a product of stress in
the crust. (RDH photo.)

Definitions either be stationary or already in motion. Newton’s second


law of motion states that

Several terms are useful in describing forces and stresses F = ma, (5–1)
mathematically. A scalar possesses only magnitude; a
vector possesses both magnitude and direction (Figure 5–2). where F is force (a vector), m is mass (a scalar), and a is
A scalar is a number (for example, the price of oil, the score acceleration (also a vector). Body forces act equally on
in a baseball game, temperature, or the thickness of a rock all parts of a body. Examples are the effect of gravity or
unit). A vector is a number with an indication of d ­ irection. electromagnetic forces on a mass. Body forces must be
For example, if we say a car is traveling northwest at considered if the behavior of fluid or ductile material
100 km per hour, we have defined a vector; if we say only is involved, but do not depend on the rheology (state of
that it is traveling 100 km per hour—with no indication “flow”) of the material under consideration. Surface forces
of direction—we have defined a scalar. In this book, vectors act on external or internal surfaces within rock masses and
in equations are indicated by bold-italic type. include forces acting along a fault or a major plate boundary.
A force (Figure 5–3) is a vector that produces a change The magnitude of a gravitational body force is propor-
in the velocity or direction of motion of a body that may tional to the amount of mass present, but the magnitude of

FIGURE 5–2  Examples of


scalars and vectors.
Viscosity Force and stress
(on a surface)

Temperature Volume Temperature


Acceleration
gradient

Scalars Vectors
Earth's
Ocean
Speed Time gravity Earth’s currents
field magnetic
field
Mass Mantle
Length
Velocity convection
flow
114 |  Mechanics: How Rocks Deform

70 kg 70 kg Mariana
Trench Sea level
1m
1m

10 km
Force Force

(a)
Surface
Stress 1m
1m

Stress

10 km
FIGURE 5–3  Difference between forces and stresses. Two persons Crust
weighing 70 kg each stand on one foot on the edge of a box so
that all of the weight is concentrated on the heel of the shoe. The
person to the left is wearing a shoe with a narrow heel, but the
shoe on the person on the right has a wide heel. Note that the same
force is applied to the box by both people, but the applied stress is Mantle
much greater with the person to the left because it is concentrated (b)
in a smaller area.
FIGURE 5–4  Stress on a body 10 km under water (a) or 10
km deep in the crust (b), assuming the stresses are distributed
a surface force is independent of the surface area affected. isotropically. The differences would arise principally from the
weight of the column of water (a) or rock (b) above the 1-m2 area
For example, the same force is exerted by the heel of a shoe
at depth.
regardless of whether the person is wearing street shoes,
field boots, or cowboy boots, or is a woman in spike heels,
but the force on the heels is distributed over a smaller area
If the same body were buried 10 km deep in the continen-
beneath the lady in spike heels than beneath the same
tal crust, the lithostatic stress could be calculated by sub-
lady wearing field boots. Moreover, a surface force can
stituting the average density of the column of rock above
be resolved into mutually perpendicular components:
the body, assumed to be 2,750 kg m−3. Substituting for
one normal to the surface, and one or two parallel to the
seawater density in equation 5–3, we obtain the lithostatic
surface. This relationship will reappear when we discuss
value of stress
normal and shear stress.
Force may be converted to stress by dividing by the 104 m × 2,750 kg m−3 × 9.8 m s−2 =
area (A) affected by the force (Figure 5–4). For example, 269.5 × 106 kg m−1 s−2 = 269.5 MPa. (5–4)
the force on a body submerged at a depth of 10 km near the
bottom of the Mariana Trench would equal the weight of Stresses may be tensional (pulling apart) or compressional
the mass of water on it, expressed as (pushing together). Stress acting on a surface may be
resolved into two vector components: normal stress, σn,
F = weight of water acts perpendicular to a reference surface; shear stress, τ,
= density of sea water (ρ) × acceleration of acts parallel to a surface. Stress vectors acting across planes
gravity (g) × height (h) × area (A in m2) = ρghA of zero shear stress are principal stresses, commonly
= 1,030 kg m−3 × 9.8 m s−2 × 104 m × 1 m2 distinguished as three principal normal stress components,
= 1.009 × 108 kg m s−2 (5–2) σ1, σ2, and σ3, of σn. The three principal normal stress
components are oriented perpendicular to each other, and
or
σ1 ≥ σ2 ≥ σ3. Their directions are principal directions of stress.
104 m × 1,030 kg m−3 × 9.8 m s−2 = The planes that contain the principal stresses are principal
100.9 × 106 kg m−1 s−2 = 101 MPa. (5–3) planes of stress. Differential stress is the difference between
the maximum (σ1) and minimum (σ3) principal normal
(1 Pa = 1 pascal = 1 kg m−1 s−2; 1 MPa = 106 pascals. stresses (σ1 − σ3); see equations 5–15 through 5–18 later in
1 megapascal is the standard unit of stress in the Earth; the chapter, and related text. Mean stress is (σ1 + σ2 + σ3)/3.
it also is equal to 10 bars or 9.8 atmospheres of pressure.) Deviatoric stress is the nonhydrostatic component of stress,
Stress | 115

Earth’s surface vertical stress σ (Figure 5–6), the stress across a small part
of the plane can be written as

∆F
Depth constant σ= , (5–5)
∆A
σ
where A is area. If we assume this segment of the plane is
infinitesimally small (equivalent to the statement of stress
at a point, discussed in the next section),
(a) Horizontal plane P 
∆F (5–6)
σi = lim
∆A→ 0 ∆A
Earth’s surface
or

dF
σ= (5–7)
dA
Depth variable Equation 5–6 indicates that stress on a plane σ is a vector
quantity—from the product of a vector (force) and a scalar
σ (1/area)—and that the stress on each plane can be expressed
as a unique set of normal and shear stress vectors (the latter
also known as tractions). Stress on any plane we choose
Horizontal plane P in a mass of rock, particularly near the surface, is likely
(b)
to vary from place to place on the plane, either because
FIGURE 5–5  Stress vectors acting on a horizontal plane the amount of overburden varies or because the plane is
beneath the surface below smooth topography (a) and irregular
inclined to the surface (Figure 5–7). At depths greater
topography (b). The vertical arrows represent a randomly
oriented stress, σ. Magnitude of stress on the horizontal plane than a few kilometers in the Earth, stress on a horizontal
also depends on the product of density (ρ), force of gravity (g), plane is related to the density and height of the column of
and the height (h), ρgh, of the column of rock above any point rock above it (ρgh = density × gravity × height), although
along the horizontal plane (P). at depths of a few kilometers, variations in the amount
of stress can occur, even within the plane. Determining
the stress on an inclined plane is more difficult and is
best accomplished using tensors (see Nye, 1957). The
expressed as total stress with mean stress subtracted from area of the plane, the density of the column of rock, and
the normal stress components. Permanent deformation the height of the column are all involved, but the angles
results if the differential stress exceeds the strength of the between the plane and the principal stress directions must
rock. The strength of a material is the stress required to also be known. Because the plane is inclined, the height of
cause permanent deformation.
A lithostatic state of stress occurs where normal stress is
the same in all directions in the Earth; a hydrostatic pres‑
sure is the confining stress acting on a body submerged
in water at a known depth. In the case of a body buried Earth’s surface
in the Earth, the weight of the column of rock (instead of
water) per unit area above it is called the lithostatic pres‑
sure (Figure 5–4). Under both lithostatic and hydrostatic
conditions, one-third of the sum of all three principal
stress components equals the mean stress. Shear compo- σn σ
nents will still exist under any deviation from lithostatic α
conditions (Figure 5–5).

Stress on a Plane
FIGURE 5–6  Stress vectors acting on an inclined plane, such
Now that we have defined the terms used in describing as a fault surface, several kilometers deep in the crust. σ is the
and measuring stress, we should consider ways of using vertical component of stress resolved into normal (σn) and shear (τ)
them. If we define a plane P in a mass of rock subjected to a components.
116 |  Mechanics: How Rocks Deform

the column of rock varies along the plane. On the plane, Mohr Circle Derivation
the stress (σ) may be resolved into components of normal
stress (σn) and shear stress (τ). The Mohr circle forms a simple and useful means of both
If we assume that our 1-m2 area in Figure 5–7 is a dif- visualizing and calculating normal and shear stresses
ferent plane now inclined 45° to the horizontal (for con- acting on planes of any orientation. Here we derive the
venience here), we should be able to calculate the vertical Mohr circle from the mathematical relationship between
stress on the plane (Figure 5–7a). Because the plane is the principal stresses in a rock mass, and between the
now inclined above the 1-m2 area, the area of the plane normal and shear stresses on a plane or surface within
is greater—1.41 m2 (= √2 m2). The vertical force, using the mass.
F = ma = volume × density × acceleration of gravity, is If we define an inclined plane P having unit area,
and a randomly oriented stress σ acts on it (Figure 5–8),
F = 104 m3 × 2,750 kg m−3 × 9.8 m s−2 we can derive several expressions for the normal (σn)
= 2.7 × 108 kg m s−2. (5–8) and shear stress (τ) on the plane. If a triangular segment
ABC represents a small prismatic element with two sides,
Conversion of force to stress involves dividing force by AC and BC, perpendicular to each other, and AC makes
the area of 1.41 m2, which yields a stress on the inclined the angle θ with AB (which lies in the larger element P),
plane of 191 MPa. Note that the area has increased, and so AC and BC are also assumed to be perpendicular to the
the normal stress on the inclined surface has decreased, greatest and least principal (normal) stress components—
because the same force is distributed over a larger area. σ1 and σ3. If A represents the area of the hypotenuse face
(Shear stress on the plane, however, must increase from P (one unit wide) of the prismatic element, and σn and τ
zero to a nonzero value.) are normal and shear stresses acting on this surface, we
Because any plane will have a unique stress vector can derive several useful relationships from the prismatic
resolvable into normal and shear components, the normal element ABC. Assume that the prism ABC is infinitesi-
and shear components of stress (Figure 5–7b) can also be mally small, so that the weight is negligible compared to
calculated from the forces acting on it. We can also assume that the prism
is in a state of static equilibrium: all forces acting to move
σn =  σ cos 45° (5–9)
the prism in any direction are countered by equal and op-
= 191 MPa × 0.707
posite forces. The same is true for forces acting to move the
= 135 MPa
prism in either direction horizontally. (These statements
and embody Newton’s third law of motion.) We have already
defined the area of the hypotenuse of the prism (plane P)
τ =  σ sin 45° (5–10) as unity, so the area of the left (vertical) face of the prism
= 191 MPa × 0.707 is 1 × sin θ, and the area of the base (horizontal face) is
= 135 MPa, 1 × cos θ. The forces acting on the three faces are the
stresses multiplied times the area of each face (force =
where σ is the traction on the plane. Note that we are con- force/area × area). If we sum the forces acting in both the
sidering only a uniaxial vertical stress here; much greater horizontal (FH) and vertical (FV) directions (they must sum
complexity would be introduced if we considered a tri- to zero in both directions),
axial case including horizontal stresses that result from
the horizontal resistance to expansion due to vertical ΣFH = 0 = Fx − N cos θ − S sin θ, (5–11)
loading.
and

45° σ ΣFV = 0 = Fy + S cos θ − N sin θ. (5–12)


σ
n
=
Pa

σn
M

45° N and S are normal and shear forces, respectively. The force
5
13

191 MPa = σ on the left side of the prism (directed parallel to the x-axis
τ
41 =
2
1. rea
m

in an xy coordinate system) is Fx = σ1 cos θ; that on the


13

1m
A

1m base (directed parallel to the y-axis) is Fy = σ3 sin θ. If the


M
Pa

1m force terms are rewritten using stress components, we can


=
τ

(a) (b) (c) write expressions for normal and shear stresses acting on a
plane at an angle θ to σ1 in terms of the principal stresses.
FIGURE 5–7  (a) Vertical stress acting on an inclined plane with
When ΣFH = 0,
an area of 1.41 m2. (b) Resolution of normal and shear stress com-
ponents of the vertical stress in (a). (c) Calculation of the magni-
tude of stresses. σ1 cos θ − σn cos θ − τ sin θ = 0. (5–13)
Stress | 117

Solving for σn and then τ based on calculations—computers were not available


then. Consequently, it was advantageous to introduce an
τ sin θ = σ1 cos θ – σn cos θ imaginary double angle to replace the real angle.
σ1 cos θ − τ sin θ Recall the trigonometric relationships sin 2θ = 2 sin θ
σn = (5–13a) cos θ, cos 2θ = 1 – 2 sin2 θ, and cos 2θ = 2 cos2 θ – 1, and
cos θ the identities
and
1 − cos 2θ 1 + cos 2θ .
σ1 cos θ − σn cos θ sin 2 θ = and cos 2 θ =
ττ = . (5–13b) 2 2
sin θ
Substituting in equations 5–15 and 5–16 to replace the
When ΣFV = 0, single-angle with double-angle trigonometric terms, we
obtain
σ3 sin θ – σn sin θ + τ cos θ = 0. (5–14)
σ1 + σ3 σ1 − σ3
Solving for σn and then τ σn = + cos θ (5–17)
2 2

σ3 sin θ − ττ cos θ (5–14a) and


σn =
sin θ
σ1 − σ3
and τ τ= sin 2θ. (5–18)
2
σn sin θ − σ3 sin θ
ττ = , (5–14b) Note that in these equations, θ is the angle between σ and
cos θ
the plane P (Figures 5–8 and 5–9). Equations 5–17 and
combining equations 5–13b and 5–14b, 5–18 define the x and y coordinates of a circle—the Mohr
circle—centered on the x-axis in a Cartesian coordinate
σ1 cos θ − σn cos θ σn sin θ − σ3 sin θ system where the axes are the normal σn (x) and the shear τ
= , (y) stresses. Here the general (x, y) = (σn, τ) coordinates for
sin θ cos θ
the circle are (x-axis value of the center + radius × cos 2θ,
cross multiplying radius × sin 2θ). Therefore, the radius of the circle for this
graphical construction is (σ1 − σ3)/2, and the x-axis value
σ1 cos2 θ – σn cos2 θ = σ3 sin2 θ – σn sin2 θ, of the center is (σ1 + σ3)/2 (Figure 5–9). Given θ, σ1, and σ3,
values of σn and τ may be calculated for any plane P that is
collecting terms normal to the σ1 σ3 plane.
A Mohr circle allows us to determine the normal and
σ1 cos2 θ + σ3 sin2 θ = σn sin2 θ + σn cos2 θ
shear stresses across any plane that is normal to two of the
= σn (sin2 θ + cos2 θ),
principal stresses. Although we have done this here in two
and, because sin2 θ + cos2 θ = 1, dimensions, some very important relationships will be
produced from this type of analysis. In three dimensions,
σn = σ1 cos2 θ + σ3 sin2 θ. (5–15) the three Mohr circles that define the normal and shear
stresses on planes normal to two principal stresses bound
Combining equations 5–13a and 5–14a, all of the possible normal and shear stresses. DePaor
(1986a) represented normal and shear stress with a graphic
σ1 cos θ − ττ sin θ σ3 sin θ − ττ cos θ
= method employing an “orthonet.” His technique provides
cos θ sin θ a way to obtain values for σn and τ without trigonometry.
The angle θ is measured between σ3 and the plane
solving for τ, including substituting sin2 θ + cos2 θ = 1,
in the Mohr construction. The relationship may also be
τ = (σ1 – σ3) cos θ sin θ. (5–16) stated in terms of the angle α, which is the complement of
θ, or the angle between σ3 and plane P in Figure 5–8. These
Otto Mohr, a German engineer who in 1882 developed angles reappear in Mohr circles as 2θ and 2α. Consider
the construction that bears his name, exhibited considerable the face of a cube of rock in Figure 5–10a in which σ1 and
insight when he realized that if equations 5–15 and 5–16 were σ3 are 27 and 5 MPa, respectively, and plane A is oriented
rewritten to incorporate double angles, determinations 20° clockwise from σ3. What are σn and τ acting on
of σn and τ could be made for all cases graphically from a plane A? We could solve equations 5–17 and 5–18 or
limited number of experimental measurements. He found construct a solution using the Mohr circle (Figure 5–10b).
a graphical solution to the problem much easier than one The state of stress is represented by the circle that
118 |  Mechanics: How Rocks Deform

FIGURE 5–8  (a) Relationships between N = σn


a randomly oriented force are resolved α
into normal stress σn, and shear stress A
τ, components within a plane P having Area of end = 1 sin θ θ S=τ
an area dA (or ∆A) on which the force is or ∆A sin θ Plane P
Area = 1
applied, and resolved principal normal
stress components σ1 and σ3, and the Fx = σ1 cos θ
angles α and θ within the prism. α
(b) 2D representation of the un it
C B 1
relationships expressed in 3D in (a).
(a) Area of base = 1 cos θ
or ∆A cos θ
Fy = σ3 sin θ
y

Randomly
oriented
σ stress
σn
Fx = σ1 cos θ Resolved
θ τ normal and
shear stress
components
∆A cos θ

∆A

α
x
τ∆A sin θ

Fy = σ3 sin θ
(b)

intersects the τ = 0 axis at 27 and 5 MPa; the point on shear stress (τ) is 8 MPa. Planes oriented at a high angle to
the Mohr circle that represents the σn and τ acting on σ1 have high normal stresses and low shear stresses acting
plane A is found by plotting the point made by a 2θ (40°) on them. Based on the geometry of the Mohr circle, it is easy
rotated counterclockwise from the σ1 side of the τ = 0 axis to see that the plane of maximum shear stress is oriented
(Figure 5–10b). The normal stress (σn) is 24.5 MPa, and the 45° to σ1 and will equal the radius of the circle (σ1 − σ3)/2.

FIGURE 5–9  Mohr circles plotted on σ σ1 – σ3


and τ axes, the envelopes, and the (σn, τ) = ( σ +2 σ
1 3
cos 2θ ,
2
sin 2θ )
relationships between 2α and 2θ.
σ1 + σ3 σ1 – σ3 State of
σ= + cos 2θ stress on
τ 2 2
plane P

σ1 – σ3
2

σ1 – σ3

τ= ( 2
sin 2θ )

0 σn
σ3 σ1

(σn, τ) = ( σ +2 σ , 0 )
1 3

2 × angle
between σ3
and plane
Stress | 119

τ FIGURE 5–10  Physical space (a) and


σ1 = 27 MPa 10 Plane A Mohr space (b) illustrating the state of
stress acting on plane A.
σn = 24.5 MPa
τ = 8 MPa

Pla 2θ = 40°
θ = 20° ne A 0 σn
σ3 = σ3 10 20 σ1 30
5 MPa

–10

(a) (b)

Mohr Construction ease, by constructing tangents to the circles both above


and below the σn axis (Figure 5–12c). The Mohr envelope
separates unfailed (no permanent deformation) from
Now we can use experimental results to calculate values of failed (permanently deformed) regions in the diagram; the
material properties. A cylindrical sample of rock is placed unfailed region lies within the envelope (Figure 5–12d).
in the test cylinder of a triaxial test apparatus (Figure 5–11a). A Mohr circle that does not intersect the envelope indi-
The specimen is jacketed in a thin sleeve of a soft metal cates that the sample did not rupture with the values of σ1
such as copper or thermal plastic, to protect it from a fluid and σ3 employed in that run. Put another way, if the Mohr
(such as a low-viscosity oil) that is used to vary the confin- circles do not intersect the Mohr envelope, the sample
ing pressure externally. In an axial compression test, the undergoes no permanent deformation (Figure 5–12d).
confining pressure is equal to both σ2 and σ3 because pres- Curvature of the Mohr envelope is frequently related
sure is being applied equally around the cylinder. The load to inherent properties of the material (Figure 5–13).
on the ends of the cylinder is usually increased until the Brittle materials tend to produce relatively straight Mohr
material ruptures (Figure 5–11b). This load is a measure of envelopes with steep slopes. With increasing confining
the axial stress σ1. Plotting each value of σ1 and σ3 on the stress (σ3), even a brittle material behaves ductilely on
σn (horizontal) axis permits construction of a Mohr dia- the megascopic scale. If the materials are ductile, they
gram for the limestone at failure (Figures 5–12a, b), and undergo some permanent, nonrecoverable strain before
determines the diameter of the Mohr circle. If the rock is rupture (Chapter 7). With an increase in ductility, the
loaded to the rupture point, a new cylinder of the same Mohr envelopes become flattened, resulting in an over-
material is required for each run at increasing confining all curvature. In general, the greater the curvature, the
pressure σ3. When the sample ruptures, a shear fracture greater the amount of ductile strain before rupture, but
develops in one of two orientations [with both shear frac- Griffith materials (Chapter 9), which are actually brittle,
tures developing at the same angle with respect to σ1 and σ3 also produce curved envelopes. Their properties are re-
(Figure 5–12b)]. Table 5–1 contains values for successive lated to the amounts of dilation of the preexisting micro-
runs to failure on cylindrical specimens of limestone using crack population.
an axial load apparatus. Each successive run was conducted The angle 2α, the conjugate shear angle, is related to the
at a higher confining pressure (σ3), and runs 1 to 4 each coefficient of internal friction μi (or tan ϕ, defined below in
required a greater differential stress, σd, for fracturing to equation 5–22), and α is the angle between σ1 and the frac-
occur. The Mohr failure envelope can be constructed with ture that forms during rupture. The relationship is

90 − ϕ = 2α. (5–19)
TABLE 5–1
TEST RUN σ3 (kg cm−2) σ1 (kg cm−2) (Here α is defined in degrees, as in Figures 5–9 and 5–12b.)
As the radii of the Mohr circles increase and approach a
1 0 750 limit (with increasing stress difference and confining pres-
2 250 1,750 sure), the slope of the envelope flattens as a function of
increased ductility, and 2α approaches a maximum value
3 500 2,400 of 90°. Therefore, the maximum shear angle α is 45°. This
angle is most commonly attained (and sometimes ex-
4 1,050 3,550
ceeded) in ductile materials.
120 |  Mechanics: How Rocks Deform

τ
Fixed platten State of
10
stress
at failure
Dial gauge
Baldwin load Output to
cell recording
potentiometer t1 t2 t3 t4
0 σn
10 20
Output to To interstitial
recording fluids pump
potentiometer increasing σdd

To packing –10
Newhall pressure
packing intensifier
sleeve (a)
Insulation σ1 = 17 MPa τ σn = 4 MPa
State of
Thermo 10 τ = 7 MPa
couple θ = 60° stress
at failure
Calrod σ3 =
tubular Rock 0 MPa
heating specimen
elements 2θ = 120°

0 σn
10 20
σ1 = 17 MPa
2θ =
2θ = –120°
-120°
To confining θ = -60°
–60°
pressure
σ3 =
intensifier
0 MPa
and 500 –10 σn = 4 MPa
MPa pump τ = –7 MPa
Movable platen (b)
τ Mohr
envelope

Test Test
Hydraulic jack (σ1) run 4 run 4
σ33 σ1
(a) 0 σn
1,000 2,000 3,000 4,000

(c)
τ
Failed

Unfailed
0 σn

Failed

(d)

(b) (1) (2) FIGURE 5–12  (a) Successive Mohr circles as the differential
stress (σd) increases over time (t) as the hydraulic jack is applied.
FIGURE 5–11  (a) Triaxial test apparatus. Axial stress (σ1) is applied (b) Physical space and Mohr space at time of failure. Based on the
with a hydraulic jack. The axial load is measured directly by the load principal stresses and the orientation of the shear fractures, the
cell. The dial gauge provides a direct measure of strain as a percent- values for σn and τ can be determined from the Mohr circle.
age of shortening of the rock cylinder. (From D. T. Griggs, F. J. Turner, (c) Plots of successive Mohr circles for the four experimental runs
and H. C. Heard, 1960, Geological Society of America Memoir 79.) in Table 5–1 taken to rupture. Failure envelope connects the
(b) Limestone cylinder before (1) and after (2) being tested. [Photos stress state at failure for the successive experiments. (d) Mohr
from Vertek Material Properties Laboratory (www.rocktestinglab.com).] diagram showing unfailed (stable) and failed (unstable) regions.
Stress | 121

τ 6,000
φ)
n
ta φ = 45°
4,000 σn σ1 – σ3
+
(τ 0 2
+
2,000 τ= 2θ

φ

kg cm–2
σn σn
τ0 σ3 σ1 + σ3 σ1
τ= 2
–2,000 +

0 +
σ
–4,000 n ta
n
φ)
–6,000

0 2,000 6,000 10,000 14,000


(a) (b) kg cm–2
τ

τ 3,000 n φ) 22°
+ σ n ta
+ (τ 0
τ= φ = 22° – 43°
2,000

1,000 2θ

kg cm–2

σn 0 2α 2α σn
σ3 σ1
σ1 + σ3
2
–1,000

–2,000

–3,000

0 2000 4000 6000 8000


(c) (d) kg cm–2

FIGURE 5–13  Mohr diagrams for brittle [(a) and (b)], and more ductile [(c) and (d)] materials. Mohr diagrams for Oil Creek sandstone
(b) and Blaine anhydrite (d), deformed at 24° C and 0–2,000 atmospheres pressure. (From M. K. Hubbert and D. G. Willis, 1957, American
Institute for Mining, Metallurgical, and Petroleum Engineers Transactions, v. 210.)

Several relationships within Mohr diagrams are


summarized in Figure 5–14. Note that the τ axis sepa-
Amontons’ Law and the
rates compressional from tensional normal stress fields Coulomb–Mohr Hypothesis
(Figure 5–14a). Materials with no tensile strength, such as
dry sand, have an envelope that terminates at the origin
French physicist Guillaume Amontons suggested at a
(Figure 5–14b). Most geologic materials have compressive
scientific meeting in 1699 that a direct proportional
strengths (breaking strength under compressive stress)
relationship exists between F, the shear force necessary
much greater than their tensile strengths (Figure 5–14c).
for sliding along a contact surface, and W, the force
Very few materials have tensile and compressive strengths
perpendicular to the surface, expressed as an equality
even approximately the same, but mild carbon steel
(Figure 5–14d) is one that does, and a horizontal envelope
results that closes in the tensile field. F = μs W, (5–20)
122 |  Mechanics: How Rocks Deform

τ
τ
Zero
tensile
strength
Tension Compression

–σn σn
σn

(a)

τ
(b)

σn

σn

(c) (d)

FIGURE 5–14  (a) Axes for Mohr diagrams showing extensional (−σn) and compressional (+σn) fields. (b) Mohr diagram for quartz sand,
a material with no tensile strength. (c) Mohr diagram for an average rock that exhibits tensile strength at low stress, and greater com-
pressive strength and possibly some ductility at higher stress. (d) Mohr diagram for a mild carbon steel, where tensile and compressive
strengths are about the same, yielding an almost horizontal Mohr envelope that ultimately closes into the tensile field.

and today known as Amontons’ first law, where μs is the value. The Coulomb equation is an equation of a straight
coefficient of sliding friction along the surface (Jaeger line that approximates straight-line segments of the Mohr
et al., 2007). The coefficient of sliding friction is a measure envelope (Figure 5–12c). It is commonly expressed today as
of the resistance of a material to sliding along a surface.
If we divide both sides of equation 5–20 by the surface area |τs| = τ 0 + σn tan ϕ (5–23)
(A), it becomes or
|τs| = τ 0 + σn μi.
τs = μs σn. (5–21)
Here, τ 0 is the cohesive strength of the material at zero
Another French physicist, Charles A. Coulomb, normal stress, and μi = tan ϕ; ϕ is the angle of internal
recognized the linear relationship between shear and friction, and tan ϕ is the coefficient of internal friction.
normal stress implied by Amontons’ first law. From In 1900, Mohr generalized that shear strength is a
that, he ­derived another relationship that today bears his function of normal stress,
name, the Coulomb criterion of failure, proposed in 1773.
It states that the absolute value (either a positive or negative |τs| = f(σn), (5–24)
number) of shear strength τs is the sum of the inherent
where |τs| is the absolute value of shear strength—the resis-
shear strength S0 and the coefficient of internal friction μi,
tance of a material to shear stress. The location of the Mohr
or static friction, multiplied by the normal stress σn‑,
envelope in σnτ space is directly related to the strength
|τs| = S0 + μi σn. (5–22) of the material. Mohr’s hypothesis and the derivation of
the equations for resolved shear and normal stresses on a
Equation 5–22 predicts that fracturing will occur when the plane (equations 5–14a and 5–14b) are both incorporated
shear stress on a plane (such as a fault plane) reaches a critical in the Mohr construction (Figure 5–9).
Stress | 123

Stress Ellipsoid σ3 = σz

z
A graphic means of showing the relationships between the σ
principal stresses is the stress ellipsoid (Figure 5–15). It is
a triaxial ellipsoid in which the greatest, intermediate, and
least principal axes are σ1, σ2, and σ3. Principal planes in σ1 = σx
the stress ellipsoid contain the principal axes. The σ1 − σ2, x
σ2 − σ3, and σ1 − σ3 planes are principal planes. The σ1 − σ3
σ2 = σy
plane represents the maximum stress difference. All non-
principal planes are shear planes. y

FIGURE 5–16  Principal normal stresses oriented parallel to the


edges of a cube in x-y-z space. Note the position of the cube rela-
Stress at a Point tive to the positive ends of the axes.

components of shear stress lying within the three faces of


Having considered stress on a plane in two dimensions, the cube defined as planes xy, yz, and xz. The symbol σxx,
we can now consider stress in three dimensions. Principal for one of the principal normal stresses, indicates that this
normal stresses, designated σ1, σ2, and σ3 (σ1 ≥ σ2 ≥ σ3), is a normal stress acting perpendicular to the yz plane and
may be thought of as oriented parallel to coordinate axes parallel to the x-axis. The components of shear stress are
x, y, and z, so that σ1 corresponds to σx, σ2 to σy, and σ3 to then expressed as τxy, τyz, τxz, τyx, τzy, and τzx. Any direction
σz (Figure 5–16). These normal stresses may also be con- in the three-dimensional space is defined by the normal
sidered parallel to the edges and normal to the faces of a to the plane that contains the direction and a line in the
cube. If the cube is reduced to infinitesimal size, a stress σ plane. Thus, two lines define the orientation in three di-
applied to one of the planes that defines the cube may be mensions. Similarly for the cube, the notation xz means in
considered as being applied at a point 0 (Figure 5–17a). We the plane normal to x in the z direction; yx means in the
may also assume, using Newton’s third law of motion, that plane normal to y in the x direction.
the stresses on all sides of the cube balance and cause no A tensor is a mathematical tool for defining and
rotation or translation, regardless of the size of the cube manipulating a group of quantities, where each quantity
and regardless of whether or not the stresses are normal is represented by a magnitude, and most have an
or shear stresses. The cube should be very small so that
both shear and normal stresses balance. The stress σ de- +z σ
fines a second-rank tensor that may be expressed in both
normal and shear components, σn and τ for any plane. In
three-dimensional space, nine components (Figure 5–17b)
are required to describe the stress σ at point 0. They in-
clude the three components of normal stress, σxx, σyy, and
σzz, oriented parallel to the three coordinate axes, and six 0 +x

σ1 σ1
Shear
plane +y
(a) σzz
+z σ
τzx
τzy

τyz τxz
σ3 σ3 σxx

σ2 σ2 +x
τxy
σyy τyx
+y
(b)
(a) (b)
FIGURE 5–17  (a) Randomly oriented stress σ applied to an
FIGURE 5–15  (a) The stress ellipsoid, a triaxial ellipsoid in which the infinitesimally small (point size) reference cube in x-y-z space.
axes are the principal stresses σ1, σ2, and σ3. (b) Planes of maximum (b) Enlarged reference cube shows resolution of nine shear and
shear stress are always parallel to σ2 and ideally at 45° to σ1 and σ3. normal stress components.
124 |  Mechanics: How Rocks Deform

accompanying “direction.” (Note that the direction does beginning of this discussion. If the x-y-z coordinate system
not have to be related to real coordinates in space or time.) is oriented parallel to the principal stresses, no shear stress
Tensors have the same magnitude in any coordinate system, will be present on the faces (all values of τ = 0). The same
but the values of the components depend on the choice of result can be accomplished by rotating the cube in the stress
coordinate system or “space.” Tensors have different ranks: field until only normal stresses are present on the faces.
a zero-rank tensor is a scalar and has only one component; The stress tensor for zero shear stresses may be written as
an example is air temperature at a particular time. A first-
rank tensor is a vector and has three components; a wind  σ 0 0   σ 
 xx   xx 
current is a first-rank tensor if it is thought of as a single    
tensor quantity at a point. A second-rank tensor relates  0 σ yy 0  =  σ yy . (5–26)
   
sets of vectors to each other and has nine components.  0 0 σzz   σ 
   zz 
As we saw earlier components of shear and normal stress at
a point form a second-rank tensor. Tensor rank describes The three normal stresses are thus called the principal
the minimum number of directions necessary to describe stresses, with σxx = σ1, σyy = σ2, and σzz = σ3. They act per-
the dimensions under consideration. Thus the number of pendicular to planes so that the tensor is transformed from
components for a particular rank depends on the number second to first rank.
of dimensions in the space being considered. For a first-
rank tensor, the number of components equals the number
of dimensions in the space (2D = 2, 3D = 3, 4D = 4, . . .).
A tensor may be used to describe a physical quantity by Measuring Present-Day
referring it to an appropriate coordinate system. The elastic
constants relating stress and strain (Chapter 6) comprise a Stress in the Earth
fourth-rank tensor. The number of components in a tensor
may be determined from 3n, where n is the rank, and the rank Knowledge of the orientation and magnitude of present-
equals the number of letters or numbers in the subscript for day stress in the Earth is important because it provides
a particular element of the tensor. The symbol τyz for a shear clues about the nature of active faults, information about
stress tells us that it is part of a second-rank tensor. what structures might become more active in the future,
All nine components of the second-rank tensor for and insight into the kinematics of plate motion. Knowledge
stress at a point σij, where i and j take the values of x, y, and of stress orientations bears directly on our understanding
z, may be arranged in matrix form as of seismic activity and its causes (Zoback and Zoback,
1980; Zoback et al., 1980). Present-day stress in the Earth
 σ ττxy xy ττxzxz 
 σxx
xx can be measured by placing a strain gauge in rock and
  recording the in situ elastic strain. Such measurements
 ττyx yx σ
σyy yy ττyzyz  (5–25)
  are most frequently carried out by drilling a hole into
 ττ 
 zx zx ττzy zy σ σzzzz  bedrock, inserting the instruments, and recording the
  amounts and orientations of elastic strain. Two techniques
This matrix contains the components of a stress tensor in are commonly used for measuring in situ stress: overcoring
terms of axes x, y, and z. The rows (horizontal) from top and hydraulic fracturing, although other data, such as
to bottom refer to stresses on a face normal to a particular well-bore breakouts, can provide stress orientation.
coordinate axis, and columns (vertical) refer to stress Overcoring involves drilling a hole 3 to 4 cm in diameter,
directions parallel to coordinate axes (Figure 5–17b). and then drilling a larger core (12 to 15 cm) with the same
The columns (vertical) from left to right represent the center as the smaller hole (Figure 5–18; Hooker and Bickel,
directions of stress components parallel to the x, y, and 1974). Before coring the larger hole, an instrument called a
z-axes, respectively. The left column represents stresses dilatometer (a stain gauge) is inserted into the smaller hole
acting parallel to the x-axis, the middle column the to permit measurement of the expansion (relaxation) of
stresses acting parallel to the y-axis, and the right column the rock mass as the larger hole is cored. Changes in shape
the stresses acting parallel to the z-axis. If we designate of the small hole are recorded as it is overcored by the
the elements of the tensors i and j—again with i and j dilatometer. The orientation of the small hole is known,
taking values of x, y, and z—we can state that the tensor so changes in shape are measured as the cross section
in equation 5–25 is symmetric because τij = τji (that is, relaxes from a circle to an ellipse. This provides a measure
elements i and j may be reversed without changing the of the orientation of the present-day stress ellipsoid. If
values of the components, a requirement of the condition several measurements can be made at right angles at the
of equilibrium; the subscript ij also indicates that this is same locality of overcores in different orientations—as is
a second-rank tensor). Symmetry is a requirement of the possible in a mine, tunnel, or quarry—a very good measure
condition of force balance, or equilibrium, assumed at the of the orientation of the stress ellipsoid may be obtained.
Stress | 125

FIGURE 5–18  Overcoring technique


for measuring in situ stress. The strain
Overcore gauge rests inside the small borehole
Inner hole and measures in situ stress—while the
Rock (drilled first) overcore is being drilled—by recording
the amount and direction of change
from a circular to an elliptical borehole.

15–20 cm 3–4 cm
3-4 cm

Cable
to Rock
recorder
Strain
gauge

Attempts have also been made to calculate the magnitude Hydraulic


of the principal stresses using this technique. line for
injecting
Hydraulic fracturing involves drilling a vertical hole, fluid
sealing off a part of the hole with packers, and increasing
the hydraulic pressure in the sealed-off part of the hole until
the wall of the hole fractures (Figure 5–19; e.g., Zoback and
Haimson, 1982). Both the amount of pressure needed to Packer
produce fractures and the orientations of hydraulic fractures
are measured to provide values of both the magnitude of
minimum horizontal stress at the site and the orientation of
Pressure is
maximum and minimum horizontal principal stresses. It increased
is possible to make several measurements in the same hole until the rock
if it is drilled to a minimum depth of 300 to 400 m. The fractures

Hydrostatic
pressure
hydraulic fracturing method requires that the hole drilled
be nearly vertical, for it is assumed that one of the principal Bedrock
normal stresses is vertical and that the hydraulic fracture in
a vertical borehole propagates perpendicular to σ3 (Hubbert
and Willis, 1957.) Zoback and Zoback (1980) concluded
that field data also support the relationship between Packer
hydraulic fracture propagation and σ3.
Numerous measurements of in situ stress using dif-
ferent techniques around the world have been compiled
into a worldwide stress distribution map (Heidbach et
al., 2008). Such maps permit identification of areas or do-
mains of common maximum principal stress orientation
(Figure 5–20). 3–5 cm
In situ stress in a tectonically active area has been measured drillhole
by Zoback et al. (1980) in a series of wells about 240 m deep FIGURE 5–19  Hydraulic fracturing method for measuring in situ
drilled as far as 34 km from the San Andreas fault. They stress. Larger diameter drill holes can be used but the pressures
also made several measurements in a 1 km deep well about required to fracture the rocks increase as the radius of the drill
4 km from the fault. Their goal was to evaluate variations in hole increases by a power-law relationship. (Modified from
R. O. Kehle, Journal of Geophysical Research, v. 69, © 1964
the magnitude of the principal stresses with depth and to by the American Geophysical Union.)
calculate the magnitude of shear stress along the fault. They
used hydraulic fracturing to measure stress and obtained the where Pb is breakdown (fracture formation) pressure, Pp
magnitudes of Shmax (maximum horizontal stress = σ1) and is pore pressure, and T is the tensile strength of the rock
Shmin (minimum horizontal stress) by modifying an equation mass. Pb is measured directly as the pressure necessary
originally derived by Hubbert and Willis (1957): to hydraulically break the rocks. Pp is calculated from the
pressure of a column of water (using ρgh) at the depth at
Pb = 3Shmin− Shmax− Pp + T, (5–27) which hydraulic fracturing was carried out, assuming that
126 |  Mechanics: How Rocks Deform

Crustal Stress
Regime Method
Normal fault Breakouts
(extensional)
Focal
Strike-slip fault mechanism
Thrust fault Overcoring,
(compressional) hydrofracturing
Unknown

0 400 800
kilometers

FIGURE 5–20  Partial data set for measurements and estimates of orientations of principal stress directions in the continental United
States, northern Mexico, and southern Canada. Heavy dashed lines indicate areas of common orientation of present-day maximum
(compressional) and minimum (extensional) principal stress. Arrowheads of large red arrows indicate whether stress produces strike-slip,
extensional, or compressional structures. (Modified from O. Heidbach, M. Tingay, A. Barth, J. Reinecker, D. Kurfeß, and B. Müller, The World
Stress Map database, 2008, http://www.world-stress-map.org/.)

the pores in the rocks are interconnected and communi- bars) at 750 to 850 m. Studies by Zoback et al. (1987)
cate with the surface. The tensile strength of the rock mass based on in situ stress measurements in the Cajon Pass
can be estimated (as here) or measured on core samples scientific borehole, other stress measurements along the
obtained from the well. The magnitude of Shmax is mea- fault nearby, and orientations of folds in Pleistocene sedi-
sured directly, and its orientation (compass bearing) is ob- ments immediately adjacent to the fault indicate that Shmax
tained from orientations of hydraulic fractures in the well. is oriented nearly perpendicular to the San Andreas
The value of Shmax is then calculated from equation 5–27 fault northeast of Los Angeles. Thus, there is almost no
and is assumed to be horizontal. shear stress along the fault near the surface in this area:
Results of measurements in the two areas studied shear stress equals zero.
(Table 5–2) indicate that the maximum principal hori- Plotting principal stress orientations on a map
zontal stress, Shmax, is oriented about 45° from the strike enables us to relate the present-day stress in the Earth
of the San Andreas fault. Shear stress, determined by to plate motion (Figure 5–20). The nearly uniform N 70°
plotting the maximum and minimum principal stresses E orientation of stress fields in the eastern United States
on a Mohr diagram, increases from about 2.5 MPa is thought to be related to “ridge push” from the Mid-
(25 bars) at depths of 150 to 300 m to about 8.0 MPa (80 Atlantic Ridge, and the markedly different orientations
Stress | 127

TABLE 5–2  Hydrofracture Data along the San Andreas Fault


FRACTURE
DISTANCE BREAK-
SAN DOWN OPENING PORE TENSILE DIRECTION OF
DEPTH ANDREAS PRESSURE PRESSURE PRESSURE Shmin Shmax STRENGTH τmax MAXIMUM
WELL (M) FAULT (KM) (BARS) (BARS) (BARS) (σ3) (σ3) σ2 (BARS) (BARS) COMPRESSION

1 167 4 200 69 17 73 133 45 131 30 N 4° W


1 196 4 209 74 20 77 138 53 135 31 N 1° E
2 338 4 109 63 34 74 125 91 46 26 N 43° W
2 561 4 163 130 56 150 264 152 33 57 N 20° W
2 787 4 192 124 78 183 346 213 68 82 N 19° W
3 80 2 144 24 8 23 38 18 120 8 N 20° W
3 185 2 250 73 19 56 73 43 177 6 N 23° W
4 167 4 139 47 17 51 89 45 92 19 N 83° E
4 230 4 164 85 23 83 140 62 79 29 N 14° W

(From Zoback, Tsukahara, and Hickman, Journal of Geophysical Research, v. 85, © 1980 by the American Geophysical Union. Used by permission.)

in western states are related to interaction between of seismic monitoring equipment, GPS stations, and a sci-
the Pacific and North American plates (Figure 1–2; entific drill hole (SAFOD—San Andreas Fault Observa-
Richardson et al., 1979). tory at Depth). Numerous stress measurements have been
In addition to direct measurements of stress, the ori- made on the surface and several were made in the SAFOD
entation of maximum and minimum principal horizontal pilot hole, a 2 km deep hole drilled southwest of the San
stresses can be obtained from well-bore breakouts (elon- Andreas fault near Parkfield, California, into a region near
gations) that form after a hole is drilled. A hole changes the fault that produces frequent small earthquakes. Stress
shape and fractures as the walls relax, producing curved measurements in this hole again indicate low shear stress
fractures and causing rock to spall from the walls of the parallel to the fault, but large mean stress—approximately
hole. These fractures form parallel to the orientation of twice the lithostatic stress there (Chéry et al., 2004). The
the principal stress directions, but the magnitude of stress high angle made by Shmax with the fault (85°) near San
cannot be determined. Mount and Suppe (1987, 1992) used Francisco is consistent with an average angle of 68 + 7°
breakouts from drill holes near the San Andreas fault to over 400 km of the fault in California, again suggesting the
independently conclude that σ1 is perpendicular to the San Andreas has low frictional strength throughout this
fault, and that shear stress along the fault surface is zero. region (Townend and Zoback, 2004).
Interest in the stress and other properties of the San
Andreas fault has resulted in a major part of the Earth- We end our survey of stress and its mechanical basis. We
Scope Project being dedicated to study of this fault in the turn in the next several chapters to a discussion of strain—
Plate Boundary Observatory, which consists of an array the presumed effect of stress.

ESSAY  The Earthquake Cycle

Whenever a large earthquake (ML ≥ 7) occurs somewhere in each other, or pulling apart from each other (see Figure 1–2).
the world, that earthquake probably occurs on a fault that Earthquakes occur along segments of large active faults
has been known to have produced large earthquakes for like the San Andreas in California, the North and East Ana-
hundreds or even thousands of years. Large active faults tolian faults in Turkey, and the Cascadia subduction zone in
produce earthquakes of different magnitudes and frequen- the Pacific Ocean off Washington and Oregon. For an earth-
cies. These faults frequently form plate boundaries that sep- quake of any magnitude to occur, stress has to accumulate
arate two plates that are colliding head-on, grinding past and be released suddenly, producing elastic rebound and
128 |  Mechanics: How Rocks Deform

ESSAY  continued

earthquake waves. Once the accumulated stress—elastic they culminate with a larger earthquake. During the weeks
strain energy—is released, the fault begins to accumulate and even months that follow a major earthquake, there will
stress to a point where movement will occur again. This is be numerous “aftershocks,” commonly of smaller magnitude
the simplest statement of the earthquake cycle, but this proc- than the largest event. Aftershocks frequently bring down
ess is not simple or we would today be able to predict when poorly constructed buildings and other structures weakened
and where earthquakes will occur. The period of stress ac- during the major earthquake. The fault then returns to an-
cumulation is commonly a quiet time along the fault lasting other period of quiescence as stress and elastic strain energy
many years to centuries, although small earthquakes may build along the fault once more.
occur during this period. As stress continues to accumulate A number of different factors influence the time interval
over months to years, however, the frequency and magni- between large earthquakes along a fault. These include the
tude of earthquakes increases. The larger earthquakes may rate of stress (elastic strain) accumulation, presence of water
or may not be “foreshocks,” depending on whether or not or weak materials like clay along the fault that may weaken a
particular segment of the fault, orientation of the fault rela-
125°W
JUAN 120°W 115°W tive to the maximum principal stress, σ1, and other factors
DE FUCA that we do not understand.
Cascadia subduction zone

PLATE OR
ID The time or “recurrence” interval between large earth-
Mt. quakes may range from a few tens of years, or even less,
Shasta
(4317 m) along some major active faults in the world to several centu-
Basin and
ries along some faults located in the interiors of continents
Range
NORTH far from plate boundaries. Large earthquakes (ML ≥ 7) occur
40°N Province 40°N

GORDA
at a rate of approximately 18 per year worldwide (USGS
UT
AMERICAN
Sie

PLATE NV Earthquake Statistics website; http://earthquake.usgs.gov),


rra

CA
and commonly not on the same fault or even the same part
Ne

PLATE
vad

of the same fault. There has not been a large earthquake


Gr

aM
ea

SF Mt. Whitney
(4418 m) in the San Francisco area since the ML = 7.8 earthquake in
tV

1
tns
all

1906. The 1989 Loma Prieta earthquake located just to the


ey

2
Death
Valley
(–86 m) south had a magnitude of 6.9, and it, with its accompany-
Sa

LV
ing aftershocks, filled in an area along the San Andreas fault
n

k f.
rloc
(a “seismic gap”) where there had been relatively few his-
An

dr
35°N eas Ga Mojave
Desert 35°N
s
California and Vicinity faul
t toric earthquakes, indicating that part of the San Andreas
Earthquakes 4
(1769–2007)
3 AZ had completed the earthquake cycle in 1989 (Figure 5E–1).
LA Salton Trough
Magnitude (ML)
PACIFIC Other parts of the San Andreas in southern California are
6.0–6.4
PLATE thought to be “locked” (still accumulating stress as elastic
6.5–7.4 C
bo alifo
rd
strain energy), despite the damaging ML = 6.6 San Fernando
er rnia
7.5–9.0 lan
d
Mexico
earthquake in 1971 and the ML = 6.7 Northridge earthquake
Gulf
0 50 100 200 in 1994 in the Los Angeles area (Smith and Sandwell, 2006).
125°W kilometers 120°W 115°W An important exception to the point made here occurred
in 1811 and 1812 in eastern Arkansas, southeastern Missouri,
FIGURE 5E–1  San Andreas and related faults in California and western Tennessee where the four large New Madrid
and adjacent region, together with ML = 6.0 and larger historical
earthquakes. The San Andreas fault terminates to the northwest earthquakes occurred over a period of less than two
into one or more transform faults and the Cascadia subduction months (U.S. Geological Survey) (Figure 5E–2). An estimated
zone; to the southeast it joins the remnants of the East Pacific ML = 7.9 earthquake occurred here on December 16, 1811,
Ridge in the Gulf of California. Note that the San Andreas also has an estimated ML = 7.0 earthquake on December 17, 1811, an
a splay, the Garlock fault, that trends eastward and terminates in estimated ML = 7.6 earthquake on January 23, 1812, and then
the Mojave Desert. 1—1906 ML = 7.8 San Francisco earthquake.
2—1989 ML = 6.9 Loma Prieta (“World Series”) earthquake. a fourth earthquake on February 7, 1812, with an estimated
3—1994 ML = 6.7 Northridge earthquake. 4—1971 ML = 6.6 magnitude of 8.0 (Johnston and Schweig, 1996). These
San Fernando earthquake. Despite the smaller magnitude of magnitudes are estimates because the modern seismograph
earthquakes 2, 3, and 4, they produced major damage in the had not been invented yet, and additional analysis by Hough
populated areas where they occurred. SF—San Francisco. (2001, Table 1) has estimated these earthquakes to have
LA—Los Angeles. LV—Las Vegas. Lines with solid teeth are active
subduction zones. Lines with open teeth are dormant or inactive slightly lower magnitudes. Nonetheless, never before or since
subduction zones. Lines without ornamentation are strike-slip have there been four major earthquakes occurring so close
faults. (Modified from J. C. Crowell, 1987, Episodes, v. 110.) together in time or space. Other anomalous facts about this
Stress | 129

90°W series of earthquakes are that the earthquakes did not occur
Charleston close to a plate boundary, there is no identifiable fault that
actually broke the surface, and subsequent measurements of
the continued seismicity in this area indicate the earthquake
New
Madrid source is more than 5 km below the surface. The New Madrid
MO
2/7/1812
KY seismic zone continues to be the most active in the eastern
AR TN United States, although there have been no large (ML ≥ 7)
Reelfoot Lake
earthquakes since the 1811–1812 series. Estimates of large
ent
Reelfoot fault
eam
1/23/1812 earthquake recurrence in this zone are on the order of 400 y
l lin

Caruthersville
(e.g., Tuttle and Schweig, 1995; Johnston and Schweig, 1996).
thee

ift The New Madrid earthquakes fall into the general class of
Boo

MO
tr TN
intraplate earthquakes—in other words, those not associated
oo

36°N Blytheville AR 36°N


er
elf

t
Re

rif
Riv

ot with plate boundaries.


lfo
pi

si p
e

sis
Re

References Cited
Mis

Earthquakes Hough, S. E., 2001, Triggered earthquakes and the 1811–1812 New Madrid,
12/16–17/1811 1811–2012
Marked Magnitude (ML)
Central United States, earthquake sequence: Bulletin of the Seismological
Tree <1.0 Society of America, v. 91, no. 6, pp. 1,574–1,581.
1.0–1.9
2.0–2.9 Hough, S. E., Armbruster, J. G., Seeber, L., and Hough J. F., 2000, On the
3.0–4.9
modified Mercalli intensities and magnitudes of the 1811–1812 New Madrid
Memphis 5.0–6.9
earthquakes: Journal of Geophysical Research, v. 105, pp. 23,839–23,864, doi:
TN
0 25 50
≥7.0 10.1029/2000JB900110.
AR
kilometers MS 90°W Johnston, A. C., and Schweig, E. S., 1996, The enigma of the New Madrid
FIGURE 5E–2  Distribution of earthquakes in the New Madrid earthquakes of 1811–1812: Annual Reviews of Earth and Planetary Science,
seismic zone. Note that the earthquake array defines several fault v. 24, p. 339–384.
segments, with only the northwest-trending Reelfoot (a thrust with Smith, B. R., and Sandwell, D. T., 2006, A model of the earthquake cycle along
the teeth on the hanging wall) identified on the map. The brown the San Andreas fault for the past 1000 years: Journal of Geophysical Re-
normal faults (ticks on the hanging wall) belonging to the Reelfoot
search, v. 111, B01405, doi: 10.1029/2005JB003703.
rift are early Palezoic faults that have been identified using
aeromagnetic data (Chapter 4). Earthquake data from the National Tuttle, M. P., and Schweig, E. S., 1995, Archeological and pedological evi-
Earthquake Information Center (NEIC) of the U.S. Geological dence for large earthquakes in the New Madrid seismic zone, central United
Survey. Locations of lineaments and rifts from Johnston and States: Geology v. 23, pp. 253–256.
Schweig (1996). Location of Reelfoot fault from Hough et al. (2000).

Chapter Highlights
• Stress in the Earth’s crust produces deformation; when • The Mohr circle is a powerful and simple graphical tool
stress exceeds the strength of a material permanent de- for determining normal (σn) and shear stresses (σs) acting
formation occurs. on planes.
• Compressional stress can be thought of as “pushing • Experimental deformation of materials leads to a failure
together,” resulting in a decreased volume, whereas envelope that separates states of stress in which a mate-
tensional stress involves “pulling apart,” frequently rial will deform vs. not deform.
producing a volume increase. • Stress can be measured in situ using a variety of
• The state of stress at a point can be resolved into three techniques such as overcoring and hydraulic
mutually perpendicular principal stresses (σ1, σ2, and σ3) fracturing.
that define the stress ellipsoid.

Questions
1. Why do we commonly deal with stress, rather than force, in 2. Calculate the lithostatic value of stress on a fault plane
the earth sciences? inclined at 30° and located at a depth of 7 km in oceanic
crust (ρ = 3,000 kg m−3).
130 |  Mechanics: How Rocks Deform

3. Given principal stresses with magnitudes of 90 and 50 MPa, 5. What does the shape of the Mohr envelope tell us about
what will the state of stress be (σn and τ) on a joint surface the strength of the material being deformed?
(Chapter 9) oriented 40° to σ1? 6. At low to moderate differential stresses, materials typically
4. Use the data presented below to determine the values of σ1 do not fail along planes that experience the highest shear
and σ3 and σn and τ acting on the two planes oriented 80° stress (θ = 45°). Why might that be the case?
and 30° with respect to σ1. 7. What is actually happening in materials that produce
straight Mohr envelopes?
Physical Space
8. A mine adit is opened at a depth of 2 km to extract gold
σ1 a = σ1 from quartz veins in granite. Experimental data indicate
30° that the cohesive strength (τ 0) of the granite is 20 MPa, its
a = 80°
coefficient of internal friction (μi) is 0.6, and its density (ρ) is
σ3 σ3 σ3 σ3 2,600 kg m−3. What is the state of stress in the gold mine?
Use the Coulomb fracture criterion (|τs| = τ 0 + σn μi) to de-
termine if there is a danger of rock bursts in the mine adit.
σ1 σ1 9. Why are earth materials generally much stronger under
compression than tension?
10. Hydrofracture measurement of principal stresses along the
𝝉 San Andreas fault at a depth of > 600 m in a drill hole by
Mohr Space
Zoback et al. (1980) yielded a value for σ3 of 141 bars and
20
for σ1 of 258 bars. Determine the value of maximum shear
10 stress at that point.
–σn σn
10 30 50

10

20
units in MPa
–𝝉

Further Reading
De Paor, D. G., 1986, A graphical approach to quantitative Means, W. D., 1976, Stress and strain: New York, Springer-Verlag,
structural geology: Journal of Geological Education, v. 34, 339 p.
p. 231–236. A clearly written introduction to the concepts of stress and strain
A graphical method is described for use in structural geology, through the principles of continuum mechanics. Mathematical
determining normal and shear stress without trigonometry. concepts and derivations are presented understandably, using
Engelder, T., 1994, Deviatoric stressitis: A virus infecting the problems (with solutions) related to geologic situations. Empha-
earth science community: EOS, v. 75, p. 209–212. sizes elastic and viscous strain.
A tongue-in-cheek discussion of deviatoric stress that is both Means, W. D., 1990, Kinematics, stress, deformation and mate-
educational and entertaining. rial behavior: Journal of Structural Geology, v. 12, p. 953–971.
Fuchs, K., and Müller, B., 2001, World stress map of the Earth: An excellent summary of the relationships between stress and
A key to tectonic processes and technological applications: strain in experimentally deformed rocks. Mohr circles are pre-
Naturwissenschaften, v. 88, p. 357–371. sented as two-dimensional tensor quantities. Also includes a
Discusses the world stress map and the techniques by which useful glossary of terms.
these data were collected. Zoback, M. L., 1992, First- and second-order patterns of stress
Hubbert, M. K., 1951, Mechanical basis for certain familiar in the continental United States: Journal of Geophysical Re-
geologic structures: Geological Society of America search, v. 97, p. 11,703–12,013.
Bulletin, v. 62, p. 355–372. Contains a number of papers on stress patterns in different parts
This is a classic study relating experimentally produced struc- of the world, along with a map by Mary Lou Zoback depicting
tures to both theory and real tectonic structures as observed in the worldwide state of stress from measurements on all the
the field. It also presents a clear, simple, and concise derivation of major plates.
the Mohr circle equations.

You might also like