Ge162 Kanamori
Ge162 Kanamori
Ge162 Kanamori
1. Introduction (2 hours)
1
4. Seismic Tomography (2 hours)
2
Primary Text Books
1. Lay, T., and T. C. Wallace, Modern Global Seismology, Academic Press, San Diego,
1-517, 1995.
Other References
6. Aki, K., and P.G. Richards, Quantitative Seismology, 2nd Edition, 685 pp., University
Science Books, Sausalito, 2002.
3
8. Shearer, P., Introduction to Seismology, Cambridge University Press, New York, 260,
1999.
4
Ge 162 Seismology
1. Introduction
Fig. 1.1
5
1.2 Source of an Earthquake
6
Fig. 1.3
The pattern of crustal deformation can be studied in detail using geodetic methods
(traditional ground-based method, GPS, SAR) and seismological methods.
7
Fracture and Frictional Sliding
Stress (τ)
Stick-slip
Stable sliding
time
Figure 1.4
F=ksΔl (1-1)
8
σ = με (1-2)
holds between stress σ and strain ε . Here μ is called the elastic constant. If the rock
is harder, then μ is larger. We have different elastic constants for shear deformation and
volumetric deformation (volume change). In general, for simple deformable bodies, like
rocks, metals etc, we need two elastic constants μ and k, rigidity and incompressibility
(also called bulk modulus), for shear and volumetric deformations, respectively. For
most seismological problems, these two elastic constants, and the density ρ are most
important. For most crustal rocks, the representative values are:
An important question is "How strong is the Earth's crust?". Many geodetic and
seismological studies have demonstrated that the change in strain (deformation)
associated with an earthquake ranges from 3x10-5 to 3x10-4, or, in terms of stress, this
corresponds to 10 to 100 bars (i.e., 1 to 10 MPa, or 10 to 100 atmospheric pressure). If
we try to break an intact piece of rock, we normally need a few kbar stress. This suggests
that an earthquake occurs on a pre-existing weak plane (fault), and frictional sliding
appears to be a more appropriate model for an earthquake.
Ground Motion
When an earthquake occurs, the ground shakes. The motion of the ground is
given by the displacement u(t) as a function of time, t, in 3 directions, usually, UD, NS,
and EW. If we take the time derivative of u(t), we get the velocity of ground motion
v(t ) = u (t ) , and if we differentiate it again, we get ground-motion acceleration
a (t ) = v(t ) = u (t ) .
The ground motions near the source of an earthquake are measured with
geological, geodetic and seismological methods (Figure 1.5).
9
For large earthquakes,
u is 1 to 20 m
v is 10 cm/sec to 3 m/sec
a is 0.1 to 20 m/sec2 (10 m/sec2 is about 1 g)
10
boundary, two types of waves exist. The first wave is mainly due to volume change, and
is called compressional wave or P wave, and the second type of wave is caused by shear
deformation, and is called shear wave or S wave. The P and S wave velocities, α and β,
respectively, are given by
k + (4 / 3) μ μ
α= and β = (1-3)
ρ ρ
11
12
Fig. 1.6 Seismograms of the 1995 Kobe, Japan, and the 1999 Izmit, Turkey,
earthquakes.
As shown above, the velocities of these waves are different, i.e., in general α >
β > CL > CR. Thus, at a station some distance from the source, the P wave arrives first,
which is followed by the S wave. Then the large amplitude Love wave and Rayleigh
wave arrive.
Seismologists study these seismic waves in detail to determine the earthquake
source parameters (the size, type of faulting etc), and the structure of the Earth.
One useful relation is that between the S-P time, t S− P , (time interval between P
and S waves) and the distance, Δ. Since t S− P is given by
13
Δ Δ Δ ⎛α ⎞
tS − P = − = ⎜ − 1⎟ (1−4)
β α α⎝β ⎠
from which
α
Δ= tS − P (1-5)
⎛α ⎞
⎜ β − 1⎟
⎝ ⎠
For the shallow part of the crust (i.e., Δ<1000 km), α is about 6 km/sec, and α / β is
1.732, so that this relation gives
14
Pendulum and Simple Mechanical Seismograph
Consider a simple pendulum (a string with a small mass hanging from it). Hold
one end of the string and let the mass swing in a vertical plane. The natural period of the
pendulum, T0, is given by,
15
l
T0 = 2π (1-7)
g
where g is the acceleration of gravity, 9.8 m/sec2. Thus if l=1 m, then the period is about
2 sec.
Suppose we have an earthquake, and the ground starts shaking horizontally.
Since you are standing on the ground (i.e., fixed to the ground), you will be shaken with
the ground. If the ground shakes very gradually, say at period T >>T0, then the mass will
move with you so that you cannot use it as a reference point (i.e., everything moves in the
same way). In this case, this pendulum is not good as a seismometer. However, if the
ground shakes very rapidly, e.g., T<<T0, then the mass tends to stay at the same place, if
not completely. This is the very principle of a horizontal seismometer, a seismometer
that measures horizontal motion. You can measure the motion of ground with respect to
the mass which is approximately stationary. Thus, if T<<T0, this pendulum is a good
seismometer, and if you record the motion of the ground (i.e., you) with respect to the
mass, you can have a seismogram.
If we use a spring with a mass hanging vertically, we can measure the vertical
ground motion with the same principle.
In the real seismometer, we need to attach a device to magnify the motion and
damp out the resonance (damper), but the basic principle is the same. The Wiechert
seismograph and the Wood-Anderson seismograph used in California are all of this type.
It is difficult to record very long-period ground motions with these simple
mechanical seismographs, because it is difficult to build a stable pendulum with a very
long natural period.
where x is the motion of a reference point (i.e., mass) of the seismograph with respect to
the ground, and y is the ground motion displacement. h is the damping constant,
ω 0 = k / m is the natural angular frequency of the seismograph ( ω0 = 2π / T0 , T0 is the
natural period), and V is the static magnification. The response of a mechanical
seismograph is completely described by these 3 constants.
16
Vx x
y = exp(iω t ) (1-9)
the output is
Vω 2
x= exp(iω t ) (1-10)
−ω 2 + 2ihω 0ω + ω 02
Vω 2
Hˆ (ω ) = (1-11)
−ω 2 + 2ihω 0ω + ω 02
17
In order to increase the sensitivity of the instrument, and improve the response at
long-period, various developments have been made. These seismographs use an electro-
magnetic sensor (moving coil etc), and a galvanometer. The examples are: Galitzin,
Benioff short-period, Benioff long-period, Press-Ewing, and Benioff strain seismographs
(Figure 1.8).
18
Until recently seismograms written on paper were the standard data for most
seismological research and routine reporting. While these analog records are still useful
for various research purposes, these instruments are limited in two respects. First, since
the movement of the pendulum is mechanically limited by the physical size of the
instrument, it is not possible to record very large ground motions; i.e., the dynamic range
is limited. Second, as long as recordings are made on paper (analog recording), the
dynamic range is limited by resolution in visually reading the records, which is normally
1/1000, i.e., 60 db.
To remove these limitations, modern seismographs adopt a force balance
mechanism and digital recording system. In the force balance mechanism, the output
signal from the transducer is amplified and fed back to a device that holds the mass at the
original unperturbed position (Figure 1.9). The strength of the signal (usually measured
in voltage) is proportional to ground motion. With some filters in the feedback circuit, it
is possible to make the output proportional to acceleration, velocity or displacement of
ground motion, at least over a certain frequency band.
In this type of instruments, there is virtually no displacement of the mass, and the
dynamic range can be increased. Also, with an appropriate feedback system, the
response can be adjusted relatively easily.
19
The response of the standard broad-band instruments used in seismology (often
called the VBB system) is approximately flat for a ground-motion velocity over a wide
frequency band (e.g. 7 Hz to 0.0033 Hz (300 sec)) (Figures 1.10 and 1.11). The broad-
band instruments usually have a 140 db (107) dynamic range.
20
21
Seismographs designed to record very strong ground motion are called strong-
motion seismographs and are used in earthquake engineering. Modern strong-motion
seismographs have a force balance mechanism with voltage output proportional to
ground-motion acceleration.
Recent developments in solid state electronics made it possible to build stable
force balance seismographs; they are now widely used in the world for research and
routine monitoring.
22
Ge 162
In the early days of seismology, the structure of the Earth was determined mainly
by using ray theory. When the wavelength of seismic waves is sufficiently short (i.e., if
the period is sufficiently short), we can treat seismic waves as a geometrical ray, just as
we do in geometrical optics.
In ray theory, the most fundamental is Snell's Law, which is illustrated in Figure
1.
23
Snell’s Law
i1 sin i1 sin i2
=
v1 v2
v1
v2
i2
i
sin i( z )
= p = const
v( z )
r sin i (r ) r sin i (r )
v(z) =p =p
z v(r ) v( r )
Fig. 1Figure 1
Suppose a ray is incident from a medium with a wave speed v1 on a medium with
a speed v2 . Let the incident and emergent angles be i1 and i2 . Then the Snell's law is
given by
sin i1 sin i2
= (1)
v1 v2
This can be shown easily from the geometry of the two triangles OAB and OBC
shown below.
24
If the wave speed changes continuously with z (i.e., depth) as v(z), then (1) can be
written as
sin i
= p =constant for a given ray (2)
v
For a spherical geometry as shown in Figure 1, the Snell's law can be written as
rk sin ik rk +1 sin ik +1
= (3)
vk vk +1
and
25
r sin i (r )
=p (4)
v( r )
Figure 2 shows various seismic rays in the Earth's interior. P and S waves are
denoted by P and S, respectively. Other symbols are:
26
27
P, S, Pdif, Sdif pP, sS, PP, SS PcP, ScS ScP, PcS
Storchak, D., Schweitzer, J., and Bormann, P., The IASPEI Standard Phase List, Seismological Research Letters,
74, 761-772, 2003.
28
Some Examples
29
Figure 5 shows ray paths in the Earth's interior for 3 representative velocity structures,
and the corresponding travel time curves.
30
31
Figure 6 shows 3 regions in the Earth's interior and the corresponding travel time
curves.
Figure 7 shows the crustal structure for ocean and continent, and corresponding
travel time curves.
32
33
2.1.3 Travel-Time Curve (LW pp.213-217)
Figure 8 shows the travel times observed at many stations and reported to the
International Seismological Center (ISC).
34
35
Figure 9 shows the same with the phase names labeled.
36
For a spherically symmetric structure, the distance, Δ , traveled by a seismic ray
with a ray parameter p, and the travel times, T , can be computed easily. Referring to the
figure below,
dr v2 p2
= cos i = 1 − sin 2 i = 1 − 2 (5)
ds r
ds sin i 1 vp
dΔ = = 2 dr (6)
r r v2 p2
1− 2
r
37
Then,
∫
r0
r
T =2 dr (7)
rp v r 2 − v2 p2
∫
r0
vp
Δ=2 dr (8)
rp r r − v2 p2
2
where r0 is the radius of Earth and rp is the radial distance of the deepest point of the
ray. These integrals are fundamental in the seismological ray theory (e.g., Herglotz
Wiechert Method).
In the early days of seismology, the structure of the Earth was determined from
the travel-time data, as shown in the upper figure of Figure 10. This figure shows one of
the standard laterally homogeneous model, which is used for various seismological
studies.
The bottom figure in Figure 10 shows an example of 3-D Earth structure
determined by more recent studies using seismic body waves, surface waves and normal-
mode data (Helffrich and Wood, 2001). Tomographic methods are used for determining
these structures.
38
References
Helffrich, G. R., and B. J. Wood, The earth's mantle, Nature, 412, 501-507, 2001.
Anderson, D. L., Top-down tectonics, Science, 293, 2016-2018, 2001.
39
Ge 162 Plate Motion and Great Earthquakes
Earthquakes occur in the Earth's crust and mantle due to stresses caused by global
plate motion. The actual pattern of stress distribution is probably very complex, but we
expect that the activities of great and large earthquakes must reflect the global plate
motion.
40
The world greatest earthquakes occur at subduction zones (e.g., 1960 Chilean
earthquake, and the 1964 Alaskan earthquake), but not every subduction zone has
seismicity. With this caveat in mind, we investigate the level of seismic activity and
plate motion. Ideally, the seismic activity along a subduction zone should be defined by
the energy release per unit length along the subduction zone, and unit time, i.e.,
41
1 L T
e=
LT ∫ ∫
0 0
ER dldt
where L and T are the length of the subduction zone and the time period involved,
respectively.
Unfortunately, the available seismic record is too short to compute this. So, we
subduction zone as a parameter that represents e for that subduction zone. Then , it is
Mw ∝V
where V is the convergence rate. However, the plot of M w versus V does not show any
obvious trend. This suggests that other factors may be controlling seismicity. Another
42
Then, we can try a 3-parameter regression between M w , V and T. The result is
shown in the following figure. The horizontal axis shows the observed M w and the
43
(Ruff, L., and H. Kanamori, Seismicity and the subduction process, Phys. Earth Planet.
Inter., 23, 240-252, 1980)
If this regression is valid, this provides a useful method for assessing the seismic
potential of subduction zones for which no great earthquake has occurred. This pattern
suggests that the subduction zones where a relatively young plate is subducting at a
44
relatively fast rate are more likely to have great earthquakes, and those with an old plate
subducting at a moderate rate are less likely to have great earthquakes. The end-member
subduction zones are the Chilean type and the Mariana type, shown below.
(Uyeda, S., and H. Kanamori, Back-arc opening and the mode of subduction, J. Geophys.
Res., 84 (B3), 1049-1061, 1979)
45
Another interesting implication of this correlation is the seismic potential of the
Pacific Northwest (i.e., Oregon-Washington coast). The Juan de Fuca plate is subducting
beneath the states of Oregon and Washington. The background seismicity there is very
low, as shown below, and until mid 1980's, it was generally believed that the seismic
potential in the Pacific Northwest is low (i.e., great earthquakes are unlikely). However,
the age of the Juan de Fuca plate is very young, about 10 My, and it is subducting at a
rate of 3 cm/year. Thus, in view of the regression relation shown above, one would
expect a large, M w =8.5 to 9, earthquake there. This suggestion motivated the interest of
Geological evidence for regional submergence and evidence for large tsunami which
occurred in 1700 [Satake et al., 1996] now seem to have convinced most people, which
seems to have led to upgrading of building code in the area. This is a good example in
46
(Heaton, T., and H. Kanamori, Seismic potential associated with subduction in the
northwestern United States, Seismol. Soc. Am. Bull., 74 (3), 933-941, 1984)
47
(see, Atwater, B. F., and others, Summary of coastal geologic evidence for past great
earthquakes at the Cascadia subduction zone, Earthquake Spectra, 11, 1-18, 1995)
48
(R. S. Yeats, Living with Earthquakes in the Pacific Northwest, Oregon State University
Press, 1998)
References
Brune, J., Seismic moment, seismicity, and rate of slip along major fault zones, J.
Geophys. Res., 73, 777-784, 1968.
Satake, K., Shimazaki, K., Tsuji, Y. and Ueda, K., Time and size of a giant earthquake in
Cascadia inferred from Japanese tsunami records of January 1700, Nature, 379, 246-249,
1996.
49
Ge 162
where the function f ( ) gives the travel time between the source and station i.
50
Here, ro is the hypocenter location (i.e., the location of the beginning of an earthquake),
ri is the location of the i-th station, v(r ) is the wave speed which is in general a function
[( xi − xo ) 2 + ( yi − yo ) 2 + ( zi − zo ) 2 ]1/ 2
f (ro , ri , v(r )) = (2)
v
t1 , t2 , t3 , t4 , ...t N .
51
This problem is a nonlinear problem even for the simplest case for a
homogeneous medium.
To solve a nonlinear problem like this, we start from a first approximation
Then writing
and taking the first order terms in δ xo etc and δ to , we set up linear equations for δ xo etc
and δ to as,
∂f (ri ) ∂f (ri ) ∂f ( ri )
ti − ti0 = δ x0 + δ y0 + δ z0 + δ t0 (5)
∂xo 0 ∂yo 0 ∂zo 0
(i=1, 2. 3, .....N), where ti0 is the arrival time at station i computed for the first
approximation (i.e., ti0 = to0 +(travel time computed for the first approximation)).
This problem can be solved by the method of least squares. By iterating this, we
can determine ro = ( xo , yo , zo ) and to which best fit the observed travel times. More
details will be discussed in the practice session.
52
Earthquake Magnitude (LW, pp. 379-385)
where A is the amplitude of the observed seismic waves (body waves, surface waves, or
unspecified), and f (Δ ) is an amplitude attenuation curve as a function of distance
determined for specific type of waves.
For example, in case of the traditional local magnitude M L , the amplitude A is the
amplitude of the Wood Anderson seismogram in mm, and f (Δ) is given by a table, or by
a nomogram such as that shown in Figure 2.
53
In case of the surface-wave magnitude, M S , A is the ground-motion amplitude of
54
logER = 1.5M S + 4.8 (joule) (8)
M 0 = μ DS (9)
where D is the fault offset, S is the fault area, and μ is the rigidity of the crust
surrounding the fault. The unit of M 0 is N-m. Unlike other magnitude scales, M W
represents a specific earthquake source parameter, the overall static size of an earthquake
given by M 0 . The relation between M w and M 0 is given by
As shown in Figure 3, the fault motion on a vertical strike slip fault would
produce compressional and dilatational quadrants in the Earth's crust.
55
The pattern of compression and dilatation can be detected by the first motion of P
waves. In the compressional quadrant P wave is up and in the dilatational quadrant, it is
down. As viewed from above, the sense of the first motion (up or down, or compression
or dilatation) alternates in quadrant. The planes separating the compressional quadrant
and the dilatational quadrant are called the nodal planes. The fault plane coincides with
56
one of the nodal planes. The other nodal plane is called the auxiliary plane. With this
method alone, we cannot distinguish the fault plane from the nodal plane, i.e., the fault
plane can be either one of the nodal planes.
Thus, from the observations of P wave first-motion data from many stations
surrounding the source we can determine the geometry of the faulting. The pattern thus
determined is usually referred to as "Mechanism of Earthquake". Faults with different
types (e.g., strike slip fault, thrust fault, and normal fault) produce different radiation
patterns of P waves. This can be easily seen, if we consider a small sphere surrounding
the source. This sphere is called the focal sphere (Figure 3).
The radiation pattern is three dimensional, and the surface of the focal sphere is
divided into quadrants of compression and dilatation. (This can be best understood using
a worn-out tennis ball with compressional quadrants painted dark.)
We need to show this three-dimensional pattern on the focal sphere on a piece of
paper. Since the pattern is point symmetric with respect to the center of the focal sphere,
we need to show only the pattern on a hemisphere. It is conventional to show the lower
focal hemisphere, but in rare cases, the upper hemisphere or the side hemisphere is
shown. We use a standard projection method, most commonly the equal-area
stereographic projection, to project the lower hemisphere to a flat horizontal plane
(Figure 4).
57
Some examples are shown in Figure 5. The projected diagram is called the
mechanism diagram. As shown in Figure 5, the stereographic mechanism diagram is
intuitive for understanding the geometry of faulting (more details in the practice session).
58
Figure 6 shows the mechanism of large earthquakes along the Circum-Pacific
belt. Most of them are low-angle thrust mechanisms which are consistent with subduction
of the Nazca and the Pacific plates beneath the South American, the North American, and
the Eurasian plates. The normal fault events represents tensional failure within the
oceanic plate upon bending caused by subduction.
59
Ge 162 Practice Session 1 Locating Earthquakes
60
Data
Table-1 is the travel time data obtained using "STP" which is a SCEC (Southern
California Earthquake Center) tool to extract earthquake data (parameter data and
waveform data). (For details of STP, see the SCEC Web site). The data are rearranged
in (x, y) coordinate (in km) with the Pasadena station (Latitude=34.1484°, Longitude=-
118.1711° ) as the origin. The origin time of the arrival times is arbitrary. Table-1 gives
only the first 10 stations. The actual data are in loc_dat_1 (all the data and program files
are in a FTP site on ftp.gps.caltech.edu /home/ftp/pub/hiroo/ge162.dir).
Table 1.
Earthquake-1
Reference Station (Origin of (x,y)) PAS 34.1484 -118.1711
Locate the earthquake taking the following steps. Ideally, you should write your
own program to carry out 2 to 6, but if you find it difficult to do so, you can use a simple
program, eqloc.f . To do 7, you will need to use this program. In case you use this
program, try to follow the steps taken in the program.
61
1. Use the first approximation, (0.0, 0.0, -10.0, 0.0).
2. Compute the travel times and the partial derivatives in a homogeneous medium with
v=6 km/s which is a good average for the shallow crust. Refer to equation (1) to (5) in
class note 2.2.
62
which we write as
Am = d (3)
and m and d are column vectors containing the parameters to be determined and the
data, respectively, i.e.,
63
⎛ t1 − t10 ⎞
⎜ 0 ⎟
⎛ δ x0 ⎞ ⎜ t2 − t2 ⎟
⎜ ⎜ t3 − t30 ⎟
⎜ δ y0 ⎟⎟ ⎜ ⎟
m= , and d =⎜ . ⎟ (5)
⎜ δ z0 ⎟
⎜⎜ ⎟⎟ ⎜ . ⎟
⎝ δ t0 ⎠ ⎜ ⎟
⎜ . ⎟
⎜t − t0 ⎟
⎝ N N⎠
4. Determine m .
AT Am = AT d (6)
m = ( AT A) −1 AT d (7)
and the error estimates are determined by the variance of the data and the diagonal
elements of the inverse matrix of the normal equation (6 ). Usually, we write the
uncertainty in mi by Δmi , and compute it by
∑ (t − t ) /( N − N ) ,
N
Δmi = cii j
c 2
j p i=1, 2, 3 (8)
j =1
64
where N p is the number of parameters (here 4), cii are the diagonal elements of ( AT A) −1 ,
x0 = 0 + δ x0
y0 = 0 + δ y0 (9)
z0 = −10. + δ z0
t0 = 0 + δ t0
6. Iterate 2, 3, 4, 5.
i_eqloc
loc_dat_1 : Name of the travel-time data file
0.0 0.0 -10.0 0.0 : 1st approx.
c_eqloc
10 : maximum number of interations
half_space
: name of the structure
s-cal.pvel
half_space
half_space
65
1 : number of layers
9999. 6.0 : layer thickness, α
s-cal.pvel
southern cal. P structure, with a slightly low surface velocity
5
1.0 4.0
3. 5.5
23.4 6.3
5. 6.8
9999.0 7.8
7. If the second line of c_eqloc is replaced by s-cal.pvel, it will use a more general
subroutine which computes the travel times etc for a layered model given by s-cal.pvel.
Try eqloc.f with s-cal.pvel. Output of eqloc is in o_eqloc.
9. Remove the data for which ti is less than 5 sec, and locate the event.
Example output
o_eqloc
10
southern california P structure, with a
1.000 4.000
3.000 5.500
23.400 6.300
5.000 6.800
9999.000 7.800
loc_dat_1
0.000 0.000 -10.000 0.000
Earthquake-1
PAS 34.148 -118.171
x0= 6.778 0.319
y0= -4.360 0.378
66
z0= -10.676 1.159
th0= 0.017 0.054
34.148 -118.171 -10.000 0.000
x0= 6.790 0.316
y0= -4.186 0.381
z0= -8.184 1.079
th0= 0.027 0.054
34.109 -118.098 -10.676 0.017
x0= 6.782 0.308
y0= -4.172 0.367
z0= -7.645 1.180
th0= 0.030 0.048
34.111 -118.098 -8.184 0.027
x0= 6.782 0.307
y0= -4.167 0.365
z0= -7.533 1.217
th0= 0.032 0.047
34.111 -118.098 -7.645 0.030
...........
station data
...........
0.3273 RMS of residuals
67
Data
Table-1 is the phase data obtained using "STP" which is a SCEC tool to extract
earthquake data (parametric data and waveform data). (For details of STP, see the SCEC Web
site). The data contain station names, first motion data (C or D), quality (Q, ignore in this
problem), azimuth, and take-off angle. The azimuth and take-off angle, ih, (measured from
downward vertical) are computed using a standard southern California structure. Table-1 lists
only selected 10 stations. The actual data are in mech_dat_1 in
/home/ftp/pub/hiroo/ge162.dir/practice2.dir.
1. Compute the radial distance on a mechanism diagram using the equal-area projection,
68
Remember that if ih is larger than 90°, then 180° must be added to the azimuth (i.e., the
station must be plotted in the opposite azimuth.), and ih must be changed to 180°-ih.
2. Plot the first-motion (filled circle for compression and open circle for dilatation) on a
mechanism diagram. (Plot the data for the 10 stations listed in Table-1 manually, or with your
own program).
i_mplotr_2
Earthquake(mech)-1 : Job ID
f : fault(f) or moment tensor(m)
85. 180. 350. : dip, rake, and fault strike
mech_dat_1 : file name of the data
(For plotting only the first-motion data, the 2nd and 3rd lines are irrelevant.)
c_mplotr_2
7.0 0.1 10. 5.0 0 0 1 0 : these parameters control the type and
style of the plot
(change plot_opt parameter (2nd from the last) only. 1 for the first-motion
data only (in this case, the fault parameters are ignored), 3 for the data +
nodal lines)
69
4. The final step is to determine the mechanism by drawing two orthogonal nodal planes so that
they divide the compressional and dilatational stations. Usually there are always some
inconsistent stations, but try to find the best solution. Many methods have been developed, but
here try a few mechanisms. mplotr_2.f draws 2 nodal lines corresponding to the fault
mechanism given by dip, δ , rake, λ , and fault strike, φ f (run mplotr_2 with plot_opt=3 in
70
5. Assume that the fault strike is -45° from N. Plot the mechanism diagrams for (1) right-
lateral vertical strike slip, (2) vertical dip slip (north-east side down), (3) thrust fault
dipping 20° NE, (4) normal fault dipping 45° NE, and (4) an oblique-slip mechanism
(e.g., add some right-lateral component to (3)). (This problem has nothing to do with
71
the first-motion data given in mech_dat_1. You can just draw a sketch of mechanism
diagrams, or run mplotr_2.f with
c_mplotr_2
7.0 0.1 10. 5.0 0 50 4 0
.)
Ge162
3. In continental interiors.
Figures 1 and 2 show seismicity in the world and California, and Figures 3 and 4 show
72
73
The world largest earthquakes occur along subduction zones (e.g., the 1960 Chilean
earthquake, Mw=9.5, the 1964 Alaskan earthquake, Mw=9.2). More than 75 % of the seismic
energy release takes place there. Most of these events represent slip on the interface between a
subducting oceanic plate and an overriding plate (Figure 3). In these zones deep focus
74
75
Most earthquakes along ridge-transform systems are shallow and relatively small (10% in
energy release). The events on ridges have normal-fault mechanism, and those along the
transform boundaries have strike-slip mechanisms. Transform fault events are generally larger;
occasionally the magnitude reaches 8 (e.g., the 1906 San Francisco earthquake, Mw≅8 ). No deep
very diffuse. Partly because of their proximity to major population centers, large intra-
76
continental events are often very devastating (e.g., the 1976 Tanshang earthquake, the
zones (e.g. 1970 Colombia, d=653 km, Mw=8.1; 1954 Spain, d=640 km, M=7.8, 1994
Bolivia, d=635 km, Mw=8.3). The largest recorded deep focus earthquake is the 1994
Bolivia earthquake.
77
A pronounced peak in the energy release is seen at a depth of 600 km, just before
Figure 6 shows the temporal variation of seismicity. The energy release during the
period from 1952 to1965 dominates. This peak is a result of five large subduction-zone
earthquakes in the Pacific (1952 Kamchatka, 1957 Aleutian Is., 1960 Chile, 1964 Alaska,
78
1965 Aleutian Is.). The energy release rate is not uniform in time, and fluctuates on a
1/4 of the energy released in volcanic eruptions, and 0.05 % of the terrestrial heat flow.
79
80
81
2.3.4 Magnitude-Frequency Relation (Gutenberg-Richter Relation)
log N ( M ) = a − bM
82
The results obtained for many regions indicate that the value of b (called b value)
is approximately equal to 1.
2.3.5 Aftershocks
83
After a large earthquake (main shock), many smaller earthquakes occur near the
epicenter of the earthquake. The decay of aftershock activity follows the Omori's law
given by
K
n(t ) =
t +c
where n(t ) is the number of aftershocks larger than a given magnitude per unit time. A
K
n(t ) =
(t + c) p
3.1.1 Stress
Body Force
The body force f is defined by a force per unit mass in a medium (Figure 1).
Then, the body force per unit volume is ρ f where ρ is the density of the medium.
84
Then the force acting on a volume element dV is
ρ fdV
∫ ρ f dV
V
Surface Force
85
The surface force is the force distributed over a surface of the body, either internal
or external. Usually it is defined by the force per unit area, f S . Then the force acting on a
surface element dS is f S dS .
Stress
Consider a deformed elastic body in equilibrium (see Figure 1). Let dS be a surface
element at P which divides the medium on the + side and the - side. A unit normal vector
n ( n = 1 ) is taken from the - side to the + side. In equilibrium, the force, F+ , exerted by
the + side on dS should be balanced by the force, F− exerted by the - side on dS , i.e.,
F+ + F− = 0 .
F+
f n = lim
dS →0 dS
f n is a vector (often called a stress vector, or traction) and its dimension is force/area.
components of f n by ( f1n , f 2 n , f3n ) . Note that f n is a function of not only the location
of P but also the orientation of n . Hence, in order to specify the stress at P uniquely, we
need two vectors n and f n .
This situation can be understood more easily in the simple example shown in
Figure 2.
86
Consider an elastic beam with one end, AB, clamped at the wall. Then, apply a
force F uniformly on the surface S at the other end BC. Let us consider the stress at P. In
(a), we consider the stress acting on S1 that is perpendicular to the axis of the beam. In
this case, it is obvious that
( f1n , f 2 n ) = ( F / S1 , 0)
In (b), we consider S2 which is parallel to the axis. Suppose we cut the beam in two
parts along S2. The beam will be still in equilibrium without change in shape. That is, there
( f1n , f 2 n ) = (0, 0)
Stress Tensor
87
For simplicity, we consider a 2-dimensional problem. The section shown in
Figure 3 depicts a 2-D medium extending to infinity in x3 direction (perpendicular to the
face of the paper).
Consider a surface element dS1 normal to the x1 axis. We call the medium on the
+ x1 side M+ and that on the - x1 side, M- (Figure 4). Let σ 11 and σ 21 be the x1 and
88
Note that the first subscript denotes the component, and the second, the direction
of the normal to the surface. σ ij ' s with any i and j can be defined similarly.
n1 : x1 component of n , n1 = cos θ
n2 : x2 component of n , n2 = sin θ
89
Then the forces acting on BOA are given as follows.
x1 component x2 component
force on S1 −σ 11S1 −σ 21S1
force on S2 −σ 12 S 2 −σ 22 S 2
force on S σ 1n S σ 2n S
In equilibrium, the total force should vanish. Since S1 = S cos θ = Sn1 , and
S 2 = S sin θ = Sn2 , we obtain
σ 1n = σ 11n1 + σ 12 n2
σ 2 n = σ 21n1 + σ 22 n2 (1)
or, in matrix notation,
90
⎛ σ 1n ⎞ ⎛ σ 11 σ 12 ⎞ ⎛ n1 ⎞
⎜ ⎟=⎜ ⎟⎜ ⎟ (2)
⎝ σ 2 n ⎠ ⎝ σ 21 σ 22 ⎠ ⎝ n2 ⎠
σ 12 = σ 21 (3)
⎛ σ 1n ⎞ ⎛ σ 11 σ 12 σ 13 ⎞ ⎛ n1 ⎞
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ σ 2 n ⎟ = ⎜ σ 21 σ 22 σ 23 ⎟ ⎜ n2 ⎟ (4)
⎜σ ⎟ ⎜σ ⎟⎜ ⎟
⎝ 3n ⎠ ⎝ 31 σ 32 σ 33 ⎠ ⎝ n3 ⎠
91
know the stress tensor ( σ ij ) at P, we can calculate, using (4), the stresses on any surface
In the above equilibrium analysis, we ignored body forces compared with surface
forces. This is justified because if we consider a small volume around P with a linear
dimension da, then the total body force is proportional to da3 while the total surface force
is proportional to da2. Hence as da → 0, the body forces can be ignored.
and the other perpendicular to n , σ nt (see Figure 8). σ nn is called the normal stress, and
always possible to choose a Cartesian coordinate system ( x1' , x2' , x3' ) for which
σ i ' j ' = 0 (i ' ≠ j ') . In other words, for this new coordinate system ( σ i ' j ' ) is a diagonal
matrix. The non-zero diagonal components, σ 1'1' , σ 2'2' , and σ 3'3' are called the principal
92
stresses, and x1' , x2' , and x3' axes, the principal axes. It can be shown that
Equations of Motion
The equation of motion for a small part of it, ABCD (length dx1 ) is given by,
93
ρ Sdx1u1 = ρ f1Sdx1 + σ 11 ( x1 + dx1 ) S − σ 11 ( x1 ) S
where ρ is the density, u1 is the displacement, and f1 is the body force. Expanding
σ 11 ( x1 + dx1 ) around x1 , and retaining only the first order terms in dx1 , we obtain,
∂σ 11
ρ u1 = ρ f1 + (5)
∂x1
3 ∂σ ij
ρ ui = ρ fi + ∑ (i=1,2,3) (6)
j =1 ∂x j
These are the equations of motion expressed in terms of the stress components.
Tensor Notation
In tensor notation, if any suffix occurs twice in a single term, it is to be put equal
to 1, 2, and 3 in turn and the results are to be added. For example,
3
aii = a11 + a22 + a33 = ∑ aii
i =1
3
aip bpj = ∑ aik bkj
k =1
94
Also, we use ,j to denote differentiation by x j . For example,
∂ui
= ui , j
∂x j
⎧1 if i = j
δ ij = ⎨
⎩0 if i ≠ j
Note that δ ii = 3 .
ρ ui = ρ fi + σ il ,l (7)
Boundary Conditions
From the definition of the stress, it is evident that the normal and the tangential
stresses should be continuous across any surface. In particular, at the free surface, there is
no force acting on it; hence the normal and the shear stresses should vanish there. If the
free surface is perpendicular to x3 axis,
95
σ 31 = σ 32 = σ 33 = 0
3.1.2 Strain
96
rigid body translation. Thus du is considered to represent rotation and deformation. To
the first order,
1 1
U = (U + U T ) + (U − U T ) (9)
2 2
1 1
We denote (U + U T ) by D and (U − U T ) by R. D is symmetric and R is anti-
2 2
symmetric. We will show that D represents deformation, and R represents rigid-body
rotation.
1
The elements of D, (ui , j + u j ,i ) = eij can be interpreted as follows.
2
If only e11 ≠ 0 ,
We then have, du1 = e11dx1 , du2 = 0 , and du3 = 0 . This means that the line element dx1 in
97
x1 direction is stretched by du1 = e11dx1 in x1 direction. Hence, e11 represents extension
(or contraction if e11 <0) per unit length in x1 direction. e22 and e33 can be interpreted
similarly.
Next consider e12 , and e21 (= e12 ).
i.e., du1 = e12 dx2 , du2 = e21dx1 , and du3 = 0 . As shown in Figure 11, the angle between x1
98
D is called the strain tensor (it can be shown that ( eij ) is a tensor).
e11 , e22 , and e33 represent extension or contraction, and e12 , e13 , and e23 represent shear.
⎛ 1 1 ⎞
⎜ 0 (u1,2 − u2,1 ) (u1,3 − u3,1 ) ⎟
2 2
⎜ ⎟
1 1
R = ⎜ (u2,1 − u1,2 ) 0 (u2,3 − u3,2 ) ⎟
⎜2 2 ⎟
⎜ ⎟
⎜⎜ 1 (u3,1 − u1,3 ) 1
(u3,2 − u2,3 ) 0 ⎟⎟
⎝2 2 ⎠
Define ω1 , ω 2 , and ω3 by
Then,
⎛ 0 −ω3 ω2 ⎞
⎜ ⎟
R = ⎜ ω3 0 −ω1 ⎟
⎜ −ω ω1 0 ⎟⎠
⎝ 2
99
Then, we have
i.e., du1 = −ω3dx2, du2 = ω3dx1, and du3 = 0 . As shown in Figure 12, this displacement
The relation between stress and strain is the extension of the Hooke’s law for a
spring. (i.e., F = k Δl where F is the force, Δl is the length change and k is the spring
constant.)
We assume that the material is isotropic and perfectly elastic. If the medium is
100
perfectly elastic, the stress should be expressed as a homogeneous linear function of
strain. Since there are six independent stress components and strain components, in
general there can be 6x6=36 constants. However, if the material is isotropic, we can show
that there are only two independent constants.
Consider an elastic parallelepiped shown in Figure 13. Apply a normal stress σ 11
in x1 direction. The extension in x1 direction is e11 (Figure 13). In the linear theory, e11 is
proportional to σ 11
1
e11 = σ 11
E
1
The constant of proportionality is written as , and is called the Young’s
E
modulus. Note that e11 is non-dimensional, so that E has the dimension of stress. Under
this stress, there will be contraction in x2 and x3 directions that is proportional to e11 .
Since the material is isotropic,
ν ν
e22 = −ν e11 = − σ 11 and e33 = −ν e11 = − σ 11
E E
101
ν is called the Poisson’s ratio.
If we apply σ 11 , σ 22 , and σ 33 in x1 , x2 and x3 directions simultaneously, then by
superposition,
σ 11 ν
e11 = − (σ 22 + σ 33 )
E E
σ 22 ν
e22 = − (σ 11 + σ 33 ) (10)
E E
σ 33 ν
e33 = − (σ 11 + σ 22 )
E E
(1 − 2ν )
Δ= Σ (11)
E
where
and
Σ = σ 11 + σ 22 + σ 33
102
Then, to the first order, the volume change is
Hence,
Δ = dV / V0
Thus, Δ represents relative change in the volume, and is called the volumetric
strain or dilatation.
If σ 11 = σ 22 = σ 33 = σ , then from (11), we obtain
E
σ= Δ = kΔ
3(1 − 2ν )
where
E
k= (13)
3(1 − 2ν )
103
is called the bulk modulus or incompressibility.
νE E
σ 11 = Δ+ e11
(1 − 2ν )(1 + ν ) (1 + ν )
νE E
σ 22 = Δ+ e22
(1 − 2ν )(1 + ν ) (1 + ν )
νE E
σ 33 = Δ+ e33
(1 − 2ν )(1 + ν ) (1 + ν )
νE
λ= (14)
(1 − 2ν )(1 + ν )
E
μ= (15)
2(1 + ν )
Then,
σ 11 = λΔ + 2μ e11
σ 22 = λΔ + 2μ e22 (16)
σ 33 = λΔ + 2μ e33
Next, we consider shear stress and shear strain. From Figure 15, we see that the
shear strain e12 is caused by the shear stress σ 12 :
σ 12 = 2Ge12
104
Similarly,
σ 13 = 2Ge13
σ 23 = 2Ge23
σ 12 = 2μ e12
σ 13 = 2 μ e13 (17)
σ 23 = 2μ e23
Equations (16) and (17) give the stress-strain relations in isotropic media. (16) and (17)
can be written collectively as
105
Note that, although we have introduced five elastic constants E , ν , k , λ , and μ above,
there are only two independent constants. If we choose E and ν as the basic constants
then
E νE E
k= , λ= , μ= (19)
3(1 − 2ν ) (1 + ν )(1 − 2ν ) 2(1 + ν )
λ 3λ + 2μ 2
ν= , E= μ, k = λ + μ (20)
2(λ + μ ) λ+μ 3
In the theory of elasticity, the following definitions and relations are often used.
Here, φ ( x1 , x2 , x3 ) is a scalar function and u (u1 , u2 , u3 ) and v (v1 , v2 , v3 ) are vectors.
We assume that these functions are continuous and differentiable.
⎛ ∂φ ∂φ ∂φ ⎞
⎜ , , ⎟ (21)
⎝ ∂x1 ∂x2 ∂x3 ⎠
⎛ ∂ ∂ ∂ ⎞
∇≡⎜ , , ⎟
⎝ ∂x1 ∂x2 ∂x3 ⎠
106
and write it as ∇φ .
∂ 2φ ∂ 2φ ∂ 2φ
∇ 2φ = + + = φ,ll (24)
∂x12 ∂x 22 ∂x32
7) curlgradφ ≡ 0
8) divcurlu ≡ 0
107
9) If curlu = 0 , u is called an irrotational vector, and can be written as u = gradφ .
φ is the scalar potential.
10) If divu =0, then u is called a solenoidal vector, and can be written as u = curlv .
v is the vector potential.
11) Any vector field u can be decomposed into an irrotational field u I and a
solenoidal field u II , i.e.
u = u I + u II
where curlu I =0 and divu II =0. Using 9) and 10) u can be written as
Relations 6), 7), and 8) can be easily verified. Proof of 9), 10) and 11) requires some
knowledge of Potential Theory.
σ ij , j = λΔ , jδ ij + μ (ui , jj + u j ,ij )
= λΔ ,i + μ (∇ 2ui + Δ ,i ) = (λ + μ )Δ ,i + μ∇ 2ui
where we assume that the medium is homogeneous, i.e., λ and μ are constants.
ρ ui = ρ fi + (λ + μ )Δ ,i + μ∇ 2ui
or in a vector form,
108
ρ u = ρ f + (λ + μ )graddivu + μ∇ 2u (28)
or
ρ u = ρ f + (λ + 2μ )∇ 2u + (λ + μ )curlcurlu (30)
Equations (28), (29) and (30) are among the most fundamental equations in seismology.
ρ u I = ρ f + (λ + 2μ )∇ 2u I (31)
ρ u II = ρ f + μ∇ 2u II (32)
(31) and (32) are the wave equations for the irrotational field and the solenoidal field,
respectively.
109
Ge 162 Problem #3
1. Show that the Lame's elastic constant μ=E/2(1+ν) (E: Young's modulus, ν: Poisson's
ratio) is actually equal to the shear modulus G defined by
σ ij = 2Geij (i ≠ j)
110
Consider equilibrium of a beam with the square cross section abcd in the
parallelpiped. The square abcd is now deformed into a parallelogram a'b'c'd' as shown in
Figure 2.
Show that
ob' (1 + ν )
=1 − σ (1)
ob E
oa' (1 + ν )
=1 + σ (2)
oa E
2) Now consider equilibrium of a triangular beam with the cross section aob. The normal
stress is -σ on oa and σ on ob.
Show that
where σt and σn are the shear and normal stresses on ab, respectively (see Figure 1).
3) Referring to the square abcd, this deformation can be viewed as shear deformation due
to the shear stress σt .
Show that the corresponding shear strain et is given by
ob ' E − (1 + ν )σ
tan((π / 4) − et ) = = (4)
oa ' E + (1 + ν )σ
111
Assuming that et is small (i.e. you can put sinet=et , coset=1), obtain the expression for G
by relating et to σt .
Ge 162, Problem #2
A line element PQ becomes P'Q' after deformation. The strain components for
this deformation are given by e11, e22, and e12.
Determine the unit elongation in the direction of PQ and the shear strain for the
directions of PQ and PT (PT is perpendicular to PQ). Follow the steps given below.
112
1. Referring to the geometry shown in Fig. 1, show that the difference in the length
between PQ" and PQ (elongation in PQ direction) is given by:
Then dividing this by PQ, show that the unit elongation in the direction of PQ is
given by:
∂u1 ∂u dx
( Use the relations like du1 = dx1 + 1 dx2 and 1 = cos α etc.)
∂x1 ∂x2 PQ
2. Referring to the geometry shown in Fig. 1, show that the angle through which PQ is
rotated is given by:
3. The line segment PT makes an angle α+π/2 with the x1 axis. Using the result obtained
above, show that the rotation of PT is given by:
∂ u2 2 ∂u
sin α − (e22 − e11 ) cos α sin α − 1 cos 2 α
∂ x1 ∂ x2
4. Using the results of 2 and 3, show that the shear strain for the directions of PQ and PT
(i.e. 1/2 of the change in angle between PQ and PT) is given by:
From this result we see that there are two values of α, differing by π/2, for which
shear strain vanishes. They are given by:
113
2e12
tan 2α =
e11 − e22
Ge 162 Problem #1
Consider a (infinitesimally) small prism BOA at P with the three sides parallel to
the x3 axis. Let n and t be the unit vectors normal and parallel to BA respectively (the
directions are shown in the figure.). In class, we showed that the x1 and x2 components of
stress acting on plane BA are given by,
⎛ σ 1n ⎞ ⎛ σ 11 σ 12 ⎞ ⎛ n1 ⎞
⎜ ⎟=⎜ ⎟⎜ ⎟
⎝ σ 2 n ⎠ ⎝ σ 21 σ 22 ⎠ ⎝ n2 ⎠
1. Show that the normal stress and shear stress on the plane BA are given by,
114
σ nt = σ 12 (cos 2 θ − sin 2 θ ) + (σ 22 − σ 11 ) cosθ sin θ
2σ 12
tan 2θ =
σ 11 − σ 22
the shear stress vanishes.
The directions of n and t for this θ are called the principal directions.
If we take x1' and x2' axes in n and t directions, they are the principal axes. The normal
stresses on the plane normal to x1' and x2' axes are the principal stresses.
σ nn = σ 11 cos 2 θ + σ 22 sin 2 θ
σ nt = (σ 22 − σ 11 ) cos θ sin θ
2. Refer to Fig. 2.
115
3.2 Wave Equation and Seismic Waves
Wave Equations
Following (31) and (32) of 3.1.4, and ignoring the body force, we obtain for the
irrotational field u I ( curlu I = 0 ),
u I = α 2∇ 2u I (1)
where
λ + 2μ
α= (2)
ρ
u II = β 2∇ 2u II (3)
where
116
μ
β= (4)
ρ
Equations (1) and (3) are three-dimensional wave equations which are of fundamental
importance in seismology.
We first examine the property of u I . Here we use a coordinate system (x, y, z) instead of
(xl, x2, x3), and denote the x, y, and z components of displacement by u, v, and w, respectively.
Let us consider a plane elastic wave propagating in the x direction, that is, a wave in
which u I is a function of x and t. Since all derivatives with respect to y and z are zero, we have
from curlu I = 0 ,
∂w ∂v
= 0 , and =0
∂x ∂x
which give v = w =0. (Actually v and w are constant, but constant displacement is not
important in wave propagation problems and they are set equal to 0.) Therefore, only non-zero
displacement component is u . This means, u I represents a wave in which the particle motion is
in the direction of propagation. The propagation velocity is, from (2),
117
λ + 2μ
α=
ρ
Because of this particle motion, this wave is also called longitudinal wave (Fig.1).
Fig. 1
Similarly, for u , we have from divu = 0 ,
II II
∂u
=0
∂x
which gives u = 0 . Then the non-zero components are v and w . This means, u II represents a
wave in which the particle motion is confined on the plane perpendicular to the propagation
direction; the propagation velocity is, from (4),
μ
β=
ρ
118
Because of this particle motion, this wave is called a shear wave (Fig. 2). Since
divu II = 0 , it does not involve volume change.
In seismology, the longitudinal wave is often called P wave, and the shear wave
is called S wave.
Fig. 2
In the above, we assumed that the medium is homogeneous (in addition to being
isotropic). That is the elastic constants λ and μ do not vary spatially. The equations
(28), (29), and (30) in 3.1.4 were all derived with this assumption.
119
u ( x, t ) = c 2u ′′( x, t ) (5)
u ( x, t ) = f ( x − ct ) + g ( x + ct ) (6)
where f(ξ) and g(ξ) are twice differentiable arbitrary functions of ξ. f and g
represent a plane wave propagating in positive and negative x directions,
respectively, with velocity c. For a wave propagating in the positive x direction, we
can consider a wave front x=ct+constant. The line perpendicular to the wave front
determines the path along which the wave front propagates. This line (or curve, in
general) is called a ray. Thus, in case of a homogeneous medium, we can use rays to
describe wave propagation completely (ray theory). Ray theory is more intuitive than
wave theory, and has been used very extensively in seismology.
However, in the real medium, λ and μ are usually a function of x, y, and z. The
equation of motion is consequently far more complex than (28), (29), or (30) in 3.1.4, and
we cannot obtain simple wave equations; consequently we cannot use the ray theory (rays
cannot be defined rigorously). Fortunately, however, if the medium is only weakly
heterogeneous, we can define a “ray” approximately, and use the ray theory. This is a
common practice in seismology. The question is “what is considered weakly heterogeous
?“ This problem can be discussed in detail in the Appendix to this section. Here we only
discuss this condition qualitatively.
Suppose that the wave velocity c(x) is a function of position x. Then the change in
velocity c over a wave length λ is given by
Δc = λ c ′
120
If this is much smaller than c itself, we can consider that the medium is “weakly
heterogeneous”. The condition then can be stated as
Δc / c = λ c′ / c = 2π c′ / ω << 1
ω >> c′ (7)
then we can use the ray theory to discuss wave propagation problem in a heterogeneous
medium. In this sense, the ray theory represents a high-frequency approximation.
1 ⎡ ⎛ dx ⎞ ⎤
u ( x, t ) ∝ exp ⎢iω ⎜ t − ∫ ⎟⎥ (9)
ρc ⎣ ⎝ c( x) ⎠ ⎦
Equation (9) shows that the wavefront propagates at the local velocity c(x), and the
amplitude varies as 1/ ρ c .
A similar relation can be obtained for a 3-D medium. The most important result
here is that a “wavefront” can be defined and it propagates at the local velocity, which
leads to the well-known “Snell’s Law” (see the Appendix, for more details).
If a ray is incident from a medium with velocity c1 into a medium with velocity
c2 , then the incidence angle i1 and the angle of emergence i2 are related by (Figure 3),
121
sin i1 sin i2
= (10)
c1 c2
In general, if the velocity varies vertically as c(z), then, for a given ray on a vertical plane, we
have (Figure 3)
sin i ( z )
= p =const.
c( z )
Fig. 3
Ge 162 Appendix to 3.4
1. One-Dimensional Case
∂ 2u ∂ ⎛ ∂u ⎞
ρ 2 = ⎜E ⎟
∂t ∂x ⎝ ∂x ⎠
122
∂E
does not reduce to the simple wave equation because ≠ 0 in general. We obtain
∂x
∂ 2u ∂ 2u ∂E ∂u
ρ = E +
∂t 2 ∂x 2 ∂x ∂x
u = E −1/ 2 v
∂ 2v ∂ 2 v 1 E ′2
ρ = E + v
∂t 2 ∂x 2 4 E
v( x, t ) = V ( x)eiω t
Then, we have
d 2V ⎡ 2 1 ⎛ E ′ ⎞ ⎤
2
+ ⎢k + ⎜ ⎟ ⎥ V = 0
dx 2 ⎢⎣ 4 ⎝ E ⎠ ⎦⎥
where
ω E
k= and c =
c ρ
If
123
2
⎛ E′ ⎞
⎜ ⎟ << k , or ω >> c′
2
⎝E⎠
Then
d 2V
2
+ k 2V = 0 (A-1)
dx
V = e ± ik0 x = e ∫
± i k0 dx
If k is a slowly varying function of x, we can solve this equation with the WKBJ method.
(
V = exp ±i ∫ κ dx )
Substituting this into (1), we find that κ must satisfy the equation
−κ 2 ± iκ ′ + k 2 = 0 (A-2)
1-st Approximation
Since
| κ ′2 |<< κ 2
κ (1)2 = k 2 or κ (1) = k
2nd Approximation
124
The 2nd approximation can be obtained by using κ (1) for the second term of (A-
2). Then, the 2nd approximation
( )
1/ 2
k′ ⎞
1/ 2
⎛
= ⎜⎛ k 2 ± iκ (1)′ ⎟⎞
1/ 2
κ (2)
= k ± ik ′
2
= k ⎜1 ± i 2 ⎟ (A-3)
⎝ ⎠ ⎝ k ⎠
k′
If, ω >> c′ , then << 1 , and
k2
i k′
κ (2) = k ±
2k
i k′ ⎞ ⎤ 1
∴
⎡ ⎛
V ( x) = exp ⎢ ±i ∫ ⎜ k ±
⎣ ⎝
⎟ dx =
2 k ⎠ ⎥⎦ k
exp ±i ∫ kdx ( )
1 1 ⎡ ⎛ dx ⎞ ⎤
∴ u ( x, t ) = V exp(iω t ) = exp ⎢ ±iω ⎜ t ± ∫ ⎟ ⎥ (A-4)
E ρω c ⎣ ⎝ c ⎠⎦
This is the WKBJ solution. This solution is valid under the condition ω >> c′ .
125
Then the equation of motion is given by
∂ 2v ⎛ ∂ 2v ∂ 2v ⎞ ∂v
ρ = μ ⎜ 2 + 2 ⎟ + μ′
∂t 2
⎝ ∂x ∂z ⎠ ∂z
Then, we obtain
d 2V ⎡ 2 1 ⎛ μ′ ⎞ ⎤
2
+ ⎢kβ ( z ) − k + ⎜ ⎟ ⎥ V = 0
2
dz 2 ⎢⎣ 4 ⎝ μ ⎠ ⎥⎦
where
ω μ
kβ = and β =
β ρ
We introduce χ ( z ) by
126
χ 2 ( z ) = k β2 ( z ) − k 2
If,
2
1 ⎛ μ′ ⎞
χ ( z ) >> ⎜ ⎟
2
(A-5)
4⎝ μ ⎠
then,
d 2V
2
+ χ 2 ( z )V = 0 (A-6)
dz
For (A-5) to be satisfied, at least ω >> β ′ must be satisfied. If k β2 ( z ) − k 2 =0, then this
condition is never satisfied.
V ( z) =
1
χ
(
exp ±i ∫ χ ( z )dz )
Then,
v ( x, z , t ) =
1 1
μ (k β − k 2 )1/ 4
2 ⎢⎣ ( )
exp ⎡i ω t + kx ± ∫ k β2 − k 2 dz ⎤
⎥⎦
kx ± ∫ k β2 − k 2 dz = const
∴
dz k
=±
dx kβ − k 2
2
127
at z = z0
k
= const
kβ − k 2
2
dz 1 k β
sin i = = = = k
d x+d z
2 2
kβ − k
2 2 kβ ω
1+
k2
∴
sin i ( z )
= const for a given ray.
β ( z)
Ge 162 Problem #4
128
An incident SH wave S1, reflected SH wave Sr and refracted (transmitted) SH
wave S2 can be written as follows. (The displacement has only z component, w.)
⎡ ⎛ sin j1 cos j1 ⎞⎤
w1 = exp ⎢iω ⎜ x− y − t ⎟⎥ (1)
⎣ ⎝ β1 β1 ⎠⎦
⎡ ⎛ sin j1 cos j1 ⎞⎤
wr = R exp ⎢iω ⎜ x+ y − t ⎟⎥ (2)
⎣ ⎝ β1 β1 ⎠⎦
⎡ ⎛ sin j2 cos j2 ⎞⎤
w2 = T exp ⎢iω ⎜ x− y − t ⎟⎥ (3)
⎣ ⎝ β2 β2 ⎠⎦
where
sin j1 sin j2 1
= =
β1 β2 c
129
Here, the amplitude of the incident wave is assumed to be 1. R and T are reflection and
transmission coefficients respectively. c is the phase velocity along the boundary. (The
above relations are solutions of the wave equation in each layer.)
⎡ ⎛ x cos j1 ⎞⎤
w1 = exp ⎢iω ⎜ − y − t ⎟⎥ (4)
⎣ ⎝c β1 ⎠⎦
⎡ ⎛ x cos j1 ⎞⎤
wr = R exp ⎢iω ⎜ + y − t ⎟⎥ (5)
⎣ ⎝c β1 ⎠⎦
⎡ ⎛ x cos j2 ⎞⎤
w2 = T exp ⎢iω ⎜ − y − t ⎟⎥ (6)
⎣ ⎝c β2 ⎠⎦
Two boundary conditions must be satisfied at the boundary: one for displacement and the
other for stress (traction).
1+ R = T
μ1 μ μ
− cos j1 + 1 cos j1 R = − 2 cos j2T
β1 β1 β2
130
− μ 2 β1 cos j2 + μ1 β 2 cos j1
R= (7)
μ1 β 2 cos j1 + μ 2 β1 cos j2
2 μ1 β 2 cos j1
T= (8)
μ1 β 2 cos j1 + μ 2 β1 cos j2
Determine the reflection and transmission coefficients for normal incidence (e.g., j1 =0).
− μ 2 β1 + μ1 β 2
R= (9)
μ1 β 2 + μ 2 β1
2μ1 β 2
T= (10)
μ1 β 2 + μ 2 β1
If β 2 > β1 , sin j2 exceeds 1 for j1 larger than the critical angle. In this case,
Referring to (6), briefly describe the behavior of the "transmitted" wave S2 for this case.
Also, the reflection coefficient R becomes complex. Determine the amplitude and
phase (with respect to the incident wave) of the reflected wave.
131
displacements and traction) on the discontinuity surfaces. The problem, in general,
becomes very complicated, and complex reflection, refraction and energy coupling
between P and S waves take place. The simplest case in which a homogeneous
medium is bounded at a plane boundary by another homogeneous medium having
different elastic property (Fig. 4) has been studied by many investigators.
First let us consider a plane S wave incident at the boundary from medium 1
to 2. Let β1 and β 2 be S wave velocities in media 1 and 2, respectively. We assume
132
Since SH wave does not have a component of particle motion perpendicular
to the boundary, the situation is relatively simple. As shown in Fig. 5a, when SH
wave is incident at the boundary with incident angle, j1 , reflected SH wave, Sr, and
sin j1 sin j2
=
β1 β2
133
This relation is similar to Snell’s law (10) of 3.4. Thus, we see that, even in this case,
134
Snell’s law can be used to determine the emergence angle. The amplitude ratios
S r / S1 and S 2 / S1 are functions of j1 and, of course, β 2 / β1 . When j1 exceeds
⎛β ⎞
j1,c = sin −1 ⎜ 1 ⎟
⎝ β2 ⎠
(2)
no refracted wave results (Fig. 5b). In this case, part of the wave energy is trapped
along the boundary, and total reflection occurs. At the same time phase shift occurs
on reflection. The angle j1,c is called the critical angle.
where i1 and i2 are angles of incidence and emergence of the P wave, and j1 and
j2 are angles of reflection and emergence of the SV wave. When i1
exceeds
135
⎛α ⎞
i1,c = sin −1 ⎜ 1 ⎟
⎝ α2 ⎠
no refracted P wave results, and part of the wave energy is trapped along the boundary. The
angle i1,c is the critical angle.
When SV wave is incident at the boundary, the situation is even more complicated. In
this case, refracted and reflected P and SV waves result (Fig. 5d). As before, the following
relations hold
sin j1 sin j2 sin i2 sin i1
= = = (5)
β1 β2 α2 α1
⎛β ⎞
j1,c1 = sin −1 ⎜ 1 ⎟ (6)
⎝ α2 ⎠
⎛β ⎞
j1,c2 = sin −1 ⎜ 1 ⎟ (7)
⎝ α1 ⎠
When j1 exceeds
136
⎛β ⎞
j1,c3 = sin −1 ⎜ 1 ⎟ (8)
⎝ β2 ⎠
angles.
When one medium is vacuum, the
boundary becomes a free surface, and only
reflections are to be considered. It is not
difficult to see from Fig. 5 that there is no
critical angle for incident SH and P waves,
but for incident SV wave
⎛β ⎞
j1,c2 = sin −1 ⎜ 1 ⎟ (9)
⎝ α1 ⎠
Ge 162 Problem #5
137
Dispersion of Love Waves
⎛ω ⎞ μ S
tan ⎜ S1 H ⎟ = 2 2 (1)
⎝c ⎠ μ1S1
c2 c2
where S1 = − 1 , and S 2 = 1 − , and β1 < c < β 2 .
β12 β 22
Equation (1) can be solved graphically to determine the phase velocity c for given
μ1 , μ 2 , β1, β2, and H. The roots of (1) are given by the intersection of the two curves
corresponding to RHS (right-hand side) and LHS of (1) both of which are functions of
phase velocity.
H=35 km
μ1 =3x1011 dyne/cm2, β1 =3.5 km/sec
μ 2 =7x1011 dyne/cm2, β 2 =4.6 km/sec
a) Compute the values of RHS and LHS of (1) for phase velocities
c=3.52, 3.6, 3.8, 4.0, 4.2, and 4.5 km/sec, and for periods T=20, 35, 50, and 80 sec.
b) Plot the values computed above (vertical axis) as a function of phase velocity c
(horizontal axis). Draw the two curves and find the phase velocity from the intersection
of the two curves (this gives the phase velocity of the fundamental mode).
c) Compare these phase velocities c(T) with the dispersion curve for Love waves for the
continental region (see Fig. 14 in class note 3.6).
138
Figure 1 shows the Rayleigh wave trains from an earthquake (1/2/2002, M=7.3) in the
Vanuatu Is., southwest Pacific, recorded at two TriNet stations, PAS (Pasadena) and NEE
(Needles) (The SAC files, van2.pas.lhz.sac and van2.nee.lhz.sac are in
/home/ftp/pub/hiroo/ge162.dir/practice3.dir, but they are not necessary to do this problem).
These two stations and the epicenter are almost on a great circle, and the distance between the
two stations, Δ, is 331 km. The unit of the time scale on the horizontal axis is sec (i.e., 2100 to
3100 sec).
Fig. 1
139
Figures 2 and 3 show the original (top trace) and band-pass filtered records (at periods of
about 20, 30, 40, 50, and 60 sec) of the seismograms shown in Figure 1.
Fig. 2 (PAS)
140
Fig. 3 (NEE)
The band-pass filtered records at T=60 sec are of marginal quality. Thus, we use the data
from 20 to 50 sec.
1. Determine the group velocity, U, between PAS and NEE, at periods of 20, 30, 40, and 50 sec,
by measuring the arrival times of the wave train. (Equation 57 in class handout), and plot the
results on a U-T diagram, and compare it with that for a simple crustal structure shown at the
end.
The determination of the phase velocities, c, is a bit more difficult. Figures 4 and 5 show
the plots of the harmonic components of the records at PAS and NEE. The top trace is the
original and the six traces below it are the harmonic components at periods of 20.08, 30.12,
40.96, 51.20, 60.24, and 73.14 sec.
141
In principle, the phase velocities can be determined from the phase arrival times of the
harmonic components using equation 51 in class handout. In this example, the records at PAS
and NEE have the same starting time. The difficulty is that, if only harmonic components at
discrete periods are given, it is not possible to determine the integer N. In other words, all the
peaks look exactly the same. In this exercise, we assume that the phase velocities in southern
California are approximately known within a certain range, and we determine N for each period
so that the measured phase velocities fall in this range. (Note: If all the harmonics are given, as
is the case in the real situation, we need to determine N only for one period.)
Fig. 4
142
Fig. 5
We assume that the phase velocities in southern California are within the
following ranges.
T (sec) c (km/sec)
20 3.1-3.9
30 3.0-4.3
40 3.3-4.5
50 3.3-4.5
2. Determine the phase velocities between PAS and NEE using the pairs of harmonic
waves shown below at periods of 20.08, 30.12, 40.97, and 51.20 sec, and plot them on a
c-T diagram, and compare it with the phase velocity dispersion curve for a simple
structure given at the end.
143
144
145
146
147
148
These dispersion curves are computed for a simple crustal structure:
Note:
149
In this exercise, we use only 2 stations and the path is chosen to be a great circle.
In practice, we use multiple stations and the wave is assumed to be a plane wave and the
propagation azimuth and the phase velocities are simultaneously determined. Using a
high-density network like TriNet, the accuracy can be improved significantly, and the
method can be used for tomographic inversion.
3.6 Seismic Surface Waves (LW, pp. 116-153)
Rayleigh Wave
u = u I + u II (1)
150
and u II is the solenoidal field satisfying
Since we consider only P and SV type motions propagating in x direction, u I and u II have only
x and z components (u I , wI ) and (u II , wII ) which do not depend on y. We write u I , wI , u II ,
151
⎛ x⎞
iω ⎜ t − ⎟
u I = f1 ( z )e ⎝ c⎠
⎛ x⎞
iω ⎜ t − ⎟
w = h1 ( z )e
I ⎝ c⎠
(4)
⎛ x⎞
iω ⎜ t − ⎟
u II = f 2 ( z )e ⎝ c⎠
⎛ x⎞
iω ⎜ t − ⎟
w = h2 ( z )e
II ⎝ c⎠
At this point, the phase velocity c is unknown, and we investigate whether we can find c which
satisfies all the equations and the boundary conditions for surface waves.
∂u I ∂wI ∂u II ∂wII
− = 0, + =0
∂z ∂x ∂x ∂z
from which
1c ′ 1c ′
h1 ( z ) = − f1 ( z ) and f2 ( z) = h2 ( z ) (5)
iω iω
immediately follow.
2
⎛ω ⎞ ⎛ c2 ⎞
f1′′ − ⎜ ⎟ ⎜1 − 2 ⎟ f1 = 0 (6)
⎝c⎠ ⎝ α ⎠
152
and
2
⎛ω ⎞ ⎛ c2 ⎞
h2′′ − ⎜ ⎟ ⎜ 1 − h =0
2 ⎟ 2
(7)
⎝c⎠ ⎝ β ⎠
⎛ c2 ⎞
respectively. If ⎜1 − 2 ⎟ < 0 , then h2 becomes a periodic function which does not
⎝ β ⎠
decay in the medium. We therefore require
⎛ c2 ⎞
⎜ 1 − 2 ⎟
>0 (8)
⎝ β ⎠
Then, we have
ω ω
± Sα z ± Sβ z
f1 = Ae c
, h2 = Be c
(9)
c2 c2
Sα = 1 − , Sβ = 1 −
α2 β2
153
⎛ ⎞
⎛ ωc Sα z S β ω
S β z ⎞ iω ⎜ t − ⎟
x
u ( x, z , t ) = ⎜ Ae + Be c
⎟e
⎝ c⎠
(10)
⎝ i ⎠
ω ω ⎛ ⎞
⎛ Sα S β z ⎞ iω ⎜ t − ⎟
x
Sα z
w( x, z , t ) = ⎜ − Ae c
+ Be c
⎟e
⎝ c⎠
(11)
⎝ i ⎠
These solutions must satisfy the stress-free boundary conditions at the free surface z = 0, i.e.,
⎛ ∂u ∂w ⎞ ∂w
0 = σ zz = λ ⎜ + ⎟ + 2μ , at z=0 (12)
⎝ ∂x ∂z ⎠ ∂z
⎛ ∂u ∂w ⎞
0 = σ zx = μ ⎜ + ⎟, at z=0 (13)
⎝ ∂z ∂x ⎠
Substituting (10) and (11) into (12) and (13), and setting z = 0, we have
⎛ c2 ⎞ c2
−i ⎜ 2 − 2 ⎟ A + 2 1 − 2 B = 0 (14)
⎝β ⎠ β
c2 ⎛ c2 ⎞
2 1− A − i ⎜ 2 − ⎟B = 0 (15)
α2 ⎝ β2 ⎠
2 1/ 2 1/ 2
⎛ c2 ⎞ ⎛ c2 ⎞ ⎛ c2 ⎞
⎜ 2 − ⎟ = 4 ⎜1 − 2 ⎟ ⎜ 1 − 2 ⎟
(16)
⎝ β2 ⎠ ⎝ α ⎠ ⎝ β ⎠
or
154
c6 c4 2 ⎛ 24 16 ⎞ ⎛ β2 ⎞
−8 + ⎜ 2 − − 16 ⎜1 − 2 ⎟ = 0 (17)
α 2 ⎟⎠
c
β6 β4 ⎝β ⎝ α ⎠
⎛ β2 ⎞
Since LHS of this equation becomes − 16 ⎜1 − 2 ⎟ < 0 and 1 at c = 0 and c = β ,
⎝ α ⎠
respectively, (17) has a real root at
c
0< <1 (18)
β
This satisfies the condition (8) assumed previously. Thus, we have proved that the surface
wave given by (10) and (11) exists, and that it propagates with a velocity smaller than the shear
wave velocity β . This kind of wave is called Rayleigh wave, and the propagation velocity c is
called the phase velocity. Combining (14), (10), and (11), we have
⎛ ωc Sα z ⎛ 1 c 2 ⎞ ωc Sβ z ⎞ iω ⎛⎜⎝ t − cx ⎞⎟⎠
u ( x, z , t ) = A ⎜ e − ⎜1 − 2 ⎟
e ⎟e (19)
⎝ ⎝ 2β ⎠ ⎠
⎛ ω
Sα z 1 ⎛ 1 c 2 ⎞ ωc Sβ z ⎞ iω ⎜⎝ t − c ⎟⎠
⎛ x⎞
w( x, z , t ) = −iA ⎜ − Sα e c
+ ⎜1 − ⎟e ⎟⎟ e (20)
⎜ Sβ ⎝ 2 β 2 ⎠
⎝ ⎠
1 c2
⎛u⎞ 2 β2
⎜ ⎟ = (21)
⎝ w ⎠ z =0 1 ⎛ 1 c2 ⎞
− Sα + ⎜1 − ⎟
Sβ ⎝ 2 β 2 ⎠
155
When α = 3β (this corresponds to the Poisson’s solid for which Poisson’s ratio ν =
c2 2 2
0.25), (17) has three real roots of , namely, 4, 2 + , and 2 − . The first two do not
β 2
3 3
satisfy the condition (8), and therefore do not yield a surface wave. From the last value, we have
c = 0.92 β (22)
which satisfies (8). In this case, from (9), Sα = 0.85 , and S β = 0.39 . The particle motion at the
surface for simple harmonic waves can be obtained by taking the real part of (19) and (20):
and the amplitude ratio becomes 0.68. Equations (23) and (24) show that the orbital motion is
elliptic and counter-clockwise (for a wave propagating in positive x direction) (Fig.6).
The decay of the amplitude with depth is governed by factors like
ω z ω z
Sα z 2π Sα Sβ z 2π S β
ec =e λ
and ec =e λ
where λ is the wave length. Thus short wave length components are more quickly attenuated
than long wave length components.
In the above, the medium is assumed homogeneous. Waves like Rayleigh waves also
exist in heterogeneous (usually only in z direction) medium. In this case, the phase velocity c is
in general a function of wave period (or wave length), and therefore the propagation becomes
dispersive. Curves which relate the phase velocity to the period are called phase velocity
156
dispersion curves. Calculation of such dispersion curves for a vertically heterogeneous medium
is usually made numerically,
Love Wave
As we saw in Fig. 5, SH wave is totally reflected when the incidence angle exceeds the
157
⎛β ⎞
critical angle j1,c = sin −1 ⎜ 1 ⎟ ( β 2 > β1 ) . Also, it is totally reflected at a free surface, regardless
⎝ β2 ⎠
of the incident angle. Therefore if we have a layer underlain by a half space (Fig. 7), and assume
that β 2 > β1 , we can consider a SH wave bouncing between the free surface and the boundary
without major loss of energy into the half space. This situation suggests propagation of wave
energy which is trapped within the layer.
Let us consider propagation of this kind of wave. We take the free surface and the
boundary to coincide with the plane z=0 and z=-H respectively. The half space occupies z < − H .
As before we consider a wave propagating in positive x direction. Since we consider an SH field,
the displacement vector has only y component. We let v1 and v2 be the displacements in the
layer and the half space, and look for the solutions in a form
158
⎛ x⎞
iω ⎜ t − ⎟
v1 = f1 ( z )e ⎝ c⎠
(25)
⎛ x⎞
iω ⎜ t − ⎟
v2 = f 2 ( z )e ⎝ c⎠
(26)
v1 and v2 satisfy the SH wave equation (3) in the layer and the half space respectively. Thus,
ω 2 ⎛ c2 ⎞
f1′′ + 2 ⎜
− 1⎟ f1 = 0 (27)
c ⎝ β1 2
⎠
′′ ω 2 ⎛ c2 ⎞
f 2 − 2 ⎜1 − 2 ⎟ f 2 = 0 (28)
c ⎝ β2 ⎠
ω ω
f1 = A cos S1 z + B sin S1 z (29)
c c
ω
S2 z
f 2 = Ce c (30)
where
159
c2 c2
S1 = − 1 and S 2 = 1 − (31)
β12 β 22
c2
For f2 to vanish at z → −∞ , 1 − > 0 , and we take only the term with positive sign in the
β 22
exponent of f2. At this point S1 can be either real or imaginary. f1 and f2 must satisfy the
boundary conditions at the free surface and at the boundary. At the free surface
∂v1
0 = σ zy = μ1 at z=0 (32)
∂z
At the boundary
∂v1 ∂v
v1 = v2 and μ1 = μ 2 2 at z=-H (33)
∂z ∂z
where μ1 and μ 2 are the rigidity of the layer and the half space, respectively. Substituting (29)
and (30) into (32) and (33) leads to
B=0 (34)
ω
ω − S2 H
A cos S1 H − Ce c
=0 (35)
c
ω
ω − S2 H
Aμ1S1 sin S1 H − C μ 2 S 2 e c
=0 (36)
c
160
ω μ 2 S2
tan S1 H = (37)
c μ1 S1
This is the characteristic equation. If this equation is satisfied, the surface wave exists, and
the phase velocity c is determined from (37). Since S2 is required to be real, this equation implies
that S1 is also real. Therefore, from (31),
β1 < c < β 2
(ω > β π /
2 ( β 2 / β1 )
2
)
− 1 H , more than one solution exists (Fig. 8).
As we can see from (35), (29) and (30), the amplitudes in the layer and the half space
become
161
ω
f1 = A cos S1 z (38)
c
ω
ω S2 ( z + H )
f 2 = A cos S1 He c (39)
c
It can be shown that for a fixed ω , the solution (mode) with the lowest value of c (c = c1 in Fig.
8), gives an amplitude function which does not have a zero crossing (node). This mode is called
fundamental mode. When more than one solution exists, the modes with higher phase velocities
than c1 have an amplitude function with zero crossings (nodes) (Fig. 8). These modes are called
We can consider this type of wave for more complex medium, for example, a medium
which has more than one layer or continuous velocity variation with depth. The computation of
the dispersion curves for such complex media must be made numerically,
162
163
Dispersion Curve and Structure
Dispersion curves c(ω ) or c(T ) are important when we use surface waves to determine
the structure of the medium. As we discussed earlier, both Love and Rayleigh waves are
dispersive in layered media, and the dispersion curves are determined by the structure.
Figures 9, 10, and 11 show several examples of dispersed Love and Rayleigh waves.
164
Fig. 9
165
Fig. 10a
166
Fig. 10b
167
Fig. 11
Consider a layer-over-half space (Figure 12). Let the thickness of the layer be H, and the
shear velocity in the layer and the half space be β1 and β 2 , respectively ( β 2 > β1 ). At short
period, the Love wave energy will be in the layer and the phase velocity approaches β1 . At long
period, Love wave energy penetrates into the half space. At very long period, the velocity
approaches β 2 . Thus, the phase velocity dispersion curve would look like the one shown in
Figure 12.
168
Fig. 12
Then, how would the shape of the dispersion curve change as H varies? If H decreases,
the Love wave would "feel" the higher speed half space at a shorter period. Thus, the shape of
the dispersion curves would change as shown. For a more complex multi-layered structure, a
similar pattern is expected. Thus, we can determine the structure from the shape of the
dispersion curve.
169
We can make the same argument for Rayleigh waves except that the phase velocities at
the long-period and the short-period ends are 0.92 β 2 and 0.92 β1 (for Poisson solid),
respectively.
Figures 13 and 14 show the examples of phase velocity dispersion curves for oceanic,
Fig. 13
170
Fig. 14
The structures and the data are given in the following tables.
Models
Ocean
171
160. 8.17 4.30 3.44
100. 8.49 4.60 3.53
Continent
Shield
Data
Rayleigh Wave
Period O C S
172
(sec) (km/sec) (km/sec) (km/sec)
Love Wave
Period O C S
(sec) (km/sec) (km/sec) (km/sec)
173
Phase and Group Velocity of Dispersive Waves
Surface waves are usually dispersive. That is, the phase velocity c is a function of
angular frequency ω . In the treatment of dispersive waves we need to distinguish phase
velocity c(ω ) and group velocity U (ω ) .
In order to understand the propagation of a dispersive wave train, we consider a
wave packet made up of many harmonic wave trains like
c =ω /k (41)
174
175
Consider a case in which all components are in phase at x=0, t=0. Then, we have
constructive interference at t=0, and destructive interference elsewhere leading to
negligible disturbance. Thus, we have an impulse at x=0. At some later time t and at a
distance x, a disturbance will be observable if any of the wave trains are in phase over a
frequency band. The condition for reinforcement is
kx − ω t = const.
⎛ dω ⎞
x −⎜ ⎟t = 0 (42)
⎝ dk ⎠
176
The ratio x/t gives the velocity with which the disturbance (wave group) at the frequency
band propagates, and is called the group velocity U. From (42),
dω dc
U= =c+k (43)
dk dk
+∞ +∞
g ( x1 , t ) =
∫
−∞
gˆ ( x1 , f ) exp(2π ift )df =
∫ −∞
| gˆ ( x1 , f ) | exp [i (2π ft + φ1 ) ] df (44)
where
+∞
gˆ ( x, f ) =| gˆ ( x, f ) | exp [iφ ( f ) ] =
∫ −∞
g ( x, t ) exp(−2π ift )dt (45)
If the wave train propagates in x direction without changing the amplitude, then at
x = x2 ,
+∞
g ( x2 , t ) =
∫ −∞
| gˆ ( x1 , f ) | exp {i [ 2π f (t − ( x2 − x1 ) / c( f )) + φ1 ]} f
+∞
=
∫ −∞
| gˆ ( x1 , f ) | exp ⎡⎣i ( 2π ft + φ2 ) ⎤⎦ df
177
(46)
and
φ2 = φ1 − 2π f ( x2 − x1 ) / c( f ) (47)
g ( x2 , t ) = g ( x1 , t − x / c0 ) (48)
That is, the wave propagates without changing its waveform. In a dispersive medium, c is
a function of frequency, and the waveform changes with propagation.
Since, adding 2π N (N is an integer) to φ2 does not change (46), (47) is actually
φ2 = φ1 − 2π f ( x2 − x1 ) / c( f ) + 2π N (49)
Using (49), we can determine the phase velocity from the two wave trains measured at
distances x1 and x2 by measuring the phases φ1 and φ2 at these distances. Solving (49)
for c(f), we obtain,
f ( x2 − x1 )
c( f ) = (50)
N − (φ2 − φ1 ) / 2π
or,
( x2 − x1 )
c(T ) = (51)
NT + (−φ2T / 2π + φ1T / 2π )
178
In the above, the records at the two stations are assumed to have a common origin
time. In actual computations, we may use the records (seismograms) starting at different
times; at t = t1 at x = x1 and at t = t2 at x = x2 . Then, if we let ψ 1 , and ψ 2 be the Fourier
φ2 − φ1 = ψ 2 − ψ 1 − 2π f (t2 − t1 ) (52)
f ( x2 − x1 )
c( f ) = (53)
N − (ψ 2 −ψ 1 ) / 2π + f (t2 − t1 )
If the propagation path is very long, and involves antipolar or polar passages, a small
correction is necessary to correct for the polar phase shift. In this case,
f ( x2 − x1 )
c( f ) = (54)
N + m / 4 − (ψ 2 − ψ 1 ) / 2π + f (t2 − t1 )
Once the phase velocity is determined as a function of ω (or, frequency, f), then
the group velocity U (ω ) can be computed from phase velocity using (43), i.e.,
179
1 dk 1 ω dc(ω )
= = − 2 (55)
U (ω ) dω c(ω ) c (ω ) dω
+∞ +∞
∫ ∫
1 1
g ( x, t ) = gˆ ( x, ω ) exp[i(ω t − kx)]dω = | gˆ ( x, ω ) |exp[i (ω t − kx + φ )]dω
2π −∞ 2π −∞
(56)
ω0 +Δω
∫
1
g ( x, t ) ≈ | gˆ ( x, ω 0 ) | cos(ω t − kx + φ )dω
π ω0 −Δω
⎛ ⎛ Δω ⎞⎞
⎜ sin ⎜ (t − k0′ x + φ0′ ⎟ ⎟
Δω ⎝ 2 ⎠ ⎟ cos(ω t − k x + φ )
= | gˆ ( x, ω 0 ) | ⎜ (57)
π ⎜ Δ ω ⎟
0 0 0
⎜ (t − k ′
0 x + φ ′
0 ) ⎟
⎝ 2 ⎠
where k0 and φ0 are the values of k and φ at ω = ω 0 , respectively, and k0′ and φ0′ are
dk dφ
the values of and at ω = ω 0 , respectively. Equation (57) gives a wave train
dω dω
which propagates at a speed
180
1 dω
U (ω 0 ) = = (58)
k0′ dk ω =ω0
Ge 162
Since the Earth is a bounded medium, the wave propagation problem can be
treated as a normal-mode problem.
181
The basic physics can be understood with a simple 1-D problem.
1. 1-D Problem
182
∂ 2u ∂ 2u
ρ = E
∂ t2 ∂ x2
u ( x, t ) = y ( x)eiω t (1)
then,
y′′ + k 2 y = 0 , k 2 = ρω 2 / E (2)
The solution is
y=Acoskx+Bsinkx
If both ends of the rod are free (i.e., y′ =0 at x=0 and L), then B=0 and kn=nπ/L
(n=integer). Then, the eigen functions are
yn=Acos(nπx/L)
ω n = E / ρ kn
183
The eigen functions yn are orthogonal, i. e.,
L
∫0
ρ yn ym dx ∝ δ nm (3)
(this relation can be derived directly from (2), if yn and ym satisfy the homogeneous
boundary conditions ayn + yn′ = 0 and cym + ym′ = 0 at x=0 and L.)
Example
184
This is a seismogram. This displacement can be viewed as superposition of normal
modes as follows.
Normal modes for a string with length L, clamped at both ends are given by
yn ( x) = sin(nπ x / L)
u ( x, t ) = ∑ A sin(nπ x / L) cos ω t
n =1
n n
where ω n = nπ c / L are the eigen frequencies, and c is the velocity of wave traveling in
the string.
185
∞
u ( x, 0) = ∑ A sin(nπ x / L)
n =1
n
we obtain
An =(2/L)sin(nπxs/L)
The following figure shows u ( x, t ) for x=xs=L/4, and is essentially the same as the
seismogram shown above. This equation can be viewed as superposition of normal
modes, sin(nπ x / L) cos ω n t , with the amplitudes (2 / L) sin(nπ xs / L) shown as
"spectrum".
186
1.2
1
(L/2) x spectral amplitude
0.8
0.6
0.4
0.2
0
0 1 2 3 4 5 6 7 8
n
The above result can be easily extended to 2-D and 3-D problems. For example,
free oscillations of a homogeneous rectangular membrane can be expressed by eigen
functions of the form,
187
Here, m and n determine the number of nodes in x and y directions, respectively (see the
attached figure.).
188
the order numbers l, m and n determine the number of nodes in x, y and z directions,
respectively.
The free oscillations of the Earth can be formulated similarly as free oscillations
of an elastic sphere. The only difference is that it is more convenient to use a spherical
coordinate system (r, θ, ϕ) instead of the Cartesian coordinate system (x, y, z).
For a sphere with a laterally homogeneous structure, the eigen functions can be
written as,
For a homogeneous elastic sphere, Rn(r) are Bessel functions, Θl(θ) are (associated)
Legendre functions and Φm(ϕ) are trigonometric functions. The three indices l, m, and n
determine the number of nodes in θ direction (meridional direction), ϕ direction
(longitudinal direction) and r direction (radial direction).
In general eigen frequencies depend on l, m, and n and can be written as n ω ml .
However, for a spherically symmetric, non-rotating Earth model, eigen frequencies do
not depend on m.
189
Free oscillations of the Earth can be thought of as extensions of Love and
Rayleigh waves propagating many times around the Earth. When Love waves propagate
many times around the Earth, torsional free oscillations are set up. Similarly, when
Rayleigh waves propagate around the Earth many times, spheroidal oscillations are set
up. Torsional and spheroidal oscillations (modes) with order numbers l, m, and n are
written as n Tlm and n Sml and are shown in the attached figures and tables.
190
191
192
Spheroidal Modes, Bolivian Earthquake
193
0S0 after 400,000 sec
Torsional Oscillations
194
⎛ ⎞
⎜ 0 ⎟
⎛ ur ⎞ ⎜ ⎟
⎜ ⎟ ⎜ 1 ∂ Yl m ⎟ iω t
⎜ uθ ⎟ = y1 (r ) ⎜ sin θ ∂φ ⎟ e (5)
⎜u ⎟ ⎜ ⎟
⎝ φ⎠ ⎜ ∂ Yl m ⎟
⎜ − ⎟
⎝ ∂θ ⎠
1 ∂ Yl m ∂ Yl m
Tlm = (0, ,− ) (7)
sin θ ∂φ ∂θ
are called the vector spherical harmonics. It is easy to show that Tlm
make up an orthogonal system:
2π π 4π l (l + 1)(l + m)!
∫ ∫
0 0
(Tlm ⋅ Tlm′ ′ ) sin θ dθ dφ =
ε m (2l + 1)(l − m)!
δ ll ′δ mm′ (8)
ε m = 1 if m = 0
ε m = 2 if m ≠ 0
⎛ d 2 y1 2 dy1 ⎞ d μ ⎛ dy1 y1 ⎞ ⎛ 2 l (l + 1) ⎞
μ⎜ 2
+ ⎟+ ⎜ − ⎟ + ⎜ω ρ − μ ⎟ y1 = 0 (9)
⎝ dr r dr ⎠ dr ⎝ dr r ⎠ ⎝ r2 ⎠
if we put
⎛ dy (r ) y (r ) ⎞
y2 ( r ) = μ ( r ) ⎜ 1 − 1 ⎟
⎝ dr r ⎠
195
dy1 (r ) 1 1
= y1 (r ) + y2 (r )
dr r μ
(10)
dy2 (r ) ⎛ (l 2 + l − 2) ⎞ 3
=⎜ 2
μ − ω 2 ρ ⎟ y1 (r ) − y2 (r )
dr ⎝ r ⎠ r
σ rr = 0, σ rθ = 0, σ rφ = 0
1 ∂ Yl m ιω t ∂Ym
σ rθ = y2 (r ) e and σ rφ = − y2 (r ) l eιω t
sin θ ∂φ ∂θ
d 2 y1 2 dy1 ⎛ ω 2 ρ l (l + 1) ⎞
+ +⎜ − ⎟ y1 = 0 (11)
dr 2 r dr ⎝ μ r2 ⎠
Equation (11) is one of the variations of the Bessel's differential equation. The general
solution is
1
y1 (r ) ≈ Z 1 (kr ) (12)
r l+ 2
ω
k= (13)
μ/ρ
196
In order to satisfy the boundary conditions, an appropriate kind of Bessel function
should be taken and k must take certain discrete values n k . Since (12) depends on l,
these values depend on l, and we write these values as n k l . Using (13), the eigen
frequencies n ω l can be given by
μ
ω l = n kl
ρ
n
Note that, since (11) does not contain m, the eigen values do not depend on m. Thus,
there are 2l+1 eigen functions belonging to n ω l , i.e. there is degeneracy of 2l+1 degree.
For a radially heterogeneous sphere where μ=μ(r), the solution of (9) cannot be
given by Bessel functions, but numerical integration of (10) with appropriate boundary
conditions yields eigen functions similar to (12), and eigen values n ω l . Since (10) does
not contain m, the eigen values do not depend on m.
For large l, the asymptotic expansion of Pl n gives
2 ⎡ 1 m π⎤
Pl n ≈ e mπ i ml cos ⎢ (l + )θ + π − ⎥ (14)
π l sin θ ⎣ 2 2 4⎦
∂Ylm ∂Yl m
≈ l Ylm and ≈ m Yl m
∂θ ∂φ
Spheroidal Oscillation
197
⎛ ⎞
⎜ 0 ⎟
⎛ ur ⎞ ⎛ Yl m ⎞ ⎜ ⎟
⎜ ⎟ ⎜ ⎟ iω t ⎜ ∂ Yl m ⎟ iω t
⎜ uθ ⎟ = y1 (r ) ⎜ 0 ⎟ e + y3 (r ) ⎜
∂θ ⎟e (15)
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ uφ ⎠ ⎝ 0 ⎠ ⎜ 1 ∂ Yl ⎟
m
⎜ sin θ ∂φ ⎟
⎝ ⎠
where
2π π 2π π 4π l (l + 1)(l + m)!
∫ ∫
0 0
(Sm2,l ⋅ Sm2,l' ' ) sin θ dθ dφ = ∫
0 ∫
0
(S1,ml ⋅ S1,ml' ' ) sin θ dθ dφ =
ε m (2l + 1)(l − m)!
δ ll ′δ mm′
Using the expression (15), the equation of motion for a radially heterogeneous
sphere can be reduced to a set of ordinary differential equations for y1 and y3. It can be
shown that y1 and y3 satisfy the orthogonality condition,
r2
∫ r1
ρ r 2 [ y1,n y1,n ' + l (l + 1) y3,n y3,n ' ]dr = Cδ nn '
For l>>1 and l>>m, ur and uθ become much larger than uφ , and the
198
199
Cartesian, Cylindrical, and Spherical Coordinates
200
201
202
203
Appendix 2
204
205
Appendix 3
206
207
Ge 162
Assume that T(Δ) and p(Δ) are continuous functions of Δ, as shown in Figures 1
and 2.
Fig.1 Fig. 2
dΔ r
Then, from Figure 3, tan i = r . Since p = sin i ,
dr v
208
Fig. 3
p
dΔ = dr (1)
2
⎛r⎞
r ⎜ ⎟ − p2
⎝v⎠
∫
r0
p
Δ( p) = 2 dr (2)
2
⎛r⎞
rp
r ⎜ ⎟ − p2
⎝v⎠
where the suffix 0 denotes the values at the surface (i.e., r = r0 ) and rp is the radial
distance of the deepest point of a ray.
r
Introducing a new variable, η =
v
209
η0
∫
p dr
Δ( p) = 2 dη (3)
p
r η 2 − p 2 dη
r
Δ(p) is known from observation, and η (r ) = , or v(r ) is a unknown function of r which we
v(r )
r
wish to determine. (Note that since p = sin i , p = η at the deepest point of the ray where
v
i = π / 2 .)
⎡ Δ1
⎡⎛ ⎞ ⎤ ⎤
∫
2
1 p ⎛ p ⎞
r1 = r0 exp ⎢ − ln ⎜ ⎟ + ⎜ ⎟ − 1 ⎥d Δ ⎥
⎢ (11)
⎢ π ⎢⎝ η1 ⎠ ⎝ η1 ⎠ ⎥ ⎥
⎣⎢ 0 ⎣ ⎦ ⎦⎥
The suffix 1 is used to denote values of variables at the level r1 , and let Δ1 be the value
of Δ for the ray whose deepest point is at r1 .
----------------------------------------------------------------------------------------------------------------
1
To solve this, multiply (3) by and integrate with p from η1 to η0 (see Figure
p − η12
2
4).
210
Fig. 4
The suffix 1 is used to denote values of variables at the level r1 , and let Δ1 be the value
of Δ for the ray whose deepest point is at r1 .
η0 η0 η0
∫ ∫ ∫
1 dp p dr
Δ( p) dp = 2 dη (4)
η1
p 2 − η12 η1
p 2 − η12 p
r η 2 − p 2 dη
η0 η0
∫
⎛ p⎞ −1 dΔ ⎛ p⎞
LHS= Δ( p) cosh ⎜ ⎟ − cosh −1 ⎜ ⎟dp (5)
⎝ η1 ⎠ η1 η1
dp ⎝ η1 ⎠
⎛η ⎞
Since Δ(η0 ) = 0 and cosh −1 ⎜ 1 ⎟ = cosh −1 (1) = 0 ,
⎝ η1 ⎠
∫
0
⎛ p⎞
LHS= − cosh −1 ⎜ ⎟d Δ (6)
Δ1 ⎝ η1 ⎠
To evaluate the integral of RHS, we change the order of integration (see Figure
5),
211
Fig. 5
η0 η
∫ ∫
1 dr 1 p
RHS= 2 dη dp (7)
r dη p 2 − η12 η 2 − p2
η1 η1
Introducing θ , by
p 2 − η12
sin θ =
2
η 2 − η12
we obtain,
π
η
∫ ∫
1 p 2
π
dp = dθ = (8)
p −η
2 2
η −p2 2 2
η1 1 0
Then,
∫
r0
1 ⎛r ⎞
RHS= π dr = π ln ⎜ 0 ⎟ (9)
r1
r ⎝ r1 ⎠
Hence.
Δ1 Δ1
⎡⎛ ⎞ ⎤
∫ ∫
2
⎛r ⎞ ⎛ p⎞ p ⎛ p⎞
π ln ⎜ 0 ⎟ = cosh −1 ⎜ ⎟d Δ = ln ⎜ ⎟ + ⎜ ⎟ − 1 ⎥d Δ
⎢ (10)
⎝ r1 ⎠ 0 ⎝ η1 ⎠ 0
⎢⎝ η1 ⎠
⎣ ⎝ η1 ⎠ ⎥
⎦
or,
212
⎡ Δ1
⎡⎛ ⎞ ⎤ ⎤
∫
2
⎢ 1 p ⎛ p⎞
r1 = r0 exp − ln ⎜ ⎟ + ⎜ ⎟ − 1 d Δ ⎥
⎢ ⎥ (11)
⎢ π ⎢⎝ η1 ⎠ ⎝ η ⎠ ⎥ ⎥
⎣⎢ ⎦ ⎦⎥
1
0 ⎣
------------------------------------------------------------------------------------------------------------
213
From equations (11) in class note, we obtain,
⎡ Δ1
⎡⎛ ⎞ ⎤ ⎤
∫
2
⎢ 1 p ⎛ p⎞
r1 = r0 exp − ln ⎜ ⎟ + ⎜ ⎟ − 1 d Δ ⎥
⎢ ⎥ (1)
⎢ π ⎢ η ⎝ η1 ⎠ ⎥ ⎥
⎢⎣ 0 ⎣⎝ 1 ⎠ ⎦ ⎥⎦
where
r
η (r ) = (2)
v( r )
Here, r1 is the radial distance to the bottoming point of the ray reaching the distance Δ1 ,
and
214
r1
η1 = (3)
v(r1 )
Since i = π / 2 at the bottoming point, the ray parameter for the ray reaching Δ = Δ1 is
given by
r1 r
p1 = sin i1 = 1 ≡ η1 (4)
v(r1 ) v(r1 )
Thus, if a p-Δ curve is given, then for a given Δ1 , we can determine η1 . Then,
we can integrate (1) to obtain r1 . Once r1 is determined, then v(r1 ) can be computed from
r1
v(r1 ) = (5)
η1
1. The following table shows the travel times (2nd column) which are consistent with the ISC
travel-time curve given in class handout. Then the travel time data are numerically differentiated
to compute the ray parameter p which is shown on the 3rd column. The unit of p is sec (i.e., the
unit of the distance is converted to radian in this computation.). The following table lists the data
up to Δ = 30° , and a more complete table up to Δ = 98° is given in i_p-delta_2 in
/home/ftp/pub/hiroo/ge162.dir/practice_4.dir.
Plot the travel time and p in i_p-delta_2 as a function of distance, and make sure that the
travel times are consistent with the ISC data. Also, check the consistency between t and p. An
enlarged travel-time curve is shown below. (Spot checks at Δ = 30°, 60° ,90° are sufficient.)
Table
215
8.00000 120.30000 802.14056
10.00000 147.98334 788.77106
12.00000 175.25000 773.49268
14.00000 201.87500 753.43927
16.00000 227.92500 736.25037
18.00000 253.14999 710.46747
20.00000 276.72501 627.38831
22.00000 297.45001 572.95752
24.00000 317.10001 555.76868
26.00000 336.14999 532.85071
28.00000 354.47501 518.52673
30.00000 372.45001 509.93188
........ ......... .........
216
217
2. Then, compute r1 for Δ1=0 ° to 98 ° at 2 ° intervals by carrying out the integral given
by (1)*.
4. Plot v(r1 ) as a function of r1 , and compare the result with the P-wave structure shown
in class handout (section 2.1, a numerical table, jeffreys_model, for the Jeffreys model is
in practice_4.dir).
∫ ∑( f
L
f ( x)dx ≈ i −1 + f i )Δx / 2 , where Δx = L /( N − 1) , and fi = f ((i − 1) * Δx)
0
i=2
Ge162
218
Then, assume that seismic sources (X) and stations (square) are distributed on the
sides. The velocity is a function of space. Divide the whole area into cells (i ,j). The
wave speed in cell (i ,j) is vij (or slowness sij =1/ vij ) which we wish to determine from
observations. If we assume that the ray paths are straight (this assumption is not valid in
general for heterogeneous media). The travel time from source p to station q can be
written as
where lijpq is the path length in the (i ,j) cell for the p-q source-station ray. If the
geometry and paths are fixed, as in this case, the problem is linear, but the problem is in
general nonlinear, and the problem is usually linearized, as is done in the earthquake
location problem.
If the problem is linear, or is linearized, (1) can be solved by the method of least
squares or some other inversion methods.
We renumber the cells sequentially such that
k = i + ( j − 1)nx
t pq = ∑ lkpq sk (2)
k
219
Simple examples are given in the following.
If we place the sources p=1, 2, 3, and 4, and the receivers q=1, 2, 3, and 4, as
shown, we can have the ray geometry as shown in the following figure.
220
⎛ l111 l211 l311 . . . . . .⎞ ⎛ t11 ⎞
⎜ ⎟ ⎜ ⎟
⎜ l112 l212 l312 ⎟ ⎜ t12 ⎟
⎜ l113 l213 l313 ⎟ ⎜ t13 ⎟
⎜ ⎟ ⎜ ⎟
⎜ . ⎟ ⎛ s ⎞ ⎜ t14 ⎟
⎜ . ⎟ ⎜ 1 ⎟ ⎜ t21 ⎟
⎜ ⎟ ⎜ s2 ⎟ ⎜ ⎟
⎜ ⎟⎜ ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s3 ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s4 ⎟ ⎜ ⎟
⎜ . ⎟⎜ ⎟ ⎜ . ⎟
⎜ . ⎟ ⎜ s5 ⎟ = ⎜ . ⎟
⎜ ⎟⎜ s ⎟ ⎜ ⎟
⎜ . ⎟⎜ 6 ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s7 ⎟ ⎜ ⎟
⎜ . ⎟⎜ s ⎟ ⎜ . ⎟
⎜ . ⎟ 8 ⎜ . ⎟
⎜ ⎟ ⎜⎝ s9 ⎟⎠ ⎜ ⎟
⎜ . ⎟ ⎜ . ⎟
⎜ . ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ ⎟
⎜ . ⎟ ⎜ . ⎟
⎜l ⎟ ⎜ ⎟ (3)
⎝ 144 l244 l344 . . .⎠ ⎝ t44 ⎠
This problem can be solved with the method of least squares in the same way as
that used in the earthquake location problem.
Example 1.
s1 = s3 = s5 = s7 = s9 = 2 , s2 = s4 = s6 = s8 = 1 (checker board).
221
The solution is given as follows.
checker_b
cross_hole_simple
cell s Δs
1: 1.892 0.011
2: 1.218 0.015
3: 1.890 0.011
4: 0.892 0.010
5: 2.219 0.010
6: 0.890 0.010
7: 1.891 0.011
8: 1.219 0.015
9: 1.890 0.011
RMS of residual 0.01160
Actually, for this geometry the inverse problem is fairly ill-posed, and even if no
noise is applied, the solution is unstable (i.e., the standard error is large), as shown above.
If random noise of up to 10% is added to t pq , the error becomes very large, as
shown below.
222
checker_b
cross_hole_simple
cell s Δs
1: 2.149 3.187
2: 0.646 4.242
3: 2.091 3.190
4: 1.110 2.745
5: 2.083 2.750
6: 0.687 2.743
7: 1.478 3.176
8: 2.104 4.222
9: 1.538 3.179
RMS of residual 3.31218
Example 2.
223
Then the solution becomes more stable as shown below.
224
Caveats
1. The results depend on the initial parameterization of the cells and rays.
2. In a heterogeneous medium, the rays are not straight, and ray bending must be
considered.
3. The effect of finite wave length must be considered in resolution.
4. The solutions are often regularized (damped) to avoid instability.
225
226
J. Ritsema [written communication, 2002]
227
Spectral Structure of the Structural Heterogeneity in Earth
228
A. M. Dziewonski, Global seismic tomography: past, present and future, in
Problems in Geophysics for the New Millennium, Editrice Compositori, 2000.
229
"Slab" Structure in Deep Interior
Ritsema, J., and van Heijst, H. J., Seismic imaging of structural heterogeneity in Earth's
mantle: evidence for large-scale mantle flow, Science Progress, 83 (3), 243-259, 2000.
230
"Slab" Structures
231
232
Fukao, et al., Reviews of Geophysics, 39, 291, 2001.
Ge162 Practice Session 5 (Optional) Simple Tomography
233
⎛ l111 l211 l311 . . . . . .⎞ ⎛ t11 ⎞
⎜ ⎟ ⎜ ⎟
⎜ l112 l212 l312 ⎟ ⎜ t12 ⎟
⎜ l113 l213 l313 ⎟ ⎜ t13 ⎟
⎜ ⎟ ⎜ ⎟
⎜ . ⎟ ⎛ s ⎞ ⎜ t21 ⎟
⎜ . ⎟ ⎜ 1 ⎟ ⎜ t22 ⎟
⎜ ⎟ ⎜ s2 ⎟ ⎜ ⎟
⎜ ⎟⎜ ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s3 ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s4 ⎟ ⎜ ⎟
⎜ . ⎟⎜ ⎟ ⎜ . ⎟
⎜ . ⎟ ⎜ s5 ⎟ = ⎜ . ⎟
⎜ ⎟⎜ s ⎟ ⎜ ⎟
⎜ . ⎟⎜ 6 ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ s7 ⎟ ⎜ ⎟
⎜ . ⎟⎜ s ⎟ ⎜ . ⎟
⎜ . ⎟ 8 ⎜ . ⎟
⎜ ⎟ ⎜⎝ s9 ⎟⎠ ⎜ ⎟
⎜ . ⎟ ⎜ . ⎟
⎜ . ⎟ ⎜ . ⎟
⎜ ⎟ ⎜ ⎟
⎜ . ⎟ ⎜ . ⎟
⎜l ⎟ ⎜ ⎟ (3)
⎝ 166 l266 l366 . . .⎠ ⎝ t66 ⎠
As we did in the practice session for earthquake location, this problem can be solved
using the method of least squares. (3) can be written as
Am = d (4)
where A is an 18x9 matrix and m and d are column vectors containing the parameters to
be determined and the data, respectively.
The normal equation is,
AT Am = AT d (5)
234
If AT A is not singular, the formal least-squares solution is given by
m = ( AT A) −1 AT d (6)
and the error estimates are determined by the variance of the data and the diagonal
elements of the inverse matrix of the normal equation (6 ). Usually, we write the
uncertainty in mi by Δmi , and compute it by
∑ (t
N
Δmi = cii o
( kl ) − t(ckl ) ) 2 /( N − N p ) , i=1, 2, 3,...,9 (7)
( kl ) =1
where N p is the number of parameters (here 9) and cii are the diagonal elements of
( AT A) −1 .
1. Determine the slowness si, i=1, 2,3,.....9, and the associated errors.
The matrix elements and the data are in o_lsq_mat_ts_0.1 and o_lsq_rhs_ts_0.1 in
/home/ftp/pub/hiroo/ge162.dir/practice5.dir, respectively. In this computation, the
235
length of the side is 10 km, and errors up to 10% are added to the data. (This is the same
geometry as Example 2 in class note. The numerical values of the standard errors may be
slightly different, because of the difference in the definition.)
Actually, in the real problems, the computations of the ray paths and the path
lengths in each cell are most difficult. In this problem, they are computed and the values
are put in o_lsq_mat_ts_0.1. (there are small round-off errors of the order of 0.3 %.)
Check the values for the first 3 rows of o_lsq_mat_ts_0.1.
Try to write your own program to do this problem, but if you find it difficult, you
can use the program tomo2.f provided in practice_5.dir. It takes an input file i_tomo2
which contains the names of the files for the matrix elements and the data. It also
requires c_tomo2, but this should not be changed (it has a constant for regularization of
the matrix inversion.).
2. Drop the data for source #6, and try the same.
3. Drop the data for sources #5 and #6, and try the same.
236
o_lsq_mat_ts_0.1
checker_b
two_sides
9 18 0.1000
⎛ 10.02 9.99 10.02 0.00 0.00 0.00 0.00 0.00 0.00⎞
⎜ 10.56 5.28 0.00 0.00 5.28 10.56 0.00 0.00 0.00 ⎟
⎜ ⎟
⎜ 9.05 0.00 0.00 3.03 12.01 3.03 0.00 0.00 9.05 ⎟
⎜ ⎟
⎜ 0.00 5.28 10.56 10.56 5.28 0.00 0.00 0.00 0.00 ⎟
⎜ 0.00 0.00 0.00 10.02 9.99 10.02 0.00 0.00 0.00 ⎟
⎜ ⎟
⎜ 0.00 0.00 0.00 10.56 5.28 0.00 0.00 5.28 10.56 ⎟
⎜ 0.00 0.00 9.05 3.03 12.01 3.03 9.05 0.00 0.00 ⎟
⎜ ⎟
⎜ 0.00 0.00 0.00 0.00 5.28 10.56 10.56 5.28 0.00 ⎟
⎜ 0.00 0.00 0.00 0.00 0.00 0.00 10.02 9.99 10.02 ⎟
⎜ ⎟
⎜ 10.02 0.00 0.00 9.99 0.00 0.00 10.02 0.00 0.00 ⎟
⎜ ⎟
⎜ 10.56 0.00 0.00 5.28 5.28 0.00 0.00 10.56 0.00 ⎟
⎜ 9.05 3.03 0.00 0.00 12.01 0.00 0.00 3.03 9.01 ⎟
⎜ ⎟
⎜ 0.00 10.56 0.00 5.28 5.28 0.00 10.56 0.00 0.00 ⎟
⎜ 0.00 10.02 0.00 0.00 9.99 0.00 0.00 10.02 0.00 ⎟
⎜ ⎟
⎜ 0.00 10.56 0.00 0.00 5.28 5.28 0.00 0.00 10.56 ⎟
⎜ 0.00 3.03 9.05 0.00 12.01 0.00 9.05 3.03 0.00 ⎟
⎜ ⎟
⎜ 0.00 0.00 10.56 0.00 5.28 5.28 0.00 10.56 0.00 ⎟
⎜ 10.02 ⎟⎠
⎝ 0.00 0.00 10.02 0.00 0.00 9.99 0.00 0.00
237
o_lsq_rhs_ts_0.1
checker_b
two_sides
9 18 0.1000
238
⎛ 47.28 ⎞
⎜ ⎟
⎜ 44.36 ⎟
⎜ 67.07 ⎟
⎜ ⎟
⎜ 44.95 ⎟
⎜ 39.69 ⎟
⎜ ⎟
⎜ 43.15 ⎟
⎜ 60.62 ⎟
⎜ ⎟
⎜ 51.73 ⎟
⎜ ⎟
⎜ 51.76 ⎟
⎜ 49.93 ⎟
⎜ ⎟
⎜ 49.19 ⎟
⎜ 68.50 ⎟
⎜ ⎟
⎜ 45.47 ⎟
⎜ 36.96 ⎟⎟
⎜
⎜ 51.15 ⎟
⎜ ⎟
⎜ 61.99 ⎟
⎜ 46.84 ⎟
⎜⎜ ⎟
⎜⎝ 53.53 ⎟⎠⎟
239
Ge 162
5.1.1 Static Displacement Field Due to a Single Point Force in Infinite Homogeneous
Medium (LW, pp. 323-331)
Consider an isotropic infinite homogeneous elastic medium with density ρ and elastic
constants λ and μ.
We apply a force at point O, and want to determine the displacement u at point P.
We assume that the outer boundary at infinity is constrained, i.e., u =0 at infinity.
240
2. Qualitative Solution
F = 4π r 2σ (1)
Let u(r) be the magnitude of the displacement at r, then, from the geometry shown in the
figure, du/dr gives the magnitude of the strain at r.
du F
−E ≈σ ≈ (2)
dr 4π r 2
241
(The negative sign is taken because u(r) decreases as r increases.) Here E˜ represents an
appropriate elastic constant, and is of the order of λ or μ . Integrating (4), and using
u (r → ∞) = 0 , we obtain
F
u (r ) ≈ (3)
4π Er
Actually, a rigorous derivation shows (in the next section) that, if we define
the polar coordinates as shown in the figure,
F
ur = sin θ
4πμ r
(4)
F α
uθ = (1 − ) cos θ
4πμ r 2
where
λ+μ
α=
λ + 2μ
242
Since for most solids, λ = μ and α =2/3, we see that (3) is a good approximation
of (4) and that the directions of the displacement shown in the figure are consistent with
(4).
Definition
F = lim ( ρ f δ V )
δ V →0
δ (r ) = 0 , r ≠ 0, ∫ δ (r)dV = 1
V
where V is a volume which includes the origin r=0. A convenient expression for this is
1 2
δ (r ) = − ∇ (1/ r) (5)
4π
∇ 2 (1/ r) = 0 , if r ≠ 0 .
243
We can show that
Although this looks like a very simple problem, it is not that easy. We have to start
with elasto-static equation of equilibrium. From equation (29) of 3.1, we have
where f is the force per unit mass. Consider a point force of magnitude F at the origin.
ρ f = Faδ (r ) (8)
where a is the unit vector in the direction of the force. By using (5) we have
⎛ a ⎞ ⎡ ⎛ a ⎞ ⎛ a ⎞⎤
ρ f = −F∇2 ⎜ ⎟ = − F ⎢graddiv ⎜ ⎟ − curlcurl ⎜ ⎟⎥ (9)
⎝ 4π r ⎠ ⎣ ⎝ 4π r ⎠ ⎝ 4π r ⎠ ⎦
This form may appear somewhat artificial, but it is suggested from the fact that any
displacement field can be represented by a sum of solenoidal and irrotational fields and
that the forcing term (9) is given in this form. Substituting (9) and (10) into (7), we have
244
⎛ a ⎞ ⎛ a ⎞
− Fgraddiv ⎜ ⎟ + Fcurlcurl ⎜ ⎟ + (λ + 2μ )graddiv(graddivAp − curlcurlAp )
⎝ 4π r ⎠ ⎝ 4π r ⎠
− μ curlcurl(graddivAs − curlcurlAs ) = 0
In the third term curlcurlAs is replaced by curlcurlAp , because its divergence vanishes
anyway. Similarly, graddivAp in the fourth term is replaced by graddivAs . Using the
relation
graddivAp − curlcurlAp = ∇ 2 Ap
⎛ a ⎞ ⎛ a ⎞
graddiv ⎜ − F + (λ + 2 μ )∇ 2 Ap ⎟ + curlcurl ⎜ F − μ∇ 2 As ⎟ = 0 (11)
⎝ 4π r ⎠ ⎝ 4π r ⎠
a
(λ + 2μ )∇ 2 Ap = F (12)
4π r
a
μ∇ 2 As = F (13)
4π r
F F
∇ 2 Ap = and ∇ 2 As = (14)
4π (λ + 2 μ )r 4πμr
2
Since ∇ 2 r = , we have
r
245
F
Ap = r (15)
8π ( λ + 2μ )
and
F
As = r (16)
8πμ
Using (10)
Substituting (15) and (10) into (17), we have for the i-th component of displacement
for a unit force (F=1) in j-th direction, uij ,
1 ∂ ∂r 1 ∂ ∂r 1 2
uij = − + δ ij ∇r
8π (λ + 2μ ) ∂ xi ∂ x j 8πμ ∂ xi ∂ x j 8πμ
(18)
1 ⎛ λ + μ ∂ 2r ⎞ 1
= ⎜⎜ δ ij ∇ r −
8πμ ⎝
2
λ + 2μ ∂ xi∂ x j
⎟⎟ =
8πμ
(δ ij∇ 2r − α r,ij ) = u ij
⎠
where
λ+μ
α= (2/3 for most solids) (19)
λ + 2μ
Equation (18) gives the solution of our problem. Note that uij = uij (symmetric). uij is
often called the Somigliana Tensor. We write all the components explicitly,
246
1 ⎡2 ⎛ 1 x12 ⎞ ⎤ 1 ⎛ x1 x2 ⎞
u11 = ⎢ − α ⎜ − 3 ⎟⎥ , u12 = ⎜α ⎟
8πμ ⎣ r ⎝ r r ⎠⎦ 8πμ ⎝ r 3 ⎠
1 ⎛ x1 x3 ⎞ 1 ⎡2 ⎛ 1 x22 ⎞ ⎤
u13 = ⎜α ⎟, u22 = ⎢ − α ⎜ − 3 ⎟⎥ (20)
8πμ ⎝ r 3 ⎠ 8πμ ⎣ r ⎝ r r ⎠⎦
1 ⎛ x2 x3 ⎞ 1 ⎡2 ⎛ 1 x32 ⎞ ⎤
u23 = ⎜α ⎟, u33 = ⎢ − α ⎜ − 3 ⎟⎥
8πμ ⎝ r 3 ⎠ 8πμ ⎣ r ⎝ r r ⎠⎦
Then
On x1 − x3 plane, φ =0,
F
ur = sin θ u11 + cos θ u31 = sin θ (22)
4πμ r
247
F ⎛ α⎞
uθ = cos θ u11 − sin θ u31 = ⎜1 − ⎟ cos θ (23)
4πμ r ⎝ 2 ⎠
248
Single Couple
r 2 = ( x1 − ξ1 ) 2 + ( x2 − ξ 2 ) 2 + ( x3 − ξ3 ) 2 ≡ ( xi − ξi ) 2 .
(1/2)dξ2
1 1
If we apply a force F in x1 direction at ξ 2 + dξ 2 , and F in - x1 direction at ξ 2 − dξ 2 ,
2 2
1 1
Fui1 (ξ1, ξ 2 + dξ 2 , ξ3 , x1, x2, x3 ) − Fui1 (ξ1, ξ 2 − dξ 2 , ξ3 , x1, x2, x3 )
2 2
∂u 1
= F i dξ 2 + o(d 2ξ 2 )
∂ξ 2
(24)
249
Since r 2 = ( x1 − ξ1 ) 2 + ( x2 − ξ 2 ) 2 + ( x3 − ξ3 ) 2 ≡ ( xi − ξi ) 2 ,
∂r ∂r
=− (25)
∂ξi ∂ xi
∂ uij ∂uj
=− i (26)
∂ξ k ∂ xk
∂ ui1
−F dξ 2 + o(d 2ξ 2 ) (27)
∂ x2
∂ui1
Ui = M = Mui1,2 (27')
∂ξ 2
or
∂ui1
Ui = −M = − Mui1,2 (27'')
∂x2
Thus, the displacement due to this single couple placed at the origin can be obtained by
replacing F by M in (20), taking the derivative of (20) with respect to x2, and changing
the sign.
250
⎡ x2
M ⎛ x2 x12 x2 ⎞ ⎤
U1 = − ⎢ −2 − α −
⎜ 3 + 3 ⎟⎥ (28)
8πμ
⎣ r
3
⎝ r r 5 ⎠⎦
M ⎡ x1 x x2 ⎤
U2 = − α ⎢ 3 − 3 1 52 ⎥ (29)
8πμ ⎣ r r ⎦
M ⎡ x1 x2 x3 ⎤
U3 = − α ⎢ −3 5 ⎥ (30)
8πμ ⎣ r ⎦
∂ui2
Ui = M = Mui2,1
∂ξ1
or
∂ui2
Ui = −M = − Mui2,1
∂x1
M ⎡ x2 x2 x12 ⎤
U1 = − α −3 5 ⎥ (31)
8πμ ⎢⎣ r 3 r ⎦
M ⎡ x1 ⎛ x1 x22 x1 ⎞ ⎤
U2 = − ⎢ −2 3 − α ⎜ − 3 + 3 5 ⎟ ⎥ (32)
8πμ ⎣ r ⎝ r r ⎠⎦
M ⎡ x1 x2 x3 ⎤
U3 = − α ⎢ −3 5 ⎥ (33)
8πμ ⎣ r ⎦
251
Double Couple
It is evident that for a double couple as shown in the figure, the displacements can
be given by
∂ui2
Ui = M = M (ui1,2 + ui2,1 ) (34)
∂ξ1
or
∂ui2
U i = −M = − M (ui1,2 + ui2,1 ) = − M (u1,2
i
+ u2,1
i
) (35)
∂x1
From the sum of (28) and (31), (29) and (32), and (30 and (33), we obtain
M ⎛ x2 ⎞ 2α M ⎛ x2 x12 x2 ⎞ M x2 ⎡ ⎛ x12 ⎞ ⎤
U1 = ⎜2 ⎟− ⎜ 3 − 3 ⎟ = ⎢1 − α ⎜ 1 − 3 ⎟⎥ (36)
8πμ ⎝ r 3 ⎠ 8πμ ⎝r r 5 ⎠ 4πμ r 2 r ⎣ ⎝ r 2 ⎠⎦
M ⎛ x1 ⎞ 2α M ⎛ x1 x22 x1 ⎞ M x1 ⎡ ⎛ x22 ⎞ ⎤
U2 = ⎜2 ⎟− ⎜ 3 − 3 ⎟ = ⎢1 − α ⎜ 1 − 3 ⎟⎥ (37)
8πμ ⎝ r 3 ⎠ 8πμ ⎝r r 5 ⎠ 4πμ r 2 r ⎣ ⎝ r 2 ⎠⎦
M ⎛ x1 x2 x3 ⎞ M ⎛ x1 x2 x3 ⎞
U3 = ⎜ 6α ⎟= ⎜ 3α ⎟ (38)
8πμ ⎝ r ⎠ 4πμ r 2 ⎝
5
r3 ⎠
252
For double couples with different orientations, similar expressions can be obtained.
253
M ⎛ α⎞ 2
ur = ⎜ 1 + ⎟ sin θ sin 2φ (39)
4πμ r ⎝ 2 ⎠
2
M ⎛1 α ⎞
uθ = ⎜ − ⎟ sin 2θ sin 2φ (40)
4πμ r 2 ⎝ 2 2 ⎠
M
uφ = (1 − α ) sin θ cos 2φ (41)
4πμ r 2
⎛⎛ α ⎞ 2 ⎞
⎜ ⎜1 + 2 ⎟ sin θ sin 2φ ⎟
⎛ ur ⎞ ⎜⎝ ⎠ ⎟
⎜ ⎟ M ⎜⎛ 1 α ⎞ ⎟
Or, ⎜ uθ ⎟ = 4πμ r 2 ⎜ ⎜ 2 − 2 ⎟ sin 2θ sin 2φ ⎟ (42)
⎜u ⎟ ⎜⎝ ⎠ ⎟
⎝ φ⎠ ⎜ (1 − α ) sin θ cos 2φ ⎟
⎜⎜ ⎟⎟
⎝ ⎠
⎛ α⎞
ur ≈ ⎜ 1 + ⎟ sin 2φ
⎝ 2⎠
uφ ≈ (1 − α ) cos 2φ
254
The displacement field on a circle x12 + x22 = const is also shown.
∂uik
U iSC =
∂sl
∂
Here, is the derivative in l direction, and l is taken positive in the direction towards
∂sl
∂ ∂
the force oriented in positive k direction. Then, since = lq and uik = uip k p ,
∂sl ∂ξ q
255
U iDC = uip , q (k p lq + kq l p ) (44)
Using the expressions discussed in 5.1.2 (i.e., uiSC = uip ,q k p lq ), we can write the
displacement due to a force dipole as uiDP = uip , q k p lq where k and l are the unit
vectors in the direction of force and the arm, respectively. k and l are parallel
but they can be of either opposite or the same direction, depending on the
orientation of the force (i.e., inward or outward).
Show that the force system with two dipoles (a) shown below is equivalent
to the double couple (b) (all the dipoles and the couples are with unit moment.).
Ge 162
256
Consider a fault shown in the figure. The displacement on one side is u+ and the
other side is u-, so that the displacement discontinuity is Δu=u+-u- . This is called fault
offset and is often denoted by D. Since we already know the displacement field due to a
point force and force couples, we will first try to produce, by applying forces or force
couples to the medium without dislocation, a displacement field similar to that caused by
faulting.
257
u+ − u−
F =μ S (1)
Δw
lim F Δw = μ (u + − u − ) S = μ DS
Δw→∞
M = μ DS (2)
258
make the gap very small so that Σ + and Σ − form an open surface Σ . Then weld the two
surfaces and remove the force. Now we have a displacement discontinuity
Δu = u + − u −
across Σ . This is called the elastic dislocation. Note that, for a dislocation created this
way, the traction must be continuous across Σ .
If u + and u − are tangential to Σ , Δu is called the shear dislocation.
259
4. Elasto-Static Theory of Dislocation
u k (Q) = ∫ Δu τ ν dS
Σ
k
i ij j (3)
stress on Σ due to a unit force in k-th direction at Q (the unit of τ ijk should be
understood
as stress/force=1/area ); Δui is the displacement discontinuity (dislocation) across Σ. i.e.
Δui =ui (on Σ + )-ui (on Σ − ). Also,
260
Let us consider one of the simplest cases.
Suppose we have a very small planar shear dislocation surface dΣ at P which is
perpendicular to x2 axis. We consider a displacement discontinuity in x1 direction.
261
u k (Q) = d Σ[Δu1 (τ 11k ν 1 + τ 12k ν 2 + τ 13k ν 3 )
+ Δu2 (τ 21k ν 1 + τ 22
k
ν 2 + τ 23k ν 3 )
(4)
+ Δu3 (τ ν + τ ν + τ ν )]
k
31 1
k
32 2
k
33 3
= d ΣΔu1τ 12k
Note that
τ k = μ (u1,2
12
k
+ u2,1
k
)
Thus,
This means that u k (Q) is equal to the k-th component of displacement at Q due to a
double couple of moment M = μΔu1dΣ placed at P. Since this relation evidently holds
for any dΣ, we can conclude that a shear dislocation Δus over a surface S is equivalent to
a double couple whose total moment is
M=
∫ μΔu d Σ =μΔu S
S
s s (6)
262
5. Volterra's Dislocation Theory
We first introduce the following notation for the stress strain relation:
where
Note:
Among 81 components of cijpq, most of them are zero. Nonzero components are
c1111=c2222=c3333= λ + 2μ
c1122=c2211=c1133=c3311=c2233=c3322= λ
c1212=c1221=c2112=c2121=c1313=c1331= μ
c3113=c3131=c2323=c2332=c3223=c3232= μ
263
Then, the equation of equilibrium can be written as:
264
cijpq v p, qj = − ρgi (12)
Multiplying (11) by vi and (12) by ui , subtracting and integrating over the volume V, we
obtain
∫V
(cijpq u p ,qj vi − cijpq v p ,qj ui )dV = − ∫ (ρ f i vi − ρ gi ui )dV
V
(13)
Note that
Owing to the symmetry of cijpq, the second term of RHS vanishes. Hence,
∫ V
(cijpq u p , q vi − cijpq v p ,q ui ), j dV = − ∫ (ρ fi vi − ρ gi ui )dV
V
(14)
265
∫S1 +Σ + +Σ −
(cijpq u p ,q vi n j − cijpq v p ,q ui n j )dS = − ∫ (ρ fi vi − ρ g i ui )dV
V
(15)
Due to the boundary conditions on S1, the integral on S1 vanishes. Since vi is continuous
on Σ
∫Σ + +Σ −
cijpq u p ,q vi n j dS = ∫ + cijpq u +p ,q vi n +j dS + ∫ − cijpq u −p ,q vi n −j dS
Σ Σ
In equilibrium, the traction on the Σ + side cijpqu +p ,q n +j should be balanced by that on the Σ -
266
∫
Σ + +Σ −
cijpq v p ,q ui n j dS = ∫ + cijpq v p , q ui+ (−ν j )dS + ∫ − cijpq v p , q ui− (+ν j )dS
Σ Σ
= − ∫ cijpq v p , q Δuiν j dS
Σ
Hence,
We put ρ fi = 0, and, for gi, consider a unit force in k-th direction at point Q:
ρ gi = δ ik δ (Q)
This is the Volterra's relation. τ ijk is the stress due to a unit force in k-th direction at Q,
and its unit should be stress/force = 1/area.
6. Moment Tensor
267
Applying the Velterra's relation
where
Mpq is a symmetric tensor called the moment tensor. Then (19) can be written as
⎛ M 11 ⎞
⎜ ⎟
⎜ M 22 ⎟
⎛ u1 ⎞ ⎛ u1,1
1
u12,2 1
u3,3 1
u1,2 + u12,1 u1,3
1
+ u3,1
1
u2,3 + u3,2 ⎞
1 1
⎜ 2⎟ ⎜ 2 1 ⎟
⎜ M 33 ⎟
⎜ u ⎟ = ⎜ u1,1 u2,2 + u2,1 + u3,1 u12,3 + u3,2
2 2 2 2 2 2
u3,3 u1,2 u1,3 ⎟⎜ M ⎟ (21)
⎜ u 3 ⎟ ⎜ u1,1 1 ⎟⎜ 12 ⎟
⎝ ⎠ ⎝
3 3
u2,2 3
u3,3 3
u1,2 + u2,1
3 3
u1,3 + u3,1
3
u12,3 + u3,2 ⎠⎜ M ⎟
⎜⎜ M ⎟⎟
13
⎝ 23 ⎠
268
Using the explicit expressions for cijpq, (20) can be written as,
M 11 = 2ΔS μΔu1ν 1
M 22 = 2ΔS μΔu2ν 2
M 33 = 2ΔS μΔu3ν 3
(23)
M 12 = ΔS μ (Δu2ν 1 + Δu1ν 2 )
M 13 = ΔS μ (Δu3ν 1 + Δu1ν 3 )
M 23 = ΔS μ (Δu2ν 3 + Δu3ν 2 )
M=μΔS ( Δu Tν + ν T Δu ) (23')
By taking the limit, ΔS => 0 ( ΔuiΔS is kept finite), we obtain a point dislocation
source that is represented by a moment tensor Mpq.
As a special case, consider a case where
269
ν = (0,1,0), and Δu = (Δu1 ,0,0) .
u1 = (u1,2
1
+ u12,1 ) M 12
u 2 = (u1,2
2
+ u2,1
2
) M 12
u 3 = (u1,2
3
+ u2,1
3
) M 12
( u ) = ( U )(M )
l
k
l pq (24)
where ( lU ) is the matrix on the RHS of (21) which can be computed by the known
Somigliana tensor. Combining (24) for l=1,2,3, ....L, we obtain a set of linear equations
270
with 6 unknowns Mpq and Lx3 data, and, for L>2 , the equations can be solved (unless the
problem is ill-posed) to obtain Mpq.
Consider a fault model as sketched in Figure 1. Only the foot-wall block is shown.
(ξ1, ξ2, ξ3) is the fault coordinate system (+ξ1 is in the direction of fault strike), and (x1,
x2, x3) is the geographical coordinate system (x1: North; x2: West; x3: Up).
The top surface of the block coincides with the surface of the Earth. D is the slip
vector of the hanging-wall block (not shown). δ is the dip angle of the fault plane
measured downward from the horizontal (i.e. from -ξ2 direction.). λ is the slip angle
(rake) measured counter-clockwise on the fault plane from the horizontal line as shown in
the figure (i.e. from +ξ1 direction.). ν is the unit vector normal to the fault plane. Let ΔS
be the fault area.
Place this fault model in a homogeneous elastic medium with the rigidity μ at the
origin of the Cartesian coordinates as shown in Figure 2. The strike of the fault, φ, is
measured clockwise on the free surface from x1 axis in Figure 2. We assume that ΔS is
small so that the fault can be considered a point source.
2) Determine the x1, x2, x3 components of the displacement vector D (magnitude D) and
the unit vector ν .
3) Determine the moment tensor elements Mij (i, j =1, 2, 3) for this fault model. Write
the result in terms of scalar moment M0=μDΔS, δ, λ, and φ. (The results are given in
5.1.4.) (Note that this is a shear dislocation. )
271
4) Determine the moment tensors for the Hector Mine earthquake, the Chi-Chi Taiwan
earthquake, and the Bhuj India earthquake. The fault parameters for these events are
given below ( δ (dip in degree), λ (rake in degree), φ f (strike in degree), M 0 (Seismic
Moment in N-m)).
Ge 162
272
1. Green's Function
Apply unit point force in k direction at ξ = (ξ1 , ξ 2 , ξ3 ) . Then, the i-th component
of the displacement at x = ( x1 , x2 , x3 ) is the Green' function Gik ( x , ξ ) .
ui = F k Gik (1)
3. Distributed force
4. Force Couples
273
In a homogeneous whole space, (3) and (4) can be also written as
ui
SC
= −Gi1,2 and uiDC = −(Gi1,2 + Gi2,1 ) .
This can be shown as follows. Let Gi( f ) be the i-th component of displacement
∂
due to a unit force in k direction, and be the derivative in l direction, then
∂sl
∂Gi( f ) ∂
uiSC = = lq (Gip k p ) = Gip , q k p lq (5)
∂sl ∂xq
For a source represented by force couples and dipoles with the strength (moment)
given by M kj ,
u i = M kj Gik , j (7)
5. Dislocation source
274
i
u i = ∫ τ ipq Δu pν q dS (8)
Σ
where
275
u i = M kl Gki ,l (12)
ui = M kl Gik ,l (12')
where
M kl = c pqkl Δu pν q ΔS = M lk (13)
This is the same as the expression for the displacement field due to force couples (7).
Thus, a dislocation source is equivalent to force couples.
M 11 = 2 μΔS Δu1ν 1
M 22 = 2μΔS Δu2ν 2
M 33 = 2 μΔS Δu3ν 3
(15)
M 12 = μΔS ( Δu1ν 2 + Δu1ν 2 )
M 13 = μΔS ( Δu1ν 3 + Δu3ν 1 )
M 23 = μΔS ( Δu2ν 3 + Δu3ν 2 )
276
or,
The following figure illustrates the force couples represented by each element of a
moment tensor.
277
6. Fault parameters and Moment tensor elements.
The fault parameters (δ, λ, and φf ) defined in the figure below can be transformed
to moment tensor elements by ( μ DS is assumed to be 1),
278
(The following part is not complete yet)
6. General Representation
Suppose that the medium is in equilibrium in the beginning (A). Then a failure
occurs in V0 (B), and displacement ui and stress σ ij are produced in the medium.
279
The equation for static equilibrium,
σ ij , j = 0 (16)
where
280
ui =
∫ ( −m
V0
kl ,l Gik )dV (19)
(S0 can be taken slightly outside of V0). Then, (19) can be written as
ui =
∫
V0
mkl Gik ,l dV (20)
∫ V0
mkl dV → M kl (21)
and
ui = M kl Gik ,l (22)
281
Consider an elastic medium which is in equilibrium under stress σ ij (r ) . Let S be
an open surface defined in this medium of which the stress is σ ij(0) (r ) . Relax the stress
from σ ij(0) to σ ij(1) by creating a cut on S and letting the material on either side
of S to move. Let Δu be the resulting dislocation (displacement discontinuity). Δu is
not necessarily parallel to S. We want to determine Δu , and the difference in strain
energy ΔW between the two states, before and after the crack formation.
Δε ≈ D / ( L / 2 )
282
This strain change is caused by the change in the stress Δσ = σ 0 − σ 1 , which we call
stress drop. Thus,
Δσ = μΔε = cμ D / ( L / 2 ) (1)
where c is a non-dimensional constant which depends upon the shape of the crack; it is
apparent from the above argument that c ~ 1.
The simplest case is a circular crack. If a is the radius of the crack, we can show
that L = 2a and c = 7π /16 ( λ = μ is assumed.).
The change in the strain energy ΔW is given by Sσ D . This is analogous to the change
in the strain energy of a spring (spring constant k) which is stretched from x0 to x1 . The
change in the strain energy is
283
σ0
σ0
we can prove that c = 2 / π . If the crack is not infinitely long but of finite length L, the
above relation is not exact. However, when L is very long, it is a good approximation.
For practical applications, we often take one half of the crack as shown by the
hatched portion in the figure. For this crack,
Δσ = (2 / π ) μ ( D / w)
284
σ0
σ0
The crack extends in the direction perpendicular to the uniform shear stress. In
this case we can derive
4(λ + μ ) μ
Δσ = ( D / w)
π (λ + 2 μ )
We call this crack a dip-slip crack. In mechanics, this is called Mode II crack. For the
hatched portion,
π (λ + 2 μ ) 2
ΔW = Lwσ D = w LΔσσ
4(λ + μ ) μ
and
π (λ + 2 μ ) 2
M 0 = μ LwD = w L Δσ
4(λ + μ )
285
Strain Energy, (16 / 7 μ )a 3 Δσσ (π / 2μ )w2 LΔσσ π (λ + 2 μ ) 2
w LΔσσ
ΔW = Sσ D 4(λ + μ ) μ
Moment, (16 / 7)a 3 Δσ (π / 2)w2 LΔσ π (λ + 2 μ ) 2
w LΔσ
M 0 = μ SD 4(λ + μ )
Relations between stress drop, strain energy change, offset, dimension, and moment for static
cracks. Dimensions of the fault are a radius, L length, w width; initial stress is σ 0 ; final stress is σ 1 ;
stress drop is Δσ = (σ 0 − σ 1 ) ; average stress is σ = 1/ 2(σ 0 + σ 1 ) ; average dislocation is D .
Here, we consider only a strike-slip crack (Mode III). In this case, the
displacement field has a component only in the direction parallel to the crack. Therefore,
the problem becomes 2-dimensional.
286
Initially the crack is not formed. We introduce initial displacement field
U = (0, 0, U 3 ) (7)
divU = 0 (8)
gives
curlcurlU = 0 (10)
If we put
A = curlU (11)
curlA = 0 (12)
287
From (12) we can put
A = − gradφ (14)
∇ 2φ = 0 (15)
⎛ ∂U 3 ∂U ⎞ ⎛ ∂φ ∂φ ∂φ ⎞
⎜ , − 3 , 0⎟ = ⎜ − ,− ,− ⎟ (16)
⎝ ∂x2 ∂x1 ⎠ ⎝ ∂x1 ∂x2 ∂x3 ⎠
Let us consider a case where the stress on the crack is completely released,
i.e., τ 32 = 0 on S.
∂U 3
μ =0 on S
∂x2
∂φ
=0 on S
∂x1
∂φ
Since = 0 (see (16)), φ = φ ( x2 ) . In the limit of a very thin crack x2 = 0 on S; thus,
∂x3
φ = const on S. Since φ is a potential, we can put, for the boundary condition
φ =0, on S (17)
288
Thus, our problem is now reduced to a problem of finding a solution of (15), with a
boundary condition (17). Note that this problem is equivalent to the well-known electro-
static problem of finding an electro-static field due to a perfect conductor placed in
vacuum.
------------------------------------------------------------------------------------------------------------
-
To solve this, we introduce elliptic coordinates (ξ1 , ξ 2 ) by
where ζ = ξ1 + iξ 2 . ξ1 = const gives an ellipse with a major axis a cosh ξ1 , and a minor
axis a sinh ξ1 . Thus, the elliptic crack is given by ξ1 = const, and the limiting case ξ1 =0
represents an infinitesimally thin crack of width 2a.
We write
φ = φ0 + φ ′ (20)
where φ0 represents the initial field, and φ ′ , the perturbation due to the crack. Since the
initial field is U 3(0) = A0 x2 ,
∇ 2φ ′ = 0 (23)
289
From ∇ 2φ ′ = 0 , we have (use Laplacian for the elliptic coordinates),
d 2F
=F
dξ12
Because F (ξ1 ) must be finite for ξ1 → ∞ , we take only the first term
A = A0 a
1
φ = − A0 a sinh ξ1 cos ξ 2 = − A0 a(sinh ζ + (sinh ζ )* )
2 (27)
= − A0 a Re(sinh ζ ) = − A0 a Re(cosh ζ − 1) = − A0 Re( w − a )
2 1/ 2 2 2 1/ 2
∂φ ∂ (−U 3 ) ∂φ ∂ (−U 3 )
= and =−
∂x1 ∂x2 ∂x2 ∂x1
σ0
U 3 = A0 Im( w2 − a 2 )1/ 2 = Im( w2 − a 2 )1/ 2 (28)
μ
290
is the solution of the problem.
------------------------------------------------------------------------------------------------------------
-
σ0
U 3 = A0 Im( w2 − a 2 )1/ 2 = Im( w2 − a 2 )1/ 2 ( w = x1 + ix2 ) (28)
μ
σ0
U3 = Im( x12 − a 2 )1/ 2
μ
U3 = 0 , x1 > a (29)
∴
σ 0 2 2 1/ 2
U3 = (a − x1 ) , x1 ≤ a
μ
σ0a/μ
291
⎛σ ⎞
U 3max = ⎜ 0 ⎟ a
⎝ μ ⎠
1 a σ0 2
(a − x12 )1/ 2 dx1 = (πσ 0 / 4μ )a
2a ∫− a μ
U3 = (30)
D = 2U 3 = (πσ 0 / 2 μ ) a (31)
For a partial stress drop in which stress drops from σ 0 to σ 1 , σ 0 in (31) should simply
be replaced by Δσ = σ 0 − σ 1 . If we replace a in (31) by width w , (31) leads to the
relation given in the table.
σ0 ⎛σ ⎞
U3 = Im(− x22 − a 2 )1/ 2 = ± ⎜ 0 ⎟ ( x22 + a 2 )1/ 2 (32)
μ ⎝ μ ⎠
292
σ0x2/μ
σ0a/μ
-σ0a/μ
⎛σ ⎞
Since the initial field is ⎜ 0 ⎟ x2 , the actual displacement caused by the crack is
⎝ μ ⎠
⎛σ ⎞
obtained by subtracting ⎜ 0 ⎟ x2 from (32):
⎝ μ ⎠
⎛σ ⎞ ⎛σ ⎞
U 3 = U 3 − ⎜ 0 ⎟ x2 = ⎜ 0 ⎟ ⎡⎣ ±( x22 + a 2 )1/ 2 − x2 ⎤⎦ (33)
⎝ μ ⎠ ⎝ μ ⎠
σ0a/μ
-σ0a/μ
X 1/ 2 = ( 3 / 4 ) a or a = ( 4 / 3) X 1/ 2 (34)
293
The three-dimensional displacement is shown in the figure above.
If we use this model, we can estimate the depth of the fault from the observed
decay rate of the displacement field by using (34).
294
295
4. Stress
1 1
τ 31 = 0 , and τ 32 = σ 0 x2 =σ0 (35)
(x + a )
2
2
2 1/ 2
(1 + (a / x2 ) 2 )1/ 2
296
----------------------------------------------------------------------------------------------------------
Note: (28) gives the solution as an imaginary part of an analytic function. The
derivatives of U 3 with respect to x1 and x2 can be obtained using the following relation.
Consider an analytic function Ψ ( w) = P( x1 , x2 ) + iQ( x1 , x2 ) of w = x1 + ix2 . Then
∂Ψ ∂P ∂Q ∂Ψ ∂P ∂Q ∂P ∂Q
= +i , and = +i = −i +
∂w ∂x1 ∂x1 ∂w ∂ (ix2 ) ∂ (ix2 ) ∂x2 ∂x2
∂Q ∂Ψ ∂Q ∂Ψ
Hence, is the imaginary part of , and is the real part of .
∂x1 ∂w ∂x2 ∂w
----------------------------------------------------------------------------------------------------------
297
σ0
τ 32 = 0 , | x1 |< a
(36)
1
τ 32 = σ 0 x1 , | x1 |> a
( x − a 2 )1/ 2
2
1
1 a
τ 32 = σ 0 x1 ≈ σ 0 ε −1/ 2 (36')
( x1 − a) ( x1 + a )
1/ 2 1/ 2
2
298
5.2 Elasto-Dynamic Source ---Summary of 5.2.1. and 5.2.2. ---
The derivation is more complex than that for the static case, but is similar. The
displacement field consists of near-field and far-field. The field consists of
"compressional" and "shear", propagating at P and S wave velocities, respectively. The
near-field term yields displacement between P and S waves, as shown in the figure on
page 5.
The derivation is essentially similar to that for the static case. In general, the
displacement field consists of near-field and far-field.
Far field:
299
M ′(τ )
u≈
4πρ v 3 r
Near field
M (τ )
u≈
4πρ v 2 r 2
The most important relation is given by equation 17', which gives the radiation
pattern and the amplitude of far-field P and S waves. The radiation pattern is exactly the
same as that for the static field.
For a shear faulting, S wave is much larger than P wave.
3
| max ur | ⎛ vs ⎞ 1
=⎜ ⎟ ≈
| max uφ | ⎜⎝ v p ⎟⎠ 5
Single Force
The method is essentially the same as that used for static problem.
300
∂ 2u
ρ = ρ f + (λ + 2μ ) graddivu − μ curlcurlu (1)
∂t 2
⎛ a ⎞
ρ f = Faδ (r )h(t ) = − Fh(t )∇ 2 ⎜ ⎟
⎝ 4π r ⎠
(2)
⎡ ⎛ a ⎞ ⎛ a ⎞⎤
= − Fh(t ) ⎢ graddiv ⎜ ⎟ − μ curlcurl ⎜ ⎟⎥
⎣ ⎝ 4π r ⎠ ⎝ 4π r ⎠ ⎦
Then, we obtain,
Fh(t ) ∂ 2 Ap
(λ + 2μ )∇ 2 Ap = a+ρ (4)
4π r ∂t 2
Fh(t ) ∂2 A
μ∇ 2 As = a + ρ 2s (5)
4π r ∂t
Putting
Ap = Ap a and As = As a (6)
1 ∂ Ap
2
Fh(t )
∇ 2 Ap = a+ 2 (7)
4π (λ + 2μ )r v p ∂t 2
Fh(t ) 1 ∂ 2 As
∇ 2 As = a+ 2 (8)
4πμ r vs ∂t 2
where
301
v p = (λ + 2 μ ) / ρ and vs = μ / ρ (9)
1 ∂ 2φ ( x, y, z , t )
∇ φ ( x, y , z , t ) − 2
2
= − g ( x, y , z , t ) (10)
c ∂t 2
is given by,
1 g (ξ ,η , ζ , t ± R / c)
φ ( x, y , z , t ) =
4π ∫
V R
dξ dη dζ (11)
where R 2 = ( x − ξ ) 2 + ( y − η ) 2 + ( z − ζ ) 2
⎡1 ∞
1 ∞
⎤
∫ ∫
F
Ap (r , t ) = ⎢r h(t ± r / v p ± ν )ν dν − h(t ± ν )ν dν ⎥ (12)
4πρ ⎣ 0 r 0 ⎦
302
⎡1 ∞
1 ∞
⎤
∫ ∫
F
As (r , t ) = ⎢r h(t ± r / vs ± ν )ν dν − h(t ± ν )ν dν ⎥ (13)
4πρ ⎣ 0 r 0 ⎦
1 ⎪⎧ ∂ 2 ⎡1 ∞
⎤ 1 h(t ± r / vs ) ⎪⎫
uij = ⎨
4πρ ⎩⎪ ∂xi ∂x j ⎢r
⎣ ∫0
(h(t ± r / v p ± ν ) − h(t ± r / vs + ν ))ν dv ⎥ + 2
⎦ vs r
δ ij ⎬
⎭⎪
(14)
This solution gives both outgoing and converging waves. For outgoing waves, we
take the minus sign, and rewriting it
∂ 2 r −1
1 r / vs
1 1 ⎛ xi x j ⎞
∫
xi x j
ui =
j
ν h(t −ν )dν + h (t − r / v ) + ⎜ δ − ⎟ h(t − r / vs )
4πρ ∂xi ∂x j 4πρ v 2p r r 2 4πρ vs2 r ⎝ r2 ⎠
p ij
r / vp
(15)
1 ∂ 2 r −1 r / vs
1 1 ⎛ xx ⎞
∫
xi x j
uij = νδ (t − τ −ν )dν + δ (t − τ − r / v p ) + 2 ⎜ ij
δ − i 2 j ⎟ δ (t − τ − r / vs )
4πρ ∂xi ∂x j r / vp 4πρ v p r r
2 2
4πρ vs r ⎝ r ⎠
(15')
303
where
r 2 = ( xi − yi ) 2
The 2nd and the 3rd terms decay as 1/r and represent propagating P and S waves,
respectively. These terms are called far-field. The first term decays as 1/ r 2 , and
represents displacement between P and S wave arrivals. This term is called near-field.
Reciprocity
It is evident that
uij (t ;τ ) = u ij (t ;τ ) .
uij (t ;τ ) ≠ uij (τ ; t )
but
304
The displacement field for single couples and double couples can be
obtained by differentiating (15) with respect to appropriate coordinates.
Qualitative Derivation
For a point force F(t) placed at Q, the displacement at P due to the 2nd or 3rd
term is
u ∝ F (τ ) / r
F (τ − a / v) /(r + a)
305
The first term decays as 1/ r 2 ; hence, at large r, we have M ′(τ ) / rv which is called far
field. At short distance, the first term, M (τ ) / r 2 , dominates, which is called near field.
1
Recovering the term , we obtain
4ππ v 2
M (τ ) M ′(τ )
u≈ + (16)
4πρ v r
2 2
4πρ v 3 r
The actual displacement field is more complicated because we need to include the
contribituion of the 1st term in (15') too. However, at far field, the second term in (16)
essentially gives the displacement field which propagates at either P or S wave.
306
The figure below shows the near-field (+far-field) displacement observed for the
1999 Chi-Chi, Taiwan earthquake.
Elasto-dynamic theory
Here, we consider only the far-field term. If we place a double couple with a
moment M (t ) on x1 − x2 plane as shown in the figure,
307
we have at large r (far-field),
2
1 ⎛v ⎞
ui = 2 ⎜ s ⎟ γ iγ 1γ 2 M ′(t − r / v p )
4πμ v p r ⎜⎝ v p ⎟⎠
(17)
1
+ ( −2γ iγ 1γ 2 + γ 2δ i1 + γ 1δ i 2 ) M ′(t − r / vs )
4πμ vs r
where γ i = xi / r .
308
γ 1 = sin θ cos φ , γ 2 = sin θ sin φ , γ 3 = cosθ ,
Then
⎛ ur ⎞ ⎛ sin 2 θ sin 2φ ⎞
⎜ ⎟ 1 ⎜ ⎟
⎜ uθ ⎟ = 4πρ rv 3 M ′(t − r / v p ) ⎜ 0 ⎟
⎜ uφ ⎟ p ⎜ 0 ⎟
⎝ ⎠ ⎝ ⎠ (17')
⎛ 0 ⎞
1 ⎜ ⎟
+ M ′(t − r / vs ) ⎜ 1/ 2sin 2θ sin 2φ ⎟
4πρ rvs
3
⎜ sin θ cos 2φ ⎟
⎝ ⎠
Thr first term on RHS represents P wave and the second term, S wave.
On x1 − x2 plane, θ = π / 2 and
1
ur = M ′(t − r / v p ) sin 2φ
4πρ rv 3p
uθ = 0 (18)
309
1
uφ = M ′(t − r / vs ) cos 2φ
4πρ rvs3
Note that the pattern (radiation pattern) is identical to that for the static case. The
amplitude ratio
3
| max ur | ⎛ vs ⎞ 1
=⎜ ⎟ ≈
| max uφ | ⎜⎝ v p ⎟⎠ 5
Also the time history of the displacement is given by the time derivative of the moment
time function. It is also important that the amplitude decays as 1/ r in contrast to 1/ r 2
for the static field.
The attached four figures show the waveforms computed for the following:
310
Fig. 1 and Fig. 2 show the responses in a whole space (α=6 km/sec, β=3.5 km/sec, ρ=2.6
g/cm3). The "depth" is 10 km. This means that the displacements are computed at a
level 10 km above the source. The mechanism is a N-S striking vertical strike slip with
and the seismic moment is 1027 dyne-cm. The source time function is a triangle with a
rise-time of 0.2 sec and fall-off time of 0.2 sec.
Fig. 1.
Station azimuth=90°. The transverse components are shown at 20 km distance
intervals. P-wave arrivals are aligned.
311
312
Fig. 2.
Station azimuth=45°. The vertical components are shown at 20 km distance
intervals. P-wave arrivals are aligned.
313
314
Fig. 3. and Fig. 4 show the responses in a southern California crustal model given below.
The depth is 11 km. The mechanism is a vertical strike slip and the seismic
moment is 1027 dyne-cm. The source time function is a 1/2 cycle cosine with half-width
of 0.2 sec.
315
Fig. 3. Station azimuth = 90°. The transverse components are shown at 20 km distance
intervals.
316
317
Fig. 4.
Station azimuth =45°. The vertical components are shown at 20 km distance
intervals.
318
319
5.2.3 Elastic Dislocation ---- Summary ----
The equivalence between an elastic dislocation and a double couple shown for the
static problem holds for the dynamic case, too. A point dislocation given by D(t ) over
an infinitesimally small area is equivalent to a double couple with moment
M (t ) = μ D(t ) S
ui ( x , t ) = M kl (ξ , t ) * Gik ,l ( x , ξ , t )
+∞
where * denotes convolution (i.e., h(t ) = f (t ) * g (t ) means h(t ) = ∫ f (t − τ ) g (τ )dτ ).
−∞
In the frequency domain,
uˆi ( x , ω ) = Mˆ kl (ξ , ω )Gˆ ik ,l ( x , ξ , ω )
This is formally equivalent to the expression we obtained for the static case. In practice,
both the time-domain and the frequency-domain formulations are used for inversion. The
time-domain representation is most commonly used in body-wave inversion, and the
frequency-domain representation is more commonly used in inversion studies of normal
modes.
320
5.2.3 Elastic Dislocation
The equivalence between elastic dislocation and a double couple we showed for
the static case holds for the dynamic case. In particular, the displacement field due to a
point elastic dislocation D on S, can be computed with a double couple with a moment of
M 0 = μ DS . For a finite source, we use a distribution of double couples.
+∞
u k ( y, t ) =
∫ dτ
∫ Δu ( x,τ )τ ν dS (1)
k
i ij j
−∞ Σ
where
321
Thus, for a small planar dislocation d Σ at P which is perpendicular to the x2
axis, and on which u1 , has discontinuity, Δu1 (τ )
+∞
u k ( y, t ) =
∫
−∞
Δu1 (τ )τ 12k d Σdτ (2)
where
τ 12k = μ (u1,2
k
(t ;τ ) + u2,1
k
(t ;τ )) = − μ (u1k ,2 (t ;τ ) + u2k ,1 (t ;τ ))
322
It is easy to see that τ 12k / μ is the k th component of displacement at Q at time t due to a
δ function double couple applied at τ on x1 − x2 plane.
Then it is clear from (2) that this dislocation is equivalent to a double couple of
moment μ d ΣΔu1 (τ ) .
We use the same geometry and notation as those used for the static case.
cijpq v p , qj ( x , t ) − ρ vi ( x , t ) = − ρ gi ( x , t ) (4)
Define
v p ( x , t ) = v p ( x , −t ) (6)
gi ( x , t ) = gi ( x , −t ) (7)
then
cijpq v p , qj ( x , t ) − ρ vi ( x , t ) = − ρ gi ( x , t ) (9)
323
Forming (3) x vi - (9) x ui and integrating it with t and V
+∞
∂
(vi ui − ui vi ) = vi ui − ui vi
∂t
+∞ +∞
∂
∫ ∫ −∞
dt
S1 +Σ + +Σ −
dS ⎡⎣ cijpq (vi u p ,q − ui v p ,q )n j ⎤⎦ −
∫
−∞
dt
∂t
(vi ui − ui vi )
(11)
+∞
=
∫ dt ∫ dV ρ (u g −vf )
−∞ V
i i i
+∞
∂
∫
+∞
(vi ui − ui vi ) = (vi ui − ui vi ) −∞
dt
−∞ ∂t
= vi (+∞)ui (+∞) − ui (+∞)v (+∞ ) − vi (−∞ )ui (−∞ ) + ui (−∞ )v (−∞ ) = 0
fi = 0 , and ρ gi ( x , t ) = δ kiδ ( x1 − y1 )δ ( x2 − y2 )δ ( x3 − y3 )δ (t + s )
then,
324
Substituting these into (11),
+∞
u k ( y, s) =
∫ dt ∫ Δu ( x, t )τ ν dS (13)
k
i ij j
−∞ Σ
where τ ijk = cijpq v p ,q ( x , t ) is for ρ gi ( x , t ) given by (12). From the definition, τ ijk is the ij
component of stress at time -t due to a δ function force applied in k direction at Q and at
time -s.
+∞
u k ( y, t ) =
∫ dt ∫ Δu ( x,τ )τ ν dS (14)
k
i ij j
−∞ Σ
+∞
u k ( y, t ) =
∫
−∞
dτ
∫c
Σ
ijpq u kp.q ( x , t − τ )Δui ( x ,τ )ν j dS (15)
325
+∞
u k ( y, t ) =
∫
−∞
dτ cijpq u kp.q ( x , t − τ )Δui ( x ,τ )ν j d Σ (16)
Introducing M pq by
we obtain,
+∞
u k ( y, t ) =
∫
−∞
M pq ( x ,τ )u kp.q ( x , t − τ )dτ (18)
The above integral is convolution of M pq and u kp.q . Hence taking the Fourier transform of
this, we obtain,
326
As discussed earlier, the resulting average dislocation D and Δσ are related by
Δσ s = cμ D / a (1)
For t<0 D =0. At t = 0 when the stress is relaxed the side of the crack starts
moving. After time τ , the medium is in another equilibrium state in which dislocation
327
D = aΔσ s / cμ
is produced.
The behavior between t = 0 and t = τ depends on how stress is relaxed on S. If
stress relaxation takes place smoothly, the change in D may be also relatively smooth.
The time constant τ depends on σ 0 and Δσ . To solve this problem rigorously using
elasto-dynamics is difficult. In most cases no analytic solution is available, and we have
to resort to numerical methods.
First let us consider the simplest case. Consider an infinite homogeneous elastic
medium under uniform shear stress σ 0 . At time t = 0, relax this stress over an infinite
plane S (parallel to the applied shear stress ) instantaneously. In order to see the
displacement for t ≥ 0 this problem can be replaced by the following problem.
328
The disturbance applied to the surface propagates into the medium with shear
velocity β . At time t, it propagates as far as β t . Thus, the instantaneous strain is
ε = u (t ) / β t
σ 0 = με = μ u (t ) / β t
∴
u (t ) = (σ 0 / μ ) β t (2)
This gives the displacement for infinite instantaneous crack. From (2)
u (t ) = (σ 0 / μ ) β = constant (3)
329
∂ 2u 1 ∂ 2u
= (4)
∂x 2 β 2 ∂t 2
∂u
μ +σ0 = 0 , t>0
∂x
∴ μ F ′( z ) = −σ 0 , F ( z ) = −σ 0 z / μ = −(σ 0 / μ )( x − β t )
∴ u = (σ 0 / μ )( β t − x) , t > x/β
330
At t = 0, uniform shear stress σ 0 is applied in x3 direction instantaneously over a surface
( x2 = 0 , | x1 |≤ a ). Let us consider the displacement at point O, the middle point of the
crack. For small t (before the effect of the edge reaches this point), the displacement of
this point should be the same as for infinite crack. Thus, the initial velocity should be
given by σ 0 β / μ . The effect of the edge arrives at this point at time t = a / β , and
eventually the motion stops when a final equilibrium state is achieved. This is given by
the solution of the static problem, i.e., u = σ 0 a / μ . Hence, the displacement time
function may be schematically given by the following figure.
331
One useful functional form which approximates this function is
This problem can be solved numerically. The equation of motion in the medium
is
1 ∂ 2u ∂ 2u
= (6)
β 2 ∂t 2 ∂x12
∂u
μ +σ0 = 0 ( x2 = 0 , | x1 |≤ a , t>0) (7)
∂x2
The numerical solution is shown in the figure below and the displacement at the middle
point is compared with the approximate solution given above. Note that the numerical
solution shows a slight overshoot.
(Burridge, R., Phil. Trans. Roy. Soc. London, 265, 353-381, 1969)
332
4. Finite Propagating Crack
Let us consider the same geometry as before. Instead of applying the stress
instantaneously, we apply a propagating stress,
σ = σ 0 H (t ± x1 / V ) , | x1 |≤ a (8)
where H(t) is the Heaviside step function and V is the rupture velocity. V is usually
slightly smaller than β . In this case, the final value of u is the same as before.
However, the point O starts feeling the edge effect even in the beginning, because at t =
0, the two edges are very close to this point. Thus, the motion is decelerated compared
with the previous example. The effect of the edge | x1 |= a will reach the point O at time
a (1/ β + 1/ V ) ≈ 2a / β after the stress application. Thus, the average velocity would be
σ 0a / μ
u≈ ≈ σ 0 β / 2μ (9)
2a / β
The schematic displacement time function may look like the one shown in the figure.
333
A useful functional form would be
334
The above description, however, is very simplified. In a real propagating crack,
the initial velocity could be faster than σ 0 β / 2 μ due to the stress concentration near the
crack tip. Equation (10) should be considered only approximate.
u ≈ σ 0β / μ
u ≈ 1 m/sec
335
6. Frictional Stress Release
In actual earthquake faulting, the stress on the fault plane is released against
frictional stress σ f opposing the motion.
336
It is also possible that, at a certain time, say τ 0 , the fault motion slows down due
to some locking mechanism. In this case, the time history of the stress drop on the fault
surface may be schematically given by the curve shown in the figure above.
Actual stress release pattern can be very complex as shown by figure (b) below.
(a) (b)
Since the details of the stress release are presently unknown, it is not meaningful
to consider overly detailed models.
If the stress release is simple as shown on the left, then the fault motion
337
is driven by σ 0 e = σ 0 − σ f ( σ f = σ 1 ) which is called the effective (tectonic) stress.
The particle velocity of the side of the crack (fault) will be given by σ 0 e β / μ or
σ 0 e β / 2μ depending upon whether the stress application is instantaneous or not
(or in general cσ 0 e β / μ , c=constant).
Thus, if we can determine the approximate time function of the source dislocation
function, we can determine the effective tectonic stress σ 0 e from the initial slope and the
stress drop Δσ s from the final dislocation (static field).
Referring to the figure shown above, the static stress drop of an earthquake, Δσ s ,
is defined by
Δσ s = σ 0 − σ 1 (11)
where σ 0 and σ 1 are the initial and the final stresses, respectively. This definition is
straightforward and unambiguous.
The dynamic stress drop, Δσ d , is the stress that drives fault motion.
Unfortunately, the definition of dynamic stress drop is not universal, and is
ambiguous. For the simplest stress release pattern shown by figure (a) above, it is the
same as the effective tectonic stress and is equal to the static stress drop, i e.,
Δσ d = σ 0 e (= σ 0 − σ f ) = Δσ s (= σ 0 − σ 1 ) (12)
338
However, for more complex stress release patterns in which friction changes as a
function of slip as shown by figure (b) above, we can define Δσ d by
Δσ d = σ 0 − σ f (13)
where
1 D
D ∫0
σf = σ f ds (14)
(s: slip).
Estimate the seismic moment M 0 , and the approximate duration of the source (as
viewed from ESK). It is also possible to constrain the depth, H, but, for simplicity, we
fix H, except in the last step where the effect of H on the waveform will be examined.
339
India 1/26/2001 Station ESK Displacement
Displacement, cm
Time, sec
Fig. 1
T (t ; τ 0 , tc ) (1)
where τ 0 and tc are the two time constants of the source. The area under T (t ; τ 0 , tc ) is
unity. Then, the P-wave form in a whole space is given by
340
M 0 Rθφ
u1 (t ) = T (t − r / α ; τ 0 , tc ) (2)
4πρ rα 3
In the real earth, the structure near the source and the receiver complicates the
wave form. Here, we use a simple half space. Then, we need to include the near-source
reflections pP ans sP, and the effect of the free surface near the station. Here, the near-
source effect is approximated by
The first, second, and third terms represent, the direct P, pP, and sP, respectively. a1 , a2 ,
t1 , and t2 depend on the radiation pattern, the depth, and the structure. In particular,
changing the depth has a large effect on the waveform.
The receiver effect is simply given by a scalar factor, CR . (For SH wave, CR =2.)
The geometrical spreading factor is also simplified by a scalar factor g, and 1/r in
(2) is replaced by g/ RE where RE is the radius of Earth.
rg
u2 (t ) = CR u1 (t ) * CS (t ) (4)
RE
Even for this simplified problem, computation of (4) is not that simple because of
the complexity of CS (t ) = 1 + a1δ (t − t1 ) + a2δ (t − t2 ) . A simple program pwsyn2g_asc.f is
provided in /home/ftp/pub/hiroo/ge162.dir/practice_6.dir. The output of this program is
u2 (t ) .
Run this program for various τ 0 , tc , and the depth H.. The constants, a1 , a2 , t1 ,
and t2 are computed in the program from the fault parameters and the depth. This
program computes the displacement in cm for a unit moment ( M 0 =1020 N-m).
341
(Another program pwsyn2g.f which does the same and outputs the results in SAC format
is also provided.)
3) Observed Waveform
To simulate the observed waveform at the station, we need to add the effects of
attenuation. The attenuation function, F(t; t*), for t * = 1sec is shown in Figures 2, and the
file is in practice_6.dir (o_futtm.asc (ASCII file) and o_futtm (SAC file)).
Fig. 2
342
u3 (t ) = u2 (t ) ∗ F (t ) (5)
This can be directly compared with the observed record, U (t ) , shown in Figure 1.
4) Determination of M 0
Once the waveforms are matched satisfactorily, the amplitude ratio of the
observed waveform
amplitude(U (t ))
amplitude(u3 (t ))
gives the seismic moment, M 0 , of the earthquake in the unit of 1020 N-m.
Since many approximations have been made, the waveforms cannot be matched
completely, but try to match the overall waveform and the amplitude of the first 1 cycle.
Here, we use only one station, and the mechanism (i.e., radiation pattern) is
assumed. However, if we have more than one station, we can determine the mechanism
by matching the amplitudes at different stations with the same M 0 .
In general, matching the displacement record is much easier than matching the
velocity record, V (t ) , shown in Figure 3, because high-frequency components which are
harder to model have been removed in the displacement record.
343
India 1/26/2001 Station ESK Velocity
Velocity, cm/sec
Time, sec
Fig. 3
The relevant parameters needed for the computation are in the following.
Eskdalemuir
344
Source Crustal Structure (Half Space)
For the mechanism and the takeoff angle given above, the radiation pattern factor for P-
wave at the source
g =0.35
1. Compute u2 (t ) for
Observe, how P, pP and sP phases interact. Note that the far-field waveform in a
whole space discussed in class should be a one-sided pulse, while the observed U (t ) is
two sided. The interaction between these phases is the primary reason why the observed
displacement is two-sided.
345
3. Compare u3 (t ) with U (t ) to determine M 0 . The starting time is arbitrary so that you
can shift the waveform arbitrarily in comparison.
4. Try 1, 2, and 3, for different combinations of τ 0 , and tc (fix H) and find the best
solution.
5. Compute u3 (t ) using a source time function shown below, and compare it with U (t ) .
Vary H from 5 to 35 km, and see the difference in the synthetic waveforms. (This source
time function is not a trapezoid, which means that the source model is different from the
Haskell model.)
6. (Optional)
Compute the velocity v3 (ti ) = [u3 (ti +1 ) − u3 (ti ) ] / Δt for the best model, and
compare it with V(t).
pwsyn2g.f
pwsyn2g.f takes input file i_pwsyn2g. Only the parameters in bald face need to be
changed. (do not change c_pwsyn2g.)
i_pwsyn2g
346
line 1 station id
line 5 flag for the choice of source time function (do not change)
etc
Example
ESK
2.67 6.0 0.05 0.01 0.01 0.01
66. 64. 60.
18. 321. 10. 0.35 1.0
1
6
0.0 0.0
2.0 0.0
10.0 1.0
14.0 1.0
22.0 0.0
100.0 0.0
347
The output file o_pwsyn2g contains the parameters used and other computed parameters.
convg3m.f
i_convg3m
348
line 1 file name of x(t )
line 2 file name of y (t )
line 3 file name of z (t )
Example
i_convg3m
o_Ptotalamp.asc
o_futtm.asc
o_Ptotalamp_q.asc
* convolution of xi (i = 1, 2, , l ) and y j ( j = 1, 2, , m) is
⎛ l ⎞
zk = ⎜ ∑ xi yk −i ⎟ Δt k = 1, 2, , l + m
⎝ i =1 ⎠
(if k − i ≤ 0 , then yk −i =0)
Ge 162
349
6. Retrieval of Seismic Source Parameters
1. Introduction
Far-field body waves are widely used to determine the source mechanism (fault
plane solution), seismic moment, and rupture patterns (complexity). The method is
conceptually simple, but the actual procedure involves many steps. These steps include
the calculation of the following:
M 0 Rg (Δ, h)
u ( Δ, t ) = T (t ;τ 0 , tc ) * CS (t ) * CR (t ) * q(t ) * I (t )
4πρ vc3 RE
The radiation pattern from a point double couple has been already given with
respect to the coordinates fixed to the double couple.
350
For actual applications it is convenient to express the radiation pattern by using a
geographical coordinate system. For a fault model shown in the figure, the displacement
at point P can be given by:
1 (1)
ur ( r , t ) = R P M (t − r / α )
4πρ rα 3
1 (2)
uθ (r , t ) = Rθ M (t − r / β )
4πρ r β 3
351
1 (3)
uφ (r , t ) = Rφ M (t − r / β )
4πρ r β 3
where
R P = sR (3cos 2 ih − 1) − qR sin 2ih − pR sin 2 ih (4)
3 1
Rθ = sR sin 2ih + qR cos 2ih + pR sin 2ih (5)
2 2
φ
R = − qL cos ih − pL sin ih (6)
qL = −(cos λ cos δ ) sin φ + (sin λ cos 2δ ) cos φ (7)
pL = (sin λ sin δ cos δ ) sin 2φ + (cos λ sin δ ) cos 2φ (8)
sR = sin λ sin δ cos δ (9)
qR = sin λ cos 2δ sin φ + cos λ cos δ cos φ (10)
pR = cos λ sin δ sin 2φ − sin λ sin δ cos δ cos 2φ (11)
If the azimuth of the fault strike ( x1 axis) φ f and the azimuth of the station (point
P) φs are measured clockwise from N, then φ (measured counter-clockwise from ξ1
axis) in the above formula should be given by
φ = φ f − φs (12)
By using (1) to (12) we can compute the far-field displacements of body waves
from a fault of arbitrary geometry. The station coordinates are given by (r , ih , φ ) rather
than the conventional (r , θ , φ ) .
352
3. Source Finiteness-- Haskell Model
One of the useful kinematic source models is the Haskell(1964) model. In this
model, a seismic source is represented by a rectangular fault with length L and width W.
The dislocation (fault offset), D, on the fault plane is uniform in space, and its temporal
variation is given by a linear ramp function with a rise time τ 0 .
The rupture propagation is assumed instantaneous in the width direction.
Lengthwise, the rupture propagates from one end of the fault to the other with a
uniform rupture speed V. In this sense, the rupture propagation is one-dimensional
unilateral.
Assume that the rupture propagates from A to B over a length L. If the seismic
moment per unit fault length is m, the moment per line element dl is mdl.
The far-field time function due to this line element is given by a box-car function of
width τ 0 and height mdl / τ 0 .
353
If AP is very large compared with AB, the displacement at P (station) can be given by
R ⎛ r l l cos Θ ⎞
mdls ⎜ t − − + ⎟
4πρ rvc
3
⎝ vc V vc ⎠
where r = AP, vc is either P or S wave velocity, R, the radiation pattern and V is the
rupture velocity. s(t) is defined above, and is called the local dislocation rate function.
Therefore, for a source propagating from A to B, we have
⎛ r l l cos Θ ⎞
∫
L
R
u (r , t ) = m s⎜t − − + ⎟ dl (13)
4πρ rvc3 0 ⎝ vc V vc ⎠
we obtain,
∫
tc
R M0 (15)
u ( r ,τ ) = s (τ − t1 ) dt1
4πρ rvc3 tc 0
354
we obtain
∫
tc
RM 0
u ( r ,τ ) = r (t1 ) s (τ − t1 ) dt1
4πρ rvc3 0
+∞
(16)
∫
RM 0
= r (t1 ) s (τ − t1 ) dt1
4πρ rvc3 −∞
which is convolution of r(t) and s(t). r(t) is called the rupture function.
if t1 > t2
Note that
+∞ t1 + t2
∫−∞
T (t ; t1 , t2 )dt =
∫ 0
T (t ; t1 , t2 )dt = 1 (17)
355
Thus, the pulse width at the station in the azimuth Θ is
L L
w(Θ) = τ 0 + − cos Θ (18)
V vc
sin(ωτ 0 / 2) sin(ω tc / 2)
Tˆ (ω ; τ 0 , tc ) = (19)
(ωτ 0 / 2) (ω tc / 2)
Note:
356
In applying the above results to body-wave radiation in the real earth, another
factor needs to be considered. For example, if AB is horizontal, we let the take-off angle
of the ray be ih , and the azimuth of the station from AB measured on the earth’s surface
be Θ . Then the phase velocity along the Earth’s surface is vc / sin ih . Then we replace
r / vc and vc in the argument of s(t) by T and vc / sin ih respectively, where T is the travel
time. Hence we have, instead of (14) and (15),
∫
tc
RM 0
u ( r ,τ ) = s (τ − t1 ) dt1 (20)
4πρ rvc3tc 0
where,
For example, directivity observed for the 1999 Landers earthquake is illustrated in
the following figure.
357
358
Directivity has strong influence on the distribution of strong ground motion as
shown for the 1994 Northridge earthquake.
359
360
4. Response Function at the Source and Receiver
If the source is shallow, the reflections from the free surface (and other structures)
complicates the source wave as shown by the following figure.
361
A similar situation occurs near the station too. These effects are usually included
as the source and receiver response functions.
In an elastic medium with density ρ , the elastic energy per unit volume is given
by
ε = ρu 2 (22)
φE = ρ u 2v (23)
Then elastic wave energy emitted per unit time from the source within a ring
defined by ih and ih + dih is
M
u= RP
4πρ rvc
3
362
P
Here the subscript h signifies the values of the variables at the source, and R is the
radiation pattern.
From the figure above, the energy arriving at station P (per unit time, per unit area
of wave front) is
ρ0u02 v0 (25)
Since the total energy emitted in a ring defined by ih and ih + dih arrives in a ring
defined by Δ and Δ + d Δ , we obtain from (23), (24), and (25),
6. Effect of Anelasticity
363
Usually, the anelasticity of the medium is given in terms of quality factor Q. If
we write the quality factor of P and S waves by Qα and Qβ , the amplitude attenuation of
P and S waves of angular frequency ω during the propagation can be given by
⎛ ω ds ⎞ ⎛ ω ds ⎞
exp ⎜ − ∫ ⎟ exp ⎜− ∫ ⎟
⎝ 2 Qα α ⎠ , and ⎝ 2 Qβ β ⎠ (29)
where the integral is taken along the ray path. Since Qα and Qβ are functions of depth
and therefore of the path length, s, we may write these
⎛ ω ⎞ ⎛ ω ⎞
exp ⎜ − tα* ⎟ , and exp ⎜ − t β* ⎟ (30)
⎝ 2 ⎠ ⎝ 2 ⎠
where
∫ ∫
ds ds
tα* = , and t β* = (31)
Qαα Qβ β
Empirically, tα* and t β* do not vary significantly with the distance, and tα* = 0.5 to 1 sec
and t β* = 3 to 4 sec for distance ranges Δ > 35 degrees.
These functions are shown below for tα* =1 sec and t β* =4 sec.
364
7. Instrument Response
Since seismographs have their own frequency response, they modify the
waveforms. In the following, the impulse response of the VBB and WWSSN-LP
instruments are shown.
365
366
367
8. Simplified Expressions for a Point Source
The following simplified expressions are not exact, but are useful for understanding
the basic principles for retrieval of source parameters.
In the following, we use the vertical component of P wave for illustration purposes.
We assume that the observed seismogram has been corrected for the instrument, I(t),
attenuation, q(t), and the response of the source strucuture, CS (t ) , and the response of the
receiver structure is given by a scalar CR . Then, from the first equation in this section,
we can write the vertical component of P wave, u p (t ) , as
M 0 (t ) R g
u p (t ) = CR (33)
4πρ hα h3 RE
From which,
4πρ hα h3 RE
M 0 (t ) = u p (t ) (34)
RgCR
Integrating,
4πρ hα h3 RE τp
M0 =
RgCR ∫
0
u p (t )dt (35)
368
where τ p is the duration of the P wave record.
2
τp τp ⎡ R ⎤
Eα =
∫ dS ∫ ρ hα hU (t )dt =
∫ dS ∫ ρ hα h ⎢
2
M ( t ) ⎥ dt
⎣ 4πρ hα h r
p 3 0
S 0 S 0 ⎦ (36)
⎡ 1 ⎤ τp
= ρ hα h ⎢
∫ R dS ∫
2
3 ⎥
M 02 (t )dt
⎣ 4πρ hα h r ⎦ S 0
Substituting (34),
2
16π ⎛R ⎞ 1 1 τp
Eα =
15
ρ hα h ⎜ E ⎟ 2 2
⎝ g ⎠ R CR ∫0
u 2p (t )dt (37)
where
369
1 4
4π r 2 ∫
S
R 2 dS ≡ R 2 =
15
(38)
Using (37), we can estimate Eα from the observed record at a station. For a point
source, we can show that
5
3⎛α ⎞
Eβ = ⎜ ⎟ Eα ≈ 23Eα (39)
2⎝ β ⎠
and
ER = Eα + Eβ = 24 Eα (40)
370
Fig. 1
1. Radiation pattern, R
2. Response of source and receiver structures, CS (t ) and CR (t )
3. Geometrical spreading, g (Δ, h)
4. Attenuation (Q operator), q (t )
5. Instrument response, I (t )
Fortunately, Hartzell and Heaton (1983) have removed all of these effects, and
obtained a trapezoidal displacement at point P, as shown in Figure 2.
371
Fig. 2
Fig. 3
372
+∞
∫
RM 0 RM 0
u ( r ,τ ) = r (t1 ) s (τ − t1 ) dt1 = T (τ ;τ 0 , tc ) (1)
4πρ rvc3 −∞ 4πρ rvc3
L L cos Θ
tc = −
V vc
where Θ is the angle between the rupture direction and the ray.
1. Check the dimension of the right-hand-side of equation (1) to make sure that it is the
dimension of "length".
2. We assume that the medium is homogeneous with α = 6 km/sec, β =3.5 km/sec and
ρ =2.6 g/cm3, and the radiation pattern R is equal to 1. We also assume that the fault
plane AB and P are on the same plane. The take-off angle of the ray to P is 30 degrees
(note that this is different from Θ ), and the distance is r=1000 km. The rupture
propagation is assumed to be one dimensional unilateral (A to B) with a rupture speed
V=2.5 km/sec.
373
(Note: The solution obtained by Hartzell and Heaton is slightly different from the above.)
∂ 2u ∂ 2u
ρ = E +ρf (1)
∂t 2 ∂x 2
d 2U
− ρω 2U = E +F (2)
dx 2
Let yi , i = 1,2... be the eigen functions of the homogeneous equation of (2) i.e.,
2
d yi
− ρω i2 yi = E
dx 2 (3)
374
L
∫ ρy y dx = c δ
0 n l n,l
(4)
∞
U = ∑ ai yi
i =1 (5)
where
1 L
c ∫0
ai = ρ yiUdx
(6)
−c ω an = E ∑ al ∫0 yl′′y n dx + ∫0Fy n dx
L L
2
1 L
∴ an = − 2 ∫0Fy n dx
c(ω − ω n )
2
(7)
∞ ⎡ ⎤ L
yn
∴ U = ∑ ⎢− 2 ⎥ ∫0F yn dx
n =1 ⎣ c(ω − ω n ) ⎦
2
(8)
∞ ⎡ ⎤ L
yn
∴ u = eiωt ∑ ⎢ − 2 ⎥ ∫0 Fyn dx
n =1 ⎣ c(ω − ω n ) ⎦
2
(9)
375
1 +∞ 1 iωt
2π ∫−∞ iω
H(t) = e dω
(10)
∞
⎡ y L ⎤ 1 +∞ ⎡ e iω t ⎤
∴ u(t) = ∑ ⎢ − n ∫0Fy n dx ⎥
⎦ 2 π ∫− ∞ 2 ⎥dω
⎢ −
n =1
⎣ c ⎣ i ω (ω − ω n ) ⎦
2
(11)
Since any physical system must have at least small attenuation, these poles are slightly
above the real axis (see Morse and Feshbach, p. 1334).
Carrying out the integration along the path given above, we have, for t > 0,
∞
⎤ (1 − cos ω n t )
∑ ∫ ⎡ L
u (t ) = ⎢ yn Fyn dx ⎥ (12)
n =1 ⎣ 0 ⎦ cnω n2
376
∞
∑ ∫ ⎡ ⎤ 1
L
u (t ) = ⎢ yn Fyn dx ⎥ (13)
⎦ cnω n
2
n =1 ⎣ 0
2. Torsional Oscillation
For torsional oscillations of a sphere, the eigen functions and the force are
vectors.
(15)
⎛ ⎞
⎜ 0 ⎟
⎜ ⎟ ⎛ Fr ⎞
⎜ 1 ∂Yl m ⎟ ⎜ ⎟
n yl = y1, n ⎜
m
⎟ , F = ⎜ Fθ ⎟
⎜ sin θ ∂ φ ⎟ ⎜F ⎟
⎜ ⎝ φ⎠
∂Yl m ⎟
⎜ − ⎟
⎝ ∂θ ⎠
The expression for step function excitation can be derived in a manner similar to the one-
dimensional case.
⎤ (1 − cos n ωlmt )
u (r , t ) = ∑ ⎡ m
⎢ n yl
l ,m,n ⎣
∫V
F ⋅ n ylm dV ⎥
⎦ n cl n ωl
m m2
(16)
where
c =
∫ρ ylm ⋅ n ylm dV
m
n l n
V
377
3. Spheroidal Oscillation
⎛ ⎞
⎜ 0 ⎟
⎛ Yl m ⎞ ⎜ ⎟
⎜ ⎟ ⎜ ∂ Y m
⎟
n yl = y1, n ⎜ 0 ⎟ + y3, n ⎜ (17)
m l
∂θ ⎟
⎜ 0 ⎟ ⎜ ⎟
⎝ ⎠ ⎜ 1 ∂Yl ⎟
m
⎜ sin θ ∂φ ⎟
⎝ ⎠
Then
⎤ (1 − cos n ωlmt )
u (r , t ) = ∑ ⎡ m
⎢ n yl
l ,m,n ⎣
∫ V
F ⋅ n ylm dV ⎥
⎦ n cl n ωl
m m2
(18)
Again
c =
∫ρ ylm ⋅ n ylm dV (19)
m
n l n
V
1 − cos n ωlmt
u (r , t ) = ∑
l ,m,n
⎡⎣( M : ε ) n ylm (r , t ) ⎤⎦
n cl n ω l
m m2
(20)
where ε is the strain tensor computed at the source for each l, m, and n.
378
379
2. 2001 Bhuj India earthquake
380
381
3. 1999 Russia-China Border earthquake
382
383
7. Summary of Seismic Source Parameters
1. Fault Geometry
2. Fault Dimension
384
D = Δu
M 0 = μ SD
6. Stresses
Initial stress* σ0
Final stress* σ1
Static stress drop Δσ s = (σ 0 − σ 1 ) ∝ D
Frictional stress* σ f
Effective tectonic stress (Dynamic stress drop) Δσ d = σ 0 − σ f ∝ D
Average stress* σ = (σ 0 + σ 1 ) / 2
7. Energy
385
Potential Energy Change*
(Strain Energy + Gravitational Energy) ΔW = σD S
Frictional Energy Loss* EF = σ f DS
Fracture Energy* EG
386
387
388
389
9. Complexity
390
391
10. Energy (or Moment) Magnitude
---------------------------------------------------------------------------------------------------------------------------------
From the energy budget, the radiated energy, ER , is given by
ER = ΔW − EF = DS (σ − σ f )
As mentioned above, the initial stress σ 0 , the final stress σ 1 , and the frictional stress σ f
cannot be determined directly with seismological methods, and ER above cannot be
determined. However, note that
σ 0 + σ1 σ 0 − σ1 Δσ s
ER = DS ( − σ f ) = DS ( + (σ 1 − σ f )) = DS ( + (σ 1 − σ f ))
2 2 2
392
If we assume that σ 1 is equal to σ f (i.e., the stress on the fault plane is equal to the
average frictional stress when fault motion stops), then,
1 Δσ s Δσ s
ER = DS Δσ s = DS μ = M0
2 2μ 2μ
ER = M0 /(2x104)
393
394
Ge 162
8. Physics of Earthquakes
For understanding the overall physics of earthquakes without going into details, it
is useful to investigate scaling relations between several macroscopic source parameters.
395
[closed circles: interplate, open circles: intraplate, Kanamori, H., and D. L.
Anderson, Theoretical basis of some empirical relations in seismology, Bull. Seis. Soc.
Amer., 65 (5), 1073-1095, 1975]
By definition,
M 0 ∝ Δσ s a 3 circular
Δσ s w L rectangular
2
396
3 ⎛ 16 ⎞
log M 0 = log S + log ⎜ π −3/ 2 Δσ s ⎟
2 ⎝7 ⎠
3
If Δ σ s is constant, then, log M 0 ≈ log S . From the attached figure, we see,
2
The scaling relation shown above has been extended to small earthquakes as
shown below.
397
[Modified from Abercrombie, R., and P. Leary, Source parameters of small earthquakes
recorded at 2.5 km depth, Cajon Pass, Southern California: Implications for earthquake
scaling, Geophys. Res. Lett., 20, 1511-1514, 1993]
The L ∝ M 01/ 3 scaling is seen, and this is generally interpreted as evidence for
constant stress drop, but the fairly large scatter in Δσ s should be noted.
1
ΔW = (σ 0 + σ 1 ) DS (1)
2
ΔW = ER + EF + EG (2)
where ER and EG are the radiated energy and the fracture energy, respectively, and EF
398
is the frictional energy. In this model, separation of EG and EF is somewhat arbitrary
(both represent non-radiated energy, i.e., dissipation), but, EG , and EF are commonly
defined by the hatched and cross-hatched areas shown in the figure below, respectively,
i.e.,
EF = σ f0 DS (3)
However, here the fracture energy, EG , is the energy dissipated during rupture
over a volume surrounding the fault zone.
(In this diagram, the energies are interpreted as those per unit area.)
The radiated energy, ER , is what we can measure from the radiated seismic
waves, as shown in 6.1, but because of the practical difficulty in measuring it accurately,
ER has not been fully used in seismology for the purpose of understanding the physics of
earthquakes.
Only recently, it became possible to measure ER accurately enough so that we
can investigate the physics of earthquakes in terms of energy budget.
399
The ratio of the fracture energy, EG , to the radiated energy, ER , determines the
characteristics of fracture, or an earthquake. Alternatively, the radiation efficiency
defined by
ηR = ER /( ER + EG ) (4)
can be used for the same purpose. Referring to the figure above,
.
8.2.3 Observations
General Observation
The results for large earthquakes are summarized in the following figure.
400
Radiation efficiency η R = ER /( ER + EG ) as a function of M w . The different symbols
show different types of earthquakes as described in the legend. Most earthquakes have
radiation efficiencies greater than 0.25, but tsunami earthquakes and two of the deep
earthquakes (the Bolivia earthquake and the Russia-China earthquake) have small
radiation efficiencies. (Venkataraman and Kanamori, 2004)
Except for the very large deep focus earthquake, the 1994 Bolivian earthquake,
and tsunami earthquakes, the radiation efficiency, η R , is larger than 0.25, which means
that the fracture energy for most large earthquakes, regardless of their tectonic
environment, are comparable or less than ER .
Deep Focus Earthquake
401
It is difficult to accurately determine the size of the fault plane, S, for deep focus
earthquakes. However, for the 1994 Bolivian earthquake (Mw=8.3, depth=635 km), the
largest deep-focus earthquake ever recorded, the source parameters could be determined
well enough to investigate the energy budget.
As we discussed in 5.1.5, we cannot determine ΔW itself, but we can estimate its
lower bound,
ΔW0 = Δσ s DS / 2 (6)
from Δσ s , D , and S .
The result for the 1994 Deep Bolivian earthquake (Mw=8.3) showed that
ΔW0=1.4x1018 J and ER =5x1016 J, which is only 3 % of ΔW0 ; the difference ΔW0 –
ER=1.35x1018 J, was not radiated, and must have been deposited near the focal region,
probably in the form of thermal energy. This energy 1.35x1018 J is comparable to the
total thermal energy released during large volcanic eruptions such as the 1980 Mount
Saint Helens eruption. The thermal energy must have been released in a relatively small
focal region, about 50x50 km2, within a matter of about 1 min. The mechanical part of the
process, i.e. the earthquake observed as seismic waves, is only a small part of the whole
process. Thus, the Bolivia earthquake should be more appropriately viewed as a thermal
process rather than a mechanical process.
With this much of non-radiated energy, the temperature in the focal region must
have risen significantly. The actual temperature rise, ΔT, depends on the thickness of the
fault zone, which is not known, but if it is of the order of a few cm, the temperature could
have risen to above 10,000 °C.
Shallow Earthquakes
Although the situation for shallow earthquakes may be different from that for
deep focus earthquakes, a simple calculation shows that if σf is comparable to Δσs, about
10 MPa, the effect of shear heating is significant. If the thermal energy is contained
402
within a few cm around the slip plane during seismic slip, the temperature can easily rise
to 100 to 1000 °C.
We consider a gross thermal budget during faulting under a frictional stress σf.
Let S and D be the fault area and the displacement offset respectively. Then the total
heat generated during faulting is Q=σf DS. If we assume that the heat is distributed
during seismic faulting within a layer of thickness w around the rupture plane, the
average temperature rise ΔT is given by
The figure below shows ΔT as a function of magnitude. If a fault zone is dry (no
fluid), melting may occur and friction may drop. If fluids exist in a fault zone, fluid
pressurization could occur.
403
404
The key question is how thick the fault slip zone is. Geologists have examined
many old fault zones which were formed at depths and were brought to the surface by
long-term uplift (i.e., exhumed faults). Some fault zones have a very narrow (about 1
mm) distinct slip zone where fault slips seem to have occurred repeatedly. The
Punchbowl fault, California, implies that earthquake ruptures were not only confined to
the ultracataclasite layer, but also largely localized to a thin prominent fracture surface.
They suggest that mechanisms that are consistent with extreme localization of slip, such
as thermal pressurization of pore fluids, are most compatible with their observations. In
other cases, several narrow slip zones were found but evidence shows that each slip zone
represents a distinct slip event (i.e., an earthquake). Thus, geological evidence suggests a
narrow slip zone, at least for some faults, but this question will remain debatable.
405
(Richard Sibson, Written communication, 2000)
406
If a fault zone is narrow and rough, and if the material in the fault zone behaves as
viscous fluid, it is also possible that elastohydrodynamic lubrication plays an important
role in reducing friction for large events. An interesting consequence of this is that as the
slip and slip velocity increase, the hydrodynamic pressure within a narrow zone becomes
large enough to widen the gap thereby suppressing high-frequency ground motion caused
by the fault asperities rubbing against each other. During the recent Chi-Chi, Taiwan,
earthquake, the observed ground-motion near the northern end of the fault was extremely
large (> 2.5 m/s, the largest ever recorded), but short period acceleration was not
particularly strong so that the damage to ordinary structures by shaking was minor. This
could be a manifestation of the high-speed lubrication effects. However, since this is the
only earthquake for which such large slip and slip velocity were instrumentally observed,
whether this is indeed a general behavior or not is yet to be seen.
State of Stress
The results obtained for large earthquakes suggest that the average stress level
along mature faults where large earthquakes occur must be low because of the dominant
thermal effects such as frictional melting and fluid pressurization, or of
elastohydrodynamic lubrication. Because of melting or pressurization, a fault zone is
self-organized into a low stress state. That is, even if the stress was high in the early
stage of fault evolution, it would eventually settle in a low stress state after many large
earthquakes. This state of stress is consistent with the generally held view that the
absence of heat flow anomaly along the San Andreas fault suggests a shear strength of
about 200 bars or less . The stress in the crust away from active mature faults can be high
as has been shown by many in-situ measurements of stress. The stress difference is
large, and a kbar type stress may be involved in small earthquakes, but the events are in
general so small that it is hard to determine the stress parameters accurately. What is
important, though, is that as long as the length of the fault is small, the state of stress in
407
the fault zone would not affect the regional stress drastically. However, as the fault
grows to some length (e.g. Japanese intra-plate earthquakes like Tango, Tottori, Nobi
etc), then some sort of self-organization occurs and the fault settles at a stress level
somewhat higher than that on more active plate boundaries.
408
409
In general, this distribution is expressed as
The results obtained for many regions indicate that the value of b (called b value)
is approximately equal to 1. This relation is called the Gutenberg-Richter relation,
Ishimoto-Iida relation, or simply the magnitude-frequency relation. Since
M ∝ log ER /1.5
410
equation (1) means
that is, the relation between N and ER is given by a power law. If b=1 and
ER ∝ r 3 (r: size)
Then,
411
References
Aki, K., Characterization of barriers on an earthquake fault, J. Geophy. Res., 84, 6140-
6148, 1979.
Ben-Zion, Y., and D. J. Andrews, Properties and implications of dynamic rupture along a
material interface, Bull. Seismol. Soc. Am., 88, 1085-1094, 1998.
Brace, W. F., and J. D. Byerlee, Stick slip as a mechanism for earthquakes, Science, 153,
990-992, 1966.
412
Dmowska, R., and J. R. Rice, Fracture theory and Its seismological applications, in
Theories in Solid Earth Physics, edited by R. Teisseyre, pp. 187-255, PWN-Polish
Publishers, Warzawa, 1986.
Eshelby, J. D., The determination of the elastic field of an ellipsoidal inclusion and
related problems, Proceedings of the Royal Soc. London, 241, 376-396, 1957.
Heaton, T., Evidence for and implications of self-healing pulses of slip in earthquake
rupture, Physicsof the Earth and Planetary Interiors, 64, 1-20, 1990.
Ida, Y., Cohesive force across the tip of a longitudinal-shear crack and Griggith's specific
surface energy, J. Geophys. Res., 77, 3796-3805, 1972.
Ide, S., and M. Takeo, Determination of constitutive relations of fault slip based on
seismic wave analysis, J. Geophys. Res., 102, 27,379-27,391, 1997.
Kanamori, H., Mechanics of Earthquakes, Ann. Rev. Earth & Planetary Sciences, 22,
207-237, 1994.
Kostrov, B. V., Seismic moment and energy of earthquakes, and seismic flow of rock
(translated to English), Izv. Earth Physics, 1, 23-40, 1974.
Lachenbruch, A. H., Frictional heating, fluid pressure, and the resistance to fault motion,
J. Geophys. Res., 85, 6097-6112, 1980.
Lachenbruch, A. H., and J. H. Sass, Heat flow and energetics of the San Andreas fault
zone, J. Geophys. Res., 85, 6185-6222, 1980.
Lawn, B., Fracture of Brittle Solids - Second Edition, Cambridge University Press,
Cambridge, 1-378, 1993.
Mase, C. W., and L. Smith, Pore-fluid pressures and frictional heating on a fault surface,
Pure and Applied Geophysics, 122, 583-607, 1985.
413
Mase, C. W., and L. Smith, Effects of frictional heating on the thermal, hydrologic, and
mechanical response of a fault, J. Geophys. Res., 92, 6249-6272, 1987.
Melosh, J., Acoustic fluidization: a new geologic process?, J. Geophys. Res., 84, 7513-
7520, 1979.
Melosh, H. J., Dynamical weakening of faults by acoustic fluidization, Nature, 379, 601-
606, 1996.
Rabinowicz, E., Friction and Wear of Materials, 2 edition, John Wiley & Sons, Inc., New
York, 1-315, 1995.
Scholz, C. H., The mechanics of earthquake faulting, Cambridge University Press, New
York, 1-438, 1990.
Weertman, J., Unstable slippage across a fault that separates elastic media of different
elastic constants, J. Geophys. Res., 85, 1455-1461, 1980.
Energy Budget
Brune, J. N., T. L. Henyey, and R. F. Roy, Heat flow, stress, and the rate of slip along the
San Andreas fault, California, J. Geophys. Res., 74, 3821-3827, 1969.
Dahlen, F. A., The balance of energy in earthquake faulting, Geophys. J. R. astr. Soc., 48,
239-261, 1977.
Kostrov, B. V., Seismic moment and energy of earthquakes, and seismic flow of rock
(translated to English), Izv. Earth Physics, 1, 23-40, 1974.
Orowan, E., Mechanism of seismic faulting in rock deformation, Geol. Soc. Am. Mem.,
79, 323-345, 1960.
Savage, J. C., and J. B. Walsh, Gravitational energy and faulting, Bull. Seismol. Soc. Am.,
68, 1613-1622, 1978.
414
Melting, Fluid Pressurization, Lubrication
Jeffreys, H., On the mechanics of faulting, Geol. Mag., 79, 291-295, 1942.
Kanamori, H., T. H. Anderson, and T. H. Heaton, Frictional melting during the rupture of
the 1994 Bolivian Earthquake, Science, 279, 839-842, 1998.
Lachenbruch, A. H., Frictional heating, fluid pressure, and the resistance to fault motion,
J. Geophys. Res., 85, 6097-6112, 1980.
McKenzie, D. P., and J. N. Brune, Melting on fault planes during large earthquakes,
Geophys. J. R. astr. Soc., 29, 65-78, 1972.
Sibson, R. H., Interactions between temperature and fluid pressure during earthquake
faulting -- A mechanism for partial or total stress relief, Nature, 243, 66-68, 1973.
Sibson, R. H., Kinetic shear resistance, fluid pressures and radiation efficiency during
seismic faulting, Pure and Applied Geophysics, 115, 387-400, 1977.
415
Sibson, R. H., Power dissipation and stress levels on faults in the upper crust, J. Geophys.
Res., 85, 6239-6247, 1980.
Fault-zone Structure
Chester, F. M., and J. S. Chester, Ultracataclasite structure and friction processes of the
Punchbowl fault, San Andreas system, California, Tectonophysic, 295, 199-221, 1998.
Others
Ando, M., Source mechanisms and tectonic significance of historical earthquakes along
the Nankai trough, Japan, Tectonophysics, 27, 119-140, 1975.
Kanamori, H., The state of stress in the Earth's lithosphere, in Phys. Earth's Int., Course
LXXVIII, edited by A. M. Dziewonski and E. Boschi, pp. 531-554, North-Holland Pub.
Co., Amsterdam, 1980.
416
Ge 162 9. Plate Motion and Great Earthquakes
Earthquakes occur in the Earth's crust and mantle due to stresses caused by global
plate motion. The actual pattern of stress distribution is probably very complex, but we
expect that the activities of great and large earthquakes must reflect the global plate
motion.
417
The world greatest earthquakes occur at subduction zones (e.g., 1960 Chilean
earthquake, and the 1964 Alaskan earthquake), but not every subduction zone has
seismicity. With this caveat in mind, we investigate the level of seismic activity and
plate motion. Ideally, the seismic activity along a subduction zone should be defined by
the energy release per unit length along the subduction zone, and unit time, i.e.,
418
1 L T
e=
LT ∫ ∫
0 0
ER dldt
where L and T are the length of the subduction zone and the time period involved,
respectively.
Unfortunately, the available seismic record is too short to compute this. So, we
subduction zone as a parameter that represents e for that subduction zone. Then , it is
Mw ∝V
where V is the convergence rate. However, the plot of M w versus V does not show any
obvious trend. This suggests that other factors may be controlling seismicity. Another
419
Then, we can try a 3-parameter regression between M w , V and T. The result is
shown in the following figure. The horizontal axis shows the observed M w and the
420
(Ruff, L., and H. Kanamori, Seismicity and the subduction process, Phys. Earth Planet.
Inter., 23, 240-252, 1980)
If this regression is valid, this provides a useful method for assessing the seismic
potential of subduction zones for which no great earthquake has occurred. This pattern
suggests that the subduction zones where a relatively young plate is subducting at a
421
relatively fast rate are more likely to have great earthquakes, and those with an old plate
subducting at a moderate rate are less likely to have great earthquakes. The end-member
subduction zones are the Chilean type and the Mariana type, shown below.
(Uyeda, S., and H. Kanamori, Back-arc opening and the mode of subduction, J. Geophys.
Res., 84 (B3), 1049-1061, 1979)
422
Another interesting implication of this correlation is the seismic potential of the
Pacific Northwest (i.e., Oregon-Washington coast). The Juan de Fuca plate is subducting
beneath the states of Oregon and Washington. The background seismicity there is very
low, as shown below, and until mid 1980's, it was generally believed that the seismic
potential in the Pacific Northwest is low (i.e., great earthquakes are unlikely). However,
the age of the Juan de Fuca plate is very young, about 10 My, and it is subducting at a
rate of 3 cm/year. Thus, in view of the regression relation shown above, one would
expect a large, M w =8.5 to 9, earthquake there. This suggestion motivated the interest of
Geological evidence for regional submergence and evidence for large tsunami which
occurred in 1700 [Satake et al., 1996] now seem to have convinced most people, which
seems to have led to upgrading of building code in the area. This is a good example in
423
(Heaton, T., and H. Kanamori, Seismic potential associated with subduction in the
northwestern United States, Seismol. Soc. Am. Bull., 74 (3), 933-941, 1984)
424
(see, Atwater, B. F., and others, Summary of coastal geologic evidence for past great
earthquakes at the Cascadia subduction zone, Earthquake Spectra, 11, 1-18, 1995)
425
(R. S. Yeats, Living with Earthquakes in the Pacific Northwest, Oregon State University
Press, 1998)
References
Brune, J., Seismic moment, seismicity, and rate of slip along major fault zones, J.
Geophys. Res., 73, 777-784, 1968.
Satake, K., Shimazaki, K., Tsuji, Y. and Ueda, K., Time and size of a giant earthquake in
Cascadia inferred from Japanese tsunami records of January 1700, Nature, 379, 246-249,
1996.
426
427