You are on page 1of 11

Bulletin of the Seismological Society of America, Vol. 106, No. 4, pp. 1652–1662, August 2016, doi: 10.

1785/0120150320

Source-Scaling Relations of Interface Subduction Earthquakes


for Strong Ground Motion and Tsunami Simulation
by A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio

Abstract The recording on high-resolution broadband seismic networks of several


great interface subduction earthquakes during the last decade provide an excellent
opportunity to extend source-scaling relations to very large magnitudes and to place
constraints on the potential range of source parameters for these events. At present,
there is a wide range of uncertainty in the median rupture areas predicted for any given
seismic moment by current relationships between magnitude and rupture area for sub-
duction interface earthquakes. Our goal is to develop an updated set of earthquake
source-scaling relations that will reduce this current large degree of epistemic uncer-
tainty and improve the accuracy of seismic-hazard analysis and the prediction of the
strong-motion characteristics and tsunamis of future subduction earthquakes. To
achieve this goal, we compiled a database including slip models of interface earth-
quakes that occurred worldwide with moment magnitudes ranging from M 6.75–9.1.
We characterized the seismic sources based on well-established criteria to estimate the
asperity areas as well as the average slip on the faults, and we used these parameters to
compute an updated set of magnitude scaling relations of the various characteristics of
the fault. Additionally, we followed an alternative approach to quantifying slip models
for use in developing characteristic slip models of future earthquakes. This involved
analyzing the 2D Fourier transforms of the slip functions in the compiled database and
deriving a wavenumber spectral model of the slip distribution.

Introduction
The ability to simulate the ground motions and tsunamis includes the spatial distribution of slip and slip velocity on
from subduction interface earthquakes requires reliable the fault, which is derived from strong-motion recordings
source-scaling relations for this type of event. At present, and in turn is required for the simulation of strong ground
there is a range of over a factor of 3 in the median rupture motions and tsunamis. Geodetic and tsunami data are also
areas predicted for a given magnitude by current relation- useful for providing constraints on the spatial distribution of
ships between magnitude and rupture area for interface earth- slip on the fault.
quakes. For a given rupture area, there is a range of over 0.5 For forward simulation of strong ground motions, we
magnitude unit and a factor of over 5 in seismic moment in need to characterize the earthquake source in the frequency
the median size of interface earthquakes. This study aims to band of about 0.1–10 Hz or above and for tsunamis in the
reduce this large epistemic uncertainty in the median values frequency band of 0–0.01 Hz. Our goal therefore is to de-
of the scaling relations. velop earthquake source-scaling relations for interface earth-
After a long period of quiescence following the 1964 quakes over the very broad frequency range of 0–10 Hz or
M 9.1 Alaska earthquake, several great interface earthquakes above, so that they can be applied to the forward simulation
have occurred during the last decade, including the 2001 of both strong ground motions and tsunamis. We distinguish
M 8.4 Arequipa, Peru, 2004 M 9.1 Sumatra, 2010 M 8.8 two categories of kinematic source parameters. The first cat-
Chile, and 2011 M 9.0 Tohoku earthquakes. The recording egory, consisting of outer parameters, includes relationships
of these events on modern digital seismic networks provides between seismic moment and rupture length, rupture width,
an opportunity to extend the source-scaling relations of inter- rupture area, and average displacement. These parameters
face earthquakes to very large magnitudes and to place con- are needed for predicting both ground motions and tsunamis.
straints on the potential range of source parameters for these The second category, consisting of inner parameters that de-
events. Moreover, information about the source characteris- scribe the heterogeneities of slip and slip velocity (asperities)
tics of these recent earthquakes is much more reliable and on the fault rupture surface, includes relations between seis-
useful than that of older, large earthquakes. This information mic moment and spatial and temporal distribution of slip and

1652
Source-Scaling Relations of Interface Subduction Earthquakes for Strong Ground Motion and Tsunami Simulation 1653

slip velocity on the fault. The temporal inner parameters are After studying the properties of those models, we chose the
of most importance for the prediction of strong ground mo- model proposed by Lorito et al. (2011). Comparison of the
tions, which are highly dependent on slip velocity and rupture published models reveals that a large variability exists for all
velocity, but the spatial distribution of slip is also important for of the parameters. As seen in Table 2, the variability for spe-
tsunami simulation. In the present study, we analyze the scal- cific models and parameters can be up to 50%, which indi-
ing with seismic moment (M 0 ) of rupture width (W), rupture cates that judgment should be exercised in selecting the slip
area (S), average slip (D), and maximum slip (defined as the model that is most appropriate for a specific application.
average slip of the asperities; Da ) for the outer parameters and
total asperity area (Sa ) for the inner parameters.
Following an alternative approach, we use the spatial Development of Source-Scaling Relationships
wavenumber spectrum as an additional method of describing
the heterogeneity of slip on the fault surface. We analyze the Self-Smilar Models
2D Fourier transforms of the slip functions of our database We used the updated database of finite-fault rupture
with the two dimensions being the dimension along strike models that we compiled to produce scaling relations of the
and the dimension down-dip. The Fourier transform de- various source parameters. For the relations to be self-con-
scribes the relative amplitudes of the different spatial wave- sistent among rupture area (S), total asperity area (Sa ), and
lengths in the slip model. We assume that the wavenumber average slip (D), we used only finite-fault rupture models for
spectra in the along-strike and down-dip directions to have which all three quantities were available. For maximum slip
self-similar scaling with moment magnitude M and, by per- (Dmax ) and width (W), we used all available data. For the
forming a least-squares fit to these data, we derive a wave- regressions, we applied the Levenberg–Marquardt algorithm
number spectral model of the slip distribution for use in (Levenberg, 1944; Marquardt, 1963) through the least-
characterizing future earthquakes. squares method. We initially fit the data using the slope con-
strained to yield self-similar relationships. The logarithms of
Compilation of Earthquake Rupture Model Database S and Sa (km2 ) are proportional to two-thirds of seismic mo-
ment, whereas the logarithms of D and Dmax (m) are propor-
We compiled an updated database of interface earth- tional to one-third of the seismic moment:
quakes that occurred worldwide in the major subduction
zones, with moment magnitudes ranging M 6.75–9.1. Infor- 1
EQ-TARGET;temp:intralink-;df1;313;431

mation regarding the earthquake location and magnitude, the logD  logc1   logM0 ;
3
sources used for the compilation of the database, as well the 2
adopted values of the basic source parameters used in the logS  logc2   logM0 ; 1
3
analysis can be found in Table 1.
The locations of these earthquakes are shown in Figure 1. in which c1 and c2 are the regression coefficients and M 0 is
The majority of the finite-fault rupture models of the earth- seismic moment. Base 10 logarithms are used in the regres-
quakes in Table 1 were available from the online database fi- sions throughout the article. The coefficients and the stan-
nite-source rupture model database. Other major sources used dard deviations derived are given in Table 3.
for collecting slip models and information about them in- In Figure 2, the derived relations for S and Sa (black
clude Somerville et al. (2002) and Murotani et al. (2008, solid lines) are plotted together with the data used in the
2013). To characterize asperities in these cases, we use the analysis (different symbols to account for the various data
definition given in Somerville et al. (1999): an asperity is sources, present study; Sea2002, Somerville et al., 2002;
initially defined to enclose fault elements whose slip is 1.5 Mea2008, Murotani et al., 2008; Mea2013, Murotani et al.,
or more times larger than the average slip over the fault and is 2013). The shaded area represents the 1 standard devia-
subdivided if any row or column has an average slip less than tion. Similarly, the derived relations for D and Dmax are
1.5 times the average slip. shown in Figure 3. In Figure 4, the scaling between the
When the original slip model was not available in the combined area of asperities and rupture area is shown. The
database, we used the source characterization parameters re- combined area of asperities is found to be 0.24 times the
ported in the literature. We evaluated all available rupture rupture area, close to the results of Somerville et al. (2002)
models of each earthquake to understand the uncertainty shown in Table 3. The 2004 Sumatra and 1964 Alaska
in the slip model inversion process and to identify the best events show the largest departures from the best fit line and,
constrained features of the rupture models of these earth- for now, remain unexplained. The residual analysis (ob-
quakes. When an earthquake had more than one available served—predicted plotted against M 0 ) presented in Figure 5
rupture, we usually selected the model with the largest is used to examine the regression quality and identify po-
amount of strong motion or teleseismic data for use in the tential trends in the dataset. For all parameters studied, the
analysis. However, there were cases in which the selection residuals do not exhibit any significant trends, although the
was based on judgment, as in the multiple rupture models of residuals are all zero or negative for rupture area and com-
the Maule, Chile, earthquake, which are listed in Table 2. bined area of asperities for M 9 and larger.
Table 1
1654

Earthquakes Used for Source Characterization in Present Study


Seismic
Date Moment
Region (yyyy/mm/dd) (N·m) M S (km2 ) Sa (km2 ) D (m) Da (m) W (km) Kcx (km−1 ) Kcy (km−1 ) References
19
1 Hyuga-nada, Japan 1996/12/02 1:50 × 10 6.72 179 154 0.42 0.92 — — — Yagi et al. (1999)
2 Peru 1974/11/09 5:40 × 1019 7.09 3000 600 0.54 1.16 50 — — Somerville et al. (2002)
3 Playa Azul* 1981/10/25 7:14 × 1019 7.17 2700 400 0.74 2.37 60 0.010560 0.010310 Somerville et al. (2002)
4 Zihuatanejo* 1985/09/21 1:35 × 1020 7.35 3150 1350 1.02 1.54 60 0.005160 0.005100 Somerville et al. (2002)
5 Near coast of Guerrero, Mexico* 2012/03/03 1:41 × 1020 7.37 4125 1050 0.41 0.20 100 0.010690 0.010430 Wei (Caltech, Oaxaca 2012)
6 Honshu, Japan 2005/08/16 2:00 × 1020 7.47 3584 960 0.15 0.70 72 — — Shao and Ji (UCSB, Honshu 2005)
7 Colima, Mexico* 2003/01/22 2:30 × 1020 7.51 5950 1350 0.61 1.30 85 0.011000 0.010400 Yagi et al. (2004)
8 Hyuga-nada, Japan 1968/04/01 2:50 × 1020 7.53 1377 1053 1.32 2.90 — — — Yagi et al. (1998)
9 Costa Rica* 2012/09/05 2:54 × 1020 7.54 18,000 3520 0.29 0.95 120 0.007085 0.005098 Hayes (NEIC, Costa Rica 2012)
10 East of Sulangan, Philippines* 2012/08/31 2:72 × 1020 7.56 4608 1440 0.42 1.90 90 0.006147 0.005599 Hayes (USGS, Philippines 2012)
11 Papua* 2009/01/03 2:82 × 1020 7.57 11,520 1680 0.59 2.00 96 0.008140 0.006630 Hayes (NEIC, Papua 2009)
12 Vanuatu* 2009/10/07 2:82 × 1020 7.57 4200 1680 0.87 2.05 60 0.010070 0.006136 Sladen (Caltech, Vanuatu 2009)
13 Fiordland, New Zealand* 2009/07/15 2:82 × 1020 7.57 10,752 2560 0.63 2.60 96 0.008290 0.006740 Hayes (NEIC, New Zealand 2009)
14 Nihonkai-chubu, Japan 1983/05/26 3:00 × 1020 7.58 2700 — 3.17 — — — — Fukuyama and Irikura (1986)
15 Hokkaido-Nansei* 1993/11/12 3:40 × 1020 7.62 14,000 2300 0.64 1.64 70 0.008071 0.004121 Mendoza and Fukuyama (1996)
16 Tocopilla, Chile* 2007/11/14 3:98 × 1020 7.67 18,954 7695 0.88 1.75 126 0.005000 0.002400 Sladen (Caltech, Tocopilla 2007)
17 Sanrikuki, Japan* 1994/12/28 3:99 × 1020 7.67 15,400 2600 0.71 1.95 140 0.007000 0.003261 Nagai et al. (2001)
18 Masset, Canada* 2012/10/28 4:27 × 1020 7.69 4800 1440 1.57 4.80 60 0.015029 0.004176 Shao and Ji (UCSB, Masset 2012)
19 Sanriku-haruka-Oki, Japan 1994/12/28 4:40 × 1020 7.70 2800 2800 0.71 1.93 — — — Nagai et al. (2001)
20 Kanto, Japan* 1923/09/01 7:60 × 1020 7.85 2340 2210 2.54 5.60 — 0.008057 0.005088 Wald and Somerville (1995)
21 Pagai, Indonesia* 2007/09/12 7:94 × 1020 7.87 21,875 6500 0.55 1.47 90 0.006500 0.004000 Ji and Zeng (Pagai 2007)
22 Colima, Mexico* 1995/10/09 9:67 × 1020 7.92 17,000 2800 1.18 2.80 100 0.007164 0.003320 Mendoza and Hartzell (1999)
23 Pisco, Peru* 2007/08/15 1:12 × 1021 7.97 20,736 5508 1.63 3.80 108 0.006670 0.005190 Ji and Zeng (Peru 2007)
24 Samoa* 2009/09/29 1:12 × 1021 7.97 7243 1983 3.33 8.98 49 0.010369 0.005690 Hayes (NEIC, Samoa 2009)
25 Michoacan, Mexico* 1985/09/19 1:15 × 1021 7.97 25,020 5004 1.40 2.95 139 0.004800 0.003900 Mendoza and Hartzell (1989)
26 Peru 1974/10/03 1:20 × 1021 7.99 28,000 6066 1.30 2.19 112 — — Somerville et al. (2002)
27 Nazca ridge, Peru* 1996/11/12 1:38 × 1021 8.03 36,000 9072 0.77 1.53 120 0.004300 0.003300 Spence et al. (1999)
28 Solomon Islands 2007/04/01 1:58 × 1021 8.07 21,600 6600 1.47 2.70 80 — — Ji (UCSB, Solomon Islands 2007)
29 Tokachi-Oki, Japan* 2003/09/25 1:92 × 1021 8.12 22,100 5600 1.46 3.15 170 0.004694 0.004377 Yagi (2004)
30 Central Chile* 1985/03/03 1:96 × 1021 8.13 34,425 9675 1.92 1.75 165 0.004100 0.003300 Mendoza et al. (1994)
31 Tonankai, Japan* 1944/12/07 2:40 × 1021 8.19 4000 4800 1.05 1.78 — 0.003300 0.002800 Ichinose et al. (2003)
32 Kuril Islands* 2006/11/15 3:16 × 1021 8.27 35,750 10,000 1.69 5.10 138 0.005216 0.001898 Ji (UCSB, Kuril 2006)
33 Tokachi-Oki, Japan 1968/05/16 3:50 × 1021 8.30 6800 5600 2.31 5.49 — — — Nagai et al. (2001)
34 Arequipa* 2001/06/23 3:70 × 1021 8.31 80,000 20,800 1.22 2.48 200 0.002770 0.001860 Somerville et al. (2003)
35 Nankai, Japan 1946/12/21 3:90 × 1021 8.33 52,650 — 1.98 — — — — Tanioka and Satake (2001)
36 Benkulu, Indonesia* 2007/09/12 4:47 × 1021 8.37 73,140 28,331 0.90 1.85 160 0.003772 0.002229 Ji (UCSB, Benkulu 2007)
37 Sumatra* 2005/03/28 1:17 × 1022 8.65 86,400 27,200 2.56 5.45 260 0.002895 0.002280 Shao and Ji (UCSB, Sumatra 2005)
38 Aleutian 1957/03/09 1:20 × 1022 8.65 93,750 30,000 3.10 — — — — Johnson et al. (1994)
39 Kamchatka 1952/11/04 1:50 × 1022 8.72 70,000 20,000 5.50 — — — — Johnson and Satake (1999)
40 Maule, Chile* 2010/02/27 1:55 × 1022 8.73 115,000 31,875 4.13 9.00 200 0.003232 0.002274 Lorito et al. (2011)
41 Tohoku-Oki, Japan* 2011/03/11 4:20 × 1022 9.02 81,000 18,900 10.51 22.97 180 0.002402 0.003285 Yokota et al. (2011)
42 Alaska* 1964/03/27 5:52 × 1022 9.09 225,000 30,000 4.00 11.30 265.6 0.003010 0.002441 Ichinose et al. (2007)
(continued)
A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio
Source-Scaling Relations of Interface Subduction Earthquakes for Strong Ground Motion and Tsunami Simulation 1655

Fujii and Satake (2013)


References

Ammon et al. (2005)

Figure 1.
0.001900
Kcy (km−1 )

Locations of the earthquakes listed in Table 1.


Non-Self-Similar Models
0.002401
Kcx (km−1 )

We assumed self-similar scaling laws in performing the


regression analysis in the previous section. However, there are

studies (e.g., Papazachos et al., 2004; Strasser et al., 2010; Ta-


jima et al., 2013) that suggest a departure from self-similarity
W (km)

of the rupture area and slip of the fault. We relaxed the con-

193

USGS, U.S. Geological Survey; UCSB, University of California, Santa Barbara; NEIC, National Earthquake Information Center.

straint of self-similarity and fit the data to a non-self-similar


relationship of the form
Da (m)

6.70

EQ-TARGET;temp:intralink-;df2;314;467 logY  logca   cb logM0 ; 2


Table 1 (Continued)

3.90
10.60
D (m)

in which Y corresponds to the different source parameters


and ca , cb are the regression coefficients. The coefficient val-
Sa (km2 )

27,571
40,000

ues along with the standard deviations from the regressions


are listed in Table 4.
In Figures 6–8, the regression results for the non-self-
265,237
135,000

similar functional forms are shown by the solid lines with


S (km2 )

triangles. Comparisons with the results from the self-similar


functional forms (black curves) do not exhibit substantial
differences for average slip and rupture area. To statistically
9.14
9.17
M

confirm the differences in data fit between the unconstrained


models and those that are constrained to be self-similar, we
22

7:20 × 1022

tested the null hypothesis by performing a two-sided t-test


Moment
Seismic

6:50 × 10
(N·m)

analysis on the slopes of the rupture area, asperity area, and


average slip relations (see Tables 3 and 4 for their values).
The estimated p-values, −7:3 × 10−10 , −1:94 × 10−10 , and
(yyyy/mm/dd)

4:13 × 10−8 , respectively, are quite small, so the differences


2004/12/26
1960/05/22
Date

are not statistically significant. Therefore, we prefer to retain


*Events used in the corner wavenumber analysis.

the simplicity of using the self-similar relations as found in


some of the previous studies (Somerville et al., 2002; Mur-
otani et al., 2008, 2013).

Fault Width Scaling Models


Sumatra, Indonesia*
Region

In Figure 9, the scaling of fault width (down-dip dimen-


sion of the rupture area) with respect to seismic moment is
presented. The scaling coefficients from the least-squares fit
are shown in Table 4 (model W 1 ). Both data (Fig. 9) and the
Chile

residual distribution (Fig. 10, solid symbols) justify the linear


model up to the maximum magnitude that was used in the
regression (M 9.17 for Sumatra 2004 earthquake). However,
43
44

there are studies (e.g., Blaser et al., 2010; Tajima et al., 2013)
1656 A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio

Table 2
Published Rupture Models for the Maule, Chile, Earthquake (Number 40 in Table 1) Reviewed in the Present Study
Seismic Rupture
Moment Length Width Area Trimmed Rupture Asperity Rupture Average Asperity
(N·m) (km) (km) (km2 ) Area (km2 ) Area (km2 ) Slip (m) Slip (m) Author
22
1:74 × 10 570 180 102,600 102,600 (10.8%) 23,850 (25.2%) 2.29 (44.6%) 4.15 (53.9%) Sladen (Caltech,
Maule 2010)
1:60 × 1022 540 200 108,000 86,400 (24.9%) 26,400 (17.2%) 3.87 (6.3%) 9.25 (2.8%) Hayes (NEIC,
Maule 2010)
1:78 × 1022 680 256 174,148 74,437 (35.3%) 18,904 (40.7%) 2.69 (34.9%) 10.20 (13.3%) Luttrell et al. (2011)
2:51 × 1022 600 187 112,200 84,150 (26.8%) 18,360 (42.4%) 4.05 (1.9%) 8.80 (2.2%) Shao et al. (UCSB,
Maule 2010)
1:55 × 1022 650 200 130,000 115,000 31,875 4.13 9.00 Lorito et al. (2011)

UCSB, University of California, Santa Barbara; NEIC, National Earthquake Information Center. The Lorito et al. (2011) rupture model was the
preferred one. The absolute percentage difference of each parameter with respect to that of the preferred model is denoted in parentheses.

Table 3
Self-Similar Scaling Relations, Regression Coefficients, and Standard Deviations
M 0 -Rupture Area (S) M0 -Average Slip (D) M 0 -Total Asperity Area (Sa ) S − Sa M 0 − Dmax
c2 σ c1 σ c2 σ C1 σ c2 σ
−10 −07 −11 −07
Present study 1:77 × 10 1.498 1:23 × 10 1.527 4:16 × 10 1.613 0.24 1.40 5:00 × 10 1.508
Murotani et al. (2013) 1:34 × 10−10 1.540 1:66 × 10−07 1.640 2:81 × 10−11 1.720 0.20 1.41 — —
Murotani et al. (2008) 1:48 × 10−10 1.610 1:48 × 10−07 1.720 2:89 × 10−11 1.780 0.20 1.41 — —
Somerville et al. (2002) 2:41 × 10−10 — 1:14 × 10−07 — 5:62 × 10−11 — 0.25 — — —

The equations used in the regressions are logD  logc1   13 logM 0 , logS  logc2   23 logM 0 .

indicating that beyond a certain magnitude, the fault width acteristic slip models of future earthquakes. This approach was
tends to a constant value (saturates). We tested this assump- originally described in Somerville et al. (1999) for crustal
tion by fitting a bilinear model that saturates for M > 8:4, earthquakes, but in the present study we apply the same model
consistent with the model of Tajima et al. (2013). The regres- to interface earthquakes.
sion coefficients for the bilinear model are listed in Table 4 The first step is to compute the 2D Fourier transforms of
as model W 2 and the fit is depicted with the dashed line in the slip as a function of distance for the subset of events listed
Figure 9. in Table 1 with the two dimensions being the along-strike
The standard deviation of the bilinear fit is slightly and the down-dip distance. The Fourier transform describes
smaller than that of the linear fit, basically because the bilinear the relative amplitudes of the different spatial wavelengths
model gives a better fit to the three data points at M > 9
(wavenumbers) in the slip model. Small wavenumbers are
(Alaska 1964, Sumatra 2004, and Tohoku 2011 earthquakes).
equivalent to long wavelengths and represent broad fluctua-
In Figure 10, the open symbols represent the residuals from
tions of slip over the fault surface, whereas large wavenum-
the bilinear fit, and it can be seen that the largest difference is
bers are equivalent to short wavelengths and represent local
observed only for these three points. We believe that this might
be an indication of fault width saturation. The study of Tajima fluctuations over the fault surface. The spatial sampling of
et al. (2013) also suggests saturation at a median width of the fault in the along-strike and down-dip directions control
200 km, similar to our result in Figure 9. The limited number the highest wavenumber (Nyquist wavenumber) for which
of data for M > 8:3, and the poor constraint of the fit for the slip model is complete. The corner spatial wavenumbers
M > 9:0, do not provide definitive resolution of width satu- were used to construct a wavenumber spectral model.
ration. However, we consider that width saturation at a median The slip models were resampled at 1 km spacing
width of 200 km is most likely present, but may vary from one using first degree bivariate splines and they were padded
subduction zone to another. with 0–1024 km in each direction (resulting dimension
2048 × 2048) to produce even sampling of the wavenumber
spectra. We obtained the parameters of a wavenumber spec-
Characterization Based on Corner Wavenumbers
tral model of the slip distribution in earthquakes by fitting a
So far, we have made quantitative estimates of the param- simple functional form to the wavenumber spectra of indi-
eters relating to slip models and analyzed their scaling with vidual earthquakes. We used a 2D Butterworth filter func-
seismic moment. In this section, we follow an alternative ap- tion to model the wavenumber amplitude spectrum, which
proach to quantifying slip models for use in developing char- is described by the following relation:
Source-Scaling Relations of Interface Subduction Earthquakes for Strong Ground Motion and Tsunami Simulation 1657

Figure 2. Scaling of the rupture area and the combined area of


asperities area with seismic moment, plotted together with data from Figure 4. Relationship between combined area of asperities and
various studies: (Sea2002, Somerville et al., 2002; Mea2013, Mur- rupture area; symbols are the same as in Figure 3. The color version
otani et al., 2013; and Mea2008, Murotani et al., 2008). The shaded of this figure is available only in the electronic edition.
area indicates the 1 standard deviation limits. The color version of
this figure is available only in the electronic edition.

Figure 5. Residual (observed–predicted) plots against seismic


moment for average and maximum slip, rupture area, and total area
Figure 3. Scaling of average and maximum slip with seismic of asperities. Symbols are the same as in Figure 4. The color version
moment, plotted together with data from various studies; symbols of this figure is available only in the electronic edition.
are the same as in Figure 2. The shaded area indicates the 1 stan-
dard deviation limits. The color version of this figure is available Figure 11 for two slip models (Kanto, Japan [1923] and
only in the electronic edition. Sumatra [2004]). The estimated maximum spatial wave-
numbers are listed in Table 1.
1 To verify that the corner wavenumbers estimated from
ampkx; ky  q ; 3
ky 2 2
EQ-TARGET;temp:intralink-;df3;55;188

kx 2 the regressions actually describe the decay of the amplitudes


1  Kc x
  Kc y
 
with wavenumber in each direction, we plotted the logarithm
of the normalized amplitudes and compared them with the
in which ampkx; ky is the 2D Fourier transform ampli-
simplified model
tudes, kx and ky are the wavenumbers, and Kcx and Kcy
are the corner wavenumbers, in each dimension. We fit re- ampk  1  k=Kc4 −2 ; 4
1
EQ-TARGET;temp:intralink-;df4;313;127

lation (3) using the damped least-squares method to solve


this nonlinear problem. We performed 1000 iterations and in which Kc is the maximum wavenumber in each direction.
the damping coefficient that was used had the value of For the two slip models in Figure 11, the fits are shown in
λ  2 × 1011 . An example of the procedure is shown in Figure 12. The vertical lines depict the Nyquist wavenum-
1658 A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio

Table 4
Non-Self-Similar Scaling Relations, Regression
Coefficients, and Standard Deviations
Source Parameter ca cb Σ

Rupture area (A) 1:72 × 10−09 0.620 1.481


Total asperity area (Aa ) 4:81 × 10−10 0.616 1.596
Average slip (D) 3:39 × 10−08 0.360 1.522
Maximum slip (Dmax ) 2:70 × 10−06 0.299 1.501
Width (W 1 ) 6:75 × 10−03 0.201 1.264
Width (W 2 ) 1:66 × 10−04 0.279 1.259

The equation used in the regressions is logY  logca 


cb logM 0 .

bers of the original slip models, before resampling and pad-


ding, along the x and y directions.
The corner wavenumbers Kcx and Kcy in the along-
strike and down-dip directions, respectively, were assumed
to each have self-similar scaling with M. For self-similar scal- Figure 6. Scaling relations of the rupture area from various
studies with respect to seismic moment (Sea2002, Somerville et al.,
ing, the logarithm of the corner wavenumber is proportional to
2002; Mea2013, Murotani et al., 2013; Mea2008, Murotani et al.,
one half the moment magnitude: 2008; Tea2010, Strasser et al., 2010; Bea2010, Blaser et al., 2010;
and Pea2004, Papazachos et al., 2004). The shaded area indicates the
EQ-TARGET;temp:intralink-;df5;55;490 logKcx   cx − 0:5M; logKcy   cy − 0:5M: 5 1 standard deviation limits of this study’s self-similar model. The
color version of this figure is available only in the electronic edition.
The coefficients from the least-squares fit are cx  1:75 with
σ  0:146 and cy  1:59 with σ  0:168. In Figure 13, the analysis. In cases where we had multiple source models for
logarithms of the corner wavenumbers (cx , top panel; cy , bottom a single earthquake, we used judgment to select the most rep-
panel) are plotted together with the least-squares fit. The shaded resentative one based on various criteria, such as the number
area corresponds to the 1 standard deviation. Based on the and type of data used in deriving the model.
constant width model that we presented in the previous sec- Despite the larger number of available data, including
tion, we additionally studied possible trends in cy with fault the most recent megathrust earthquakes, there are still few
width. The symbols in the lower panel of Figure 13 correspond
data to constrain the behavior of the scaling relations at very
to different fault width bins. The residual (observed–predicted)
large magnitudes (M > 8:6). This limitation is prominent in
distribution against M is shown in Figure 14. The residual plots
the study of fault width scaling, for which several researchers
verify that for the along-strike dimension there is no apparent
suggest saturation of width (i.e., the down-dip rupture width
trend in our model. Similarly for cy (down-dip direction), there
stops growing beyond a certain magnitude resulting in con-
is no dependence on fault width except for the largest three
earthquakes with width greater than 160 km, consistent with stant width). In this study, there were only three data points
the saturation of width shown in Figure 9. available at very large magnitudes. This high level of uncer-
tainty makes it difficult to resolve the presence of width sat-
uration. However, we consider that width saturation at a
Conclusions median width of 200 km is most likely present but may vary
We compiled an updated database of interface earthquakes from one subduction zone to another.
that occurred worldwide in the major subduction zones with The differences in data fit between the unconstrained
moment magnitudes ranging M 6.75–9.17. We evaluated all models and those that are constrained to be self-similar are
available rupture models for each earthquake and selected not large and are not statistically significant. Therefore, we
the ones that were based on the largest amount of strong motion prefer to retain the simplicity of using the self-similar relations
or teleseismic data. To estimate the various source parameters, (Table 3) as found in some of the previous studies (Somerville
we characterized the asperities for the original slip models based et al., 2002; Murotani et al., 2008, 2013).
on a well-established methodology (Somerville et al., 1999). The comparison of the available scaling models of rup-
We studied the scaling with seismic moment of fault ture area for both self-similar and non-self-similar functional
width, rupture area, total asperity area, and average and forms presented in Figure 6 shows that there are still large
maximum slip. In all cases, the standard deviations are com- differences between the models of the various authors. The
parable, if not smaller, than the values estimated by Murotani results from our regression analysis of a large dataset suggest
et al. (2008, 2013). A factor that might have contributed to that the Papazachos et al. (2004), Blaser et al. (2010), and
the smaller standard deviations is that we did not use more Strasser et al. (2010), models are not consistent with the
than one model of the same earthquake in the regression moment–area relation obtained from the data in Figure 6
Source-Scaling Relations of Interface Subduction Earthquakes for Strong Ground Motion and Tsunami Simulation 1659

Figure 7. Scaling relations of the average slip from various Figure 9. Scaling relation of fault width with respect to seismic
studies with respect to seismic moment. Line styles are the same moment. Symbols are the same as in Figure 2. The dashed line cor-
as in Figure 6. The shaded area indicates the 1 standard deviation responds to the bilinear model used in the regressions (model num-
limits of this study’s self-similar model. The color version of this ber 2 in figure legend). The shaded area indicates the 1 standard
figure is available only in the electronic edition. deviation limits of the linear model. The color version of this figure
is available only in the electronic edition.

Figure 8. Scaling relations of the total area of asperities from


various studies with respect to seismic moment. Line styles are the Figure 10. Width residuals (observed–predicted) with respect to
same as in Figure 7. The shaded area indicates the 1 standard seismic moment. Symbols are the same as in Figure 2. The open sym-
deviation limits of this study’s self-similar model. The color version bols denote the residuals for the bilinear model used in the regressions.
of this figure is available only in the electronic edition. The color version of this figure is available only in the electronic edition.

(although Stirling et al., 2013, prefer the Strasser et al., 2010, et al. (1999) and later by several other studies (e.g., Papaza-
model without discussing or showing comparisons with other chos et al., 2004; Murotani et al., 2008, 2013; Strasser et al.,
models). Similarly, the Papazachos et al. (2004) model is not 2010). Somerville et al. (1999) reported that, on average, in-
consistent with the moment–average displacement relation ob- terface earthquakes have rupture areas that are two or more
tained from the data in Figure 7. We believe that these discrep- times larger than those of crustal earthquakes having the
ancies arise mainly from the different datasets used in the same seismic moment. To test this assumption with our data-
derivation of the relations and, to a lesser extent, from the dif- set, we compared the scaling coefficients of Table 3 with the
ferent functional forms used by the different authors. corresponding ones reported in Somerville et al. (1999) for
The differences in rupture areas between interface and crustal earthquakes. The results presented in Table 5 show that,
crustal earthquakes were originally identified by Somerville on average, the rupture areas of interface earthquakes are ∼1:7
1660 A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio

Figure 11. (a,b) Examples of original and interpolated slip models (upper plot: Kanto, Japan [1923] and bottom plot: Sumatra [2004]).
(c,d) Spectral decay fits for along-strike and down-dip directions. The color version of this figure is available only in the electronic edition.

Figure 12. Spectral decay fits for along-strike and down-dip directions for Kanto, Japan (1923) and Sumatra (2004) slip models. The
vertical lines depict the Nyquist wavenumber of the original slip models, before resampling and padding, along the x and y directions. The
color version of this figure is available only in the electronic edition.
Source-Scaling Relations of Interface Subduction Earthquakes for Strong Ground Motion and Tsunami Simulation 1661

Table 5
Comparison of Scaling Coefficients of Interface and Crustal
Earthquakes
Parameter Interface Crustal Ratio

Rupture area 1:77 × 10−10 1:04 × 10−10 1.70


Average slip 1:23 × 10−07 3:36 × 10−07 0.37
Combined area of asperities 4:16 × 10−11 2:32 × 10−11 1.79
Fraction of fault covered by 0.24 0.22 1.09
asperities

present study results estimate larger combined asperity areas


for the same seismic moment than those in Murotani et al.
(2008, 2013), and lower than those in Somerville et al. (2002).
We also characterized the slip functions using wavenumber
spectral models of the slip distribution. We resampled and zero
Figure 13. Least-square fits of equation (4). The shaded area padded the models to produce even sampling. The fitting of the
indicates the 1 standard deviation limits. The symbols in the bot- 2D Butterworth filter function to estimate the corner wavenum-
tom panel correspond to different fault width bins. The color version bers for each direction involved a nonlinear regression procedure
of this figure is available only in the electronic edition. solving the damped least-squares equations using the Leven-
berg–Marquardt algorithm. The corner wavenumbers follow a
self-similar scaling law with moment magnitude, as can be seen
in Figure 13. The linear fit to the data does not exhibit any sig-
nificant trends based on the residual analysis presented in Fig-
ure 14, except for magnitudes larger than 9. The along-strike
coefficient cx has a very similar value compared with the one
obtained in crustal earthquakes in Somerville et al. (1999),
whereas the down-dip coefficient cy has a lower value. Hence,
for a given magnitude, interface earthquake spatial slip functions
exhibit heterogeneity in slip that is similar along strike but less
down-dip compared with crustal earthquakes. We attribute this
to the shallower dip angle and lower seismic-velocity gradient of
the plate interface compared with the crust.

Data and Resources


The majority of the finite-fault models used in the analyses
were downloaded from the online database finite-source rupture
model database (http://equake‑rc.info/SRCMOD/, last accessed
Figure 14. Residual distribution against magnitude (M) for cx
and cy . The symbols in the bottom panel correspond to different November 2015). The Kuril, 2006; Solomon Islands, 2007;
fault width bins. The color version of this figure is available only Peru, 2007; Sumatra, 2005; New Zealand, 2009; Papua, 2009;
in the electronic edition. Samoa, 2009; Philippines, 2012; and the Costa Rica, 2012,
finite-fault models were downloaded from U.S. Geological Sur-
times larger and the average slip is ∼0:4 times as large as those vey (USGS) significant earthquakes archive webpage (http://
of crustal earthquakes having the same seismic moment. earthquake.usgs.gov/earthquakes/eqinthenews, last accessed
For the rupture area, the mean values computed in this February 2016). The Sumatra 2005; Honsu, 2005; Peru,
study are very similar (within 1 standard deviation) to those 2007; Sumatra, 2007; and Masset, 2012, finite-fault models were
reported in Murotani et al. (2008, 2013) (Fig. 6, dashed- downloaded from the University of California, Santa Barbara
diamond and dashed-square lines, respectively). The compar- (UCSB) big earthquake webpage (http://www.geol.ucsb.edu/
isons of the present study results with those of Somerville et al. faculty/ji/big_earthquakes/earthquakes.html, last accessed Febru-
(2002) (dotted line) indicate smaller areas for the same seismic ary 2016). The Tocopilla, 2007; Vanuatu, 2009; and the Mexico,
moment in the new relationships. For the average slip, the 2012, finite-fault model was downloaded from the Caltech slip
mean values computed in this study are similar to but lower history database (http://www.tectonics.caltech.edu/slip_history/,
than those estimated in Murotani et al. (2008, 2013; 20% and last accessed February 2016). Some plots were created using
35% lower, respectively) and slightly higher (7%) than those the Generic Mapping Tools v.4.2.1 (www.soest.hawaii.edu/
in Somerville et al. (2002) (Fig. 7). Figure 8 shows that the gmt, last accessed November 2013; Wessel and Smith, 1998).
1662 A. A. Skarlatoudis, P. G. Somerville, and H. K. Thio

Acknowledgments The 1968 Tokachi-Oki earthquake and the 1994 Sanriku-Oki earth-
quake, Zisin 54, 267–280 (in Japanese with English abstract).
This work was supported by the U.S. Geological Survey (USGS), Papazachos, B. C., E. M. Scordilis, D. G. Panagiotopoulos, C. B. Papaza-
Department of the Interior, under USGS Award Number (P. Somerville, chos, and G. F. Karakaisis (2004). Global relations between seismic
G13AP00028). The views and conclusions contained in this document are fault parameters and moment magnitudes of earthquakes, Bull. Geol.
those of the authors and should not be interpreted as necessarily representing Soc. Greece 36, 1482–1489.
the official policies, either expressed or implied, of the U.S. Government. We Somerville, P. G., K. Irikura, R. Graves, S. Sawada, D. Wald, N. Abraham-
want to thank the two anonymous reviewers who contributed to the improv- son, Y. Iwasaki, T. Kagawa, N. Smith, and A. Kowada (1999). Char-
ment of the quality of the article with their comments and suggestions. acterizing crustal earthquake slip models for the prediction of strong
ground motion, Seismol. Res. Lett. 70, 59–80.
Somerville, P. G., T. Sato, T. Ishii, N. F. Collins, K. Dan, and H. Fujiwara
References (2002). Characterizing heterogeneous slip models for the large subduc-
tion earthquakes for strong ground motion prediction, Proc. of 11th
Ammon, C. J., J. Chen, H.-K. Thio, D. Robinson, S. Ni, V. Hjorleifsdottir, H. Japan Earthq. Eng. Symp., 20–22 November 2002, 163–166 (in Jap-
Kanamori, T. Lay, S. Das, D. Helmberger, et al. (2005). Rupture process of anese with English abstract).
the great 2004 Sumatra–Andaman earthquake, Science 308, 1133–1139. Somerville, P. G., H. K. Thio, G. Ichinose, N. Collins, A. Pitarka, and R. Graves
Blaser, L., F. Krüger, M. Ohrnberger, and F. Scherbaum (2010). Scaling re- (2003). Earthquake source and ground motion characteristics of the June
lations of earthquake source parameter estimates with special focus on 23, 2001 M w 8.4 Arequipa, Peru, earthquake, Seismol. Res. Lett. 74, 223.
subduction environment, Bull. Seismol. Soc. Am. 100, 2914–2926. Spence, W., C. Mendoza, E. R. Engdahl, G. L. Choy, and E. Norabuena
Fujii, Y., and K. Satake (2013). Slip distribution and seismic moment of the (1999). Seismic subduction of the Nazca ridge as shown by the
2010 and 1960 Chilean earthquakes inferred from tsunami waveforms 1996–97 Peru earthquakes, Pure Appl. Geophys. 154, 753–776.
and coastal geodetic data, Pure Appl. Geophys. 170, 1493–1509. Stirling, M., T. Goded, K. Berryman, and N. Litchfield (2013). Selection of
Fukuyama, E., and K. Irikura (1986). Rupture process of the 1983 Japan Sea earthquake scaling relationships for seismic-hazard analysis, Bull.
(Akita-Oki) earthquake using a waveform inversion method, Bull. Seismol. Soc. Am. 103, no. 6, 2993–3011, doi: 10.1785/0120130052.
Seismol. Soc. Am. 76, 1623–1640. Strasser, F. O., M. C. Arango, and J. J. Bommer (2010). Scaling of the source
Ichinose, G., P. Somerville, H.-K. Thio, R. Graves, and D. O’Connell (2007). dimensions of interface and intraslab subduction-zone earthquakes
Rupture process of the 1964 Prince William Sound, Alaska, earthquake with moment magnitude, Seismol. Res. Lett. 81, 941–950.
from the combined inversion of seismic, tsunami, and geodetic data, Tajima, R., Y. Matsumoto, H. Si, and K. Irikura (2013). Comparative study
J. Geophys. Res. 112, no. B07306, doi: 10.1029/2006JB004728. on scaling relations of source parameters for great earthquakes in in-
Ichinose, G. A., H. K. Thio, P. G. Somerville, T. Sato, and T. Ishii (2003). land crusts and on subducting plate boundaries, Zisin 66, 31–45 (in
Rupture process of the 1944 Tonankai earthquake (M s 8.1) from the Japanese with English abstract).
inversion of teleseismic and regional seismograms, J. Geophys. Res. Tanioka, Y., and K. Satake (2001). Coseismic slip distribution of the 1946
108, no. B10, 2497, doi: 10.1029/2003JB002393. Nankai earthquake and aseismic slips caused by the earthquake, Earth
Johnson, J. M., and K. Satake (1999). Asperity distribution of the 1952 great Planets Space 53, no. 4, 235–241.
Kamchatka earthquake and its relation to future earthquake potential in Wald, D. J., and P. G. Somerville (1995). Variable-slip rupture model of the
Kamchatka, Pure Appl. Geophys. 154, 541–553. great 1923 Kanto, Japan, earthquake: Geodetic and body-waveform
Johnson, J. M., Y. Tanioka, L. J. Ruff, K. Satake, H. Kanamori, and L. R. Sykes analysis, Bull. Seismol. Soc. Am. 85, 159–177.
(1994). The 1957 great Aleutian earthquake, Pure Appl. Geophys. 142, 3–28. Wessel, P., and W. H. F. Smith (1998). New, improved version of the Generic
Levenberg, K. (1944). A method for the solution of certain nonlinear prob- Mapping Tools released, Eos Trans. AGU 79, 579.
lems in least squares, Q. Appl. Math. 2, 164–168. Yagi, Y. (2004). Source rupture process of the 2003 Tokachi-Oki earthquake
Lorito, S., F. Romano, S. Atzori, X. Tong, A. Avallone, J. McCloskey, M. determined by joint inversion of teleseismic body wave and ground
Cocco, E. Boschi, and A. Piatanesi (2011). Limited overlap between motion data, Earth Planets Space 56, 311–316.
the seismic gap and coseismic slip of the great 2010 Chile earthquake, Yagi, Y., M. Kikuchi, S. Yoshida, and T. Sagiya (1999). Comparison of the
Nat. Geosci. 4, 173–177. coseismic rupture with the aftershock distribution in the Hyuga-nada
Luttrell, K. M., X. Tong, D. T. Sandwell, B. A. Brooks, and M. G. Bevis earthquakes of 1996, Geophys. Res. Lett. 26, 3161–3164.
(2011). Estimates of stress drop and crustal tectonic stress from the 27 Yagi, Y., M. Kikuchi, S. Yoshida, and Y. Yamanaka (1998). Source process
February 2010 Maule, Chile, earthquake: Implications for fault of the Hyuga-nada earthquake of April 1, 1968 (M JMA 7.5), and its
strength, J. Geophys. Res.: Solid Earth (1978–2012) 116, no. B11. relationship to the subsequent seismicity, Zisin 51, 139–148 (in Jap-
Marquardt, D. (1963). An algorithm for least-squares estimation of nonlinear anese with English abstract).
parameters, SIAM J. Appl. Math. 11, 431–441. Yagi, Y., T. Mikurno, J. Pacheco, and G. Reyes (2004). Source rupture proc-
Mendoza, C., and E. Fukuyama (1996). The July 12, 1993, Hokkaido Nan- ess of the Tecoman, Colima, Mexico earthquake of 22 January 2003,
sei-Oki, Japan, earthquake: Coseismic slip pattern from strong-motion determined by joint inversion of teleseismic body-wave and near-
and teleseismic recordings, J. Geophys. Res. 101, 791–801. source data, Bull. Seismol. Soc. Am. 94, 1795–1807.
Mendoza, C., and S. Hartzell (1989). Slip distribution of the 19 September Yokota, Y., K. Koketsu, Y. Fujii, K. Satake, S. Sakai, M. Shinohara, and T.
1985 Michoacan, Mexico, earthquake: Near-source and teleseismic Kanazawa (2011). Joint inversion of strong motion, teleseismic, geodetic,
constraints, Bull. Seismol. Soc. Am. 79, 655–669. and tsunami datasets for the rupture process of the 2011 Tohoku earth-
Mendoza, C., and S. Hartzell (1999). Fault-slip distribution of the 1995 Col- quake, Geophys. Res. Lett. 38, L00G21, doi: 10.1029/2011GL050098.
ima-Jalisco, Mexico, earthquake, Bull. Seismol. Soc. Am. 89, 1338–1344.
Mendoza, C., S. Hartzell, and T. Monfret (1994). Wide-band analysis of the
3 March 1985 central Chile earthquake: Overall source process and AECOM
rupture history, Bull. Seismol. Soc. Am. 84, 269–283. 915 Wilshire Boulevard
Murotani, S., H. Miyake, and K. Koketsu (2008). Scaling of characterized slip Los Angeles, California 90017
models for plate-boundary earthquakes, Earth Planets Space 60, 987–991. andreas.skarlatoudis@aecom.com
Murotani, S., K. Satake, and Y. Fujii (2013). Scaling relations of seismic paul.somerville@aecom.com
moment, rupture area, average slip, and asperity size for M ∼ 9 sub- hong.kie.thio@aecom.com
duction-zone earthquakes, Geophys. Res. Lett. 40, 1–5.
Nagai, R., M. Kikuchi, and Y. Yamanaka (2001). Comparative study on the Manuscript received 16 November 2015;
source processes of recurrent large earthquakes in Sanriku-Oki region: Published Online 31 May 2016

You might also like