• 沒有找到結果。

Robust determination of earthquake source parameters and mantle attenuation

N/A
N/A
Protected

Academic year: 2021

Share "Robust determination of earthquake source parameters and mantle attenuation"

Copied!
17
0
0

加載中.... (立即查看全文)

全文

(1)

Robust determination of earthquake source parameters

and mantle attenuation

Yen-Ting Ko,

1,2

Ban-Yuan Kuo,

1

and Shu-Huei Hung

2

Received 8 August 2011; revised 13 February 2012; accepted 14 February 2012; published 12 April 2012.

[1]

An improved inversion technique is needed to effectively separate the frequency

dependence of the source from the intrinsic attenuation of the medium. We developed a

cluster-event method (CEM) in which clusters of nearby events, instead of individual

events, pair with stations to form the basis for measurements of Q value and corner

frequency (f

c

). We assume that the raypaths from one cluster to a station share an identical

Q while each event in the same cluster is allowed for only one f

c

in the inversion process.

This approach largely reduces the degrees of freedom to achieve a robust inversion.

We use an optimization algorithm of simulated annealing to solve the nonlinear inverse

problem. The CEM was applied to events at 70–150 km depths in the Japan subduction

zone recorded by F-net. We show that the method proposed here leads to better constraints

on both source parameters and attenuation. The resultant Q’s in the mantle wedge

increase from lower than 300 beneath the arc and back-arc to greater than 600 in the

fore-arc region. The f

c

’s satisfy a self-similar scaling relationship with seismic moment

of M

0

∝ f

c3

with a best fit stress drop of 21.9

 6.9 MPa in Madariaga’s form.

This contrasts to the stress drop of 1.4

 1.1 MPa for a global data set composed of

prior measurements for crustal events. The results of this study agree with results from

previous studies, except with an upward deviation due to higher corner frequencies

and stress drops.

Citation: Ko, Y.-T., B.-Y. Kuo, and S.-H. Hung (2012), Robust determination of earthquake source parameters and mantle attenuation, J. Geophys. Res., 117, B04304, doi:10.1029/2011JB008759.

1.

Introduction

[2] Study of attenuation involves measuring the spectral content of a waveform, which is more difficult than mea-suring the traveltime. A broadband power spectrum substan-tially above the noise level is required to extract the correct spectral characteristics of the seismic signals. The earthquake source has its own frequency dependence, sometimes char-acterized by a corner frequency fc, which convolves with

path effects to compose a seismogram. These path effects include focusing/defocusing and scattering during traveling blended with intrinsic attenuation to shape the spectrum at a seismic station.

[3] One strategy to discriminate the source and path effects is to eliminate either of the two with specific tools. Examining the ratio of the spectra recorded by two stations cancels the source contribution [e.g., Roth et al., 1999; Shito et al., 2004]. In studies of earthquake sources, the spectral ratio is usually calculated between 2 events to eliminate the path effects. Examples in this scheme are the empirical Green’s function (EGF) method if one event is small enough [e.g., Mueller, 1985; Mori and Frankel, 1990; Abercrombie

and Rice, 2005; Viegas et al., 2010] and the determination of 2 corner frequencies rather than the conventional 2 parameters for source and path [e.g., Tajima and Tajima, 2007]. Although the spectral ratio mitigates the tradeoff problem, a reference station or event must be chosen, which may limit the applicability of the method in mapping large-scale attenuation structure of a region.

[4] Other approaches tackle the tradeoff problem during or after the inversion of the spectrum. The fccan be obtained

by fitting individual spectra recorded at each station, but this leads to multiple fc’s for one event. Despite fc’s are in

prin-ciple path-dependent for rupturing of a finite fault, we show in this study that the multiple fc’s determined with paths are

primarily artifacts of inversion due to tradeoffs between parameters. One way to circumvent this is to average all the fc’s over stations for each event and fix it in the second

stage of the determination of Q [e.g., Wang et al., 2010]. However, a better remedy inversion-wise is to implement the single fcrequirement in the inversion process for the

one-event, many-station system, anticipating a reduced direct-ivity effect averaged over stations. This scheme was first used by Stachnik et al. [2004] and adopted in later studies [e.g., Rychert et al., 2008; Pozgay et al., 2009]. Additional con-straints of fcincluding the existing moment-fcrelationship can

be exercised to bracket the solution [e.g., Pozgay et al., 2009]. [5] Recently, Shearer et al. [2006] proposed stacking of event spectra within bins of similar seismic moment before

1Institute of Earth Sciences, Academia Sinica, Taipei, Taiwan. 2

Department of Geosciences, National Taiwan University, Taipei, Taiwan.

Copyright 2012 by the American Geophysical Union. 0148-0227/12/2011JB008759

(2)

computing the EGF. Mayeda et al. [2007] applied the spectral ratio method on S coda and demonstrated that the path effect can be canceled more efficiently than for direct seismic phases. These improved methods, however, are not designed to extract both source characteristics and intrinsic attenuation of the medium in one inversion. In this work, we present the cluster-event method (CEM) to simultaneously determine fcand Q from“cluster”-station pairs. In addition

to enforcing fcas the single-valued parameter for each event,

the CEM inversion scheme groups raypaths based on event “clusters” to reduce the degrees of freedom of the inversion and suppress the tradeoff effects. In this study we focus on demonstration and evaluation of the method in the Japan subduction zone. A better determination of fcalso allows us

to re-examine the fc-moment scaling problem.

2.

The Issue of Tradeoff

[6] We constructed synthetic spectra to examine the contri-butions of different parameters to a spectrum. A displacement

spectrum A( f ) with a Brune-type source model [Brune, 1970] can be written as [e.g., Anderson and Hough, 1984]

Að f Þ ¼W0expðpf t*Þ 1þ ff c  2 : ð1Þ where W0¼ C⋅M0; C ¼ cUfq 4prvp3R; ð2Þ t* ¼ Z path dt Q: ð3Þ

In equations (1) and (2), f is the frequency andW0is the

low-frequency asymptote of the spectrum related to seismic moment M0 through a frequency independent constant C,

which is composed of vp, the P wave velocity near the source,

r, the density of the source material, R, the source-receiver Figure 1. Illustration of the tradeoff between t* and fcwith synthetic spectra computed from equation (1).

Both parameters act to bend the spectral amplitude with increasing frequency. Example spectral curves are for (a) t* = 0.01 and fc= 1.5, 3, 6, and 10 Hz and for (b) fc= 10 Hz, and t* = 0.2, 0.1, 0.07, and 0.03.

(c) Synthetic spectra for (0.08, 4.0) (black line, with random error) and (0.13, 6.7) (green) are indistinguish-able. The former is taken as data for the grid-search in Figure 1d. (d) The misfit map shows a curve-shaped low misfit valley typical of fitting the spectrum with t* and fc. The minimum is located at (0.13, 6.7) (green

(3)

distance, andcUfq, the spherical average of radiation patterns

(0.52 for P) [e.g., Aki and Richards, 1980; Abercrombie, 1995]. The exponential term in equation (1) describes the along-path attenuation, in which t* is the attenuation opera-tor defined in the path integral in equation (3). M0 is a

constant parameter that does not tradeoff with t* and fc

sig-nificantly and can be determined independently as described in section 6.3.

[7] High-temperature laboratory experiments [Jackson et al., 2002] as well as seismic attenuation study in mantle wedge [Stachnik et al., 2004] suggested that the attenuation operator is frequency-dependent in the form of

t * ¼ t0* fa; ð4Þ

where t0* is the t* value at 1 Hz, and a accounts for the frequency dependence of intrinsic attenuation. In this study, a is fixed at an experimental value of 0.27 [Jackson et al., 2002, 2006]. Inversion of a as a free parameter is usually non-robust, and a few studies have reported inversion results of a in the range of 0–0.4 with only marginal pre-ferences overa [e.g., Stachnik et al., 2004; Rychert et al., 2008; Pozgay et al., 2009].

[8] Figure 1 shows that, while t* and fcin principle shape

the spectral decay in different fashions, their effects can be similar and hard to distinguish in the presence of noise. In this example, the “data” is generated by blending the spectrum for t* = 0.08 and fc= 4.0, or (0.08, 4.0) with the

Gaussian random noise which mimics the scatter of data at high frequencies well. We perform a grid-search against this spectrum in the (t*, fc) space, yielding a misfit map

with a long, low-value valley, indicating a serious tradeoff between the parameters. Solutions along this valley explain the data equally well in a statistical sense, and therefore neither of the two can be determined in a robust way. The synthetic test outputs a best fit model of (0.13, 6.7), which is quite distant from the input (0.08, 4.0) (Figure 1d).

[9] Figure 2 elaborates the tradeoff with real data. We analyzed P waves recorded by 11 F-net stations for a sub-duction zone event in Japan. Fitting individual spectrum through a grid search yields 11 sets of (t*, fc) that scatter

from 0.01 to 0.31 for t* and from 1 to 8 Hz for fc. The

uncertainty interval is characterized by a misfit value 5% larger than the minimum [Sonley and Abercrombie, 2006]. The 2 extreme Q values, 89 and 2645, were determined for 2 stations, HSS and IMG, respectively, which are both located in the back-arc environment of the subduction sys-tem and are closer to each other than most of the other pairs of stations. This exemplifies the tradeoff problem in the sense that the 2 parameters are highly mutually substitutable in minimizing the misfit. Taking the average of the scattered measurements as the representative fcof the event may obviate

the over-parameterization problem [e.g., Wang et al., 2010], but a more fundamental treatment should be sought to secure a correct recovery of the physical property of the earthquake. Figure 2. Illustration of wide-spread, distinct fc’s obtained for one event. Solutions (t*, fc) were

deter-mined by individually fitting the spectra recorded by 11 F-net stations for a Japan subduction zone event (2009/12/29, depth 80 km). Red triangles are arc volcanoes in this region. Stations IMG and HSS are back-arc affiliated and spatially close, but their resolved Q values differ significantly. The uncertainty of the solution (thin contour) illustrates the tradeoff between the two parameters.

(4)

Figure 3. Test of inversion using equation (5), Stachnik et al. ’s [2004] approach. Each point at the curve is the misfit value for a given fc and the Q value found to minimize the misfit at that fc . Black arrow points out the numerical opti-mal solution in each case. (a) 10 synthetic curves without random errors for fc = 3 Hz and t* evenly distributed between 0.01 and 0.1. The input fc is correctly recovered. (b) The same set of curves as in Figure 3a but with random errors. Location of minimum is ambiguous and statistically insignificant. (c) Inversion for the real data in Figure 2 provides a one-sided constraint with a numerical minimum at the frequency upper limit.

(5)

[10] Stachnik et al. [2004, hereinafter ST] formulated a one-event, many station system in which fcis allowed for only

a single value. Equation (1) can be rearranged into the form

ln Að fð ÞÞ þ ln 1 1þ f fc  2 0 B @ 1 C A ¼ pf t* þ ln Wð Þ:0 ð5Þ

[11] For each fixed fc, equation (5) is linear for t* and ln

(W0), enabling a linearized inversion for Q and W0for all

paths. The best fc can be found to give the“global”

mini-mum misfit for all trial fc’s. The linearized equation (5) has

also been a common tool in the investigations whose major concern is earthquake properties rather than Q structures [e. g., Anderson and Humphrey, 1991; Ide et al., 2003]. We performed a synthetic test simulating a case in which one event with fc= 3 Hz is recorded by 10 stations giving rise to

10 values of t*. Solving equation (5) for a range of trial fc’s

locates the minimum misfit at fc = 3 Hz successfully

(Figure 3a). However, this minimum is by nature non-robust, and in the presence of random errors, can be ambig-uous or even absent as the misfit curve turns to an L-shape (Figure 3b). We applied the ST method to the real data analyzed above (Figure 2), and obtained a loose lower bound for fcand an open solution at higher frequencies (Figure 3c).

3.

Methodologies and Data Analysis

3.1. Determining thet* Values and Source Parameters [12] Judging from the exercises in Figure 3, merely imposing single fcdoes not guarantee a robust solution. In

this section, we describe the approach used in the cluster-event method (CEM) which we show achieves better con-straints on both source and path parameters for events in a seismically active region.

[13] Figure 4 summarizes the basic idea of the CEM. We regard the nearby events (events 1–5) as a cluster, and assume that all the raypaths from a cluster to the same station are close enough to share an identical Q while each event in the cluster, usually recorded by several stations, is allowed only for one single fc. Assuming K events and

N stations, the degrees of freedom for CEM, ST, and indi-vidual fitting are K + N, (K  N) + K, and (K  N)  2, respectively. In the example in Figure 4, there are only 8 parameters (3 Q’s and 5 fc’s) for cluster I to be determined

to fit the spectral falloff (other than the constant W0). The

ST approach requires 20 parameters, while fitting individual spectra at all stations involves 30 unknowns (a pair of Q and fc for 15 paths). The CEM gains robustness by

effi-ciently reducing the degrees of freedom of the inverse problem while obey the basic physic of earthquake sources under the point source assumption. The CEM comes not without a cost, which is a sacrifice in spatial resolution of Q when bundling many otherwise independent, path-average Q’s into one. This cost, however, becomes transparent as the spatial size of the cluster decreases and the ray-paths converge. Meanwhile, as demonstrated above, attach-ing significance to each along-path Q seems impractical or even misleading.

[14] Following equation (1), we formulate a misfit func-tion in L2 norm Q ¼XM k¼1













Dkjð f Þ  W0kjexp pf tj*   1þ ff c k  2 2 6 4 3 7 5⋅8kj













; 8kj¼ Dfkj PM k¼1 PN j¼1Dfkj ; ð6Þ

where Dkjis the observed spectrum for kth event within a

cluster recorded by jth station. To weight more strongly the contribution from the spectrum with a longer frequency range D( f ), a weighting factor8kjwhich is proportional to

D( f ) is applied to each spectrum in the calculation of the total cluster misfit. Equation (6) is cluster-based, so cluster index is omitted. Note that Q in t* along the cluster-station path is the attenuation parameter to be solved for.Q measures the deviation of the theoretical spectrum from the observed spectrum for each cluster. An optimization algorithm of simulated annealing (SA) [Kirkpatrick et al., 1983], a generic probabilistic metaheuristic for the global optimi-zation problem, is employed to seek global minimum of the misfit function in a multidimensional parameter space. A grid search can achieve the same goal but much ineffi-ciently. For more detailed descriptions of the SA method see Aarts and Korst [1989].

3.2. Synthetic Tests

[15] To compare the solutions obtained from CEM and the other approaches, we carried out a series of synthetic tests with 3 stations and 3 events. We present an example with t* = 0.03, 0.05 and 0.08, and fc= 2, 4 and 6 Hz in Figure 5.

Figure 4. Synopsis of the CEM. Nearby events (events 1–5) are grouped into a cluster and the raypaths from this cluster to the same station share the same Q while each event in this cluster is assigned only one fc.The single fcconstraint applies

only to events in the same cluster. Clusters may overlap and share events, e.g., cluster I and II share events 1 and 5, each of which may have different values determined for different clusters. These cross-cluster events provide data to examine how robustly each fcis determined.

(6)

The synthetic data are theoretical spectra with random errors (Figure 5a), and the synopsis of the three approaches under test are (Figure 5b) (1) The CEM with 6 parameters (3 Q and 3 fc) for spectral falloff; (2) The ST approach with

12 parameters (9 for Q and 3 for fc); and (3) Fitting

indi-vidual spectra to solve for 18 parameters (9 pairs of Q and fc).

[16] Figure 5c shows that the CEM returns a solution that is closest to the input model, demonstrating a strong

resolvability despite data shrouded in noise. In contrast, the other 2 approaches yield parameters deviating from the input, especially for the highly over-parameterized individ-ual spectral fitting. This occurs as a general phenomenon in a series of tests with wide choices of input combinations. In some cases, when different sets of t* and fcgive similar

shape of spectral decay, it is the noise that determines the numerical best fit. A similar conclusion was reached by Sonley and Abercrombie [2006] that a major tradeoff occurs Figure 5. Test of CEM and the two other approaches. (a) Synthetic data with t* and fc annotated.

(b) Configurations of the 3 approaches with increasing degrees of freedom: CEM, ST method, and individ-ual fitting. (c) The misfit maps for synthetic models (0.03, 4), (0.05, 4), and (0.08, 4), respectively, from left to right. The input models are denoted by yellow stars. The solutions from CEM, ST, and individual fitting are denoted by white x, blue cross, and green dot, respectively. The CEM recovers the input t* and fc

(7)

Figure 6. The CEM is applied to events in Japan subduction zone. (left) Events (white circle) and F-net stations (blue tri-angle) used, active volcanoes (red triangle), and locations of vertical cross-sections (black line). (right) Cross-sections show the subducting Pacific slab (background events in black crosses), events used (red circle) with focal depths 70 –150 km, and volcanoes and stations on the surface. Each cross-section is 200 km wide.

(8)

between t* and fc. The CEM suppresses this likelihood with

much less degrees of freedom to satisfy all the spectrum data.

4.

Data

4.1. Event Selection

[17] The CEM performs best in high seismicity regions where clusters can be easily formed. We choose the Pacific plate portion of the Japan subduction zone to apply the method, and the F-net [Okada et al., 2004] waveforms, which have been archived since 1994, were used (Figure 6). In this study we use events deeper than 70 km to both deemphasize the crustal effects and thin the ray volume from cluster to station. 87% of the paths exceed 150 km in length. These paths travel a long distance in the mantle wedge, allowing us to map the attenuation structure there and simultaneously extract the source properties of intraslab

earthquakes. The final data set consists of 243 events, 17,443 spectra, and 123 clusters. About 87, 12, and less than 1% of the paths are from events with Mw 3.5–5, 5–6,

and >6, respectively. 4.2. Criteria of Clusters

[18] For each target event, neighboring events within a certain radius from the target event is considered of the same cluster. The smaller the radius, the closer the raypaths are to each other for one station and the better the shared Q assumption is satisfied, but the “density” of the cluster is lower. Because 87% of the path lengths are 150 km or lon-ger, we adopt a radius of 30 km in this study to ensure that a predominance of data have a path length 5 times longer than the radius of the cluster. This is a compromise between the number of clusters and the validity of the similar-Q assumption. We tested the radii of 10 and 20 km and found Figure 7. Example of CEM measurements and solutions. (a) Map of the 3-event cluster (circled by

dotted line) and the event illustrated as the example (red circle with focal mechanism; date 2008/03/ 03 and depth 83 km). Twelve stations (triangle) received good P waves and 6 of them (blue triangle) are shown in Figure 7c. (b) Solutions (open circle) from ST method (left), and CEM (right). The CEM curve is constructed by the SA procedure explained in the text. (c) Example waveforms and spectra. P wave windows are shaded. Each spectrum is compared with solutions from CEM (blue curve), Stachnik et al.’s [2004] method (green curve), and individual fitting (red curves), with solutions (Q, fc) annotated on

the lower-left corner of each spectrum panel. For individual fitting, the second solution is for the averaged fcof 5.6.

(9)

the results essentially the same as for the 30 km except many events are left isolated and abandoned. The region was scanned for each event with the 30 km radius to establish as many clusters as possible to maximize the coverage of the cluster-based raypaths. The density of cluster is 3–11 events with a majority exceeding 5. In this scheme, one event can belong to several overlapping clusters as illustrated in Figure 4. However, although the “one fc” requirement is

applied to each event in one cluster, the requirement is not enforced across clusters, leaving the cross-cluster events susceptible to multiple values of fc determined from

dif-ferent clusters. This, although can be improved in the future, provides a means to test the statistical robustness of the fc determined with CEM, which will be addressed in

section 5.2.

4.3. Waveform and Spectra

[19] The vertical components of the velocity recordings were deconvolved to remove the instrument response. Only those P waves without dominating coda energy were used to minimize the interference from scattering between source and receiver. While coda waves average out the heteroge-neity of the medium efficiently, we focus on body wave phases for which the physics of the intrinsic attenuation has long been established. Examples of waveforms and corresponding spectra are shown in Figure 7. P wave onsets were manually picked for traveltimes. The displacement spectrum is calculated from a 5 s window of waveform starting 0.5 s before the P arrival time with 0.5 s long cosine tapers on both ends. The noise spectrum is calculated from a window of the same length 1 s preceding the P wave window. The 5 s window limits the minim frequency resolvable at about 0.5 Hz. In our data set, the minimum fcis

0.7 Hz for an Mw5.5 event. We found that longer windows,

e.g., 5–10 s, produce spectra that do not yield distinguish-able results after inversion. This is consistent with Sonley and Abercrombie’s [2006] note that amplitude spectra are insensitive to the window length as long as it exceeds the expected pulse width of the target phase. The spectra were smoothed using a Hanning window. To avoid aliasing, the upper limit of the frequency band is set at 80% of the Nyquist frequency of the F-net data, or 8 Hz (Figure 7). Inversion is performed only for the frequency band with a signal-to-noise ratio greater than 5.

5.

Results

[20] Figure 7b illustrates a typical constraint on fc by

CEM. While the ST inversion usually provides a loose lower limit of fc, the CEM delivers a V-shape misfit curve to better

constrain this parameter. In practice, the highly nonlinear inversion converges to a curve like that in Figure 7b in a few steps, but we repeated the SA process several times with different initial guesses to secure a global minimum. We also applied the ST method to the same data set, and the com-parison in fc between the CEM and the ST is shown in

Appendix A, section A1. 5.1. Q Structure

[21] The path-average Q measurements for intraslab events mostly map the mantle wedge of the Japan subduction zone,

which potentially constitutes a data set for follow-up investigations such as tomographic imaging. For the pur-pose of this study we describe the overall pattern of Q distribution. The first order feature is low Q values, i.e., <300, for paths sampling the mantle mostly beneath the arc and back-arc, in contrast to the high Q values of greater than 600 for paths traversing the fore-arc mantle (Figure 8). This pattern agrees with previous observations in Japan [e.g., Tsumura et al., 2000] and other subduction zones [e.g., Roth et al., 1999; Schurr et al., 2003; Stachnik et al., 2004; Rychert et al., 2008]. Low Q values have been attributed to the high temperature and probably the presence of melt retained under the arc and back-arc. The high Q values in the fore-arc are thought to characterize the low temperature “nose” of the wedge corner initially induced by decoupling between the wedge corner and the subduction. Differences in hydroxyl content may also cause variations in attenua-tion (e.g., Y. T. Ko et al., The southwestern edge of the Ryukyu subduction zone: A high Q mantle wedge, sub-mitted to Earth and Planetary Science Letters, 2011), but viable interpretations will rely upon an integration of evi-dence from attenuation and velocity observations, arc geo-chemistry, deformation experiments, and dynamic modeling [e.g., Jackson et al., 2002; Karato, 2003; Chou et al., 2009; Lin et al., 2010].

[22] In addition to the first order back-arc–fore-arc varia-tion, along-strike variations appear to exist in the measured Q values. At the two southernmost profiles, which straddle the region of strong change in trench curvature and slab dip, Q in the fore-arc is low relative to at the northern profiles. Dynamic modeling demonstrates that lateral flow may be driven by the pressure gradient induced by trench curvature and slab dip [Kneller and van Keken, 2008]. Further south, rocks of adakitic composition were reported that suggest the presence of a high-temperature subduction environment [Sugimoto et al., 2006]. Detailed tomographic imaging will map the attenuation in 3D and shed light on the interpreta-tion of this along-strike feature.

5.2. The Determination of Corner Frequency

[23] As described above, the single-fc requirement is

implemented in the inversion for each cluster, but not across the cluster, i.e., different values of fcmay be obtained for

events shared by overlapping clusters. If the intracluster source parameter is resolved robustly, those cross-cluster fc’s

should converge. This presents an opportunity to appraise the method. We calculated the statistics of the fc’s

deter-mined by CEM and individual spectrum fitting (Figure 9). The number of fc’s from CEM for each event equals the

number of clusters containing this event, and the number of fc’s from individual fitting corresponds to the number

of stations.

[24] Overall, the scatter in fcfor each event is significantly

smaller for CEM than for individual fitting. We calculate the standard deviation of fc for each event and average of the

standard deviations for all events. The average is 0.56 Hz for CEM, compared with 3.84 Hz for individual fitting. Almost 90% of the events under test have standard devia-tions < 1.3 Hz, indicating a robust determination of fcwithin

each cluster. For each event, the mean fc’s from the two

(10)

Figure 8. Path-dependent Q (color-coded) derived from CEM (left) on the map view and (right) in cross-sections of Japan subduction zone. All the paths with Q > 600 appear to sample the fore-arc mantle, while those with Q < 300 travel to the arc and back-arc.

(11)

reaching 3 Hz or greater. This deviation is large enough to bias the interpretation of Q.

6.

Discussion

6.1. Assumptions and Errors

[25] In this study, we ignore site effects, which are used to parameterize near-station geology or topography and may be isolated from the path effect. However, extracting site effects require both extra parameterizations and the spectra sensitive to surface structures. Recent studies using high density array data and crustal events indicate that the frequency-dependent site amplification increases with

frequencies and is sensitive to near-station structures [e.g., Edwards et al., 2008; Edwards and Rietbrock, 2009]. We assume that for the path length mostly greater than 150 km in this study ignoring site effects does not bias the determination of the path effect.

[26] Errors may be introduced by the assumption of point source invoked to ignore source directivity. The directivity effect has been assumed to be averaged out either with a wide azimuthal coverage of stations in teleseismic studies [e.g., Allmann and Shearer, 2009] or through the diversity of focal mechanism in mantle wedge models [Stachnik et al., 2004]. This merely reflects the difficulty in identifying the fault plane of all the events used. In Appendix A, section A2 Figure 9. Comparison of fc’s derived from (top) CEM and (bottom) individual spectrum fitting. (a) Number

of fc’s corresponding to number of clusters sharing one event (see text). (b) Red dots are the means of the fc’s

obtained from CEM and the vertical bars are one standard deviations. (c and d) Same as Figures 9a and 9b, respectively, except for individual fitting. The small standard deviations of fcfor CEM, averaged to 0.56 Hz

(12)

numbers compared with the average of 0.56 Hz error in fc

derived from cross-cluster events in section 5.2. However, one notable feature is that these errors are much smaller than the errors stemming from the tradeoff when fc is freely

determined along with path, i.e., 3.84 Hz as reported in section 5.2, implying that fixing fcin the inversion comes

with a smaller cost.

[27] Another first order feature is that the directivity effect is proportional to fc, and thus it cannot be ruled out through

event selection [Wang and Rubin, 2011]. However, the actual error must be smaller than that presented here, because (1) narrow pulses are susceptible to attenuation more strongly than wide pulses, naturally suppressing the most severe bias from upward rupture; and (2) the enhanced robustness of the CEM, along with the multievents in each cluster, likely average out large errors associated with indi-vidual events. Nonetheless, directivity effects cannot be completely eliminated and may ultimately contribute to the errors in both fcand Q.

[28] The assumption ofa = 0.27 in equation (4) could also introduce errors. To evaluate this assumption, we repeated the synthetic experiments in Figure 5 witha = 0 for the input synthetic spectra and inverted them witha = 0.1 and 0.27. We found that the maximum error in fc is <0.5 Hz. As

described in section 2, it is difficult to invert for a as an additional independent parameter. We adopt the value widely used in studies of mantle wedge [e.g., Stachnik et al. 2004; Rychert et al., 2008; Pozgay et al., 2009]. The dif-ference in Q for different a is systematic and the relative pattern remains consistent.

6.2. Estimation of Scalar Moment

[29] Other than the spectral falloff which is characterized by (t*, fc), the base-line level of the low-frequency spectral

amplitude, or W0, is another source parameter to be

deter-mined. The seismic moment is related toW0via equation (2)

[Brune, 1970]. In this work, vpand r were assumed to be

8 km/s and 3.3 g/cm3, respectively, for events at 100 km depth based on the Preliminary Reference Earth Model (PREM) [Dziewonski and Anderson, 1981].

[30] We calculate seismic moments for all stations and averaged them for each event. This estimated average seismic moment,M0, is compared with the cataloged M0provided by

NIED CMT solutions. The two sets of seismic moment are highly correlated with a correlation coefficient of 0.98 and the best fit slop of 0.95. The small discrepancy between M0 and M0may be attributed to several factors. The catalog M0

was determined with surface waves in bandwidths 20–100 s, while the bandwidth in our spectrum analysis is predomi-nantly 0.5–8 Hz. In addition, equation (2) employs sim-plified assumptions in radiation pattern and geometrical spreading.

where vsis the near-source shear wave velocity and 0.32 is

the coefficient for P wave. The stress dropDs can be esti-mated from the seismic moment and the source radius [Eshelby, 1957]

Ds ¼7M0 16 rc3

: ð8Þ

Using vs= 4.5 km/s as the near-source velocity for100 km

depth events, the estimated stress drops spread mostly over 1–100 MPa with a median at 22.5 MPa (Figure 10a).

[32] Substituting (7) into (8) to eliminate rc, the relation

between corner frequency and seismic moment reads

fc¼ 0:42vs Ds M0  1=q:

: ð9Þ

[33] We fit equation (9) to the measured fcin this study

and the cataloged M0and obtained a best fitDs of 21.9 

6.9 MPa (one standard error) with a variance reduction of 86%. In equation (9) the exponent 1/q with q explicitly set at 3 is consistent with a self-similarity scenario. If q is allowed to change with stress drops bracketed between 1 and 100 MPa, the best fit q is 3.1 with a variance reduction of 85%. As the deviation from 3 is statistically insignificant, we infer that the data are compatible with the theory, implying a self-similarity for intraslab earthquakes at100 km depths over Mw3.5–6 or 4 orders of magnitude in seismic moment.

[34] A number of studies have verified self-similarity of earthquake across a certain range of magnitude [e.g., Ide et al., 2003; Prieto et al., 2004]. Others question it as a property for all earthquake sizes [e.g., Tajima and Tajima, 2007; Mayeda et al., 2007]. Allmann and Shearer [2009] compiled a data set consisting of most measurements up-to-date (referred to as “global” data set) and suggested the presence of self-similarity over 13 orders of magnitude in seismic moment. We modified this data set by converting some of the fc’s that were originally measured from S-wave

spectra to the P-wave fc’s by multiplying a factor of 1.5 [e.g.,

Madariaga, 1976; Pozgay et al., 2009]. In this global data set, events are shallower than 60 km depths and skewed toward shallow crust above 20 km. For vs = 3.9 km/s

representing the upper crust [Allmann and Shearer, 2009], the median stress drop is 3.3 MPa. We also fit equation (10) to these data and obtained a best fitDs = 1.4  1.1 MPa and a variance reduction of 81%. Our data merge concordantly with this data set except with higher corner frequencies and higher stress drops (Figure 10b).

[35] Because biases from different methodologies exist, a cross-study comparison at present should be perceived as a proof of concept exercise before a comprehensive, equally based comparison is conducted in the future. We compare the best fit Ds constrained by the similarity relationship

(13)

Figure 10. (a) Measured fc of this study against M0 (blue dots) fitted by the theoretical relationship (blue thick line) with the best fit D s of 21.9  6.9 MPa based on Madariaga ’s [1976] source model. The best fit HW model with D s of 2.5  0.8 MPa occupies nearly the same line. Background dotted lines represent D s of 0.1, 1, 10, and 100 MPa for Madariaga [1976] (blue) and HW (golden). Inset is the histogram for D s with the same color code. (b) Merging of our data with a global data set (sources shown in the lower-left) over much wider data ranges. The blue line is the same as in Figure 10a, and the gray line correspond to best fit theoretical relationships to the global data with D s of 1.4  1.1 MPa for Madariaga [1976] and 0.25  0.1 MPa for HW. The D s annotating the lines follow the same color code. In the global data set, the higher values of Mori and Frankel [1990] (yellow open square) are estimated from the HW model (see text). The difference in D s between this study and the global data set is much larger for Madariaga ’s [1976] model than for HW although it is statistically significant for both.

(14)

equation (9) rather than the statistics directly fromDs esti-mates through equations (7) and (8), although the conclu-sions would be similar. We tested the null hypothesis that the two stress drop estimates, 21.9  6.9 MPa and 1.4  1.1 MPa, are the same, and found that the null hypothesis can be rejected at the 99% confidence level by the t-test. The alternative hypothesis that the two estimates are different implies that the faults of the intraslab earthquakes analyzed in this study break at higher stress level than that in the shallow crust. Although not for intraslab events, the depth-dependence ofDs for crustal events has been documented. Hardebeck and Aron [2009] reported a dramatic increase of Ds from 5 MPa near the surface to 50 MPa in crust deeper than 13 km in the vicinity of the Hayward fault, California. Oth et al. [2010] showed medianDs of 1.1 and 9.2 MPa for events shallower and deeper than 30 km, respectively, from

an analysis of borehole acceleration spectra in Japan. Sta-tistics were not explicitly given in these studies. The latter used the source formulation of Hanks and Wyss [1972] that entails smaller stress drops as described below.

[36] Another source dimension formulation commonly employed was proposed by Hanks and Wyss [1972, herein-after HW]:

rc¼ 2:34vp

2pfc ; ð10Þ

which is an extension from Brune’s [1970, 1971] original rupture model for S wave. The larger product of near-source velocity vpand the constant (0.37) in equation (10) than in

equation (7) renders a larger source dimension and con-sequently a smaller Ds than for Madariaga [1976]. The Figure A1. (a) Comparison in fcbetween the CEM (red dot) and the ST method (blue dot) for all events

used in this study. The shaded vertical lines denote the uncertainty range for the fcdetermined by the ST

method. (b and c) Examples of misfit curves for the ST method. In Figure A1b, for an Mw 4.1 event, the best fitting fcis tied at 1.15 Hz, in contrast to the fcof 7.9 Hz by the CEM. Figure A1c shows the typical

misfit curve with the minimum at the upper limit of the frequency range. The CEM yields an fcof about

(15)

corresponding fc – M0 theoretical relationship takes dif-ferent constants: fc¼ 0:49vp Ds M0  1=3: : ð11Þ

[37] Fitting equation (11) to our data yields aDs of 2.5  0.7 MPa for vp = 8 km/s. To apply the HW model to the

global data set, we convert the rc’s reported by Mori and

Frankel [1990] to fc’s using equation (10). The fitting to

the global data givesDs = 0.25  0.1 MPa with vp= 6 km/s.

Both fitting reduce the data variance by about 85%. The gap in Ds is now substantially narrowed from that for Madariaga [1976] and the hypothesis that the two estimates are equal cannot be rejected at any high confidence level.

[38] The HW model reconciles the apparent depth-dependent fc by the more depth-sensitive vp and a larger

constant, together suppressing the variation in Ds. This effect has not caught much attention because so far most investigations have targeted crustal events for which a con-stant near-source velocity usually suffices. Both Madariaga’s [1976] and the HW models are appealing in source studies, and distinguishing which quantifies the fault strength more appropriately is beyond the current study. It is implied that a conclusion on whether intermediate-depth, intraslab earth-quakes are “stronger” than shallow crustal events awaits not only more sampling of deep events and more vigorous analysis of measurement errors but also an appraisal of source models.

7.

Conclusions

[39] The cluster-event method uses an improved inversion scheme to simultaneously determine the corner frequency

of the source and the attenuation of the medium. We dem-onstrate that distinguishing between these two parameters is difficult in the presence of noise unless the number of unknowns is sufficiently reduced. The CEM bundles nearby events into cluster and solve for Q on a cluster-station basis, while imposing single-fcto each event of the cluster,

largely reducing the degrees of freedom of the inversion problem. The cross-cluster comparison in fc proves that

the CEM yields a robust constraint on fc, which assures

equivalently a reliable determination of Q. Applying CEM to the Japan subduction zone depicts a characteristic varia-tion in Q from mostly lower than 300 in the arc and back-arc to >600 in the fore-back-arc. This will form a basis for detailed mapping of subduction zone structures. The resolved fc’s

fit well the relationship with seismic moment in the form of M0 ∝ fc3, suggesting self-similarity of

intermediate-depth subduction zone events. We report stress drop esti-mated from fc for our data and a data set composed of

prior measurements. We found that the stress drops of this study are significantly higher than previously estimated for crustal events based on Madariaga’s [1976] formula-tion, while the difference is much suppressed if Hanks and Wyss’s [1972] model is used. We consider the results shown here as motivation for studies of earthquake sources at greater depths of subduction zones and in different tec-tonic settings.

Appendix A

A1. ST Method

[40] We applied the ST method to the same data set and show the resultant fcin Figure A1. There are some fc’s that

are constrained at a local valley of the misfit curve, although the solution could be significantly different from that of Figure A2. Plots of fc versus takeoff angle q (from rupture direction). The lengths of the fault are

(a) 5 and (b) 2 km. Thin curves are apparent fcdue to directivity. Dotted lines are the“true” fcwithout

directivity observed atq = 90o. Thick lines are average fc’s over the respective q ranges for cases A, B,

and C. The difference in fcbetweenq = 0oand the dotted line (forq = 90o) characterizes the maximum

(16)

rupture durationtais determined by ta¼ L vr  L cosq vp ðA1Þ

where vpis set at 8 km/s, vris the rupture velocity assumed

to be 0.8vs with vs 4.5 km/s, and q is the angle between

the rupture direction and the takeoff vector. For simplicity, we assume a 2 dimensional problem with the source and the stations in a vertical section to which the fault plane is per-pendicular and that all the stations are above the source with raypaths approximated by straight lines. Half of the ta

multiplied by vris assumed as the radius of the circular fault

[Boatwright, 1980; Mori and Frankel, 1990],

rc¼L2 1vr cosq vp

ðA2Þ

which is then converted to fcusing equation (7). (Equation A2)

implies that directivity effect gives rise to variable apparent fault size and fc. Figure A2 shows fcas a function ofq. We

assume that the stations form a 120 wide observation aperture above the source and that the average fcover this

aperture is the CEM solution. We tested 3 types of fault geometry. A: Horizontal fault plane rupturing toward right. In this case, the stations are located in the q range of (30–150). B: Vertical fault plane rupturing downward with the station q range of (120–240). C: Vertical fault plane rupturing upward with the stationq range of (300–60). The true fc of the source can be obtained at q = 90 where

directivity is absent. For L = 5 km the deviations of the measured fcfrom the true fcare 0.04,0.15, and 0.35 Hz for

case A, B, and C, respectively. For L = 2 km, the deviations for the 3 cases are 0.1,0.4, and 0.9 Hz, respectively. The fact that the directivity effect in fc is magnified at smaller

events can be illustrated by comparing cases at q = 0 and 90, i.e., fc(q = 0) ∝ (2/L)/(1  vr/vp), and fc(q = 90) ∝ (2/L).

The difference between them is the maximum directivity effect

dfc∝ 2 L vr vp O 2 ðA3Þ

where O2represents a second order term much smaller than (vr/vp). It is noted that not only fcitself but also the magnitude

of the directivity effect in fcare inversely proportional to the

fault size.

[42] Acknowledgments. The authors thank Rachel E. Abercrombie and an anonymous reviewer for their critical questions and constructive comments, which much improved the revised manuscript. The broadband seismic records were provided by F-net operated by National Research Institute for Earth Science. The work was supported by National Science Council of Taiwan under grant NSC 100-2627-E-001-001.

Aki, K., and P. G. Richards (1980), Quantitative Seismology, 932 pp., W. H. Freeman, New York.

Allmann, B. P., and P. M. Shearer (2009), Global variations of stress drop for moderate to large earthquakes, J. Geophys. Res., 114, B01310, doi:10.1029/2008JB005821.

Anderson, J. G., and S. E. Hough (1984), A model for the shape of the Fourier amplitude spectrum of acceleration at high frequencies, Bull. Seismol. Soc. Am., 74(5), 1969–1993.

Anderson, J. G., and J. R. Humphrey Jr. (1991), A least squares method for objective determination of earthquake source parameters, Seismol. Res. Lett., 62, 201–209.

Boatwright, J. (1980), A spectral theory for circular seismic sources; simple estimates of source dimension, dynamic stress drop, and radiated seismic energy, Bull. Seismol. Soc. Am., 70, 1–27.

Brune, J. N. (1970), Tectonic stress and the spectra of seismic shear waves from earthquakes, J. Geophys. Res., 75, 4997–5009, doi:10.1029/ JB075i026p04997.

Brune, J. N. (1971), Correction to “Tectonic stress and the spectra of seismic shear waves from earthquakes,” J. Geophys. Res., 76, 5002. Chou, H.-C., B.-Y. Kuo, L.-Y. Chiao, D. Zhao, and S.-H. Hung (2009),

Tomography of the westernmost Ryukyu subduction zone and the ser-pentinization of the fore-arc mantle, J. Geophys. Res., 114, B12301, doi:10.1029/2008JB006192.

Dziewonski, A. M., and D. L. Anderson (1981), Preliminary reference Earth model, Phys. Earth Planet. Inter., 25, 297–356, doi:10.1016/ 0031-9201(81)90046-7.

Edwards, B., and A. Rietbrock (2009), A comparative study on attenuation and source-scaling relations in the Kantō, Tokai, and Chubu region of Japan, using data from Hi-Net and KiK-Net, Bull. Seismol. Soc. Am., 99, 2435–2460, doi:10.1785/0120080292.

Edwards, B., A. Rietbrock, J. J. Bommmer, and B. Baptie (2008), The acquisition of source, path, and site effects from microearthquake record-ings using Q tomography: Application to the United Kingdom, Bull. Seismol. Soc. Am., 98, 1915–1935, doi:10.1785/0120070127.

Eshelby, J. D. (1957), The determination of the elastic field of an ellip-soidal inclusion, and related problems, Proc. R. Soc. London, Ser. A, 241, 376–396, doi:10.1098/rspa.1957.0133.

Geller, R. J. (1976), Scaling relations for earthquake source parameters and magnitudes, Bull. Seismol. Soc. Am., 66, 1501–1523.

Hanks, T. C., and M. Wyss (1972), The use of body-wave spectra in the determination of seismic-source parameters, Bull. Seismol. Soc. Am., 62, 561–589.

Hardebeck, J. L., and A. Aron (2009), Earthquake stress drops and inferred fault strength on the Hayward fault, east San Francisco Bay, California, Bull. Seismol. Soc. Am., 99, 1801–1814, doi:10.1785/0120080242. Ide, S., G. C. Beroza, S. G. Prejean, and W. L. Ellsworth (2003), Apparent

break in earthquake scaling due to path and site effects on deep bore-hole recordings, J. Geophys. Res., 108(B5), 2271, doi:10.1029/ 2001JB001617.

Jackson, I., J. D. Fitz Gerald, U. H. Faul, and B. H. Tan (2002), Grain-size-sensitive seismic wave attenuation in polycrystalline olivine, J. Geophys. Res., 107(B12), 2360, doi:10.1029/2001JB001225.

Jackson, I., U. H. Faul, J. D. Fitz Gerald, and S. J. S. Morris (2006), Contrasting viscoelastic behavior of melt-free and melt-bearing olivine: Implications for the nature of grain-boundary sliding, Mater. Sci. Eng. A, 442, 170–174, doi:10.1016/j.msea.2006.01.136.

Karato, S. (2003), Mapping water content in the upper mantle, in Inside the Subduction Factory, Geophys. Monogr. Ser, vol. 138, edited by J. Eiler, pp. 135–152, AGU, Washington, D. C., doi:10.1029/138GM08. Kirkpatrick, S., C. D. Gelatt Jr., and M. P. Vecchi (1983), Optimization

by simulated annealing, Science, 220(4598), 671–680, doi:10.1126/ science.220.4598.671.

Kneller, E. A., and P. E. van Keken (2008), Effect of three-dimensional slab geometry on deformation in the mantle wedge: Implications for shear wave anisotropy, Geochem. Geophys. Geosyst., 9, Q01003, doi:10.1029/ 2007GC001677.

Lin, S.-C., B.-Y. Kuo, and S.-L. Chung (2010), Thermomechanical models for the dynamics and melting processes in the Mariana subduction system, J. Geophys. Res., 115, B12403, doi:10.1029/2010JB007658.

(17)

Madariaga, R. (1976), Dynamics of an expanding circular fault, Bull. Seismol. Soc. Am., 66, 639–666.

Mayeda, K., L. Malagnini, and W. R. Walter (2007), A new spectral ratio method using narrow band coda envelopes: Evidence for non-self-similarity in the Hector Mine sequence, Geophys. Res. Lett., 34, L11303, doi:10.1029/2007GL030041.

Mori, J., and A. Frankel (1990), Source parameters for small events associ-ated with the 1986 North Palm Springs, California, earthquake deter-mined using empirical Green functions, Bull. Seismol. Soc. Am., 80, 278–295.

Mueller, C. S. (1985), Source pulse enhancement by deconvolution of an empirical Green’s function, Geophys. Res. Lett., 12, 33–36, doi:10.1029/GL012i001p00033.

Okada, Y., K. Kasahara, S. Hori, K. Obara, S. Sekiguchi, H. Fujiwara, and A. Yamamoto (2004), Recent progress of seismic observation networks in Japan Hi-net, F-net, K-NET and KiK-net, Earth Planets Space, 56, xv–xxviii.

Oth, A., D. Bindi, S. Parolai, and D. Di Giacomo (2010), Earthquake scaling characteristics and the scale-(in)dependence of seismic energy-to-moment ratio: Insights from KiK-net data in Japan, Geophys. Res. Lett., 37, L19304, doi:10.1029/2010GL044572.

Pozgay, S. H., D. A. Wiens, J. A. Conder, H. Shiobara, and H. Sugioka (2009), Seismic attenuation tomography of the Mariana subduction system: Implications for thermal structure, volatile distribution, and slow spreading dynamics, Geochem. Geophys. Geosyst., 10, Q04X05, doi:10.1029/2008GC002313.

Prieto, G. A., P. M. Shearer, F. L. Vernon, and D. Kilb (2004), Earthquake source scaling and self-similarity estimation from stacking P and S spectra, J. Geophys. Res., 109, B08310, doi:10.1029/2004JB003084. Roth, E. G., D. A. Wiens, L. M. Dorman, J. Hildebrand, and S. C. Webb

(1999), Seismic attenuation tomography of the Tonga-Fiji region using phase pair methods, J. Geophys. Res., 104, 4795–4809, doi:10.1029/ 1998JB900052.

Rychert, C. A., K. M. Fischer, G. A. Abers, T. Plank, E. Syracuse, J. M. Protti, V. Gonzalez, and W. Strauch (2008), Strong along-arc variations in attenuation in the mantle wedge beneath Costa Rica and Nicaragua, Geochem. Geophys. Geosyst., 9, Q10S10, doi:10.1029/2008GC002040. Schurr, B., G. Asch, A. Rietbrock, R. Trumbull, and C. Haberland (2003),

Complex patterns of fluid and melt transport in the central Andean subduction zone revealed by attenuation tomography, Earth Planet. Sci. Lett., 215, 105–119, doi:10.1016/S0012-821X(03)00441-2.

Shearer, P. M., G. A. Prieto, and E. Hauksson (2006), Comprehensive analysis of earthquake source spectra in Southern California, J. Geophys. Res., 111, B06303, doi:10.1029/2005JB003979.

Shito, A., S. Karato, and J. Park (2004), Frequency dependence of Q in Earth’s upper mantle inferred from continuous spectra of body waves, Geophys. Res. Lett., 31, L12603, doi:10.1029/2004GL019582. Sonley, E., and R. E. Abercrombie (2006), Effects of methods of

attenua-tion correcattenua-tion on source parameter determinaattenua-tion, in Earthquakes: Radiated Energy and the Physics of Faulting, Geophys. Monogr. Ser, vol. 170, edited by R. E. Abercrombie et al., pp. 91–97, AGU, Washington, D. C., doi:10.1029/170GM11.

Stachnik, J. C., G. A. Abers, and D. H. Christensen (2004), Seismic atten-uation and mantle wedge temperatures in the Alaska subduction zone, J. Geophys. Res., 109, B10304, doi:10.1029/2004JB003018.

Sugimoto, T., T. Shibata, M. Yoshikawa, and K. Takemura (2006), Sr-Nd-Pb isotopic and major and trace element compositions of the Yufu-Tsurumi volcanic rocks: Implications for the magma genesis of the Yufu-Tsurumi volcanoes, northeast Kyushu, Japan, J. Mineral. Petrol. Sci., 101, 270–275, doi:10.2465/jmps.101.270.

Tajima, R., and F. Tajima (2007), Seismic scaling relations and aftershock activity from the sequences of the 2004 mid Niigata and the 2005 west off Fukuoka earthquakes (Mw6.6) in Japan, J. Geophys. Res., 112, B10302,

doi:10.1029/2007JB004941.

Tsumura, N., S. Matsumoto, S. Horiuchi, and A. Hasegawa (2000), Three-dimensional attenuation structure beneath the northeastern Japan arc esti-mated from spectra of small earthquakes, Tectonophysics, 319, 241–260, doi:10.1016/S0040-1951(99)00297-8.

Viegas, G., R. E. Abercrombie, and W.-Y. Kim (2010), The 2002 M5 Au Sable Forks, NY, earthquake sequence: Source scaling relationships and energy budget, J. Geophys. Res., 115, B07310, doi:10.1029/ 2009JB006799.

Wang, E., and A. M. Rubin (2011), Rupture directivity of microearthquakes on the San Andreas Fault from spectral ratio inversion, Geophys. J. Int., 186, 852–866, doi:10.1111/j.1365-246X.2011.05087.x.

Wang, Y.-J., K.-F. Ma, F. Mouthereau, and D. Eberhart-Phillips (2010), Three-dimensional Qp- and Qs-tomography beneath Taiwan orogenic belt: Implications for tectonic and thermal structure, Geophys. J. Int., 180, 891–910, doi:10.1111/j.1365-246X.2009.04459.x.

S.-H. Hung, Department of Geosciences, National Taiwan University, 1 Roosevelt Rd., Sec. 4, Taipei 106, Taiwan.

Y.-T. Ko and B.-Y. Kuo, Institute of Earth Sciences, Academia Sinica, 128 Academia Rd., Sec. 2, Taipei 115, Taiwan. (byk@earth.sinica.edu.tw)

數據

Figure 2. Illustration of wide-spread, distinct f c ’s obtained for one event. Solutions (t*, f c ) were deter- deter-mined by individually fitting the spectra recorded by 11 F-net stations for a Japan subduction zone event (2009/12/29, depth 80 km)
Figure 4. Synopsis of the CEM. Nearby events (events 1 –5) are grouped into a cluster and the raypaths from this cluster to the same station share the same Q while each event in this cluster is assigned only one f c

參考文獻

相關文件

The research proposes a data oriented approach for choosing the type of clustering algorithms and a new cluster validity index for choosing their input parameters.. The

We cannot exclude the presence of the SM Higgs boson below 127 GeV/c 2 because of a modest excess of events in the region. between 115 and 127

The case where all the ρ s are equal to identity shows that this is not true in general (in this case the irreducible representations are lines, and we have an infinity of ways

The remaining positions contain //the rest of the original array elements //the rest of the original array elements.

– Each listener may respond to a different kind of  event or multiple listeners might may respond to event, or multiple listeners might may respond to 

The development of IPv6 was to a large extent motivated by the concern that we are running out of 4-byte IPv4 address space.. And indeed we are: projections in- dicate that the

To decide the correspondence between different sets of fea- ture points and to consider the binary relationships of point pairs at the same time, we construct a graph for each set

For your reference, the following shows an alternative proof that is based on a combinatorial method... For each x ∈ S, we show that x contributes the same count to each side of