Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Click Here JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B07304, doi:10.1029/2009JB007109, 2010 for Full Article Multiple scattering and mode conversion revealed by an active seismic experiment at Asama volcano, Japan Mare Yamamoto1 and Haruo Sato1 Received 7 November 2009; revised 19 January 2010; accepted 17 February 2010; published 14 July 2010. [1] Volcanoes are one of the most heterogeneous fields in the Earth’s crust, and the understanding of such inhomogeneities may provide us important information on various volcanic processes. In the previous studies on the seismic wave propagation at active volcanoes, the diffusion model has been widely used to model the energy transportation and the contribution of P and S waves and their mode conversion have not been well recognized partly due to lack of dense observations capturing spatiotemporal pattern of energy propagation. In this study, we present observational evidence of mode conversions and multiple scattering at Asama volcano, Japan, revealed by an active seismic experiment with a dense seismic network. The observed spatial distribution of propagating energy emitted from the explosive P source shows a wavefront of direct P wave and a pattern exhibiting two slopes which is indicative of multiple scattering and conversion scattering of two modes having different scattering coefficients. These facts suggest that the radiative transfer theory is more preferable than the diffusion theory to explain the observed characteristics. We thus modeled the energy propagation using the radiative transfer theory assuming multiple isotropic scattering including conversion scatterings and quantitatively estimated scattering parameters. Estimated total scattering coefficients for P‐to‐S and S‐to‐S scattering are about three times larger than that of P‐to‐P scattering, and the mean free path of S wave is about 1 km for an 8–16 Hz band. These results suggest that the mode conversion and multiple scattering have an indispensable effect on the seismic energy propagation in heterogeneous volcanic environments. Citation: Yamamoto, M., and H. Sato (2010), Multiple scattering and mode conversion revealed by an active seismic experiment at Asama volcano, Japan, J. Geophys. Res., 115, B07304, doi:10.1029/2009JB007109. 1. Introduction [2] In active volcanoes, various kinds of seismic signals including volcanotectonic earthquakes, volcanic tremors, and long‐period events have been observed [e.g., Chouet, 2003; Kawakatsu and Yamamoto, 2007]. Volcanotectonic earthquakes are brittle failures occurring inside volcanoes due to the concentration of stresses which may be caused by fluid motions such as magma migrations. Volcanic tremors and long‐period events are considered to be generated by the dynamic interactions between volcanic fluids and volcanic edifice. In spite of the difference in their generation mechanisms, all of these signals thus can be seen as manifestations of passive and active processes of volcanic fluids under active volcanoes. On the other hand, such interactions along with more large‐scale tectonic activities make the volcanic structure very complex, and volcano is considered to be one of the most heterogeneous parts of the Earth’s crust. The understanding of such inhomogeneities in volcanoes 1 Department of Geophysics, Graduate School of Science, Tohoku University, Sendai, Japan. Copyright 2010 by the American Geophysical Union. 0148‐0227/10/2009JB007109 may thus provide us important information on the various volcanic processes as well as the volcanic structure. [3] One way to infer the distribution and movement of volcanic fluids beneath volcanoes is the use of tomography for velocity and/or attenuation, which basically relies on the direct waves. The tomographic method, however, has some limitations in its resolution partly due to the wavelength of the waves used in the analyses, and it is rather difficult to image small‐scale heterogeneities like intruded dykes and fractures using the method. An alternative approach to image small‐ scale heterogeneities is the use of scattered waves. In contrast to the deterministic structure obtained by the tomographic method, the analyses of scattered waves yield stochastic images of the heterogeneous media which can be characterized by the background velocity (mean velocity) and the scattering coefficients which represent the scattering power per unit volume. [4] So far, various models of scattering process have been proposed and applied to the observed data. These models include the single scattering model [e.g., Aki and Chouet, 1975; Sato, 1977], the multiple scattering model, and the diffusion model. Among these, the diffusion model, which reflects strong multiple scattering of the seismic energy, has been widely used in the previous studies on the wave prop- B07304 1 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO Figure 1. Distribution of active sources (stars) and stations (dots). Names of some stations are annotated. The topography of Asama volcano is represented by contours at every 200 m. The inset shows the location of Asama volcano in central Japan. agation at active volcanoes [e.g., Wegler and Luhr, 2001; Wegler, 2003]. These studies indicated that the mean free path of S waves beneath volcanoes is as short as the order of hundreds of meters and suggested significant multiple scattering of energy due to the heterogeneous structure of volcanoes. Almost all of these studies, however, assume the diffusion of a single mode (usually S wave), and the contribution of P and S waves and their conversion have not been well recognized partly due to lack of dense seismic observations capturing spatiotemporal distribution of propagating energy. [5] In this paper, we use the data of an active seismic experiment at Asama volcano, Japan, conducted in 2006 and analyze the wave propagation in the shallow heterogeneous structure of the volcano to understand the scattering and conversion processes due to the heterogeneity. First, we briefly summarize the outlines of the active experiment and describe the characteristics of the observed seismograms. We then apply the multiple isotropic scattering model to explain the observed spatiotemporal distribution of energy and estimate the scattering parameters as well as the intrinsic absorption. Finally, based on the estimation of the mode conversion and multiple scattering of P and S modes, we discuss some implications to the modeling and analysis of seismic wave propagation in heterogeneous volcanic environments. 2. Active Seismic Experiment at Asama Volcano [6] Asama volcano is one of the most active volcanoes in Japan. It is a medium‐sized, andesitic composite stratovolcano in central Japan, located at the junction of the Izu‐ Marianas and NE Japan volcanic arcs. The volcano has B07304 erupted generally in vulcanian style with a few exceptional subplinian eruptions in 1108 and 1783, and the last eruption occurred in 2004. The magma movement associated with the 2004 eruption had been successfully delineated from the precise relocation of earthquakes and geodetic modeling [e.g., Takeo et al., 2006], and the detected horizontal dyke intrusion in the western part of the volcano and the vertical alignment of the hypocenters enabled us to gain insights into the magma pathway beneath the volcano. [7] To delineate the seismic velocity structure around the summit and an area of dike intrusion during the 2004 eruption of the volcano, an active seismic experiment with artificial shots and a dense seismic network was conducted in October 2006 as a part of the national project for the prediction of volcanic eruptions [Aoki et al., 2009]. The experiment was conducted with five active sources and 464 temporarily deployed seismic stations around the volcano (Figure 1). Most of the stations were deployed linearly to north–south and east–west, transversing the volcano with a typical station separation of 100 m. At each station, a seismometer (Mark Products/Sercel L‐22D, natural frequency 2 Hz) was installed on the ground surface to record seismic waves excited by chemical explosions using 250–300 kg dynamite charges at a depth of around 60 m. The seismometers are mainly vertical ones except a few three‐component ones, and their outputs were recorded by portable data loggers (Datamark LS7000SH) with a sampling rate of 4 ms according to the predefined schedules. The clock of each data logger was locked with the GPS signal and kept enough accuracy throughout the observation. [8] Figure 2a shows an example of raw seismograms observed by the east–west measurement line for the S3 shot which is located at the cross point of north–south and east– west measurement lines (see Figure 1). The seismograms from the artificial shots, emitting mainly P wave around 10 Hz, clearly show the arrivals of the direct and refracted P waves. Their travel times have been used to determine the velocity structure beneath the volcano by Aoki et al. [2009]. The amplitude of these first arrivals are relatively small, and the overall seismograms are characterized by spindle‐like envelopes having small P onsets, missing clear S onsets, and long‐lasting codas. The coda waves last for a few tens of seconds, as shown in Figure 2b, and indicate significant scattering of seismic energies due to the inhomogeneity beneath the volcano. 3. Characteristics of Propagating Seismic Energy [9] Figure 3 shows the root‐mean‐square (RMS) envelopes of seismograms from the different artificial shots at one of the stations. It is evident that the temporal decay of the coda wave amplitudes for different shot are almost parallel at long lapse time (about after 10 s). This characteristic of the coda wave decay is also observed at other stations and suggests that the seismic energy is uniformly distributed in the volume surrounding the source and stations. On the other hand, the characteristics of the early portion of the envelopes differ between artificial shots as well as between stations. Since this difference reflects the transition between the transient direct wave propagation and the quasi static energy diffusion at long lapse time, we examined the spatiotemporal distribution of 2 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 artificial shots to the original velocity seismograms and compute the band‐pass‐filtered mean square (MS) envelope. We then estimated the site amplification factor at each station by applying the coda normalization method [e.g., Aki, 1980] to a 5 s long time window starting from 15 s after the shot origin time. We finally divided the MS envelope at each station by the estimated site amplification factor and calculated temporal evolution of the mean normalized seismic energy density at each station every 0.05 s to obtain the spatiotemporal distribution of energy density. Since we are focusing on the early portion of the wave propagation, we used a linear phase symmetric finite impulse response filter rather than a zero‐phase filter to preserve the characteristic of the waveform of the wavefront and selected a very short time window of 0.05 s. [11] Figure 4 shows an example of the spatial distributions of the propagating energy at different lapse times. We used seismograms of the S3 shot observed at stations along measurement lines extended to the north, south, and east directions from the shot location. This is because the use of the multiple measurement lines surrounding the source enhances the apparent spatial density of the observation, which enables us to study the details of the spatiotemporal energy distributions. Because of the rather complicated geological background, including repeated dyke intrusions beneath the west measurement line between the S3 and S2 shots, we avoided using the data observed along the line. [12] The spatial distribution of the energy, which shows a wide range of energy difference as symbolized by the logarithmic axis, clearly exhibits the accumulation of scattered energy behind the wavefront of the P wave from the explosive source. Such accumulation of the scattered energy is a direct indication of the multiple scattering process and/or the diffusion process due to the medium heterogeneities because the single scattering process generates the energy concentration Figure 2. (a) Seismogram section of the raw vertical seismogram observed by east–west measure line for the S3 shot. The amplitude of each trace is normalized by its maximum amplitude. (b) Example of the raw seismogram, band‐pass‐ filtered (8–16 Hz) seismogram, and RMS envelope at F09 station which is located at 3390 m south of the S3 shot (see Figure 1). The envelope is obtained with a smoothing window of 0.2 s. The vertical dashed line indicates the arrival time of the direct wave. propagating seismic energy to extract the information on the wave scattering and mode conversions during the transient stage. [10] Although the temporal decay gradient of the coda wave amplitude is similar between stations as mentioned above, the absolute values of the amplitude sometimes differ from station to station. To obtain the spatial distribution of energy, we thus need to correct the effect of site amplification due to the local shallow structure and station‐specific factors such as the coupling between seismometers and the ground. So, the spatiotemporal distribution of energy density is obtained by the following procedure: We first applied an 8– 16 Hz band‐pass filter covering the dominant energy of the Figure 3. Envelopes of 8–16 Hz band‐pass‐filtered seismograms for five shots observed at U20 station. The distance to each shot is 11880 m (S1), 7370 m (S2), 3140 m (S3), 10880 m (S4), and 13660 m (S5), respectively. To see the overall behavior of the coda waves, the envelopes are obtained with longer smoothing windows of 0.5 s. Thick lines show regression lines of the envelopes calculated with 7 s long time window starting from 13 s. All these regression lines have almost the same gradient of RMS / 10(−0.090±0.002)t. 3 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 mainly around the wavefront. On the basis of the similar energy accumulation in the vicinity of the source, many previous studies consider the diffusion of the energy and estimate the diffusivity and scattering strength at active volcanoes [e.g., Matsumoto and Hasegawa, 1991; Wegler and Luhr, 2001]. [13] Another important characteristic of the observed spatial energy distribution is that each spatial distribution exhibits a pattern having two slopes as shown in Figure 4h, rather than a single smooth curve expected from the diffusion model (see also Appendix A for examples of the spatial energy density distribution expected from the diffusion model). The pattern has a steeper slope behind the wavefront of the direct P wave and has a moderate one in the vicinity of the source; the kink of the pattern is situated around the midpoint between the source and the wavefront of the direct P wave. This characteristic at the early portion of the energy propagation tends to disappear after a few seconds, and the spatial distribution becomes like a single smooth curve at long lapse time like the one in Figure 4g. [14] Such spatial distributions of the energy showing two slopes are reminiscent of the multiple scattering of two modes having different scattering coefficients. Considering the initial energy generated by the explosion of buried dynamites, which mainly generates P energy, this pattern can be seen as a transient process of the mode conversion from P to S mode and their multiple scattering. In sections 4 and 5, we thus model the observed spatial energy distribution using the radiative transfer theory, including the mode conversions, and estimate the scattering coefficients using the observed spatial pattern. 4. Spatiotemporal Energy Distribution Modeled by the Radiative Transfer Theory Figure 4. (a–g) Observed spatial distribution of energy density at difference lapse times for the S3 shot. The energy density at each station (dots) is evaluated by using a 0.05 s time window after correcting the site amplification factor which is obtained by taking the average of those estimated from the records of S4 and S5 shots. Black and gray arrows indicate the wavefront of direct P wave and the kink of the spatial energy density distribution. (h) Close‐up view of the spatial distribution at 1.2 s. As indicated by shadows, observed energy density distribution exhibits a pattern having two slopes. [15] In media having strong random heterogeneity, like the one beneath active volcanoes, the contribution of the multiple scattering to the wavefield becomes larger and the scattered wavefield cannot be approximated by a small disturbance of waves. To model the multiple scattering process, the radiative transfer theory which describes energy transportation through heterogeneous media has been frequently used [e.g., Wu, 1985; Zeng, 1993]. The radiative transfer theory deals directly with the transport of energy neglecting phase information rather than that of wave, and it is basically applicable to model scattered energies at both short and long lapse time. For the case of isotropic scattering in heterogeneous elastic media, analytical solutions to the radiative transfer equation had been derived for both one‐mode and two‐mode types [Zeng, 1993; Sato, 1994], while numerical solutions using the Monte Carlo method had been used to model more general cases like the nonisotropic scattering of vector waves [e.g., Yoshimoto, 2000; Margerin et al., 2000]. [16] Here, we follow the derivation of Sato [1994] to synthesize the time traces and the spatial distributions of the energy densities in a 3‐D infinite medium taking into account the multiple isotropic scattering including conversion scattering. We suppose a spherical radiation of P and S energies, WP and WS, from a point source, and consider P‐to‐P, P‐to‐S, S‐to‐P, and S‐to‐S isotropic scatterings by randomly and uniformly distributed scatterers. The total energy density at 4 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 PS SP SS Here, gPP 0 , g0 , g0 , and g0 are the total scattering coefficients of the four isotropic scattering modes mentioned above, and a0 and b0 are the P and S wave velocities of the background medium, respectively. GP and GS are Green’s functions for energies, which take into account the causality and the effects of geometrical spreading, attenuation due to scattering, and intrinsic absorption, and they are written as PP PS 1 r eð0 g0 þ0 g0 þbP Þt H ð t Þ t 4r2 0 0 1 r ð0 g0SS þ0 g0SP þbS Þt e GS ðx; t Þ ¼ H ð t Þ t 4r2 0 0 GP ðx; t Þ ¼ ð3Þ where r = ∣x∣ and bP, bS represent the intrinsic absorptions for P and S energies. These equations can be solved by means of the Fourier‐Laplace transform, and the energy densities of P and S waves, as well as the total energy density, can be obtained as a function of space and time [e.g., Sato and Fehler, 1998, chapter 7]. [17] Figure 5 shows an example of the spatial distribution of the energy density evaluated by the radiative transfer theory described above. In this example, we assumed that the source is an isotropic P source and substituted WS = 0 into equation (2). In the calculation hereafter, we also used the relationship between the conversion scattering coefficients Figure 5. Spatial energy distribution obtained by the radiative transfer theory considering the multiple isotropic scattering of P and S energies and their conversion scatterings in an infinite scattering medium. The source is an isotropic P energy source (WP = 1, WS = 0) with a Gaussian time function of 0.04 s. (a) Spatial distribution of the total energy at 1.2 s and those of P and S energies. The parameters used are −1 PS PP SS PP a0 = 3.0 km s−1, gPP 0 = 1/3.0 km , g0 /g0 = 2.0, g0 /g0 = 3.0, and b = 0. (b) Spatial distribution of the total energy with different scattering coefficients. The solid curve represents the energy distribution using the same parameters as Figure 5a, PP and the other two curves are obtained with gSS 0 /g0 = 6.0 PS PP (broken curve) and g0 /g0 = 4.0 (dashed‐dotted curve) while the other parameters are fixed to those of Figure 5a. arbitrary point x in the medium is written as a function of time, t, as E ðx; t Þ ¼ E P ðx; tÞ þ ES ðx; t Þ ð1Þ where EP and ES represent the energy density of P and S wave, respectively, and they are expressed as sums of coherent energy densities as ZþZ1Z Z E ðx; t Þ ¼ W G ðx; t Þ þ P P P EP ðx 0 ; t 0 Þ0 g0PP 1 þ E S ðx 0 ; t 0 Þ0 g0SP GP ðx x 0 ; t t 0 Þdt 0 dx 0 ZþZ1Z Z S 0 0 E ðx ; t Þ0 g0SS ES ðx; t Þ ¼ W S GS ðx; tÞ þ 1 þEP ðx 0 ; t 0 Þ0 g0PS GS ðx x 0 ; t t 0 Þdt 0 dx 0 ð2Þ g0SP ¼ 1 PS g ; 202 0 0 0 0 ð4Þ which is derived from the reciprocal theorem and valid for any kind of heterogeneity [Aki, 1992]. The spatial distribution of the energy density is, therefore, controlled by six PS SS parameters: a0, g 0, gPP 0 , g0 , g0 , and b. In Figure 5, with the strong heterogeneity beneath active volcanoes in mind, we used the following scattering coefficients corresponding to pffiffiffi a strong scattering medium: a0 = 3.0 km s−1, g 0 = 3, g0PP = PP SS PP 1/3.0 km−1, gPS 0 /g0 = 2.0, and g0 /g0 = 3.0. [18] As shown in Figure 5a, the spatial distribution of energy density obtained from the radiative transfer theory considering the multiple isotropic scattering and mode conversion shows the energy accumulation in the vicinity of the source and the clear wavefront of P wave. The modeled energy density distribution also shows two slopes having different inclinations and a kink between two slopes. The location of the kink coincides with the wavefront of S wave even though the source radiates only P wave. It is reasonable considering that the mode conversion from P to S occurs most efficiently where the energy density of P wave is the highest (i.e., in the vicinity of the source) and the conversion back to P from S occurs less efficiently (see equation (4)) due to the uniformity of the heterogeneity assumed in the modeling. On the other hand, the inclinations of the two slopes on both sides of the kink are controlled by the value of scattering coefficients as shown in Figure 5b. For example, a larger value of the conversion scattering coefficient, g0PS, yields steeper slope between the wavefront of direct P having smaller energy and the kink having larger energy because a larger value of gPS 0 means that the energy of the P wave emitted from the source rapidly converted to S energy which travels with slower velocity than P energy, and such P‐to‐S conversion occurs most efficiently in the vicinity of the source. Similarly, a larger value of S‐to‐S scattering coefficient, gSS 0 , yields a 5 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO Table 1. Parameters Used and Estimated in the Fitting −1 a0 (km s ) g0 1/gPP 0 (km) PP gPS 0 /g0 PP gSS 0 /g0 b (s−1) Search Range Increment Best Fit Possible Range p2.5–3.9 ffiffiffi 3 (fixed) 2.0–5.0 1.0–5.0 1.0–15.0 0.30–1.40 0.2 ‐ 0.2 0.2 0.5 0.05 2.7ffiffiffi p 3 4.2 2.8 3.0 0.65 2.5–2.9 ‐ 3.2–4.8 1.8–3.8 2.0–7.5 0.50–0.70 bell‐shaped curve in the vicinity of the source due to stronger multiple scattering of S energy. 5. Estimation of the Scattering Coefficients [19] The spatiotemporal distribution of energy density obtained by the radiative transfer theory exhibits similar characteristics to that observed at Asama volcano. In addition, each scattering coefficient has a different effect to the spatial distribution of the energy density as described in section 4. These facts suggest the possibility to estimate the scattering coefficient from the observed spatiotemporal distribution of energy based on the multiple isotropic scattering model. [20] We used the spatial distributions of energy at six time windows shown in Figures 4a–4f, and estimated the medium parameters by minimizing the misfit, 2, between the observed and model‐predicted energy densities defined by 2 ¼ Nw N s ð iÞ X X flog Eo ði; jÞ log Em ði; jÞg2 ð5Þ i¼1 j¼1 where Eo(i, j) and Em(i, j) represent the observed and modeled energy densities at station j and at time window i. Nw is the number of time windows (Nw = 6 in this case) and Ns(i) is the number of stations of which hypocentral distances are less than a0t(i) where t(i) is the middle time of time window i. [21] To calculate the modeled energy, Em, we assumed an isotropic P energy source with affi Gaussian source time qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi function given by WP(t) = 4= td2 exp(−(2t/td)2) where td = 0.08 s and considered the multiple isotropic scattering of P and S energies in 3‐D infinite elastic medium ignoring the pffiffiffi effect of the free surface. We further assumed g 0 = 3, and as PS a consequence the value of gSP 0 /g0 is fixed as 1/6 from equation (4). The values of the intrinsic absorption for P and S energies are also assumed to be the same (bP = bS ≡ b). The remaining unknown parameters are thus P wave velocity (a0), PS PP isotropic scattering coefficients and their ratio (gPP 0 , g0 /g0 , SS PP g0 /g0 ), initial P energy (WP), and intrinsic absorption (b). Although these parameters are basically free parameters in the fitting, the range of the scattering parameters was chosen based on the scattering amplitude of each scattering process on the basis of the Born approximation [e.g., Sato and Fehler, 1998, chapter 4]. As discussed also in section 6, the scattering pattern of each scattering mode strongly depends on the correlation distance, a, of the heterogeneity and wave number, k. In the case of scattering of elastic wave in a medium having exponential autocorrelation, the ratio of scattering PP amplitudes, gSS 0 /g0 , obtained by integrating the angular‐ dependent scattering amplitude over whole solid angle varies between 4 ∼ 11 for ak < 1 depending on the value of ak, B07304 whereas the other ratios take more or less constant values of PP SP PP gPS 0 /g0 = 1 ∼ 3 and g0 /g0 ∼ 1. Using these results as a rough PP guide, we selected broader parameter range for gSS 0 /g0 than other parameters (Table 1). [22] In the estimation of the parameters, as the first step we ignored the effect of the intrinsic absorption because its contribution appears just as the constant shift of energy density on the spatial distribution at fixed time as indicated by equation (3) and it is rather difficult to resolve the trade‐off between the intrinsic absorption, b, and the initial energy, WP, using the short time interval used here. Under these assumptions, we searched the best parameter set so that the residual 2 becomes minimum by the grid search method. In Table 1, along with the best fit parameter set, errors of the estimated parameters obtained by the bootstrap method with 100 bootstrap samples are also listed. [23] The spatial distributions of energy density modeled with the best parameters are shown as the broken lines in Figure 6 along with the observed ones. Although we only used the time windows from 1.0 to 1.5 s in the estimation of model parameters, the spatial energy distributions at longer lapse times are also shown. It is seen that in each time window used for the estimation, the observed spatial pattern having the wavefront and two slopes are quantitatively well modeled by the multiple isotropic scattering model. Although there are small differences around the wavefront, they may be partly due to the assumption in the radiative transfer model that the wandering effect of the travel time fluctuations of different rays [e.g., Lee and Jokipii, 1975] is not taken into consideration. [24] In contrast to the good agreement at short lapse times, the modeled energy density distributions drawn by the broken lines show almost constant bias from the observed ones at longer lapse time. To explain the constant bias, we next took into consideration of the effect of the intrinsic absorption in the modeling of the energy density distribution. The effect of the intrinsic absorption is comparably small at short lapse time; we thus fixed the values of the scattering coefficients and the P wave velocity estimated above and search the optimal value of the intrinsic absorption, b, and the initial energy, WP. [25] The obtained value of b using the time windows starting from 5.0, 10.0, 15.0 sec and the distance range within 6 km is 0.65 sec−1, and the corresponding spatial distribution of energy density is shown as the solid lines in Figure 6. The obtained value of b corresponds to the quality factor Q of Q ∼ 100 using the relationship b = Q−1w where w is the central angular frequency of the frequency band (w = 2p × 12 Hz in this case). Here, it is noticed that the value of b can be constrained only using the longer lapse time windows where the multiple scattering of S energy dominates as discussed in section 6 and Appendix A. The obtained value of b therefore mainly reflects the attenuation for S energy, bS, and it is rather difficult to constrain the attenuation for P energy, bP, using the present method. In that sense, the methodology to separate the effects of the scattering and attenuation described here resembles the multiple lapse time analysis [e.g., Fehler et al., 1992; Hoshiba, 1991] in which several time windows are used to discriminate the scattering and intrinsic attenuation of a single mode. It is also noted that the absorption parameter, b, includes the effects of not only the intrinsic absorption but 6 of 14 B07304 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO the following relationship between the isotropic scattering coefficients: lP ¼ 1= g0PP þ g0PS ; lS ¼ 1= g0SS þ g0SP ; tP ¼ lP =0 ; tS ¼ lS =0 ð6Þ 6. Figure 6. The modeled spatiotemporal distribution of energy using the best fit parameters. The broken lines represent the energy distributions assuming no intrinsic absorption (b = 0), whereas the solid lines take into account the effect of the intrinsic absorption. Black and gray arrows indicate the wavefronts of P and S waves. also the window effect corresponding to the one‐way energy leakage to larger depth. The systematically small energy at the vicinity of the source shown in Figure 6 is thus partly due to the upward migration of a high‐velocity anomaly around the S3 shot detected by Aoki et al. [2009], which causes the rapid escape of the energy from the volume considered. [26] The values of isotropic scattering coefficients and corresponding mean free paths and mean free times derived from the best fit parameters are summarized in Table 2. Here, to calculate the mean free paths (lP, lS) and corresponding mean free times (tP, tS) of P and S waves, we used Discussion [27] Using the multiple isotropic scattering model including conversion scattering between P and S energies, we could estimate quantitatively the scattering coefficients beneath Asama volcano from the spatiotemporal distribution of energy density observed by the dense seismic network of an active source experiment. We also demonstrated that we could estimate the scattering coefficients and the intrinsic absorption separately by using the different time windows in a manner slightly different from the multiple lapse time analysis [e.g., Fehler et al., 1992; Hoshiba, 1991]. The estimated isotropic scattering coefficient in the 8–16 Hz band for P‐to‐P scattering is 0.24 km−1 and those of S‐to‐S and P‐to‐ S scattering are 3 and 2.8 times larger than that of P‐to‐P scattering. These scattering coefficients indicate that the mean free paths of P and S waves are about 1 km in the frequency band, and they correspond to the mean free times of about 0.4 s and 0.8 s, respectively. [28] The obtained value of the mean free path is much shorter than the typical value of the usual Earth’s crust [e.g., Sato and Fehler, 1998], and it may reflect the strong heterogeneity at shallow depth of the active volcano. Very short mean free path beneath active volcanoes has been also reported by previous studies. Using the diffusion model for single mode, Wegler and Luhr [2001] estimated a mean free path of the order of 0.1 km for S wave at shallow depth of the Merapi volcano in the frequency band of 4–20 Hz from an active seismic experiment. Using the multiple lapse time analysis, Londono [1996] estimated a mean free path of about 10 km in the frequency band of 6–24 Hz at Nevado Del Ruiz volcano, Colombia, using volcanotectonic earthquakes. [29] The model used in this study is, of course, more or less too simplified, and we need to investigate several points including the effects of the existence of the free surface and the nonisotropic scattering as well as the three‐dimensional structure of the volcanic edifice. These subjects would be addressed by using numerical methods like the finite difference modeling and Monte Carlo simulation in the near future. Nevertheless, the characterization of the strong scattering Table 2. Parameters Derived From the Best Fit Parameters for an 8–16 Hz Band Property Symbol Value Background P wave velocity Background S wave velocity Intrinsic absorption P‐to‐P isotropic scattering coefficient P‐to‐S isotropic scattering coefficient S‐to‐P isotropic scattering coefficient S‐to‐S isotropic scattering coefficient Mean free path of P wave Mean free path of S wave Mean free time of P wave Mean free time of S wave a0 b0 Q gPP 0 gPS 0 gSP 0 gSS 0 lP lS tP tS 2.7 km s−1 1.56 km s−1 100 0.24 km−1 0.67 km−1 0.11 km−1 0.71 km−1 1.10 km 1.21 km 0.41 s 0.78 s 7 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 Figure 7. Energy partitioning modeled by the multiple isotropic scattering model in an infinite scattering medium using the estimated best fit parameters. The source is an isotropic P energy source (WP = 1, WS = 0) with a Gaussian time function of 0.04 s. (a) Time traces of the total (solid lines), P (broken lines), and S (dashed‐dotted lines) energy densities and the ratio between P and S energy densities at two different hypocentral distances. Thin broken lines represents the theoretical energy ratio at equilibrium state. (b) Spatial distribution of energy densities and their ratio at two fixed times. phenomena beneath the volcano and the quantitative estimation of the scattering coefficients enable us to have a few implications important for the interpretations and modelings of the seismic waves observed at active volcanoes. These implications include the energy partitioning between P and S modes, the effect of the scattering and intrinsic attention on the amplitude of direct waves, and the scattering characteristics of volcanic environments. 6.1. Energy Partitioning [30] It has been considered that the coda waves consist of incoherent scattered waves and S energy dominates in the coda waves. This is because the P‐to‐S conversion scattering occurs more effectively than S‐to‐P scattering (see equation (4)) and S‐to‐S scattering is generally stronger than P‐to‐P scattering. On the basis of such considerations, many previous studies assumed the scattering and diffusion of S energy to explain the observed energy propagation even when the source itself is the P energy source like artificial shots. For example, from the experiment with a marine air gun source, Matsumoto and Hasegawa [1991] suggested a large contribution of S waves converted from P waves in seismograms recorded in a volcanic area. It is thus meaningful to discuss the transition of the energy partition among P and S modes based on the scattering parameters estimated in this study. [31] As described in section 5, the radiative transfer theory considering the multiple isotropic scattering predicts energy densities EP and ES individually as functions of space and time once the scattering parameters are given. Figure 7a shows an example of time traces of the total energy density Et (= EP + ES), EP, and ES at two different hypocentral distances. Here we assumed a P energy source and used the scattering coefficients estimated in this study. The intrinsic absorption is neglected for simplicity. The modeled energy densities clearly show that the S energy density becomes larger than P energy density just after the arrival of the direct wave which is mainly formed by P energy. It is also shown that the local energy balance at a fixed observation point comes close to an equipartition about 1 s after the direct wave arrival. These results suggest that the incoherent energy in coda waves near the source is dominated by S energy after a few seconds regardless of the source radiation in a strongly homogeneous medium and support the assumption of single‐ mode diffusion process used by Wegler and Luhr [2001] and other previous studies, although the use of the single‐mode diffusion model tends to yield stronger scattering strength as discussed in Appendix A. [32] It should also be noted that the modeled ES/EP ratio once exceeds the value of equipartition in the full space pffiffiffi ES/EP ≡ rS/P = 2g 30 (= 10.39 for g 0 = 3) represented by thin broken lines and that it takes much longer time to steady this local energy balance down to the equilibrium one. For 8 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 the dependence of the duration on hypocentral distance are qualitatively consistent with the establishment of equipartition shown in Figure 7a. [34] Of course, the discussion above is based on the scattering coefficients obtained under the assumption of the multiple isotropic scattering model. So, we need to have more direct evidence of the energy partitioning by independent ways like the wavefield decomposition by array observations [e.g., Shapiro et al., 2000] to verify the discussions and indications obtained here. Figure 8. Examples of the observed three‐component MS envelopes at two different hypocentral distances (see Figure 1). The solid lines represents the vertical component, and the broken and dashed‐dotted lines represent two horizontal components. The energy of each trace is individually normalized by using the coda level of the component. example, at a hypocentral distance of 1500 m, it takes about 8.8 s to fall within ±5% of the value of equipartition, and the duration is about 13 times longer than the mean free time of S wave (see also Appendix A for different examples). Similar slow convergence toward equipartition is also reported by Margerin et al. [2000], in which they numerically studied the multiple scattering of elastic waves using the Monte Carlo method. They found that equipartition is established after at least 15 mean free times at the hypocentral distance of one mean free path in the Rayleigh regime where scattering anisotropy is weak. Figure 7b shows the spatial distributions of P and S energies at fixed time and gives an account for the phenomenon; that is, the strong multiple scattering of S energy in the vicinity of the source prevents the S energy to diffuse outward, whereas multiple scattering of P energy is less effective and P energy propagates more easily. [33] The signature of the rapid mode conversion can also be seen on the observed three‐component seismograms at Asama volcano. Figure 8 shows examples of the MS envelopes of the three‐component data which are band‐pass filtered between 8 Hz and 16 Hz and individually normalized using the coda level. As shown in Figure 8, energy densities of the horizontal components rapidly grow up just after the arrival of the direct wave and show almost the same temporal decay as that of the vertical component after a few seconds from the shot origin time. Furthermore, the duration to reach the same decay rate seems to become longer as hypocentral distance increases. Although the energies of P and S modes and those of vertical and horizontal components obviously have different physical meanings [e.g., Margerin et al., 2009], the observed same decay rate at short lapse time and 6.2. Attenuation of the Direct P Waves [35] The distribution of the intrinsic attenuation is one of the most important keys to infer the thermal state and composition of Earth’s media as well as the existence of fluid. The observed attenuation is, however, caused by both the intrinsic attenuation and the attenuation due to scattering, and it is needed to separate the contributions of these two different attenuation mechanisms. One way to separate these two attenuation mechanisms is the use of the multiple lapse time analysis of the whole S waves as mentioned before. While such analyses yield the averaged intrinsic and scattering attenuation over some volume surrounding the raypath, analyses using the direct wave give information of more localized volume. So, it is also meaningful to examine the amplitude of direct waves using the scattering and attenuation parameters estimated in this study. [36] Figure 9 shows the distance dependence of the observed peak amplitudes of the direct wave which are measured after applying the 8–16 Hz linear phase filter to the raw velocity seismograms. As shown in Figure 9, the peak amplitude of the direct wave decays as distance increases, and the decay of the peak amplitude is almost inversely propor- Figure 9. Distance dependence of the observed and modeled amplitude of direct waves. Solid circles represent the peak amplitude of the observed direct P wave in an 8– 16 Hz band, and solid and broken lines represent the peak amplitude obtained from the modeled energy density assuming the parameters estimated in this study. The modeled peak amplitude is obtained by taking the square of the calculated energy density from an isotropic P energy source. The broken line represents the amplitude decay due to the intrinsic absorption (b = 0.65), whereas the solid line takes into account the effects of both the scattering attenuation and intrinsic absorption. 9 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 Asama volcano. Different parameters quantifying the medium heterogeneity are the characteristic scale and strength of the randomness. For that purpose, the correlation distance, a, and the magnitude of fractional fluctuation, , of the medium have been widely used as a statistical measure of the randomness [e.g., Sato and Fehler, 1998, chapter 4]. Of course, the scattering coefficients and these statistical parameters (a and ) are not independent, and it is possible to relate each other once the statistical characteristic of the medium is assumed. [39] Figure 10 shows the ratios of scattering coefficients for different correlation distance, a, and wave number, k, in case of an isotropically random medium having exponential autocorrelation. Even though each scattering process generally exhibits a nonisotropic scattering pattern, we obtained the values of the ratios shown as solid lines by integrating the angular‐dependent scattering coefficients over whole solid PP angle. That is, the ratio gPS 0 /g0 is calculated from Figure 10. Plot of the ratios of scattering coefficients against normalized wave number, ak. The medium is assumed to have exponential autocorrelation with correlation pffiffiffi length of a. We assumed g 0 = 3, and the fractional fluctuation of P and S wave velocities are assumed to be the same. Solid lines represent the ratios obtained by integrating the scattering coefficient of each scattering mode over whole solid angle, and broken lines represent the ratios of integrated momentum transfer scattering coefficients. tional to the square of the distance (a / r−2) instead of the distance itself (a / r−1) expected from the geometrical spreading of the spherical body wave. This relatively strong distance dependence of the amplitude suggests that both intrinsic and scattering attenuation affect the amplitude of the direct wave. [37] To explain the observed distance dependence of the direct wave amplitude, we calculated the amplitude using the estimated scattering parameters and the multiple isotropic scattering model. In the modeling, we first calculated the time trace of energy density at every hypocentral distance assuming a Gaussian source pulse of 0.08 s and then measure the peak energy density of the direct wave. We then obtained the peak amplitude of the direct wave by taking the square root of the measured peak energy density. Figure 9 also shows the distance dependencies of the modeled peak amplitudes assuming two cases: one only considers the contribution of the intrinsic attenuation (broken line) and the other takes into consideration the effect of both the intrinsic and scattering attenuation (solid line). It is shown that it seems not to be possible to explain the observed distance dependence by merely considering the intrinsic attenuation because of the very short travel time of the direct wave. In contrast, the total contribution of the intrinsic and scattering attenuation well capture the distance dependence of inverse square of distance. Although the use of the isotropic scattering model is not suitable to model the direct wave, nevertheless, these results suggest that the scattering attenuation has an indispensable effect on the modeling of not only coda waves but also direct waves in heterogeneous volcanic environments. 6.3. Scattering Coefficients and Medium Heterogeneity [38] In this study, we estimated the isotropic scattering coefficient as a measure of the medium heterogeneity beneath g0PS ¼ g0PP R R 1 2 PS g ð 4 0 R R0 1 2 PP 0 g ð 4 0 ; Þ sin d d ; Þ sin d d ð7Þ where y is the angle from the propagation direction of the incident wave and z is the angle in the plane perpendicular to PP SS PP the propagation direction. Other ratios gSP 0 /g0 and g0 /g0 are also calculated in a similar manner. Here, we used the scattering amplitudes calculated by the Born approximation assuming the same fractional fluctuation both for P and S PP waves. As shown by solid lines in Figure 10, the ratio gSS 0 /g0 varies greatly depending on the value of ak, whereas the other ratios are more or less constant in the range of ak < 1. The ratios are also dependent on the value of g 0 (not shown here), but the overall characteristic of the ratios is almost the same. The regime where scattering process becomes isotropic corPP PS responds to the range ak 1, and the ratios gSS 0 /g0 and g0 / PP g0 are around 11 and 3 in that range, respectively. When multiple scattering is dominant, scattering process can be effectively described by isotropic scattering even if each scattering process is nonisotropic. In Figure 10, the ratios of the equivalent isotropic coefficients (momentum transfer R R 1 2 scattering coefficient gm ≡ 4 0 0 (1 − cos y)g(y, z) sin ydydz) calculated like gmPS ¼ gmPP R R 1 2 ð1 cos 4 0 R R0 1 2 ð 0 1 cos 4 0 Þg PS ð ; Þ sin d d Þg PP ð ; Þ sin d d ð8Þ are also shown as broken lines. Although the ratios of momentum transfer scattering coefficients, gm, show slightly different values than the ratios of the total scattering coefficients, g0, due to the suppression of the strong forward scattering, overall characteristics of the ratios is kept unchanged. PP PS PP [40] Our estimation of gSS 0 /g0 and g0 /g0 at Asama volcano are 3.0 and 2.8, respectively. These values are almost the same as those of ak ∼ 1. This seems to somehow contradict our assumption of isotropic scattering which is expected in the regime ak 1. To figure out the puzzle, systematic estimation of the scattering coefficients at different frequency bands would be needed in the future. In addition, to gain details of the scattering process and the medium 10 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO parameters, it may also be necessary to examine the scattering patterns and the ratios between scattering coefficients in the case of different velocity fluctuations for P and S waves which is expected in a volcanic environment where volcano‐ specific inhomogeneities such as fluid‐filled cracks exists. 7. Conclusion [41] From an active seismic experiment with a dense network, we found an observational evidence of mode conversions and multiple scattering in the shallow heterogeneous structure at Asama volcano, Japan. Spatial distributions of observed energy density suggest the dominant P‐to‐S conversion scattering at short lapse time and following S‐to‐S multiple scattering. Assuming the multiple isotropic scattering model including conversions scattering in an infinite uniformly random medium, we modeled the observed spatiotemporal energy distribution on the basis of the radiative transfer theory and quantitatively estimated the scattering coefficients. The mean free paths for P and S waves are estimated as short as about 1 km for the 8–16 Hz band. These values of the mean free paths are much shorter than values for the usual Earth’s crust, and we may say that the scattering coefficients are useful parameters to quantify the medium heterogeneity beneath active volcanoes. [42] From the estimated scattering coefficients, we obtain the following implications which might be important in the analyses of seismic signals using the seismic interferometry [e.g., Wapenaar et al., 2008] and the attenuation tomography. The local energy partitioning between P and S energies modeled with the obtained scattering coefficients approaches the theoretical equipartition in the full space within a few seconds after the arrival of the direct wave, although it takes much more time to reach equipartition (about 10–15 times longer than the mean free time of S wave). The amplitude of direct wave is controlled by the scattering attenuation EP ðk; sÞ ¼ ES ðk; sÞ ¼ B07304 Appendix A: Comparison With the Diffusion Model [44] When multiple scattering dominates like in this study, we may expect that the wave propagation becomes like a diffusion process because the spatial distributions of the directional distribution of P and S energy densities become smooth and their angular dependence is very small. In this section, using the estimated scattering parameters at Asama volcano, we thus compare the characteristics of the diffusion model with the multiple isotropic scattering model we used in this study. [45] In the case of the dominant multiple scattering, the propagation of P and S energy densities from an impulsive isotropic radiation of P and S energies (WP and WS) are described by a set of diffusion equations as [e.g., Trégourès and van Tiggelen, 2002] @ EP DP DEP þ g0PS 0 EP þ bEP g0SP 0 ES ¼ WP ðxÞðt Þ @t @ ES DS DES þ g0SP 0 ES þ bES g0PS 0 EP ¼ WS ðxÞðt Þ @t ðA1Þ where we further suppose that the temporal change is small. Diffusion constants for P and S energy densities, DP and DS, are related to the isotropic scattering coefficients as 1 0 3 g0PP þ g0PS 1 0 DS ¼ 3 g0SS þ g0SP DP ¼ ðA2Þ Coupled equations (A1) can be solved by using the Fourier‐ Laplace transform in space and time, and P and S energy densities are obtained as s þ b þ DS k 2 þ 0 g0SP WP þ 0 g0SP WS þ 0 g0SP þ DP k 2 þ DS k 2 s þ b2 þ b0 g0PS þ b0 g0SP þ bDP þ bDS þ 0 DS g0PS þ 0 DP g0SP k 2 þ DP DS k 4 s þ b þ DP k 2 þ 0 g0PS WS þ 0 g0PS WP þ 0 g0SP þ DP k 2 þ DS k 2 s þ b2 þ b0 g0PS þ b0 g0SP þ bDP þ bDS þ 0 DS g0PS þ 0 DP g0SP k 2 þ DP DS k 4 s2 þ 2b þ 0 g0PS s2 þ 2b þ 0 g0PS ðA3Þ rather than the intrinsic absorption in the frequency range we analyzed. [43] The strong scattering characteristics estimated in this study and the implications above suggest that mode conversion and multiple scattering have an indispensable effect in the modeling and analysis of seismic wave propagation in a heterogeneous volcanic environment. Further observation and analysis using, e.g., the three‐component seismic arrays may give us more information on the scattering and conversion processes, and the characterization of the stochastic small‐scale heterogeneity using the multiply scattered wave along with that of the deterministic structure using the direct waves undoubtedly help us to infer more details of the existence and transportation of volcanic fluids beneath active volcanoes. [46] Figure A1 shows the comparison of spatial distributions of energy densities at different lapse times obtained by the multiple isotropic scattering model with mode conversions (solid lines) and the diffusion model (broken lines) for P energy source (Figure A1a) and S energy source (Figure A1b). The values of the four isotropic scattering coefficients and P and S wave velocities are the same as the best fit parameters obtained in this study (see Table 1), and the intrinsic absorption is neglected for simplicity. The corresponding diffusivities for the diffusion model are DP = 1.0 km2/s and DS = 0.63 km2/s. As clearly seen in Figure A1, the diffusion solution breaks the causality and shows the propagation of energy ahead of the wavefront of P energy which is properly evaluated by the multiple isotropic scattering model and the acausal propagation of the energy results 11 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO B07304 Figure A1. Spatial distribution of energy densities at different lapse times modeled by the multiple isotropic scattering (solid lines) and the coupled diffusion model (broken lines) using the estimated best fit parameters. (a) Energy distributions in the case of P energy source. (b) Energy distributions in the case of S energy source. in the gentle slope of the energy distribution especially in the case of P energy source. The result implies that the diffusion model yields stronger scattering strength than the multiple isotropic scattering model for the same spatial energy distribution and suggests the adequacy of the use of the multiple isotropic scattering model in this study. [47] It is also meaningful to see the difference in the energy balance between P and S energy densities obtained by the two models. The spatial distributions of the energy density ratio at different lapse times calculated with the same parameters are shown in Figure A2. In the case of P energy source (Figure A2a), similarly to Figure 7, the spatial distribution of energy density ratio modeled by the multiple isotropic scattering model (solid lines) shows a steep slope corresponding to the P‐to‐S conversion scattering behind the wavefront of P wave, and the local energy density ratio in the vicinity of the source once exceeds the theoretical ratio of equipartition in the full space represented by thin dotted lines, and reach to equipartition. On the other hand, the spatial energy density distribution modeled by the diffusion model (broken lines) shows more smooth spatial variation and slowly changes with time because the energy flux is proportional to the gradient of the spatial distribution of energy density. In the case of S energy source (Figure A2b), the spatial energy density distributions obtained by the two models are more or less similar except the existence of the Gaussian peak corresponding to the wavefront of S wave. The local energy density ratio behind the S wavefront shows a relatively flat spatial variation and monotonically decreases approaching equipartition. [48] The local energy density ratio is found to approach the theoretical equipartition regardless of the energy balance at the source due to the existence of conversion scatterings, although it takes a long time to achieve equilibrium. The temporal changes of the local energy balance at fixed hypocentral distances are shown in Figure A3 for P and S energy radiations. In each plot of Figure A3, as a comparison, the energy ratio of total P and S energies in the full space modeled by the multiple isotropic scattering model, which is determined from the Fourier transform at wave number k = 0, is also shown as a dashed‐dotted line. In the case of P energy radiation shown in Figure A3a, the local energy density ratio obtained by both the multiple scattering model and the diffusion model increases and becomes larger than the theoretical ratio. The ratio then decreases with increasing lapse time. It takes about 10 s for the local energy density ratios obtained by these two models to reach within a 5% error of the theoretical ratio, while it takes only about 3 s for the total energy ratio to reach within the 5% error. As the estimated mean free time of S wave at the volcano is about 0.7 s, the time interval needed to steady down to the error range corresponds to about 14 times as large as the mean free time for the local energy density ratio and about 4 times as large as the mean free time for the total energy density ratio in the full space. At lapse 12 of 14 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO Figure A2. Spatial distribution of energy density ratios at different lapse times modeled by the multiple isotropic scattering (solid lines) and the coupled diffusion model (broken lines) using the estimated best fit parameters. Thin broken line represents the theoretical ratio of the equipartition. (a) Energy distributions in the case of P energy source. (b) Energy distributions in the case of S energy source. Figure A3. Temporal change of energy density ratios at different hypocentral distances modeled by the multiple isotropic scattering (solid lines) and the coupled diffusion model (broken lines) using the estimated best fit parameters. Thin broken line represents the theoretical ratio of the equipartition, and dashed‐dotted line represents the energy ratio in the full space. (a) Energy distributions in the case of P energy source. (b) Energy distributions in the case of S energy source. 13 of 14 B07304 B07304 YAMAMOTO AND SATO: MULTIPLE SCATTERING AT ASAMA VOLCANO times larger than about 7 s the temporal decay of the local energy density ratios of the two models are almost the same, and we may say that the coupled diffusion model is a good approximation of the multiple isotropic scattering model with conversion scatterings. In the case of S energy radiation (Figure A3b), the local energy density ratios by the two models monotonously decrease with lapse time and they show almost the same temporal decay at lapse times larger than about 7 s. It again takes about 10 s for the local energy ratios obtained by the two models to reach within the 5% error, whereas about 3 s of lapse time is needed for the total energy ratio to reach the error range. The mean free time of S wave in short periods is about 25 s in the lithosphere, which is 35 times longer than the mean free time at the volcano we analyzed. If the scaling is applicable, these results suggest that it takes about 350 s for the local energy density ratio to reach the theoretical equipartition. [49] Acknowledgments. We are grateful to the participants of the active seismic experiment at Asama volcano. We would especially like to thank Yosuke Aoki and Minoru Takeo for all the effort they put into the project. Careful and constructive comments from the Associate Editor and two anonymous reviewers significantly improved this manuscript. This study was partly supported by the G‐COE program of Tohoku University “Global Education and Research Center for Earth and Planetary Dynamics.” References Aki, K. (1980), Attenuation of shear‐waves in the lithosphere for frequencies from 0.05 to 25 Hz, Phys. Earth Planet. Inter., 21(1), 50–60, doi:10.1016/0031-9201(80)90019-9. Aki, K. (1992), Scattering conversions P to S versus S to P, Bull. Seismol. Soc. Am., 82(4), 1969–1972. Aki, K., and B. Chouet (1975), Origin of coda waves: Source, attenuation and scattering effects, J. Geophys. Res., 80(23), 3322–3342, doi:10.1029/ JB080i023p03322. Aoki, Y., et al. (2009), P wave velocity structure beneath Asama volcano, Japan, inferred from active source seismic experiment, J. Volcanol. Geotherm. Res., 187(3–4), 272–277, doi:10.1016/j.jvolgeores.2009.09.004. Chouet, B. (2003), Volcano seismology, Pure Appl. Geophys., 160(3–4), 739–788, doi:10.1007/PL00012556. Fehler, M., M. Hoshiba, H. Sato, and K. Obara (1992), Separation of scattering and intrinsic attenuation for the Kanto‐Tokai region, Japan, using measurements of S wave energy versus hypocentral distance, Geophys. J. Int., 108(3), 787–800, doi:10.1111/j.1365-246X.1992.tb03470.x. Hoshiba, M. (1991), Simulation of multiple‐scattered coda wave excitation based on the energy conservation law, Phys. Earth Planet. Inter., 67(1–2), 123–136, doi:10.1016/0031-9201(91)90066-Q. Kawakatsu, H., and M. Yamamoto (2007), Volcano seismology, in Earthquake Seismology, Treatise Geophys., vol. 4, edited by H. Kanamori, chap. 13, Elsevier, New York. B07304 Lee, L. C., and J. R. Jokipii (1975), Strong scintillations in astrophysics. II. A theory of temporal broadening of pulses, Astrophys. J., 201, 532–543. Londono, J. M. (1996), Temporal change in coda Q at Nevado Del Ruiz volcano, Colombia, J. Volcanol. Geotherm. Res., 73(1–2), 129–139, doi:10.1016/0377-0273(95)00084-4. Margerin, L., M. Campillo, and B. V. Tiggelen (2000), Monte Carlo simulation of multiple scattering of elastic waves, J. Geophys. Res., 105(B4), 7873–7892, doi:10.1029/1999JB900359. Margerin, L., M. Campillo, B. A. V. Tiggelen, and R. Hennino (2009), Energy partition of seismic coda waves in layered media: Theory and application to Pinyon Flats Observatory, Geophys. J. Int., 177(2), 571–585, doi:10.1111/j.1365-246X.2008.04068.x. Matsumoto, S., and A. Hasegawa (1991), Estimation of relative site effect, coda Q and scattering strength from the records obtained by a large airgun experiment, Phys. Earth Planet. Inter., 67(1–2), 95–103, doi:10.1016/ 0031-9201(91)90063-N. Sato, H. (1977), Single isotropic scattering model including wave conversions: Simple theoretical model of the short period body wave propagations, J. Phys. Earth, 25, 163–176. Sato, H. (1994), Multiple isotropic scattering model including P‐S conversions for the seismogram envelope formation, Geophys. J. Int., 117(2), 487–494, doi:10.1111/j.1365-246X.1994.tb03946.x. Sato, H., and M. C. Fehler (1998), Seismic Wave Propagation and Scattering in the Heterogeneous Earth, Springer, New York. Shapiro, N. M., M. Campillo, L. Margerin, S. K. Singh, V. Kostoglodov, and J. Pacheco (2000), The energy partitioning and the diffusive character of the seismic coda, Bull. Seismol. Soc. Am., 90(3), 655–665, doi:10.1785/ 0119990021. Takeo, M., Y. Aoki, T. Ohminato, and M. Yamamoto (2006), Magma supply path beneath Mt. Asama volcano, Geophys. Res. Lett., 33, L15310, doi:10.1029/2006GL026247. Trégourès, N. P., and B. A. van Tiggelen (2002), Generalized diffusion equation for multiple scattered elastic waves, Waves in Random Media, 12(1), 21–38, doi:10.1088/0959-7174/12/1/302. Wapenaar, K., D. Draganov, and J. O. A. Robertsson (2008), Seismic Interferometry: History and Present Status, Geophys. Reprint Ser. 26, Soc. of Explor. Geophys., Tulsa, Okla. Wegler, U. (2003), Analysis of multiple scattering at Vesuvius volcano, Italy, using data of the TomoVes active seismic experiment, J. Volcanol. Geotherm. Res., 128(1–3), 45–63, doi:10.1016/S0377-0273(03)00246-4. Wegler, U., and B.‐G. Luhr (2001), Scattering behaviour at Merapi volcano (Java) revealed from an active seismic experiment, Geophys. J. Int., 145(3), 579–592, doi:10.1046/j.1365-246x.2001.01390.x. Wu, R.‐S. (1985), Multiple scattering and energy transfer of seismic waves ‐ separation of scattering effect from intrinsic attenuation: I. Theoretical modeling, Geophys. J. R. Astron. Soc., 82(1), 57–80, doi:10.1111/j.1365-246X.1985.tb05128.x. Yoshimoto, K. (2000), Monte Carlo simulation of seismogram envelopes in scattering media, J. Geophys. Res., 105(B3), 6153–6162, doi:10.1029/1999JB900437. Zeng, Y. (1993), Theory of scattered P and S wave energy in a random isotropic scattering medium, Bull. Seismol. Soc. Am., 83(4), 1264–1276. H. Sato and M. Yamamoto, Department of Geophysics, Graduate School of Science, Tohoku University, Sendai 980‐8578, Japan. (sato@zisin. geophys.tohoku.ac.jp; [email protected]) 14 of 14