Title. Author(s)Aoyama, Hiroshi; Oshima, Hiromitsu. CitationGeophysical Research Letters, 35(6): L Issue Date Doc URL. Rights.

Similar documents
Fig. 1. Joint volcanological experiment on volcanic structure and magma supply system in Japan.

Precursory tilt changes of small phreatic eruptions of Meakan-dake volcano, Hokkaido, Japan, in November 2008

Seismic signature of fluid motion in a shallow conduit system beneath Aso volcano, Japan

68. Izu-Torishima. Summary. Latitude: 30 29'02" N, Longitude: '11" E, Elevation: 394 m (Ioyama) (Elevation Point) (68.

Monthly Volcanic Activity Report (July, 2012)

Reexamination of moment tensors for initial motion of explosion earthquakes using borehole seismograms at Sakurajima volcano, Japan

Real-time inversion of VLP source functions at Stromboli Volcano, Italy.

Seismic inversion analysis of the 2014 and 2015 Kuchinoerabujima volcanic eruptions, using F-net broadband seismometers

volcanic tremor and Low frequency earthquakes at mt. vesuvius M. La Rocca 1, D. Galluzzo 2 1

km. step. 0.5km. Ishihara km. al., Rayleigh. cavity. cavity

Haruhisa N. (Fig. + ) *+ Graduate School of Environmental Studies, Nagoya University, Furo-cho, Chikusa-ku, Nagoya.0. 20*+ Japan.

Monthly Volcanic Activity Report (March 2013)

19. Esan Continuously Monitored by JMA

Dynamic Triggering Semi-Volcanic Tremor in Japanese Volcanic Region by The 2016 Mw 7.0 Kumamoto Earthquake

Continuously Monitored by JMA

Continuously Monitored by JMA. Latitude: 34 23'49" N, Longitude: '13" E, Elevation: 432 m (Miyatsukayama) (Spot elevation measured by JMA)

Seismic quantification of the explosions that destroyed the dome of Volcán de Colima, Mexico, in July August 2003

Monthly Volcanic Activity Report (April 2013)

Magnitude frequency distribution of volcanic explosion earthquakes

4.Mashu. Summary. Latitude: 43 34'20" N, Longitude: '39" E, Elevation: 857 m (Kamuinupuri) (Elevation Point) (4. Mashu)

Monthly Volcanic Activity Report (February 2016)

Very-Long-Period Volcanic Tremor at Stromboli, Italy

Continuously Monitored by JMA. Overview of Kuchinoerabujima taken from the East on July 23, 1996 by the Japan Meteorological Agency

Temporal evolution of a hydrothermal system in Kusatsu-Shirane Volcano, Japan, inferred from the complex frequencies of long-period events

Effect of the Emperor seamounts on trans-oceanic propagation of the 2006 Kuril Island earthquake tsunami

Continuously Monitored by JMA. Latitude: 24 45'02" N, Longitude: '21" E, Elevation: 169 m (Suribachiyama) (GSI Measuring Point)

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

Continuously Monitored by JMA. Latitude: 34 13'10" N, Longitude: '11" E, Elevation: 572 m (Tenjosan) (Triangulation Point - Kozushima)

LETTER Earth Planets Space, 56, , 2004

Monthly Volcanic Activity Report (November 2015)

EAS 116 Earthquakes and Volcanoes

Scaling relationship between the duration and the amplitude of non-volcanic deep low-frequency tremors

Characteristics of dilatational infrasonic pulses accompanying low-frequency earthquakes at Miyakejima Volcano, Japan

Latitude: 39 57'28" N, Longitude: '15" E, Elevation: 1,613 m (Hachimantai) (Triangulation Point)

Monthly Volcanic Activity Report (August 2015)

,**2, /,** ,*** 1, +3.* +.,**1 0,**

Relation between single very long period pulses and volcanic gas emissions at Mt. Asama, Japan

Depth estimation of fumarolic gas source deduced by fume pressure measurement

Monthly Volcanic Activity Report (March, 2011)

Seismic source mechanism of degassing bursts at Kilauea Volcano, Hawaii: Results from waveform inversion in the s band

Monthly Volcanic Activity Report (April 2015)

Introduction to Volcanic Seismology

Ascending seismic source during an explosive eruption at Tungurahua volcano, Ecuador

Continuously Monitored by JMA. Overview of Niigata-Yakeyama taken from the north side on September 29, 2003 by the Japan Meteorological Agency

Moment tensor inversion of near source seismograms

What Are Recorded In A Strong-Motion Record?

GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L19604, doi: /2004gl020366, 2004

Aseismic slip and low-frequency earthquakes in the Bungo channel, southwestern Japan

Continuing deflation by fumaroles at Kuju Volcano, Japan

Latitude: 34 31'13" N, Longitude: '45" E, Elevation: 508 m (Miyatsukayama) (Triangulation Point - Toshima)

Advanced Workshop on Evaluating, Monitoring and Communicating Volcanic and Seismic Hazards in East Africa.

PEAT SEISMOLOGY Lecture 12: Earthquake source mechanisms and radiation patterns II

DETAILED IMAGE OF FRACTURES ACTIVATED BY A FLUID INJECTION IN A PRODUCING INDONESIAN GEOTHERMAL FIELD

INTRODUCTION TO VOLCANIC SEISMOLOGY

Application of Phase Matched Filtering on Surface Waves for Regional Moment Tensor Analysis Andrea Chiang a and G. Eli Baker b

Evaluation of recent activity at Satsuma-Iwojima Felt earthquake on June 8, 1996

Journal of Volcanology and Geothermal Research

Aira/Sakura-jima. Kyushu, Japan N, E; summit elev m. All times are local (= UTC + 9 hours)

σ set = σ drop / N drop. (1)

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law

THREE-DIMENSIONAL FINITE DIFFERENCE SIMULATION OF LONG-PERIOD GROUND MOTION IN THE KANTO PLAIN, JAPAN

Continuous Caldera Changes in Miyakejima Volcano after Hiroyuki HASEGAWA, Hiroshi P. SATO and Junko IWAHASHI

VOLCANO MONITORING PRACTICAL. Hazard alert levels established for communication at Mt. Pinatubo

Latitude: 43 25'03" N, Longitude: '52" E, Elevation: 1,692 m (Maruyama) (Triangulation Point)

Continuously Monitored by JMA. Latitude: 36 47'55" N, Longitude: '33" E, Elevation: 2,578 m (Shiranesan) (Elevation Point)

RELATION BETWEEN RAYLEIGH WAVES AND UPLIFT OF THE SEABED DUE TO SEISMIC FAULTING

HEAT AND MASS TRANSFER PROCESSES AFTER 1995 PHREATIC ERUPTION OF KUJU VOLCANO, CENTRAL KYUSHU, JAPAN

Multi-parameter investigations at Fuego and Santiaguito volcanoes

MAGMATIC, ERUPTIVE AND TECTONIC PROCESSES IN THE ALEUTIAN ARC, ALASKA

Characteristics of Volcanic Activity at Sakurajima Volcanoʼs Showa Crater During the Period 2006 to 2011

Journal of Volcanology and Geothermal Research

Supporting the response to the 2018 lower East Rift Zone and summit collapse at Kīlauea Volcano, Hawaiʻi

Effects of Surface Geology on Seismic Motion

Kuchierabujima Volcano, Japan detected by repeated precise leveling surveys

Problems Associated with Activity Assessment, Dissemination of Information, and Disaster Response During the 2000 Eruption of Miyakejima

Effects of subsurface structures of source regions on long period ground motions observed in the Tokyo Bay area, Japan

A resistivity cross-section of Usu volcano, Hokkaido, Japan, by audiomagnetotelluric soundings

A - Piton de la Fournaise activity

Fault Length and Direction of Rupture Propagation for the 1993 Kushiro-Oki Earthquake as Derived from Strong Motion Duration

Fracture induced shear wave splitting in a source area of triggered seismicity by the Tohoku-oki earthquake in northeastern Japan.

Latitude: 42 49'36" N, Longitude: '41" E, Elevation: 1,898 m (Ezo-Fuji) (Elevation Point)

Three Dimensional Simulations of Tsunami Generation and Propagation

Continuously Monitored by JMA

An autocorrelation method to detect low frequency earthquakes within tremor

Source Characteristics of Large Outer Rise Earthquakes in the Pacific Plate

Eruptive fracture location forecasts from high-frequency events on Piton de la Fournaise Volcano

Name: Page 1. 2) What do increases and decreases in RSAM data indicate about magma movement inside the volcano?

An intermediate deep earthquake rupturing on a dip-bending fault: Waveform analysis of the 2003 Miyagi-ken Oki earthquake

Extending the magnitude range of seismic reservoir monitoring by Utilizing Hybrid Surface Downhole Seismic Networks

Rapid magnitude determination from peak amplitudes at local stations

Deep low-frequency earthquakes near the downward extension of the seismogenic fault of the 2000 Western Tottori earthquake

Spatio-temporal variation in slip rate on the plate boundary off Sanriku, northeastern Japan, estimated from small repeating earthquakes

Chapter 2 Multivariate Statistical Analysis for Seismotectonic Provinces Using Earthquake, Active Fault, and Crustal Structure Datasets

Geophysical comparison of the three eruptions in the 20th century of Usu volcano, Japan

Global Positioning System(GPS) Global Positioning System(GPS) is a new surveying technique applying space technology designed by the United States

Tsunami waveform analyses of the 2006 underthrust and 2007 outer-rise Kurile earthquakes

SUPPLEMENTAL INFORMATION

MEASUREMENT OF SURFACE DEFORMATION ASSOCIATED WITH HYDROFRACTURE

Depth-dependent slip regime on the plate interface revealed from slow earthquake activities in the Nankai subduction zone

Title. CitationJournal of Volcanology and Geothermal Research, 215- Issue Date Doc URL. Type. File Information.

Transcription:

Title Tilt change recorded by broadband seismometer prior Japan Author(s)Aoyama, Hiroshi; Oshima, Hiromitsu CitationGeophysical Research Letters, 35(6): L06307 Issue Date 2008-03 Doc URL http://hdl.handle.net/2115/59834 Rights Copyright 2008 American Geophysical Union. Type article File Information Aoyama-GRL2008.pdf Instructions for use Hokkaido University Collection of Scholarly and Aca

GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L06307, doi:10.1029/2007gl032988, 2008 Tilt change recorded by broadband seismometer prior to small phreatic explosion of Meakan-dake volcano, Hokkaido, Japan Hiroshi Aoyama 1 and Hiromitsu Oshima 2 Received 18 December 2007; revised 13 February 2008; accepted 18 February 2008; published 22 March 2008. [1] We discovered an anomalous very long period (VLP) signal (duration: about 60 s) in the broadband seismic record of the precursory swarm of a small phreatic explosion at the Meakan-dake volcano, Hokkaido, Japan, in 2006; this signal was masked by a short-period tremor in the original velocity waveform, but a considerable DC step appeared in the displacement trace. Since such a step results from a tilt change, we interpreted the observed trace as being a combination of rotational (tilt) and translational (displacement) motions. The apparent displacement step in the VLP signal indicates subsidence of the summit crater. Owing to limited data, we attempted deflating the Mogi source and simulating the seismic record including both motions. When only rotational motion is considered, the simulated waveform explains well the general features of the observed trace. However, the combined trace has incompatible phases, suggesting that a more complicated source process should be considered. Citation: Aoyama, H., and H. Oshima (2008), Tilt change recorded by broadband seismometer prior to small phreatic explosion of Meakan-dake volcano, Hokkaido, Japan, Geophys. Res. Lett., 35, L06307, doi:10.1029/2007gl032988. 1. Introduction [2] Phreatic explosions are triggered by interactions between magma and ground water at shallow depths; they are widely observed at many volcanoes. However, despite the frequency of field observations of phreatic explosions, it is very difficult to measure the precursory phenomena of such small explosions using physical monitoring techniques. Moreover, whether the precursors of phreatic explosion are entirely nonexistent or too small to record has been the subject of considerable debate. [3] Broadband seismic monitoring on and around active volcanoes has become popular in recent years. Furthermore, the high sensitivity of broadband seismometers has led to the frequent detection of anomalous very long period (VLP) signals around volcanoes [e.g., Kawakatsu et al., 2000; Kumagai et al., 2003; Nishimura et al., 2002; Chouet et al., 2003]. Some of these VLP signals have been recorded by using multiple seismometers, and waveform analyses has been attempted to reveal their excitation processes. 1 Institute of Volcanology and Seismology, Graduate School of Science, Hokkaido University, Sapporo, Japan. 2 Usu Volcano Observatory, Institute of Volcanology and Seismology, Graduate School of Science, Hokkaido University, Sobetsu, Japan. Copyright 2008 by the American Geophysical Union. 0094-8276/08/2007GL032988 [4] Using a broadband seismometer, we monitored a small phreatic explosion that occurred at the Meakan-dake volcano on 21 March 2006. The explosion was associated with two precursory seismic swarms; we discovered an anomalous VLP signal in the data of the first swarm. In this report, we describe the VLP signal observed about one month prior to the phreatic explosion and discuss its implications as a precursor to the explosion. 2. Explosions of Meakan-Dake and Precursory Swarms [5] The Meakan-dake volcano is an andesitic stratovolcano; it stands on the southwestern rim of Akan Caldera in eastern Hokkaido. Many fumaroles and hot springs are distributed on and around the volcano. The documented eruptions since 1957 have all been phreatic explosions that occurred at the Pon-machineshiri summit crater (Figure 1). The Meakan-dake volcano has produced four phreatic explosions in the last 20 years (Figure 2). On the occasion of the 1996 explosion, precursory seismic swarms similar to the 2006 activity were observed [Okada et al., 1997]. The 1996 eruption opened a new small crater on the southern rim of the Pon-machineshiri crater but did not affect the neighboring fumarolic activity. In addition, an increase in seismic activity was also observed just before the 1988 eruption [Okada et al., 1997]. [6] The small phreatic explosion that occurred in the early morning on 21 March 2006 caused the creation of new small pits around the northwestern perimeter of the Akanuma summit crater, which occupies the northern half of the Pon-machineshiri crater (Figure 1). Before the explosion, three separate earthquake swarms were observed by a continuous monitoring network: The first swarm continued for about five days from 18 to 22 February, comprising more than 1250 earthquakes; this swarm occurred after two years of seismic quiescence. About 20 days later, on 11 March, the second swarm began, and subsequently, the earthquakes increased in frequency once again one day prior to the phreatic explosion on 21 March. [7] Institute of Seismology and Volcanology (ISV), Hokkaido University, operates a few seismic stations at the western foot of the Meakan-dake volcano (Figure 1). During the active period in 2006, a Guralp CMG-40T (0.033 50 Hz) was installed at the Nonaka Onsen spa (hereinafter, MEA), which is about 2.0 km west-northwest of the Akanuma crater. The sensor outputs were digitized by a Hakusan LT8500 data converter (22-bit precision) oper- L06307 1of7

Figure 1. Contour map of Meakan-dake volcano showing location of seismic station (black circle). Contours represent 50-m elevation intervals. Dark and light gray spots are crater areas made by the 2006 and 1996 explosions, respectively. Inset shows location of Meakan-dake volcano. ating in continuous mode at 200 samples s 1, and were then transmitted to ISV thorough a telephone line. 3. Tremor Overlapped With VLP Signal and Earthquake [8] The first swarm comprised small earthquakes and two volcanic tremors. The volcanic tremors occurred at around 0552, 19 February, and at 0550, 20 February. The largest earthquake during this swarm overlapped with the first tremor, and soon after the first tremor, the number of earthquakes began to decrease. [9] We discovered an anomalous VLP signal in the seismic trace of the first tremor observed at MEA. Since the velocity amplitude of the earthquake was about 4 times larger than that of the tremor, this earthquake is remarkable in the original data (Figure 3a). On the other hand, the displacement waveforms obtained by time integration obviously show different features (Figure 3b): a large step appears in both horizontal components; moreover, the offset in the east-west component is about twice as large as that in the north-south component. This VLP signal continues for about 60 s, and the remarkable offset remains in the displacement waveforms. Since the pass-band frequency of the CMG-40T at MEA is restricted to above 0.033 Hz, this long-period offset does not reflect the actual displacement. 4. Estimation of Tilt Change From Seismogram [10] When a seismometer leans slowly, it is expected that a VLP signal is recorded in its horizontal components due to the effects of gravitational acceleration. Rodgers [1968] analyzed the pendulum response of a traditional inertial seismometer to a tilt change and obtained the relation between the displacement response and tilt response as a function of the angular frequency of the input signal. Recently, Graizer [2006] described the complete equation of motion for pendulums and summarized the results of shaking tests on accelerographs. These reports suggest that the VLP signal observed by the CMG-40T seismometer in the present study is quite likely to be the pendulum response to a tilt change. [11] The gravitational force experienced due to a tilt change acts as additional acceleration on pendulums [Graizer, 2006]. When the tilt change is small, the additional acceleration can be approximated by gy, where g is the gravitational acceleration and y is the tilt angle. Therefore we can expect the following relationship to hold between the tilt step response function C T (w) and the velocity impulse response function C v (w) for the CMG-40T: C T ðwþ ¼ gy w 2 C vðwþ; where w is the angular frequency of the input signal. C v (w) can be calculated using the distribution of poles and zeros, p i, z i (i =1 3), for the CMG-40T as, ð1þ ð C v ðwþ ¼ 0:314 iw z 1Þðiw z 2 Þðiw z 3 Þ ðiw p 1 Þðiw p 2 Þðiw p 3 Þ : ð2þ With this relation, we can calculate the theoretical amplitude of the response to a tilt step change of 1 mrad, jc T (w)j, as shown in Figure 4a. As shown, a constant amplitude response is expected in the frequency range below 0.033 Hz, which is the natural frequency of the CMG-40T. The lower frequency limit in equation (2) gives us a theoretical value of 0.223 mm/mrad (= 4.48 mrad/mm). This amplitude can be used as a transfer factor between the apparent displacement due to tilting and the angle of tilt change. Equation (1) has the same implication at that of equation (41) from Rodgers [1968]. 2of7

Figure 2. Bar graphs indicating number of earthquakes counted by JMA Sapporo: (top) daily frequency since 1973; (middle) daily frequency since 2003; (bottom) hourly frequency during February March 2006. Triangles indicate occurrence of explosions. [12] Thus, to confirm the actual tilt response of the CMG- 40T as expressed in equation (1) and the transfer factor, we performed a simple test on the tilt response of the seismometer [Aoyama, 2008]. A seismometer and tiltmeter are set up together on the same thick metal plate, which is placed on a thin spacer. With this setup, when the spacer under the edge of the plate is pulled out, an instantaneous tilt change occurs for the metal plate, with negligible associated translational motion; thus this test affords the tilt step response of the instruments. For this test, we used a Model 701-2 biaxial tiltmeter (Applied Geomechanics Inc.) to monitor the absolute value of the applied tilt change. [13] Example waveforms obtained from the CMG-40T and tiltmeter in this test are shown in Figure 4b. Figure 4b confirms that the test actually generated a step change in the tilt angle. The resultant velocity amplitude spectra of the horizontal component seismograms are shown in Figure 4a. Except for the noisy frequency range above 1 Hz, the experimental spectra and theoretical response are quite similar. In the frequency band below 0.003 Hz, the amplitude is independent of the frequency, while above this value, it is proportional to w 2. The transfer factor experimentally determined from the measured tilt angle and apparent displacement is 4.01 mrad/mm for this example. We repeated this test eight times using different step amplitudes for the tilt change and obtained an average transfer factor of 4.3 mrad/mm with standard deviation of 0.25 mrad/mm, which corresponds well to the theoretically expected value of 4.48 mrad/mm. This result confirms that the angle of tilt change can be estimated from the seismogram by a CMG-40T insofar as the inclination of the seismometer is small. [14] In Figure 3b, the apparent amplitudes of displacement offset, dn and de, are about 0.08 mm and 0.19 mm, respectively, and therefore the amplitude ratio between the north-south and east-west components is dn/de 0.42. This means that the downward dipping vector is approximately toward east-southeast (Figure 3c). The summit crater is located about 2.0 km east-southeast of MEA. The offset in the displacement seismogram can be regard as subsidence around the summit area during the first tremor. By using the theoretical transfer factor of 4.48 mrad/mm, the final tilt change is calculated to be 0.94 mrad from the apparent amplitude (0.21 mm) of the displacement offset. 5. Discussion [15] What occurred beneath the volcano before the explosion is intriguing problem, both in terms of volcano research and explosion forecasting. Although the data we now have are very limited in the number of components 3of7

Figure 3. (a) Horizontal ground velocity at MEA. (b) Displacement traces obtained by simple integration. (c) Maximum downward tilt direction estimated from displacement traces. measured at one station, the seismic data does suggest subsidence of the volcano during the volcanic tremor. [16] Volcanic uplift and subsidence are often explained by volumetric changes in a chamber under the volcano. One of the simplest and the most conventional models for determining the volume change associated with volcanic activity is the Mogi model [Mogi, 1958; Yamashina, 1986]. [17] The expected tilt change g due to volume change in the ground is given by, g ¼ 9 Dr DV; ð3þ 4 pðd 2 þ r 2 2:5 Þ where DV is the volumetric change at the source; D, the source depth relative to the station; and r, the horizontal distance between the source and station. According to the hypocenter relocation analysis performed by Sapporo District Meteorological Observatory, Japan Meteorological Agency (JMA Sapporo), the source area of the first swarm is located under the western perimeter of the Akanuma crater at a depth around 800 m below sea level (J. Funasaki et al., unpublished data, 2006). Here we assume D to be 1500 m and r to be 2000 m and let g be 0.94 mrad; then, the volume decrease required to cause the subsidence of the volcano during the tremor is estimated to be about 4.3 10 4 m 3. [18] However, when such deflation occurs, both tilt change and translational motion are caused. The expected radial displacement u r due to a Mogi source is given by, u r ¼ 3 r DV : ð4þ 4 pðd 2 þ r 2 1:5 Þ The volume decrease produces a translational motion of 1.3 mm. It is expected that the seismometer also recorded this translational motion. [19] Here we simulate the combined waveform, considering both translational and rotational motion. The assumed time function for the volume change at the source is shown in Figure 5a. Then, the tilt change and radial displacement can be expressed as shown in Figures 5b and 5e, respectively. For rotational motion, the apparent displacement due to tilt change with the CMG-40T tilt response is shown in Figure 5c; its time derivative yields the apparent velocity trace shown in Figure 5d. For translational motion, Figure 5f shows the radial ground velocity, and Figure 5g shows the velocity output from the CMG-40T. The combined velocity waveform (Figure 5h) can be obtained by summing Figures 5d and 5g; integrating this combined waveform yields the simulated displacement waveform shown in Figure 5i. When only rotational motion is considered, the simulated waveform (Figure 5c) explains well the general features of the observed trace (Figure 3b). On the other hand, the combined displacement trace in Figure 5i has incompatible phases when compared with Figure 3b. However, a more complicated source model such as crack deflation or volume change of dual sources may be able to better explain this phenomenon. 4of7

Figure 4. (a) Amplitude response curve of CMG-40T. Broken line is tilt step response to 1 mrad. Solid lines express velocity amplitude spectra calculated from tilt step experiment data. (b) Waveform examples from the tilt test: (top) outputs of the biaxial tiltmeter; (middle) velocity outputs of CMG-40T; (bottom) displacements of the horizontal components calculated from the velocity traces. [20] Finally, to interpret what went on under the volcano during the tremor, we return to the original displacement waveform of the CMG-40T. Just after the beginning of the tremor, a small step about 0.01 mm can be seen in both horizontal components. Since the duration of this step is on the order of a few seconds, it may correspond to the translational displacement at MEA. However, despite the low-cut frequency response of the CMG-40T, a DC component remains in the displacement trace after the small step. This suggests that the tilt change began right after the small step. Then about 10 s later, the largest earthquake occurred, and soon after, the VLP signal appeared. These processes suggest the following chain of events: an initial rapid pressure disturbance in a fluid chamber induced the large earthquake close to the chamber, and then the large volumetric change began. Phase change of the volcanic fluid and fluid migration may be candidates for the volumetric change estimated here. 6. Summary [21] A small phreatic explosion occurred at the Meakandake volcano on 21 March 2006, accompanied precursory earthquake swarms. We discovered an anomalous VLP signal in a seismogram obtained at the foot of the volcano during a tremor in the first swarm, which occurred about one month before the explosion. The VLP signal could be interpreted as a tilt change due to the subsidence of the volcano. The offset in the displacement trace could be explained by an isotropic chamber deflation of about 4.3 10 4 m 3. Then, we simulated the seismogram of the CMG-40T, considering both rotational and translational motions. However, the inconsistency between the simulated 5of7

Figure 5. Simulation procedure for synthetic seismogram: (a) assumed volume change of the source; (b) tilt change at the station; (c) apparent displacement due to tilt change considering tilt response of CMG-40T; (d) apparent velocity waveform; (e) translational displacement; (f) translational velocity; (g) velocity waveform considering velocity response of CMG-40T; (h) superposed velocity waveform; (i) simulated displacement seismogram. and observed waveforms suggests that a more complicated source model should be considered. [22] It should be emphasized that the seismic data before the small phreatic explosion suggests the existence of precursory activity of volcanic fluid. Broadband seismic observation is very useful not only for studies on source processes but also for precise monitoring of volcanic activity. However, the data presented herein warns of the possibility of contamination due to tilt change, especially for long-period signals. [23] Acknowledgments. We are grateful to J. Funasaki and the staff of JMA Sapporo for providing us the results of relocation analysis. Technical staff M. Okayama and M. Takada kindly permitted the use of a lathe for the laboratory experiments on the CMG-40T and 701-2. References Aoyama, H. (2008), Simplified test on tilt response of CMG40T seismometers (in Japanese with English abstract), Bull. Volcanol. Soc. Jpn., 53, 35 46. Chouet, B., P. Dawson, T. Ohminato, M. Martini, G. Saccorotti, F. Giudicepietro, G. De Luca, G. Milana, and R. Scarpa (2003), Source mechanisms of explosions at Stromboli Volcano, Italy, determined from moment-tensor inversions of very-long-period data, J. Geophys. Res., 108(B1), 2019, doi:10.1029/2002jb001919. Graizer, V. (2006), Tilts in strong ground motion, Bull. Seismol. Soc. Am., 96, 2090 2102. Kawakatsu, H., S. Kaneshima, H. Matsubayashi, T. Ohminato, Y. Sudo, T. Tsutsui, K. Uhira, H. Yamasato, H. Ito, and D. Legrand (2000), Aso94: Aso seismic observation with broadband instruments, J. Volcanol. Geotherm. Res., 101, 129 154. Kumagai, H., K. Miyakawa, H. Negishi, H. Inoue, K. Obara, and D. Suetsugu (2003), Magmatic dike resonances inferred from very-long-period seismic signals, Science, 299, 2058 2061. 6of7

Mogi, K. (1958), Relation between the eruptions of various volcanoes and the deformations of the ground surfaces around them, Bull. Earthquake Res. Inst. Univ. Tokyo, 36, 99 134. Nishimura, T., S. Ueki, T. Yamawaki, S. Tanaka, H. Hashino, M. Sato, H. Nakamichi, and H. Hamaguchi (2002), Broadband seismic signals associated with the 2000 volcanic unrest of Mount Bandai, northeastern Japan, J. Volcanol. Geotherm. Res., 119, 51 59. Okada, H., A. Suzuki, T. Maekawa, H. Y. Mori, and Y. Nishimura (1997), Preliminary report on the November, 1996 eruption at Mt. Meakan-dake, Hokkaido (in Japanese with English abstract), Geophys. Bull. Hokkaido Univ., 60, 131 144. Rodgers, P. W. (1968), The response of the horizontal pendulum seismometer to Rayleigh and Love waves, tilt, and free oscillations of the Earth, Bull. Seismol. Soc. Am., 58, 1384 1406. Yamashina, K. (1986), Stress fields and volcanic eruptions (in Japanese with English abstract), Bull. Volcanol. Soc. Jpn., S30, 101 119. H. Aoyama, Institute of Volcanology and Seismology, Graduate School of Science, Hokkaido University, N10W8, Kita-ku, Sapporo, Hokkaido, 060-0810, Japan. (aoyama@uvo.sci.hokudai.ac.jp) H. Oshima, Usu Volcano Observatory, Institute of Volcanology and Seismology, Graduate School of Science, Hokkaido University, Tatsuka 142, Sobetsu, Hokkaido, 052-0106, Japan. (oshima@uvo.sci.hokudai.ac.jp) 7of7