Вы находитесь на странице: 1из 14

Bulletin of the Seismological Society of America, Vol. 97, No. 2, pp. 591604, April 2007, doi: 10.

1785/0120060095

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico
by F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Abstract Colima city is the capital of the Mexican federal state of the same name. It is located close to the Pacic coast and is subjected to a large seismic risk. We present a microzonation study in this city, based on microtremors using single-station and array measurements. We applied horizontal-to-vertical spectral ratios (HVSR) analysis to single-station measurements at 310 sites within the city, concentrating measurements in zones that were damaged by the January 2003 (M 7.4) earthquake. The results show that a seismic zonation based exclusively on single-station microtremor measurements is not a reliable alternative when the local geology is complex and site effects are not the result of a single-impedance contrast. For this reason, we applied two independent analysis techniques to array measurements of microtremors: the spatial autocorrelation (SPAC) method and the refraction microtremor (ReMi) method. We used linear arrays to record 25-sec microtremor windows at eight sites within the city, which were analyzed with those two techniques. The result of both techniques of analysis is a phase-velocity dispersion curve, which can be inverted to obtain a shallow S-wave velocity prole. Two of the sites were the location of shallow (50 m) boreholes, where P- and S-wave velocity proles were measured using a P-S suspension log. The phase-velocity dispersion curves obtained from the ReMi and SPAC analyses of the microtremor records showed very good agreement. The velocity proles inverted from the phase-velocity dispersion curves showed good agreement with the suspension logging measurements at one of the two sites where they were available and poor agreement at the other site. The transfer functions computed from the inverted soil proles are in good agreement with previous estimates of local amplication from spectral ratios analysis of earthquake records. Our results are compatible with previous indications of site effects and explain the failure of single-station microtremor measurements when the concept of dominant frequency loses its meaning. Finally, we propose an estimate of local site amplication at the city of Colima, which will be useful for future predictions of ground motion at this city.

Introduction
Damage distribution during large earthquakes is frequently controlled by site effects. Subsoil impedance contrasts can signicantly amplify the shaking level, as well as increase the duration of strong ground motion. The larger cities around the world have already been the subject of microzonation studies, where the different levels of groundmotion amplication are measured throughout the city. However, especially in developing countries, a very signicant effort has yet to be made. Seismic microzonation has been based on observational studies, where ground-motion amplication is measured by using spectral ratios of small events (Borcherdt, 1970; Cha vez-Garc a et al., 1990). In recent years, though, a wealth of studies have been based on ambient vibration 591 (microtremor) records, given the ease and low cost with which these data can be obtained in regions of moderate to low seismicity. In early studies, the site resonant frequency was deduced from spectral ratios of microtremor records (used in the same way as earthquake records, e.g., Kagami et al., 1986; Seo, 1992), or it was taken to be the frequency of the peak of the Fourier amplitude spectrum of horizontal components (e.g., Kobayashi et al., 1986; Gutie rrez and Singh, 1992). Later, the use of spectral ratios computed between horizontal components relative to the vertical component recorded simultaneously (HVSR) became very popular (e.g., Nakamura, 1989; Lermo and Cha vez-Garc a, 1994; Field and Jacob, 1995, among many others. See, for example, the review article by Bard, 1999.). It is now gen-

592 erally recognized that the HVSR technique provides a reliable estimate of the resonant frequency. Some authors have also shown that, in some cases, the amplitude of that ratio is a good estimate of the site maximum-amplication value relative to bedrock motion. A consensus regarding the use of HVSR to estimate maximum amplication does not yet exist. Microzonation efforts require inexpensive techniques, such as HVSR. It is important, however, to understand the limitations of HVSR, and establish some guidelines to know when the results from this technique can be considered reliable. Recent articles have shown many examples where this technique was used with prot (e.g., Toshinawa et al., 1997), but it is clear that it is not a cure-all. A few reports show examples where HVSR was not useful (e.g., Volant et al., 1998), or where the dominant frequency was correctly determined but the associated amplitude was ineffective to estimate the relative local amplication (Malagnini et al., 1996). It seems clear that when we observe amplication due to a single, large impedance contrast between a soft soil layer and its basement, HVSR is reliable. It is when amplication is caused by more complex local geology that the usefulness of HVSR becomes problematic and the question of its reliability is posed more acutely. In this context, the city of Colima is an interesting case study. Colima is located near the Pacic coast of Mexico (see Fig. 1). This city is the capital of the federal state of the same name. Because its current population is only about onehalf million, Colima has not received much attention from the seismological community. However, this city is located close to an active subduction zone and has been affected

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

repeatedly by destructive earthquakes. For example, Colima state was affected by the Tecoma n earthquake of 21 January 2003, which caused 21 casualties and about 90 million U.S. dollars in damage (Cenapred, 2003), most of which occurred in the capital city. Unfortunately, it was not possible to correlate observed damage with ground motion, as no strongmotion station was in operation at the time of that earthquake, and the collapsed structures showed blatant design errors, making it impossible to use them to estimate the differences in ground-motion intensity throughout the city. Colima is located on a thick (about 800 m) sequence of volcanic deposits, consisting of a mixture of avalanches, lahar deposits, and reworked volcanic sediments. Previous seismic experiments have measured amplication due to site effects as large as a factor 6 between 1 and 3.5 Hz (Gutie rrez et al., 1996). Even if large earthquakes do occur in this region, the seismicity rate is much lower than that observed further south along the subduction zone. This makes it difcult to base microzonation efforts on earthquake records obtained with temporary networks. Moreover, the two previous attempts at microzonation of Colima (Lermo et al., 1991; Gutie rrez et al., 1996) produced contradictory results. In this article we analyze ground motion within Colima city and its relation with subsoil ground conditions. We reappraise the results of previous experiments and have made additional measurements. We measured microtremor records using single-station measurements at 310 points within the urban area. Given the small size of the city, this means a large density of measurements throughout. In addition, we used microtremor measurements recorded using an array of

Figure 1.

Location of Colima city, Mexico, and its regional geology.

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

593

geophones. These data were analyzed by using two different techniques: the SPatial AutoCorrelation (SPAC) method (Aki, 1957) adapted to measurements using single-station pairs (Cha vez-Garc a et al., 2005, 2006), and the refraction microtremor method (ReMi) introduced by Louie (2001). Our purpose is to assess the usefulness of HVSR in the volcanic geology of Colima city, both as standalone method and when complemented by microtremor array measurements. We compare our results with those from the previous studies. Our results conrm that in a volcanic environment the usefulness of HVSR decreases. In our case study, HVSR suggests that some site effects are present, but it is unable to constrain their spatial distribution. The use of the two array techniques is more fruitful. We obtain phase-velocity dispersion curves from which shear-wave velocity proles are inverted. These proles are consistent with results from suspension logging at two sites. Our results agree with previous site-amplication estimates and allow us to propose a family of 1D soil proles throughout the city. We do not observe a close relation between surface geology and site response. This probably means that the differences between the outcropping geological formations make sense in terms of the emplacement process but do not reect signicant variations in the mechanical properties of the volcanic deposits. For this reason, we are unable to separate zones within the city with homogeneous expected shaking, but our results allows us to explain the observed effects with a model that can be used to predict ground motion for future large earthquakes.

Background
Geology Colima state is located 32 km to the south of the Colima Volcanic Complex (CVC), which itself is in the western Trans-Mexican Volcanic Belt (TVB). The CVC consists of three andesitic stratovolcanoes (Cantaro, Nevado, and Fuego de Colima), which dene a volcanic chain with a north south orientation as the result of the migration of volcanic activity due to the subduction of the oceanic plate beneath the American continent. The Fuego de Colima is one of the most active volcanoes in Mexico. The city of Colima is built over the volcanic sequences produced by the CVC, which overlay a late cretacic limestone basement outcropping east and west of the city. These volcanic sequences include materials of different ages and from different depositional processes. Geologists have identied and dated four avalanche deposits aged 1800 to 2500 years, and many more uviolaharic and debris ow deposits in between. Three types of deposits can be mapped within the city (Fig. 2). Volcanic Debris Avalanche. These are massive deposits that consist of andesitic rubble, mainly between 5 and 20 cm diameter, but with some boulders as large as 1 m. These deposits have great thickness and cover large areas. They are

Figure 2. Surface geology within Colima city. The main streets and rivers are shown with solid lines as reference. The urban zone is delimited by the rings formed by the main streets.

produced by the total or partial collapse of volcanic edices of the CVC. The blocks are cemented by small quantities of a clay and sand matrix, and present characteristic irregular cracks. Within Colima, avalanche deposits crop out in its northern half. Their total thickness is about 600 m. Volcanic Debris Flows. These are massive deposits (on the order of several meters) consisting of andesitic, rounded blocks within a compact sandy matrix. They are the result of the transportation of the avalanche deposits by subsequent water ows. The thickness of these deposits around Colima is between 20 and 30 m, and they crop out to the south of the city.

594 Lahars, Lacustrine Sand, and Gravel Deposits. These are stratied deposits a few centimeters to a few meters thick, consisting of andesitic blocks within a sandy matrix. They are about 50 m thick within the city, where they crop out mainly in the western half. Previous Microzonation Studies The importance of the city of Colima and the past occurrence of large earthquakes have spurred previous attempts at microzonation of the city. The rst one was carried out by Lermo et al. (1991). They estimated amplication at four sites from standard spectral ratios (Borcherdt, 1970) using data from a single, small earthquake. They found a dominant period of 0.22 sec with an amplication factor of 2 on the avalanche deposits (downtown), and a dominant period of 0.15 sec with an amplication of 4 on the uvial deposits. However, their reference station was located north of the city, on volcanic deposits similar to those that crop out at Colima city, making their amplication values untrustworthy. In addition, Lermo et al. (1991) measured microtremors at 36 sites, along two perpendicular lines across the city and estimated the dominant period as that for which Fourier spectra of the microtremor measurements had their maximum. They found values between 0.25 and 0.33 sec on the volcanic avalanche, and above 2 sec on the uvial deposits. Figure 3a reproduces the dominant period map presented by Lermo et al. (1991). A few years later, and in part because of the occurrence of a large (M 7.9) event in 1995, Gutie rrez et al. (1996) carried out a second attempt at the microzonation of Colima. These authors installed a temporary seismic network of digital PRS-4 seismographs by Lennartz, coupled to threecomponent 1-Hz sensors. They successfully recorded a few small events (M 4.5). They used a seismic station on limestone as reference (10 km to the east of the city) to evaluate relative amplication using spectral ratios. Their empirical transfer functions show signicant amplication on the avalanche deposits (up to a factor of 6), distributed about a wide-frequency band, without well-marked peaks. Sites on uvial deposits showed smaller amplication (between a factor of 2 and 5) and dominant periods between 0.3 and 0.8 sec. In addition, Gutie rrez et al. (1996) measured microtremors at 57 sites, and estimated the dominant period from peak Fourier amplitude spectra. They observed dominant periods of about 0.3 sec, similar to those of Lermo et al. (1991) and proposed a second dominant period map, where contours were drawn (Fig. 3b). Finally, Gutie rrez et al. (1996) measured P- and S-wave velocities at two shallow (50 m) boreholes by using suspension logging. The measurements for the rst 20 m were unreliable, however.

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Analysis Techniques, Data Used, and Results


In the next paragraphs, we describe in brief the three techniques we use to analyze microtremor records. They are

(a) Dominant-period values in seconds measured at Colima city by Lermo et al. (1991) from microtremor measurements. These authors did not draw contours from their measured values. (b) Contour map of the dominant period in seconds measured at Colima city by Gutie rrez et al. (1996) from microtremor measurements.

Figure 3.

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

595

the HVSRs, an innovative approach to the SPAC method, and the ReMi method. In addition, we present the experiments and the results obtained in each case.
HVSR

Spectral ratios of horizontal components relative to the vertical recorded simultaneously have been widely used to determine site response from ambient-vibration records. The two previous microtremor studies in Colima interpreted dominant period from Fourier amplitude spectra maxima. In addition the number of measurements made was small. For those reasons, we recorded single-station microtremors at 310 sites within the Colima urban zone (open and lled circles in Fig. 4). A Kinemetrics K2 recorder with triaxial accelerometers was used. At each site 3 to 5 min of ambient vibration was recorded. From each record, we selected a 1-min window in which the records showed the smallest number of transitory signals and the noise appeared most stationary. We computed the spectral ratio between horizontal and vertical motion from the selected window. The horizontal components were combined by using a simple mean. Despite the careful window selection, most of the records did not produce a useful HVSR; the resulting ratios showed no clear peak. From the 310 measured sites, only 125 (lled circles in Fig. 4) produced HVSR where a peak could be identied, and from which a value of dominant period and maximum amplication were determined. Figure 5 shows an example of an HVSR with a clear peak and another that was rejected because no clear peak could be identied. The 125 retained values of dominant period and maximum relative amplication, as measured from the HVSR, were used to draw the contours shown in Figure 6. The maximum amplication values vary between 2 and 5, although the vast majority are smaller than a factor of 2. These values are consistent with the maximum amplication estimates by Gutie rrez et al. (1996), given that HVSR of microtremor records usually underestimates amplication (Bard, 1999), but are clearly not very signicant. Moreover, the contours of dominant period values are not correlated with surcial geology and do not coincide with either the map of Lermo et al. (1991) or that of Gutie rrez et al. (1996) shown in Figure 3. We have used this standard technique with as many sites as possible within Colima, and are convinced that Figure 6 reects limitations in the HVSR method. These limitations must result from the absence of a clear-cut impedance contrast as the origin of the local amplication, something also apparent in the transfer functions determined by Gutie rrez et al. (1996).
SPAC

Figure 4. Location of the points where singlestation microtremor measurements were carried out. A total of 310 points are shown (open circles), from which only 125 were retained for the determination of dominant period and local amplication (lled circles). Stars indicate the location of SPAC/ReMi measurements. The location of the two shallow boreholes is indicated with open squares.

at the same interstation distance, r, and sampling many different azimuths at the recording site. The correlation coefcients, q(r, x), as a function of frequency x, are computed as the normalized cross-correlation between all station pairs separated a distance r and averaged over all azimuths, h. Aki (1957) showed that q(r, x) 1 2p (r 0, x)

2p

(r, h, x) dh

(1)

J0

c(x)
rx

Aki (1957) proposed the SPAC method almost 50 years ago. As presented in that publication, the method requires ambient-noise records obtained in a circular array of stations, with one station at the center. This geometry allows the computation of the cross-correlation between many station pairs

where (r 0, x) is the average autocorrelation function at the center of the array, (r, h, x) is the cross-correlation function between the records obtained at coordinates (r, h) and the record obtained at the center of the circle, c(x) is the phase velocity at frequency x at the site, and J0() is the

596

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Figure 5. Examples of the results obtained using HVSR with single-station microtremor measurements. (a) The result for a site where a clear peak can be observed at a period slightly greater than 0.2 sec. (b) The result for a site where no signicant peak can be observed.

Figure 6.

Contour maps of dominant period in seconds (a) and relative amplication derived from 125 single-station microtremor measurements, analyzed using HVSR (b).

Bessel function of rst kind and order zero. In this equation, the only unknown is the phase velocity, c(x), which can be obtained from the inversion of the correlation coefcients. The subsoil structure can be deduced from the inversion of the phase-velocity dispersion curve following standard procedures (e.g., Herrmann, 1987). The details of the method have been presented in several publications (e.g., Asten, 1976; Chouet et al., 1998). Cha vez-Garc a et al. (2005) presented an extension of SPAC, in which phase-velocity dispersion curves were obtained from data recorded using a temporary seismic array

with a very irregular geometry. The basic difference with respect to Akis (1957) approach was to substitute the temporal averaging for the azimuthal averaging required by the method. Cha vez-Garc a et al. (2005) showed a comparison between correlation coefcients computed for a singlestation pair with those computed using an azimuthal average at approximately the same interstation distance. The results indicated that the substitution of temporal averaging for the aziuthal average required by the SPAC method is valid. The good results obtained led the same authors to apply SPAC with an array of stations as different as possible from a circle,

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

597

a line of stations (Cha vez-Garc a et al., 2006). The results were again very good, further supporting the use of the SPAC temporal averaging method. We carried out measurements at eight locations throughout the city (shown as stars in Fig. 4, coinciding with singlestation measurements sites), sampling the different surcial formations. We used an Oyo Geospace DAS-1 exploration seismograph with a 24-bit dynamic range and a line of 12, vertical-component, 4.5-Hz natural frequency geophones. The sampling rate was 2 msec. This system had a at response for velocity between 4.5 and 250 Hz. At each location, the geophones were installed with a 6-m distance between them, giving a total length of 66 m, and ve time windows of about 25 sec of ambient vibration were recorded. We veried in the eld that the power spectral density for all 12 traces was comparable, thus ruling out the possibility of including signals in the analysis that were not common to the whole geophone spread. Following Cha vez-Garc a et al. (2006), we considered all possible station pairs to compute correlation coefcients. The recorded data were baseline corrected and tapered over 10% of their duration. They were then ltered using a set of 38 Butterworth bandpass lters, 1 Hz wide, between 3 and 40 Hz. The correlation coefcient for each frequency was computed by using the ltered traces as the average of the normalized, zero-lag cross-correlation for eight 3-sec windows extracted from the ltered records. These computations were repeated for all possible station pairs for each site, and the results at the same interstation distance were averaged for all ve 25-sec microtremor windows recorded. An analysis of the range of validity of the measurements (Rodr guez and Cha vezGarc a, 2006) indicated that our results are reliable in the range from 5 to 20 Hz. Figure 7 shows an example of the results. The mean and standard deviation correlation coefcients are given as a function of frequency for all station pairs analyzed from the records at location Parque (see Fig. 4). In the SPAC method, each interstation distance is useful to constrain phase velocities for a different wavelength. As we treat each station pair independently, data from a single linear array give us results for the 11 different interstation distances shown in Figure 7. We observe that, in all cases, the coefcients follow the shape of a zero order, rst kind Bessel function, as they should according to equation (1). This suggests that it is correct to assume the equivalence between the azimuthal averaging included in the initial proposal of SPAC (Aki, 1957) and the temporal averaging proposed in Cha vezGarc a et al. (2005, 2006), where a more detailed validation has been presented. A similar result was presented by Ohori et al. (2002) using microtremor measurements obtained using T-shaped arrays, although these authors do not explain how they circumvented the requirement of the azimuthal average. It may be surprising that we get good results from the linear SPAC method using only ve 25-sec window mea-

surements. Cha vez-Garc a et al. (2005) used several days of continuous microtremor measurements, whereas Cha vezGarc a et al. (2006) analyzed 30-min windows of ambient noise. The length of the records necessary to be able to substitute temporal averaging for the azimuthal average required by the SPAC method has not been established and most likely it is site dependent. Relative to our previous articles, we note that the frequencies analyzed in this article are higher, implying many cycles even in short time windows. In addition, we have averaged the results of all station pairs at the same interstation distance. This means that, for each window recorded by one of our arrays, the result for 6 m interstation distance, for example, was obtained as the average of 11 correlation coefcients between different station pairs. We must mention, however, that the good results obtained with such short time windows here may be not representative of other geologic or geographic settings. ReMi The ReMi method, introduced by Louie (2001), is based on the p-f (ray parameter-frequency) transformation described by McMechan and Yedlin (1981), applied by Mokhtar et al. (1988), and programmed in Herrmann (1987). This transformation permits the separation of the different waves composing the records obtained in an array of stations, according to their different apparent velocity through the array. The Louies innovation was in the application of this transformation to ambient vibration records obtained using a standard exploration seismograph, without any seismic source. The p-f transformation allows stacking all the recorded waves according to their apparent wavelength. If a large component of the recorded wave eld consists of Rayleigh waves (for vertical-component geophones) it is possible to identify their phase-velocity dispersion as a function of frequency from the image produced in the p-f plane. The interpretation of the images obtained from the ReMi method, however, is not straightforward. The maximum values in the image would correspond to the Rayleigh phase-velocity dispersion curve if the microtremor wave eld were traveling in the direction of the linear array of geophones. The recorded wave eld, however, includes Rayleigh waves propagating with similar power in many different directions. If this were not the case, we would not have obtained good results from the SPAC method, where a requisite is the presence of similar energy propagating in different directions. Thus, the condition that brings about the success of SPAC makes the interpretation of the results from ReMi more problematic. Stephenson et al. (2005) proposed to choose, for each frequency, the average value between the slowness where the power density is maximum and the slowness value for which the power density basically becomes zero. This point corresponds to the slowness value for which the spatial coherence between the records becomes insignificant (Rodr guez and Cha vez-Garc a, 2006). Another possibility is choosing the peak of the derivative with respect

598

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Figure 7. Example of the correlation coefcients as a function of frequency computed for the records obtained at site Parque (Fig. 4). Each diagram shows the average (symbols) and standard deviation (error bars) of the correlation coefcients computed for station pairs at the indicated distance. The value of dx is the distance between geophones, equal to 6 m.

to slowness of the ReMi image on the large slowness ank of the peak. The data used for ReMi were the same records employed for the SPAC analysis; ve 25-sec ambient vibration records (with a 2-msec sampling) recorded using the exploration seismograph at eight sites within the city (see Fig. 4). The ReMi images obtained from each 25-sec window were stacked to improve the signal-to-noise ratio. To facilitate the picking of the dispersion curve, we smoothed the stacked image on the p-f plane with two successive rectangular windows, the rst was 11 points long in the frequency direction, and the second was 5 points long in the slowness direction. Figure 8 shows, for example, the resulting ReMi images obtained at the two borehole sites in Colima. The open squares indicate the manual pick of the slowness-frequency curve following the suggestion by Stephenson et al. (2005). The lled squares show the machine picks at the maximum, for each frequency, of the derivative of the image with respect to slowness. We observe good agreement between the two picks, which were made independently. A similar agreement was obtained for all eight sites in the frequency range where the results can be considered reliable. Figure 8 shows clearly the limits of the method; we obtain useful phase-velocity

dispersion only in the frequency range from 4 to 18 Hz, in agreement with Louie (2001).

Discussion
We derived contour maps of dominant period and maximum amplication using the results from HVSR. The dominant period values and the maximum-amplication values are similar to those observed in previous studies (Lermo et al., 1991; Gutie rrez et al., 1996). However, all three dominant-period maps are different, and none of them shows a good correlation with surcial geology (compare Fig. 2, 3, and 6a). We observe that the amplication values are small (90% of the values are smaller than a factor 3); more than half of the measurements points indicated no amplication at all. However, spectral ratios of earthquake records have shown that local amplication within Colima attains a factor of 6, although no clear resonant frequency could be identied. We thus conclude that single-station microtremor measurements are not useful to evaluate site effects in the city of Colima. The reason is probably the complexity of the volcanic geology. Surface geology shows four different types of volcanic sediments, without obvious spatial rela-

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

599

Figure 8. Examples of images obtained in the p-f (horizontal slownessfrequency) plane by using the ReMi method. (a) Image corresponding to the measurements at UCOL (see Fig. 4). (b) Image corresponding to the measurements at STAB (see Fig. 4). The solid squares show the result of the automatic pick; for each frequency the slowness for which the derivative of the image with respect to slowness is a maximum. The open squares indicate the choice of phase-velocity values for each frequency, using the criterion proposed by Stephenson et al. (2005).

tions among them. These relations, in addition, may vary with depth, as it is known that the thickness of the deposits may change over short distances. Finally, the distinction between types of volcanic sediments is made in terms of their depositional mechanism, which may be not closely related with the average S-wave velocity within the deposits. This is clearly a situation where HVSR is not useful to evaluate site effects. The evaluation of the results of SPAC and ReMi has to follow a different line. The rst check is the comparison between the phase-velocity dispersion curves obtained from the two methods. Figure 9 shows, for example, the comparison between the dispersion curves selected from the ReMi images (those hand picked) and the phase-velocity dispersion curves derived from the SPAC method at UCOL and STAB sites. We observe very good agreement in the frequency range from 5 to 17 Hz. For frequencies smaller than 5 Hz, the results from the SPAC method are not reliable.

Wavelength at this frequency is about 112 m, close to double the array length, making the measurement of phase differences unreliable (Cha vez-Garc a et al., 2005). For frequencies greater than 18 Hz for UCOL or 16 Hz for STAB, phase velocities from the SPAC method increase with frequency, which is clearly unphysical. At these high frequencies, wavelengths become shorter than 16 m and we approach the limits imposed by the fundamental sampling theorem. However, contrary to the ReMi results, the mean phase-velocity dispersion curve derived from the SPAC method suffers no ambiguity, and an error bar can be estimated through the inversion process of the correlation coefcients (see the details in Cha vez-Garc a et al., 2005). A similar agreement between ReMi and SPAC was obtained for all eight sites, in the frequency range where the results can be considered reliable. We are interested in site amplication, however, and phase-velocity dispersion curves cannot be our nal result.

600

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Comparison between phase-velocity dispersion curves at locations UCOL (a) and STAB (b). The gray circles show the phase velocity determined with the manual picking from the ReMi images. Open squares with error bars show the mean values and the standard errors determined from the SPAC measurements. The solid line shows the phasevelocity dispersion curve computed from the S-wave velocity prole inverted at the corresponding location using the open squares as input data.

Figure 9.

We have inverted the phase-velocity dispersion curves derived from the SPAC method, rst, because the difference between ReMi and SPAC dispersion curves is smaller than the error bars. Second, the inversion procedure we use, that included in Herrmann (1987), takes into account the standard deviation of the data points. For the inversion of the phase-velocity dispersion curves we have arbitrarily xed the layer thicknesses, small enough to give exibility to the inversion, but accepting that our data are not enough to constrain the model completely (the inversion results are not unique, and phase-velocity dispersion has low vertical resolution and is more sensitive to vertically averaged elastic properties than interfaces). Density was set to 1.8 g/cm3, and the Poissons ratio was xed to 0.25. We inverted exclusively for S-wave velocity. The frequency range in which our phase-velocity dispersion values are reliable implies that we are able to constrain the S-wave velocities in the upper 30 to 40 m depths. The nonlinear inversion procedure is replaced by a linearized stochastic least-squares problem,

which is iterated until the changes to the model in any one iteration are small, and the analyst judges the computed dispersion curve to be close to the data. An example of this judgment is shown in Figure 9 for sites UCOL and STAB. The solid line corresponds to the dispersion curve computed for the nal prole inverted from the data for those two sites. From the inversion of phase-velocity dispersion curves, we obtained velocity proles at the eight sites where linear measurements were made. However, it is possible to verify these proles only at two locations, UCOL and STAB, where Gutie rrez et al. (1996) measured S-wave velocity using a suspension log. Their measurements, down to 50 m depth, are shown by the gray circles in Figure 10. Suspension log (SL) measurements at UCOL show a very large scatter in the upper 22 m and no clear layering. At STAB, no measurement could be made for the upper 10 m because of the very poor quality of the signals (Gutie rrez et al., 1996). The solid lines in Figure 10 show the nal model derived from the inversion of phase-velocity dispersion at these two sites. We observe good agreement at UCOL, especially if we ignore the scattered points above 22 m depth. The results for STAB show a much poorer agreement; the S-wave velocities derived from phase-velocity dispersion are consistently higher than the SL measurements. It would seem that we need only to decrease S-wave velocities between 10 and 45 m depth to get an improved t. However, when we do that, the phasevelocity dispersion computed from that prole does not match the observed dispersion curve at all. We have to accept that there is no S-wave velocity prole that could simultaneously t the observed dispersion curve at STAB and the SL measurements. The more likely reason is the different nature of the measurements. A borehole measurement is a punctual measurement that may not be representative of a signicant volume of the subsoil (especially if it comprises large boulders, where the hitting of a large boulder or hitting just between boulders can greatly change the results). Phasevelocity dispersion measurements, on the contrary, allow the estimation of the average properties for the larger volume averaged by the surface waves. It is clear that more extensive measurements of S-wave velocity are badly needed at Colima. To compute local amplication, we have extrapolated the surcial stratigraphy obtained from the inversion of phase-velocity dispersion curves. Based on geologic considerations, we assumed a half-space at 800 m depth. The lack of a clear resonant frequency in our HVSR results, together with the broadband character of the amplication observed in the earthquake spectral ratios by Gutie rrez et al. (1996) argue against a signicant clear-cut impedance contrast. For this reason, we extrapolated our results, assuming a very smooth S-wave velocity gradient between 40 and 800 m depth. The topmost 140 m of the nal shear-wave velocity proles for all eight sites are shown in Figure 11. On these proles, only the topmost 40 m are constrained and the deeper structure was extrapolated. We observe surcial velocities between 200 and 400 m/sec, which increase to 600

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

601

Figure 10. Shear-wave velocity proles for locations UCOL (a) and STAB (b). The gray circles show the values measured using a suspension log in 50 m depth boreholes by Gutie rrez et al. (1996). The solid lines show the velocity prole obtained from the inversion of the phase-velocity dispersion curves observed using SPAC at the corresponding location. to 900 m/sec at 40 m depth. This velocity increase occurs gradually in several layers. Although the shallow structure is well constrained on average (the averaging imposed by the surface waves), the precise location of the shallow interfaces cannot be well resolved by dispersion curve inversion. For this reason, we have not tried to classify the sites based on measures like Vs30 . This approach would require additional S-wave measurements, preferably using methods with larger resolution at shallow depths. We do not observe any obvious correlation between the velocity proles and surcial geology. A likely explanation is that a geologist may differentiate between geologic types based on the different deposition mechanisms. However, it is far from evident that the mechanical properties of a deposit would change signicantly because the shape of the blocks in a given matrix goes from angular to more or less round. Different geologic types may have similar S-wave velocities, whereas this parameter may vary within a single volcanic deposit because of its heterogeneity. It is clear that we do not have the data for a more rened comparison (for example, no geologic proles are available from the shallow boreholes of Gutie rrez et al., 1996). Clearly, a more systematic study of S-wave velocity distribution within Colima is badly needed to better understand the relation between surface geology and site response. Finally, we computed transfer functions for vertical incidence of shear waves on the proles shown in Figure 11. The results are given in Figure 12. Lacking data, we have neglected anelastic attenuation; therefore, amplication is overestimated for frequencies larger than 3 Hz. Maximum amplication attains of about a factor 5, in excellent agreement with the amplication factors observed by Gutie rrez et al. (1996) using spectral ratios of small earthquake records. The frequency of the amplication peaks varies among sites and, similar to the soil proles, it is not easy to relate them to surface geology. Moreover, the transfer functions show peaks related to the resonance of different layers or that could result from the contribution of two or more layers. We cannot ascribe a large signicance to the precise frequency of the resonant peaks in Figure 12 and therefore we do not claim that those transfer functions faithfully reect local amplication at our eight sites. The layering in the models is poorly constrained because of the low vertical resolution of surface-wave dispersion. In addition, the inversion is not unique. Finally, other observations argue against a clear resonant peak: the failure of the HVSR measurements to identify one, the differences between the dominant period maps produced by different studies, and the broadband amplication observed in earthquake spectral ratios. This, plus the impossibility of relating computed amplication to surface geology, makes it difcult to propose a microzonation map for Colima based on our results. We do observe, however, that

602

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez

Figure 11. Shear-wave velocity proles inverted from phase-velocity dispersion curves observed using the SPAC method for the eight sites indicated with stars in Figure 4. Given the frequencies for which phase dispersion was observed, these proles are reliable only down to 40 m depth. Below that depth, the proles are an extrapolation.

seismic amplication is fairly homogeneous throughout the city, and occurs in the same frequency band.

Conclusions
We have presented a site effect study in the city of Colima based on HVSR, SPAC, and ReMi microtremor methods. Single-station HVSR measurements were carried out at 310 sites within the Colima urban zone. These data were analyzed by using horizontal-to-vertical spectral ratios (e.g., Lermo and Cha vez-Garc a, 1994). However, a resonant peak could be identied at only 125 of the measurement sites. Moreover, the amplitude of this peak was very small, and the congured isoperiod map is not correlated with surface geology because the site effects at Colima are not the result of fairly homogeneous soft sediments overlaying a fairly ho-

mogeneous bedrock, where amplication would be due to the impedance contrast across a single interface. Clearly, in this study, when the geology is complex and seismic-motion amplication cannot be readily tied to a single resonant frequency, HVSR cannot provide a reliable estimate of site effects. Thus, the disagreement between previous studies based on microtremors at Colima city was not the result of an insufcient number of measurement points. We have shown that the limitations of single-point measurements can be partially overcome with array measurements of microtremors. We obtained good results using the SPAC method with a linear array, supporting the use of this method with array geometries different from a circle and conrming previous studies that have used this method. The results from the SPAC method were validated by comparison with results analyzed with the ReMi method. Both methods provided very similar phase-velocity dispersion curves. We inverted those curves to obtain shallow shear-wave velocity proles throughout the city. The inverted proles were compared with shear-wave velocity proles measured with suspension log measurements at two locations, UCOL and STAB. The agreement is good at one location and poor at the other; at the STAB site, the dispersion curves observed either with ReMi or SPAC are incompatible with the S-wave velocity prole measured using a suspension log. Indeed, borehole velocity measurements, although more reliable, may not be representative of the values that may affect surface-wave propagation, especially if the subsoil volcanic sediments include large boulders as in Colima city. The presence of large heterogeneities is suggested by the scatter of the suspension log measurements for the upper 20 m at UCOL, and the lack of measurements for the upper 10 m at STAB. Our nal results are well constrained for the topmost 40 m and show surface velocities between 200 and 400 m/sec, increasing to 600900 m/sec. This large velocity increase, however, occurs gradually, in several layers, and not across a single interface. Moreover, the inverted soil proles show a large variability throughout the city, with no obvious correlation with surcial geology. This could be anticipated given the intrinsic variability common in volcanic deposits. Colima appears then as a good example of where surface geology is a poor proxy for site characterization. The fact that site amplication is due to several layers with gradually increasing velocity explains the failure of HVSR to identify a resonant peak and is consistent with the failure of previous studies to identify resonant frequencies either with microtremor measurements or using spectral ratios of earthquake records. We extrapolated the shallow proles inverted from the dispersion curves down to 800 m, based on geologic considerations. We computed transfer functions for vertically incident shear waves on the complete soil proles. The computed level of amplication is similar to that observed by Gutie rrez et al. (1996) from spectral ratios of earthquake records. The presence of different shallow impedance con-

Site Effects in a Volcanic Environment: A Comparison between HVSR and Array Techniques at Colima, Mexico

603

Figure 12.

Transfer functions computed for vertical, plane, and shear-wave incidence on the velocity proles shown in the preceding gure. Attenuation was neglected.
This research was supported by Conacyt, Mexico, through contract SEP2003-C02-43880/A.

trasts brings about several peaks in the transfer functions with similar amplitude. Therefore, even if the global amplication is important, it is distributed in a wide frequency range, and the frequency of the rst resonant peak varies largely throughout the city. In conclusion, our results allow us to integrate previous indications of site effects and explain the failure of singlestation microtremor measurements when local geology is complex, a problem that can be overcome, in part, using array measurements. Our results indicate that it is not worthwhile to make a microzonation of the city. Seismic amplication level is similar all over the city, and the concept itself of resonant frequency loses its meaning in this geologic context. The best approach at the moment is to consider a homogeneous amplication factor of 6 for the frequency band between 0.2 and 5 Hz. We are convinced that this is the best that can be proposed at the moment, and is in agreement with all the data that have been analyzed to date in Colima. This estimate will have to be validated when a strong motion network operates in this city, but for now it can be used to predict seismic risk.

References
Aki, K. (1957). Space and time spectra of stationary stochastic waves, with special reference to microtremors, Bull. Earthquake Res. Inst. Tokyo Univ. 25, 415457. Asten, M. W. (1976). The use of microseisms in geophysical exploration, Ph.D. Thesis, Macquarie University. Bard, P.-Y. (1999). Microtremor measurements: a tool for site effect estimation?, in The Effects of Surface Geology on Seismic Motion, Recent Progress and New Horizon on ESG Study, K. Irikura, K. Judo, H. Okada, and T. Sasatani (Editors), Vol. 3, Balkema, Lisse, The Netherlands, 12511279. Borcherdt, R. D. (1970). Effects of local geology on ground motion near San Francisco Bay, Bull. Seism. Soc. Am. 60, 2961. Cenapred (2003). El sismo de Tecoman, Colima del 21 de enero de 2003 (Me 7.6), Direccion de Investigacion, Centro Nacional de Prevencion de Desastres, Mexico (in Spanish). Cha vez-Garc a, F. J., G. Pedotti, D. Hatzfeld, and P.-Y. Bard (1990). An experimental study of site effects near Thessaloniki (Northern Greece), Bull. Seism. Soc. Am. 80, 784806. Cha vez-Garc a, F. J., M. Rodr guez, and W. R. Stephenson (2005). An alternative to the SPAC analysis of microtremors: exploiting stationarity of noise, Bull. Seism. Soc. Am. 95, 277293. Cha vez-Garc a, F. J., M. Rodr guez, and W. R. Stephenson (2006). Subsoil structure using SPAC measurements along a line, Bull. Seism. Soc. Am. 96, 729736. Chouet, B. A., G. De Luca, G. Milana, P. B. Dawson, M. Martini, and R. Scarpa (1998). Shallow velocity structure of Stromboli volcano, Italy, derived from small-aperture array measurements of Strombolian tremor, Bull. Seism. Soc. Am. 88, 653666. Field, E. H., and K. Jacob (1995). A comparison and test of various site

Acknowledgments
We thank Abel Cortes for the time he spent explaining the volcanic deposits of Colima valley to us. We also thank Juan Tejeda Ja come for his continuous support and help throughout the different stages of the eld work. The comments by two anonymous reviewers and the Associate Editor, A. McGarr, helped us to improve our manuscript. Signal processing beneted signicantly from the availability of SAC (Goldstein et al., 1998).

604
response estimation techniques, including three that are non reference-site dependent, Bull. Seism. Soc. Am. 85, 11271143. Goldstein, P., D. Dodge, M. Firpo, and R. Stan (1998). Electronic seismologist: whats new in sac2000? Enhanced processing and database access, Seism. Res. Lett. 69, 202205. Gutie rrez, C., and S. K. Singh (1992). A site effect study in Acapulco, Guerrero, Mexico: comparison of results from strong motion and microtremor data, Bull. Seism. Soc. Am. 82, 642659. Gutie rrez, C., K. Masaki, J. Lermo, and J. Cuenca (1996). Microzonicacio n s smica de la ciudad de Colima, Cuadernos de Investigacio n No. 33, Me xico (in Spanish). Herrmann, R. B. (1987). Computer Programs in Seismology, Saint Louis University, 7 vols. Kagami, H., S. Okada, K. Shiono, M. Oner, M. Dravinski, and A. K. Mal (1986). Observation of 1 to 5 second microtremors and their application to earthquake engineering, part III: A two-dimensional study of site effects in S. Fernando valley, Bull. Seism. Soc. Am. 76, 1801 1812. Kobayashi, H., K. Seo, and S. Midorikawa (1986). Estimated strong ground motions in the Mexico city due to the Michoacan, Mexico earthquake of September 19, 1985 based on characteristics of microtremor. Part 2, Report on microzoning studies of the Mexico earthquake of September 19, 1985, The Graduate School of Nagatsuta, Tokyo Institute of Technology, Yokohama, Japan. Lermo, J., and F. J. Cha vez-Garc a (1994). Are microtremors useful in site effect evaluation?, Bull. Seism. Soc. Am. 84, 13501364. Lermo, J., J. Diaz de Leon, E. Nava, and M. Macias (1991). Estimacio n de periodos dominantes y amplicacio n relativa del suelo en la zona urbana de Colima, presented at IX Congreso Nacional de Ingenier a S smica, Manzanillo, Colima (in Spanish). Louie, J. N. (2001). Faster, better: shear-wave velocity to 100 meters depth from refraction microtremor arrays, Bull. Seism. Soc. Am. 91, 347 364. Malagnini, L., P. Tricarico, A. Rovelli, R. B. Herrmann, S. Opice, G. Biella, and R. de Franco (1996). Explosion, earthquake, and ambient noise recordings in a pliocene sediment-lled valley: inferences on seismic response properties by reference- and non-reference-site techniques, Bull. Seism. Soc. Am. 86, 670682. McMechan, G. A., and M. J. Yedlin (1981). Analysis of dispersive waves by wave eld transformation, Geophysics 46, 869874. Mokhtar, T. A., R. B. Herrmann, and D. R. Russell (1988). Seismic velocity and Q model for the shallow structure of the Arabian shield from short-period Rayleigh waves, Geophysics 53, 13791387.

F. J. Cha vez-Garc a, T. Dom nguez, M. Rodr guez, and F. Pe rez


Ohori, M., A. Nobata, and K. Wakamatsu (2002). A comparison of ESAC and FK methods of estimating phase velocity using arbitrarily shaped microtremor arrays, Bull. Seism. Soc. Am. 92, 23232332. Nakamura, Y. (1989). A method for dynamic characteristics estimation of subsurface using microtremor on the ground surface, Q. Rep. Railway Tech. Res. Inst. 30, no. 1, 2533. Rodr guez, M., and F. J. Cha vez-Garc a (2006). Comparison of two array techniques to determine soil structure from microtremor presented at 1st European Conference on Earthquake Engineering and Seismology, Geneva, 38 September, abstract 341. Seo, K. (1992). A joint work for measurements of microtremors in the Ashigara valley, in Int. Symp. Effects of Surf. Geol. on Seismic Motion, Vol. I, ESG, Odawara, Japan, 4352. Stephenson, W. J., J. N. Louie, S. Pullammanappallil, R. A. Williams, and J. K. Odum (2005). Blind shear-wave velocity comparison of ReMi and MASW results with boreholes to 200 m in Santa Clara valley: implications for earthquake ground-motion assessment, Bull. Seism. Soc. Am. 95, 25062516. Toshinawa, T., J. J. Taber, and J. J. Berrill (1997). Distribution of ground motion intensity inferred from questionnaire survey, earthquake recordings, and microtremor measurementsa case study in Christchurch, New Zealand, during the 1994 Arthurs Pass earthquake, Bull. Seism. Soc. Am. 87, 356369. Volant, P., F. Cotton, and J.-C. Gariel (1998). Estimation of site response using the H/V technique. Applicability and limits on Garner valley downhole array dataset (California), in Proc. XIth Eur. Conf. on Earthq. Engrg., Paris, 611 September, Bisch, Labbe, and Pecker (Editors), Balkema, Lisse, The Netherlands. Instituto de Ingenier a UNAM, Ciudad Universitaria Coyoaca n, 04510, Me xico D.F., Mexico (F.J.C.-G., M.R.) Observatorio Vulcanolo gico Universidad de Colima Colima, 28045, Colima, Mexico (T.D.) Facultad de Ingenier a Civil Universidad de Colima Colima, 28045, Colima, Mexico (F.P.) Manuscript received 28 April 2006.

Вам также может понравиться