Sie sind auf Seite 1von 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/307442989

The Relationship between Ambient Vibration H/V and SH Transfer Function:


Some Experimental Results

Article  in  Seismological Research Letters · September 2016


DOI: 10.1785/0220160113

CITATIONS READS
8 401

7 authors, including:

El Hadi Oubaiche Jean-Luc Chatelain


Centre National de Recherche Appliquée en Génie Parasismique CGS Institute of Research for Development
48 PUBLICATIONS   217 CITATIONS    112 PUBLICATIONS   2,222 CITATIONS   

SEE PROFILE SEE PROFILE

Mustapha Hellel Djamel Machane


University of Science and Technology Houari Boumediene Centre National de Recherche Appliquée en Génie Parasismique CGS
24 PUBLICATIONS   157 CITATIONS    81 PUBLICATIONS   228 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Seismic Hazard View project

Flood risk at Algiers region View project

All content following this page was uploaded by Jean-Luc Chatelain on 31 October 2017.

The user has requested enhancement of the downloaded file.


SRL Early Edition


E

The Relationship between Ambient Vibration


H/V and SH Transfer Function:
Some Experimental Results
by El Hadi Oubaiche, Jean-Luc Chatelain, Mustapha Hellel,
Marc Wathelet, Djamel Machane, Rabah Bensalem, and
Abderrahmane Bouguern
ABSTRACT

The horizontal-to-vertical spectral ratios (HVSRs) of ambient The origin of the HVSR peak has been the subject of
vibrations are commonly used to observe soil resonance several studies, either theoretical analyses or numerical experi-
frequencies, which are revealed by HVSR curve peaks. These ments based on synthetic seismic noise. Nakamura (1989,
resonances have been explained either in terms of S-wave trans- 2000, 2009) linked the HVSR peak to the S-wave transfer func-
fer function or in terms of Rayleigh-wave ellipticity. In this tion. This hypothesis has been challenged by several authors
study, ambient vibration recordings have been carried out next (e.g., Lachet and Bard, 1994; Kudo, 1995; Bard, 1998), who
to nine boreholes in the eastern Mitidja basin (Algeria), all of related HVSR peaks to the Rayleigh-wave ellipticity, as already
which have been characterized by downhole geophysical sur- suggested by Nogoshi and Igarashi (1971).
veys. Using velocity profiles obtained from the downhole sur- Some studies based on synthetic noise show that, for soft
veys, we compare the frequency of the second HVSR peak (f HV ) and hard layers shear-wave impedance (Z) contrasts over 4,
to frequencies obtained with (1) the time-averaged velocity (f T ), the HVSR peak can be explained by the horizontal polarization
(2) depth-averaged velocity (f D ), (3) the SH transfer function of the Rayleigh-wave fundamental mode, whereas for contrasts
(f SH ), and (4) the fundamental-mode Rayleigh-wave ellipticity lower than 4, this explanation breaks down (e.g., Konno and
(f E ). We find that f SH , f T , and f D fit well with f HV , whereas Ohmachi, 1998; Bonnefoy-Claudet et al., 2008). In a study based
this is not the case for f E , implying that the HVSR peak fre- on synthetic ambient vibrations, Bonnefoy-Claudet et al. (2008)
quency is better explained by the SH transfer function peak than showed that (1) for high-impedance contrasts (Z > 4), the seis-
by the Rayleigh-wave ellipticity. mic noise is composed of both fundamental Rayleigh- and Love-
wave modes; (2) for impedance contrasts 3 < Z < 4, the fun-
damental Rayleigh wave is not predominant, and the Love mode
Online Material: Tables of site coordinates, borehole soil mod- strongly dominates; and (3) for lower impedance contrasts
els, borehole and vibration data results, velocity contrasts, and (Z < 3), the contribution of the Rayleigh mode is insignificant
relative frequency differences. and the simulated noise is composed of the fundamental Love
waves and S waves. Bonnefoy-Claudet et al. (2008) have shown
INTRODUCTION that, for a set of 136 sites, the S-wave velocity contrast exceeds the
value of 3 in only 32% of the cases; that is, in most cases, Rayleigh
The ambient vibration method was proposed by Kanai (1957) waves are not involved.
in engineering seismology to evaluate soft-soil amplification. Nevertheless, the relationship between the HVSR peak and
This method is now widely used in microzonation and site- the Rayleigh-wave ellipticity is commonly invoked when esti-
effect studies (e.g., Chávez-García and Bard, 1994; Guéguen mating the subsurface structure, shear-wave velocity, and sedi-
et al., 2000; Alfaro et al., 2001; Duval et al., 2001; Guillier ment thickness (e.g., Yamanaka et al., 1994; Fäh et al., 2001;
et al., 2004; Panou et al., 2005; Chatelain, Guillier, and Parvez, Malischewsky and Scherbaum, 2004; Tuan et al., 2011;
2008; Bensalem et al., 2010; Hellel et al., 2010). Most of these Hobiger et al., 2013), as well as for determining P- and S-wave
studies are based on the single-station horizontal-to-vertical velocity models by inversion of Rayleigh-wave dispersion curves
spectral ratio (HVSR) method (Nogoshi and Igarashi, 1971; from ambient vibration array data (e.g., Aki, 1957; Horike,
Nakamura, 1989), used to obtain the fundamental soil resonance 1985; Tokimatsu, 1997; Okada, 2003; Wathelet et al., 2004;
frequency, corresponding to the HVSR curve peak frequency. Di Giulio et al., 2006; Zor et al., 2010). In these studies, HVSR

doi: 10.1785/0220160113 Seismological Research Letters Volume 87, Number 5 September/October 2016 1
SRL Early Edition

▴ Figure 1. Satellite photograph (from Google) showing the location of the nine boreholes where ambient vibration and downhole surveys
have been carried out. Borehole locations are indicated by the filled circles numbered from 1 to 9 (coordinates are listed in Ⓔ Table S1).

peaks are interpreted in terms of Rayleigh-wave ellipticity, no did not reach the seismic bedrock, so the total thickness of the
matter the shear-wave velocity contrast at depth. overlying sediments cannot be established.
In most cases, HVSR curves present a single peak due to We show below that the frequency of the second HVSR
the resonance of a soft sedimentary layer overlying a hard-rock peak (f HV ), the SH transfer function (f SH ), the time-averaged
layer. However, in some circumstances a second peak appears at velocity (f T ), and the depth-averaged velocity (f D ) fit well
a higher frequency than the fundamental peak (e.g., Guéguen with f HV , but not with the fundamental-mode Rayleigh-wave
et al., 1998). This second peak appears when two sediment ellipticity (f E ), indicating that the HVSR peak frequency is bet-
layers with significant shear-wave velocity contrasts (e.g., a ter explained by the SH transfer function peak than by the
soft sediment layer overlying stiffer consolidated sediments) Rayleigh-wave ellipticity.
overlie harder seismic bedrock. In such a case, the fundamental
frequency is produced by both sediment layers acting as a DATA ACQUISITION AND PROCESSING
single-layer overlying bedrock, while the second peak reflects
the resonance frequency of the upper sediment layer. In the Borehole Data
Mitidja basin (Algeria), Oubaiche et al. (2012) found that We performed ambient vibration recordings adjacent to nine
boreholes located about 10–25 km east of Algiers and distrib-
the second H/V peak was more consistently developed than
uted over a zone covering about 100 km2 (Fig. 1; Ⓔ Table S1,
the fundamental peak, presumably because of stronger lateral
available in the electronic supplement to this article).
changes in the depth and/or the velocity contrast associated with
These boreholes were selected because each of them had
the sediment–basement interface. Using experimental data, they been characterized by downhole geophysical surveys (Fig. 1, Ⓔ
showed that the amplitude of the second peak is clearly related Table S1) during a seismic microzoning study (Japan Interna-
to the velocity contrast between the topmost soft sediment layer tional Cooperation Agency–Centre National de Recherche Ap-
and the underlying stiffer consolidated sediments, no matter pliquée en Génie Parasismique [JICA-CGS], unpublished report,
what the properties of the underlying bedrock. 2006; see Data and Resources). Two of these boreholes have been
In this work, we present an experimental study of the re- added to the seven boreholes previously used by Oubaiche et al.
lationship between these second HVSR peaks and both the (2012) who studied the relationship between the HVSR curve
shear-wave transfer function and the Rayleigh-wave ellipticity. amplitude peak and the SH velocity contrast at depth.
As in Oubaiche et al. (2012), the first HVSR peak cannot be At the nine boreholes, two layers are directly observed:
studied using the available borehole data because the boreholes (1) a surficial soft sedimentary layer comprising various pro-

2 Seismological Research Letters Volume 87, Number 5 September/October 2016


SRL Early Edition

▴ Figure 2. P-wave (gray line) and S-wave (black line) velocity profiles from the downhole measurements at the nine borehole surveys.

portions of clay, sandy clay, sand, and altered sandstones, with a (V S ) velocities. As described in Oubaiche et al. (2012), the seis-
thickness (H soft ) varying from 15 to 35 m and (2) an under- mic source was a 5 kg hammer, and the source-borehole offset
lying layer of consolidated sandstones (Ⓔ Table S2, Fig. 2; was 5 m. At each of the 1-m-spaced locations inside the bore-
Oubaiche et al., 2012). The thickness of the second layer is not hole, recordings were obtained with 28 Hz geophones, for a
known because the bottom of this layer is not reached at any of vertical hit and two horizontal hits in opposite directions.
the nine boreholes. Borehole data were processed as in Oubaiche et al. (2012):
the P wave generated by the vertical shot is extracted on the
Downhole Data first arrival of the vertical-component seismogram, and the
Downhole surveys were performed at the nine boreholes to S wave generated by the two opposite horizontal shoots is ex-
investigate both the compression-wave (V P ) and shear-wave tracted by superposing the two horizontal components with

Seismological Research Letters Volume 87, Number 5 September/October 2016 3


SRL Early Edition
opposite polarities. The P- and S-wave velocities of the soft RESULTS
layer and of the consolidated sandstones (Fig. 2, Ⓔ Table S2)
are calculated with the P- and S-wave vertical travel times, Downholes
obtained after correction for the travel-time obliquity. From the downhole surveys (Fig. 2, Ⓔ Table S2), the upper-
most soft sediment layer thickness varies from 15 m at SC5 to
36 m at SC8, with a depth-averaged shear-wave velocity (V D )
Ambient Vibration Data ranging from 264 m=s at SC3 to 458 m=s at SC8 and a time-
Ambient vibrations were recorded next to each of the nine averaged shear-wave velocity (V T ) ranging from 228 m=s at
boreholes, with a CityShark station (Chatelain et al., 2000, SC3 to 413 m=s at SC8 (Ⓔ Table S3). In the underlying con-
2012) coupled to a three-component 5 s Lennartz seismometer solidated sandstone layer, the shear-wave velocity (V stiff ) varies
at a 200 samples=s sampling rate for 15 min and following the from 550 m=s at SC5 to 1364 m=s at SC4. Using V D , V T , and
recommendations of Chatelain et al. (2008). H soft in the generic formula f  V =4H leads to f D in the 2.0–
As in Oubaiche et al. (2012), the open-source Geopsy soft- 4.2 Hz range and f T in the 1.7–4.0 Hz range (Ⓔ Table S3).
ware (see Data and Resources) has been used for processing
ambient vibration signals to obtain the HVSR curves, using HVSR
25 s time windows. Before processing, a 5% cosine taper was HVSR curves show two frequency peaks (Fig. 3). The first peak
applied at both ends of each window. Before averaging, each ranges from 0.7 Hz at SC4 to 1.5 Hz at SC9. This peak is the
individual HVSR curve was smoothed using the Konno and natural frequency of the deepest sediment-based interface not
Ohmachi (1998) method, with a constant of 40. reached in the boreholes (Oubaiche et al., 2012). The second
HVSR curves show two frequency peaks, indicating the peak ranges from 1.8 Hz at SC3 to 4.7 Hz at SC7 (Fig. 3, Ⓔ
presence of two interfaces at depth. The first peak reflects the Table S3), corresponding to the resonance frequency of the
resonance frequency of a thick sediment layer resting on marls, surficial soft sediment layer, as already demonstrated for the
which, with S-wave velocities around 2000 m=s, constitute seis- seven sites studied in Oubaiche et al. (2012). This study relates
mological bedrock (Harbi et al., 2007). The second peak fits to the second HVSR peak, as in Oubaiche et al. (2012).
well with an uppermost layer composed of soft sediments over-
lying stiffer sandstones (Oubaiche et al. 2012). Again, as the SH Transfer Function and Rayleigh-Wave Ellipticity
depth to basement is unknown at our nine study sites, no eva- Peaks
luation of the fundamental peak (around 1 Hz) can be under- The transfer function and the Rayleigh-wave ellipticity curve
taken. Therefore, the focus of this study is the second peak peaks (Fig. 3, Ⓔ Table S3) are close to the HVSR curve peaks,
(above ∼1:8 Hz). which range from 1.8 to 4.7 Hz. SH transfer function peaks
range from 2.4 Hz at SC3 to 4.7 Hz at SC7, and Rayleigh-wave
ellipticity peaks range from 2.6 Hz at SC3 to 5.2 Hz at SC5. At
SH Transfer Function and Rayleigh-Wave Ellipticity two-third of the sites (1, 2, 3, 4, 5, and 7) f SH is closer to f HV
Curves than f E , whereas f E is closer to f HV than f SH at site 9; and, for
The soft-sediment mass densities obtained from the drilling both sites 6 and 8, f SH and f E are equal (Fig. 3).
data vary from 1800 to 2160 kg=m3 and those of the sand-
stone from 2060 to 2160 kg=m3 (Ⓔ Table S2). SH transfer DISCUSSION
function and fundamental-mode Rayleigh-wave ellipticity
curves were calculated with quality factors Q P and Q S fixed to Difference between Frequencies
50 and 25 for the overlying soft sediments and to 100 and 50 The comparison of HVSR curves with both SH transfer function
for the sandstone, as commonly used (e.g.,Wang et al., 1994; Di and Rayleigh-wave ellipticity curves shows that f SH is consis-
Giulio et al., 2003; Bonnefoy-Claudet et al., 2006). tently closer to f HV than f E (Fig. 3). The average relative differ-
Multilayer V S profiles (Fig. 2, Ⓔ Table S2) obtained with ences (Ⓔ Table S3) between f HV and f T , f D , and f SH , taking
the downhole tests are used to calculate the vertical incidence Δf  jf −f HV j, in which f is f T , f D , and f SH , Δf T =f HV av 
SH transfer function curves with the gpsh application included 8%4%, Δf D =f HV av  8%  4%, and Δf SH =f HV av 
in the Geopsy package. Multilayer V P and V S profiles (Fig. 2, 12%  9% are similar to the relative f HV standard deviation
Ⓔ Table S2) are used to calculate the Rayleigh-wave ellipticity average σ HV =f HV av , which amounts to 12%  3%. It is not
curves with the gpell application, also included in the Geopsy the case for Δf E =f HV av, because, with a value of 26%  16%,
package. it is about two times larger than σ HV =f HV av .
SH transfer function, Rayleigh-wave ellipticity, and HVSR
curves are obtained between 0.5 and 20 Hz, all computed and Variation with the Velocity Contrast
displayed with a 100-sample logarithmic distribution so as to The HVSR peak frequency is well correlated with both f T and
insure that results are comparable. For additional accuracy, the f D obtained from downhole data, because Δf T =f HV and
frequency peaks were not extracted from the graphics but Δf D =f HV do not vary with the shear-wave velocity contrast
directly from the Geopsy output files of HVSR , SH transfer at depth (Fig. 4a,b, Ⓔ Table S4), to the contrary of what
function, and Rayleigh-wave ellipticity. has been observed for the peak amplitudes at the seven first

4 Seismological Research Letters Volume 87, Number 5 September/October 2016


SRL Early Edition

▴ Figure 4. Relative frequency differences variation with the


velocity contrast between the surficial soft sediment and the
underlying sandstone layers. Relative frequency difference of
(a) the frequency from time-averaged velocity Δf T = f HV versus
velocity contrast (V stiff = V T ), (b) the frequency from depth-aver-
aged velocity Δf D = f HV , (c) the SH transfer function peaks
(Δf SH = f HV ), and (d) of the Rayleigh-wave ellipticity peaks
(Δf E = f HV ) versus the velocity contrast (V stiff = V D ). The vertical gray
bars represent the standard deviations of Δf T = f HV , Δf D = f HV ,
Δf SH = f HV , and Δf E = f HV listed in Ⓔ Table S4.

boreholes by Oubaiche et al. (2012), which are closely linked to


the shear-wave velocity contrast.
The relative difference Δf SH =f HV increases with the veloc-
ity contrast, whereas no clear relationship can be evidenced with
Δf E =f HV (Fig. 4c,d, Ⓔ Table S4).

Correlation between Frequencies


Four weighted linear least-square regressions have been calcu-
lated between the dependent parameters f HV , considering a
▴ Figure 3. (a) Horizontal-to-vertical spectral ratio (HVSR), SH
weight w  1=σ HV 2 , and the independent parameters f T ,
transfer function, and Rayleigh-wave ellipticity curves obtained
at the nine sites of Figure 1. Gray line, HVSR; black line, SH trans-
f D , f SH , and f E (Fig. 5). We calculated the p-value associated
fer function; dashed line, Rayleigh-wave ellipticity. Gray circle, with each correlation and set the threshold to the commonly
black triangles, and crosses represent the picked value of the used value of 0.05, which corresponds to a confidence level
HVSR peak frequency, the SH transfer function, and the Ray- of 95%. Because the p-values obtained for the correlations
leigh-wave ellipticity, respectively. (b) Frequencies picked on between f HV and, respectively, f T (1:16 × 10−5 ), f D
(a) at each site. The gray bars represent f HV standard deviations. (1:57 × 10−4 ), f SH (8:87 × 10−4 ), and f E (0.039) are below
SH transfer function and Rayleigh-wave ellipticity curves are ob- the chosen threshold, the coefficients of determination can
tained using the borehole data (Fig. 2, Ⓔ Table S2). Because only be considered as significant.
the second HVSR peak is studied, and for clarity of the figure, SH Given the coefficient of determination and slopes ob-
transfer function and Rayleigh-wave ellipticity curves are calcu- tained, much better matches are obtained between f HV and
lated without estimating the thickness of the underlying stiff layer f T (slope a  1:1; coefficient of determination R2  0:94),
nor the bedrock shear-wave velocity; that is, the fundamental f D (a  0:97; R2  0:88), and f SH (a  0:95; R2  0:81),
peak is not calculated. than between f HV and f E (a  0:63; R2  0:47).

Seismological Research Letters Volume 87, Number 5 September/October 2016 5


SRL Early Edition

▴ Figure 5. Weighted linear least-square regressions between HVSR peaks frequencies (f HV ) and frequencies (a) from time-averaged
shear-wave velocity (f T ), (b) from depth-averaged shear-wave velocity (f D ), (c) of SH transfer function peak frequencies (f SH ), and (d) of
Rayleigh-wave ellipticity peak frequencies (f E ). The black line represents linear fit and dashed gray lines represent the upper and lower
control limits for 95% level regression confidence.

CONCLUSION Algiers earthquake scenario project, which has not been pub-
lished and cannot be released to the public. Ambient vibration
An experimental study of ambient vibration HVSR and data were collected for this study. Data can be obtained from
downhole velocity profiles has been conducted at nine sites the corresponding author. Geopsy software is available at www.
over a zone of about 100 km2 in the Dar el Beida region geopsy.org (last accessed June 2015).
(Algeria) to study the relationship between HVSR peak fre-
quency and both the shear-wave transfer function and the ACKNOWLEDGMENTS
Rayleigh-wave ellipticity frequencies. The downholes are not
deep enough to allow us to study the soil fundamental peak This study is part of a cooperative project between the Centre
frequency near 1 Hz on HVSR curves. Therefore, we focused National de la Recherche Appliquée en Génie Parasismique
on the second HVSR peak, in the 1.8–4.7 Hz range, produced (CGS) in Algiers (Algeria) and the French Institut de
by shear-wave velocity contrasts in the 1.8–4.5 range between Recherche pour le Développement (IRD). We thank Michel
a topmost soft sediment layer and an underlying consolidated Campillo for very fruitful discussions on the diffuse-field
sediment layer. theory. We thank the Service de Coopération et d’Action
Very good fits are found between these HVSR peak Culturelle (SCAC) of the French Embassy in Algiers for their
frequencies and the frequencies obtained with the generic for- financial support. We thank all of the CGS and ISTerre people
mula f  V =4H using both depth-averaged and time-aver- who participated in the data acquisition. This study was funded
aged velocities in the surficial soft sediment layer. The best by CGS and IRD. We thank Michael Bevis for assisting us with
fit is observed with time-averaged velocities, whereas Oubaiche our English.
et al. (2012) reported that, for HSVR peak amplitudes, the best
fit is obtained with depth-averaged velocities. REFERENCES
A comparison between these HVSR peak frequencies and
both SH transfer function and Rayleigh-wave ellipticity peak Aki, K. (1957). Space and time spectra of stationary stochastic waves,
frequencies shows a strong fit with the SH transfer function with special reference to microtremors, Bull. Earthq. Res. Inst.
35, 415–456.
frequency, whereas no relationship can be evidenced with the Alfaro, A., L. Pujades, X. Goula, T. Susagna, B. M. Navarro, F. J. Sanchez,
Rayleigh-wave ellipticity frequency. and J. A. Canas (2001). Preliminary map of soil’s predominant peri-
This experimental study favors the claims of Nakamura ods in Barcelona using microtremors, Pure Appl. Geophys. 158,
(1989, 2000, 2009), Konno and Ohmachi (1998), and Bonnefoy- 2499–2511.
Claudet et al. (2008) that associated the HVSR frequency peak Bard, P.-Y. (1998). Microtremor measurements: A tool for site effect es-
timation? Proc. of the Second International Symposium on the Effects
with the SH transfer function rather than with Rayleigh-wave of Surface Geology on Seismic Motion, Yokohama, Japan, Vol. 3,
ellipticity. 1251–1279.
Bensalem, R., J.-L. Chatelain, D. Machane, E. H. Oubaiche, M. Hellel, B.
Guillier, M. Djeddi, and L. Djadia (2010). Ambient vibration tech-
DATA AND RESOURCES niques applied to explain heavy damages caused in Corso (Algeria)
by the 2003 Boumerdes earthquake: Understanding seismic ampli-
fications due to gentle slopes, Seismol. Res. Lett. 81, no. 6, 928–940,
Borehole and downhole data were obtained from the Japan doi: 10.1785/gssrl.81.6.928.
International Cooperation Agency–Centre National de Re- Bonnefoy-Claudet, S., C. Cornou, P.-Y. Bard, F. Cotton, P. Moczo, J.
cherche Appliquée en Génie Parasismique (JICA-CGS, 2006) Kristek, and D. Fäh (2006). H/V ratio: A tool for site effects

6 Seismological Research Letters Volume 87, Number 5 September/October 2016


SRL Early Edition
evaluation. Results from 1-D noise simulations, Geophys. J. Int. 167, Horike, M. (1985). Inversion of phase velocity of long-period microtre-
no. 2, 827–837. mors to the S-wave velocity structure down to the basement in
Bonnefoy-Claudet, S., A. Köhler, C. Cornou, M. Wathelet, and P. Y. Bard urbanized areas, J. Phys. Earth 33, 59–96.
(2008). Effects of Love waves on microtremor H/V ratio, Bull. Kanai, K. (1957). The requisite conditions for the predominant vibration
Seismol. Soc. Am. 98, no. 1, 288–300. of ground, Bull. Earthq. Res. Inst. Tokyo Univ. 23, 457–471.
Chatelain, J.-L., P. Guéguen, B. Guillier, J. Fréchet, F. Bondoux, J. Konno, K., and T. Ohmachi (1998). Ground-motion characteristics
Sarrault, P. Sulpice, and J.-M. Neuville (2000). CityShark: A user- estimated from spectral ratio between horizontal and vertical com-
friendly instrument dedicated to ambient noise (microtremor) re- ponents of microtremor, Bull. Seismol. Soc. Am. 88, 228–241.
cording for site and building response studies, Seismol. Res. Lett. Kudo, K. (1995). Practical estimates of site response. State-of-art report,
71, 698–703. Proc. of the Fifth International Conference on Seismic Zonation, Nice,
Chatelain, J.-L., B. Guillier, F. Cara, A.-M. Duval, K. Atakan, P.-Y. Bard, France, 17–19 October 1995.
and the WP02 SESAME Team (2008). Evaluation of the influence Lachet, C., and P.-Y. Bard (1994). Numerical and theoretical investiga-
of experimental conditions on HVSR results from ambient noise tions on the possibilities and limitations of Nakamura’s technique, J.
recordings, Bull. Earthq. Eng. 6, 33–74, doi: 10.1007/S10518- Phys. Earth 42, 377–397.
007-9040-7. Malischewsky, P., and F. Scherbaum (2004). Love’s formula and H/V-ratio
Chatelain, J.-L., B. Guillier, P. Guéguen, J. Fréchet, and J. Sarrault (2012). (ellipticity) of Rayleigh waves, Wave Motion 40, no. 1, 57–67.
Ambient vibration recording for single-station, array and building Nakamura, Y. (1989). A method for dynamic characteristics estimation of
studies made simple: CityShark II, Int. J. Geosci. 3, no. 6A, 1168– subsurface using microtremor on the ground surface, Q. Rep.
1175, doi: 10.4236/ijg.2012.326118. Railway Tech. Res. Inst. 30, no. 1, 25–30.
Chatelain, J.-L., B. Guillier, and I. A. Parvez (2008). False site effects: The Nakamura, Y. (2000). Clear identification of fundamental idea of
Anjar case, following the 2001 Bhuj (India) earthquake, Seismol. Nakamura’s technique and its applications, Proc. of the 12th World
Res. Lett. 79, 698–703, doi: 10.1785/gssrl.79.6.816. Conference on Earthquake Engineering, Auckland, New Zealand,
Chávez-García, F. J., and P.-Y. Bard (1994). Site effect in Mexico City 30 January–4 February 2000, Paper 2656.
eight years after the September 1985 Michoacan earthquake, Soil Nakamura, Y. (2009). Basic structure of QTS (HVSR) and examples of
Dynam. Earthq. Eng. 13, no. 4, 229–247. applications, in Increasing Seismic Safety by Combining Engineering
Di Giulio, G., C. Cornou, M. Ohrnberger, M. Wathelet, and A. Rovelli Technologies and Seismological Data, M. Mucciarelli, M. Herak, and
(2006). 2-D small aperture arrays for velocity profiles estimation J. Cassidy (Editors), NATO Science for Peace and Security, Series C:
using ambient seismic noise in a small-size alluvial basin (Colfiorito, Environmental Security, Springer Sciences, Dordrecht, The Nether-
Italy), Bull. Seismol. Soc. Am. 96, no. 5, 1915–1933, doi: 10.1785/ lands, 33–51.
0120060119. Nogoshi, M., and T. Igarashi (1971). On the amplitude characteristics of
Di Giulio, G., A. Rovelli, F. Cara, R. M. Azzara, F. Marra, R. Basili, and microtremor (part 2), J. Seismol. Soc. Japan 24, 26–40 (in Japanese
A. Caserta (2003). Long-duration asynchronous ground motions with English abstract).
in the Colfiorito plain, central Italy, observed on a two-dimen- Okada, H. (2003). The Microtremor Survey Method, M. W. Asten (Editor),
sional dense array, J. Geophys. Res. 108, no. B10, 2486–2500, Geophysical Monograph Series No. 12, Society of Exploration
doi: 10.1029/2002JB002367. Geophysicists, Tulsa, Oklahoma.
Duval, A.-M., S. Vidal, J.-P. Méneroud, A. Singer, F. De Santis, C. Ramos, Oubaiche, E. H., J.-L. Chatelain, A. Bouguern, R. Ben Salem,
G. Romero, R. Rodriguez, A. Pernia, N. Reyes, et al. (2001). D. Machane, M. Hellel, F. Khaldaoui, and B. Guillier (2012).
Caracas, Venezuela, site effect determination with microtremor, Experimental relationship between ambient vibration H/V peak
Pure Appl. Geophys. 158, 2513–2523. amplitude and shear-wave velocity contrast, Seismol. Res. Lett.
Fäh, D., F. Kind, and D. Giardini (2001). A theoretical investigation of 83, no. 6, 1038–1046.
average H/V ratios, Geophys. J. Int. 145, no. 2, 535–549. Panou, A., N. Theodulidis, P. Hatzidimitriou, K. Stylianidis, and C.
Guéguen, P., J.-L. Chatelain, B. Guillier, and H. Yepes (2000). An indi- Papazachos (2005). Ambient noise horizontal-to-vertical spectral
cation of the soil topmost layer response in Quito (Ecuador) using ratio in site effects estimation and correlation with seismic damage
HVSR spectral ratio, Soil Dynam. Earthq. Eng. 19, 127–133. distribution in urban environment: The case of the city of
Guéguen, P., J.-L. Chatelain, B. Guillier, H. Yepes, and J. Egred (1998). Thessaloniki (northern Greece), Soil Dynam. Earthq. Eng. 25,
Site effect and damage distribution in Pujili (Ecuador) after the no. 4, 261–274, doi: 10.1016/j.soildyn.2005.02.004.
28 March 1996 earthquake, Soil Dynam. Earthq. Eng. 17, 329–334. Tokimatsu, K. (1997). Geotechnical site characterization using surface
Guillier, B., D. Machane, E. H. Oubaiche, J.-L. Chatelain, Y. Ait Meziane, waves, Proc. of the First International Conference on Earthquake
R. Bensalem, F. Dunand, P. Guéguen, M. Hadid, M. Hellel, et al. Geotechnical Engineering, Vol. 3, 1333–1368.
(2004). Résultats préliminaires sur les fréquences fondamentales et Tuan, T., F. Scherbaum, and P. G. Malischewsky (2011). On the relation-
les amplifications de sols, obtenus par l’étude du bruit de fond, sur la ship of peaks and troughs of the ellipticity (H/V) of Rayleigh waves
ville de Boumerdes–Algérie, Mémoires du Service Géologique de and the transmission response of single layer over half-space models,
l’Algérie 12, 103–114 (in French). Geophys. J. Int. 184, no. 2, 793–800.
Harbi, A., S. Maouche, F. Vaccari, A. Aoudia, F. Oussadou, G. F. Panza, Wang, Z., R. Street, and E. Woolery (1994). Q S estimation for
and D. Benouar (2007). Seismicity, seismic input and site effects in unconsolidated sediments using first-arrival SH critical refractions,
the Sahel—Algiers region (North Algeria), Soil Dynam. Earthq. J. Geophys. Res. 99, 13,543–13,551.
Eng. 27, no. 5, 427–447. Wathelet, M., D. Jongmans, and M. Ohrnberger (2004). Surface-wave
Hellel, M., J.-L. Chatelain, B. Guillier, D. Machane, R. Bensalem, E. H. inversion using a direct search algorithm and its application to
Oubaiche, and H. Haddoum (2010). Heavier damages without ambient vibration measurements, Near Surf. Geophys. 2, no. 4,
site effects and site effects with lighter damages: Boumerdes City 211–221.
(Algeria) after the May 2003 earthquake, Seismol. Res. Lett. 81, Yamanaka, H., M. Takemura, H. Ishida, and M. Niwa (1994). Character-
37–43, doi: 10.1785/gssrl.81.1.37. istics of long-period microtremors and their applicability in explo-
Hobiger, M., C. Cornou, M. Wathelet, G. Di Giulio, B. Knapmeyer- ration of deep sedimentary layers, Bull. Seismol. Soc. Am. 84, no. 6,
Endrun, F. Renalier, and N. Theodoulidis (2013). Ground structure 1831–1841.
imaging by inversions of Rayleigh wave ellipticity: Sensitivity analy- Zor, E., S. Özalaybey, A. Karaaslan, M. C. Tapırdamaz, S. Ç. Özalaybey,
sis and application to European strong-motion sites, Geophys. J. Int. A. Tarancıoğlu, and B. Erkan (2010). Shear wave velocity structure
192, no. 1, 207–229. of the İzmit Bay area (Turkey) estimated from active-passive array

Seismological Research Letters Volume 87, Number 5 September/October 2016 7


SRL Early Edition
surface wave and single-station microtremor methods, Geophys. J. Mustapha Hellel
Int. 182, no. 3, 1603–1618. Faculté des Sciences de la Terre, de la Géographie et de
l’Aménagement du Territoire
El Hadi Oubaiche1 Université des Sciences et de la Technologie Houari Boume-
Djamel Machane dienne (USTHB)
Rabah Bensalem BP 32 El Alia Bab Ezzouar
Centre National de Recherche Appliquée 16111 Alger, Algérie
en Génie Parasismique (CGS)
Rue Kaddour Rahim Prolongée Abderrahmane Bouguern
Hussein Dey BP 252 Faculté des Hydrocarbures et de la Chimie
16040 Alger, Algérie Université M’Hamed Bougara de Boumerdes (UMBB)
eoubaiche@yahoo.com Rue de l’indépendance
35000 Boumerdès, Algérie
Jean-Luc Chatelain
Marc Wathelet Published Online 13 July 2016
Université Grenoble-Alpes
Institut de Recherche pour le Développement (IRD) – ISTerre 1
Also at Faculté des Hydrocarbures et de la Chimie, Université M’Hamed
CS 40700 Bougara de Boumerdes (UMBB), Rue de l’indépendance, 35000 Bou-
38058 Grenoble, CEDEX 9 merdès, Algérie.
France

8 Seismological Research Letters Volume 87, Number 5 September/October 2016

View publication stats

Das könnte Ihnen auch gefallen