You are on page 1of 34

Constraining site response and shallow

geophysical structure by ambient noise


measurements and 1D numerical
simulations: the case of Grevena town (N.
Greece)
Marios Anthymidis, Nikos Theodoulidis,
Alexandros Savvaidis & Constantinos
Papazachos

Bulletin of Earthquake Engineering


Official Publication of the European
Association for Earthquake Engineering

ISSN 1570-761X

Bull Earthquake Eng


DOI 10.1007/s10518-012-9378-3

1 23
Your article is protected by copyright and
all rights are held exclusively by Springer
Science+Business Media B.V.. This e-offprint
is for personal use only and shall not be self-
archived in electronic repositories. If you
wish to self-archive your work, please use the
accepted author’s version for posting to your
own website or your institution’s repository.
You may further deposit the accepted author’s
version on a funder’s repository at a funder’s
request, provided it is not made publicly
available until 12 months after publication.

1 23
Author's personal copy
Bull Earthquake Eng
DOI 10.1007/s10518-012-9378-3

ORIGINAL RESEARCH PAPER

Constraining site response and shallow geophysical


structure by ambient noise measurements and 1D
numerical simulations: the case of Grevena town
(N. Greece)

Marios Anthymidis · Nikos Theodoulidis ·


Alexandros Savvaidis · Constantinos Papazachos

Received: 21 April 2011 / Accepted: 16 August 2012


© Springer Science+Business Media B.V. 2012

Abstract A large number of earthquake studies using both empirical and theoretical
approaches clearly depict the strong correlation of resulting damage and local geology. This
correlation forms the scientific basis of the subsurface soil structure studies for site response
evaluation, since adequate knowledge of the subsoil geotechnical and geophysical proper-
ties can lead to realistic seismic hazard estimation through appropriate modeling of strong
seismic motion. In this framework, ambient noise analysis and one dimensional numerical
simulation modeling have been performed for the town of Grevena (Northwestern Greece)
in order to study both the subsurface soil structure, as well as its expected effect on seismic
motions. The horizontal to vertical spectral ratio (HVSR) method was implemented on an
almost uniform grid of 60 single station ambient noise measurements inside the urban area,
while the noise array technique was applied at four selected sites of the study area. The
HVSR curves show complex patterns, occasionally with double HVSR peak frequencies,
with the higher one HVSR frequency showing a good correlation with the recent Holocene
clay-dominant formation. Using these results, as well as information on surface geology and
existing geotechnical data, a microzonation of the Grevena town is attempted, indicating
zones of similar site response to ground motion. Numerical simulation of ambient noise with
synthetic recordings suggests that the complex features observed in several HVSR curves
could be attributed to the impedance contrast due to the quite shallow (Holocene clays to Pli-
ocene–Pleistocene sands) and a deeper (Pliocene–Pleistocene sands to Oligocene–Miocene
bedrock) sedimentary formation.

Keywords Ambient noise · Array technique · HVSR method · Shear wave velocity ·
Site effects · Subsurface structure · 1D numerical simulation

M. Anthymidis (B) · C. Papazachos


Division of Geophysics, Department of Geology, Aristotle University of Thessaloniki,
54124 Thessaloniki, Greece
e-mail: manthymi@geo.auth.gr

N. Theodoulidis · A. Savvaidis
Earthquake Planning and Protection Organization, Institute of Engineering Seismology and Earthquake
Engineering (ITSAK), Thessaloniki, Greece

123
Author's personal copy
Bull Earthquake Eng

1 Introduction

Local geology can significantly affect the main seismic ground motion characteristics such as
amplitude, frequency content and duration, amplifying them in the overwhelming majority
of cases. The main consequence of this phenomenon is that in large urban environments,
usually built on relative recent geological formations (e.g. Quaternary sedimentary basins),
constructions are more prone to damage in case of a strong earthquake. Seismic damage
has significant social, political and economic impact to urban areas, hence knowledge of
geological and geophysical subsurface structure (usually shear wave velocity variation with
depth) and its effect on strong seismic motion are important factors for efficient seismic risk
mitigation.
One of the most reliable approaches for site effect evaluation is to rely on observed
earthquake recordings. The typical procedure involves the simultaneous recording of strong
ground motion using sensors placed on different sites (stiff geological formations, soft sed-
iment deposits, etc.) during an earthquake. The spectral ratio of the horizontal component
recorded on soft soil formations to the equivalent component recorded on hard rock (reference
site) provides the experimental transfer function, namely the amplification of seismic ground
motion on sediments with respect to the reference site. The main prerequisite of this approach
is that the distance of the seismic source should be much greater than the distance between
the recording stations (Borcherdt 1970; Steidl et al. 1996). The aforementioned analysis of
earthquake recordings is the so called standard spectral ratio (SSR) method (Borcherdt 1970).
In many cases it is not possible to locate an appropriate reference site, hence site effects char-
acterization needs to be performed using a single station analysis. In this case, the spectral
ratio of the horizontal to vertical component, accounting only for the shear wave part of the
recording (Langston 1979; Lermo and Chávez-García 1993), represents the receiver function
(RF) and often provides the fundamental resonance frequency ( f 0 ) of the upper sedimentary
layers, as well as an approximate lower estimate of the ground motion amplification factor
(A0 ) at this frequency.
The main disadvantage of the previous methods (SSR and RF) is the large time period
needed in regions with low or moderate seismicity to acquire earthquake recordings with sig-
nal to noise ratio exceeding a predetermined threshold, in order to obtain satisfactory results
(Field and Jacob 1995). Several studies (Chávez-García et al. 1990; Duval et al. 1994; Kudo
1995; Bour et al. 1998; Riepl et al. 1998; Bindi et al. 2000; Bard et al. 2004 among them)
have shown that the horizontal to vertical spectral ratio (HVSR) method based on ambient
noise recordings can be used to reliably evaluate f 0 , though A0 values derived from ambient
noise recordings, underestimate the real amplification of seismic ground motion (Haghshenas
et al. 2008). In spite of this limitation, the HVSR method is widely used for microzonation
studies because of several important advantages, such as the relatively low cost, the minimal
requirements for equipment and human power, it’s easy and quick application in urban areas,
the fast data processing, reliability of the results and its passive character.
An alternative to experimental techniques is to employ theoretical simulation approaches
that take advantage of all available subsurface geotechnical and geophysical models, since
geological and geotechnical surveys (boreholes, laboratory tests in core samples) in com-
bination with active geophysical methods (crosshole and downhole measurements, seismic
reflection and refraction, geoelectrical tomography) provide quite accurate estimates of the
subsoil geophysical properties. During the last decades, information of subsurface structure
is also derived from passive methods employing ambient noise measurements. Array tech-
niques applied on ambient noise recordings have been proved to be a useful tool for the
estimation of one dimensional (1D) models of shear wave velocity (Vs ) from the inversion of

123
Author's personal copy
Bull Earthquake Eng

surface waves’ dispersion curves contained in the ambient noise wavefield (Aki 1957; Asten
1978). Compared to geotechnical and geophysical surveys, ambient noise analysis with array
technique offers several important advantages (Satoh et al. 2001), especially for large thick-
ness sedimentary basins, such as easy application in urban environments, no requirements
for artificial seismic source and deep soil properties investigation with an extremely lower
application cost.
The determined subsurface structure can not only be used for numerical simulation of seis-
mic motions but can also be indirectly validated using the same approach. Such an approach
employs the subsurface structure model with specific soil properties (number and thickness
of layers, Vs distribution, damping factors) derived from the survey and a virtual set of
sources randomly distributed in space and time, in order to generate synthetic recordings
using numerical simulation methods. Analysis of these synthetic recordings can provide the
target site ground motion response for the specific subsurface structure model. Therefore,
the comparison of the determined synthetic recordings with experimental noise results can
provide information on the subsurface model validity and its reliability. Certainly, since
the upper layers of the crust are the most heterogeneous and complex (e.g. anisotropic),
it is not often possible to obtain a perfect similarity between synthetic and experimental
recordings.
In the present study, ambient noise methods and 1D numerical simulation modeling for
subsurface structure model validation were applied for the town of Grevena (Northwestern
Greece). The study area is characterized by moderate seismicity, with quite large recurrence
periods of large magnitude mainshocks (Papazachos 1990). The most recent destructive
earthquake occurred on 13th of May, 1995 and caused considerable seismic damage to the
surrounding area of Grevena town. The results obtained in the present study can be used
for microzonation purposes of Grevena town, as well as for the appropriate design of its
expansion plan.

2 Geological setting

The study area belongs to the Subpelagonic geological zone, which covers a narrow part of
Greece, having a dominant NW–SE development direction, parallel to the central mountain
ridge (Pindos) of Greek mainland (Mountrakis 1985). The Subpelagonic zone petrology con-
sists of Mesozoic age carbonate rocks (mainly limestones) and ophiolite complexes coupled
with related sedimentary formations, typical of deep sea environment. These Subpelagon-
ic rocks correspond to the geological bedrock formation of the area and are covered by
very thick Paleogene–Neogene sediments (the maximum thickness locally exceeds 5.000 m)
of the Mesohellenic Graben. Emplacement of Mesohellenic Graben’s sediments occurred
during Oligocene–Miocene and they are mainly of molassic type, with their structure char-
acterized by alternations of deep and shallow sea sediments with interjections of lacustrine
and fluvial phases. Mesohellenic Graben sedimentary formations were not affected from the
compression stages of Alpine orogenetic cycle (which took place during Paleocene–Eocene
of Cenozoic era in the Greek region) and hence no compression deformation is present
(folds, thrust and reversed faults). Nevertheless, they appear slopped and influenced mainly
from normal faults, due to the later extension tectonic incidents. These Tertiary Mesohel-
lenic Graben sediments are covered by recent Pliocene–Pleistocene fluvial and lacustrine
deposits, as well as the most recent Quaternary (Holocene) land deposits of relative small
thickness.

123
Author's personal copy
Bull Earthquake Eng

Fig. 1 Schematic geological map of the study area. The solid line rectangle depicts the area shown in Figs.
4, 5 and 6, whereas the dashed line rectangle the area presented in Fig. 2

The geology of the study area (Fig. 1) shows that Grevena town is built upon the very
recent Holocene formations which were deposited at the edges of the river’s bed crossing the
town. The underlying Pliocene–Pleistocene sedimentary formations are identified over the
fringes of Grevena town, mainly to the northern and southwestern part of the broader study
area, which also exhibits the maximum altitude (∼600 m). The older (Oligocene–Miocene)
Mesohellenic Graben’s formations crop out in the south, west and northwest of the town,
at an approximate distance of 5 Km from the town center. This formation is considered as
the effective bedrock formation of the region (since the actual bedrock, corresponding to
Subpelagonic limestones, is located at a depth of several kilometers) and consists of a series
of cohesive conglomerates and sandstones. In summary, Grevena town is built on a typical
but complex sedimentary basin, where hard rock formations (effective geological bedrock)
are covered by two younger and softer sedimentary layers of more recent geological age,
hence their response to seismic motion can potentially influence the seismic risk assessment
of this urban area.

3 Geophysical and geotechnical data

No geophysical survey results (active or passive) have been previously published for the
town of Grevena, hence geophysical parameters concerning the subsurface structure proper-
ties, such as distribution of seismic wave velocities with depth, were not available. However,
data from geotechnical surveys were collected, mainly from small depth drillings (typical

123
Author's personal copy
Bull Earthquake Eng

Fig. 2 Positions of boreholes and array technique application sites in the study area. Boreholes with prefix
D, G or Gb correspond to geotechnical drillings for road or building construction, whereas those with prefix
WD for water supply. Geological formations are in accordance with Fig. 1

∼25 m) for building and road construction, as well as relative large depth boreholes (100–
170 m) for water supply purposes. The main geotechnical borehole information available con-
cerned the lithology derived from borehole cores and classification of soils according to the
American Unified Soil Classification System (AUSCS), underground water level and data
from standard penetration tests (SPTs), unconfined compression tests and undrained triaxial
compression tests (U-U). Information from water drills was limited to the subsoil lithology
and the underground water level. The total number of the available boreholes was 23 and
their positions are presented in Fig. 2. Boreholes with prefix D and G/Gb in Fig. 2 correspond
to geotechnical drillings for building and road construction, respectively, while those with
prefix WD to water drills. A typical example of the subsurface lithology, the underground
water level and SPT data (for the geotechnical boreholes) is illustrated in Fig. 3 for three
boreholes within the study area.
The subsurface geological structure derived from the available geological and geotechni-
cal data for the study area is in very good agreement with the surface geology. The superficial
Holocene sedimentary layers consist mainly of dark brown clays and silts, locally with the
presence of coarse grain materials such as sand, gravels and pebbles. The deeper Pliocene–
Pleistocene sedimentary layers consist of a dense mixture of sand and gravels, with fine
grained connective material and frequent intersection of round ophiolite-calcareous pebbles.
The deeper formations of the Mesohellenic Graben (cohesive base conglomerates and sand-
stones) correspond to the effective bedrock formations of the area. Similar formations have
been detected only in three boreholes (D5, G3 and G4 in Fig. 2), at depths ranging between
12.5 and 15 m but are completely absent from the stratigraphy column of water drills which
provide information down to 170 m. Therefore, it is an open question whether the presence of

123
Author's personal copy
Bull Earthquake Eng

Fig. 3 A typical example of available information for three boreholes positions. In all cases, the subsurface
structure lithology is illustrated, together with the shallow water level and SPT data (when available)

the cohesive formations at small depths is due to fault activation or, most probably, it concerns
slide rocks (olistoliths) or even some isolated layers that locally interject the stratigraphic
column.
According to the geological structure derived from boreholes information (Fig. 3), a pro-
gressive depth graduation of the sedimentary materials is observed, from fine grained (at
the superior Holocene sedimentary formations) to coarser grained at lower layers (Pliocene–
Pleistocene formations). A consequence of this pattern is that the limit of the two upper
sedimentary layers is not always clear. Therefore, the velocity (or impendence) contrast
between the two formations may vary from site to site, depending on the presence of fine and
coarse grain materials in the upper and lower layers, respectively. This can have important
consequences, since if the impendence contrast is not high enough, ambient noise methods
may not be successful, as in these cases (low impendence contrast) the energy carried out by
surface waves in the wavefield of ambient noise is significantly low.
A probable geological structural model for the study area can be compiled using all the
available geological and geotechnical data. The superficial clay sedimentary layer (typically
Holocene) displays a relative constant thickness, ranging from 3 to 12 m, for the whole area
where the Grevena town is settled. The underlying sedimentary layer consists of a dense
mixture of sand and gravel with medium to cohesive conglomerates (typically Pliocene–
Pleistocene). The thickness of this formation seems to present considerable variations, with
values in the range of 100–200 m. Both sedimentary layers lay on the stiff Mesohellenic
Graben bedrock formations, consisting mainly of cohesive conglomerates and sandstones.
The upper surface of this effective bedrock layer probably displays strong undulations and
discontinuities due to faulting activity. In summary, the geological subsurface structure of
the study area is likely to be considered as a three-dimensional (3D) sedimentary basin, with
the geophysical properties of the three main geological formations varying in both horizontal
and vertical directions.

123
Author's personal copy
Bull Earthquake Eng

4 Ambient noise analysis

Ambient noise methods are exploited in four main ways regarding the study of local site
conditions (Bard 1999), which employ: Absolute Spectra, Site to Reference Spectral Ratios,
Horizontal to Vertical Spectral Ratio (or HVSR) and Array Technique. In the present study
we focus on the application of HVSR method and array technique on ambient noise record-
ings, which provides information mainly on the response of the upper sedimentary layers to
ground motion (mainly fundamental resonance frequency, f 0 , estimates), whereas the array
technique can be applied to determine the Vs profile of the subsurface structure.
An important issue for the application of HVSR method and array technique in the study
area was the evaluation of the impact of diurnal and seasonal variations of ambient noise.
Systematic variations of the noise level are observed during the day in the frequency band
of 2–50 Hz that is mainly related to human activity (Fyen 1990). Since such ambient noise
variations have been identified mainly during day working hours (08:00–16:00), all ambient
noise data were collected during the remaining hours of the day.

4.1 HVSR method application

The theoretical background behind the application of HVSR method on ambient noise mea-
surements is not yet fully clarified among the scientific community. Interpretation of the
HVSR method results is based on two main assumptions concerning the composition of
ambient noise wavefield. According to the first assumption, the presence of surface waves in
ambient noise wavefield is dominant (Nogoshi and Igarashi 1971). Thus, the HVSR curve
is correlated with the ellipticity of Rayleigh waves, which mainly affects the vertical com-
ponents. The second assumption considers that the ambient noise wavefield consisting of
both body and surface waves; hence the derived spectral ratio can be associated with the site
transfer function (Nakamura 1989, 2000). In spite the discrepancy of the theoretical back-
ground, both assumptions agree that the implementation of HVSR method on ambient noise
data recorded with a single station is capable to provide the fundamental resonance frequency
( f 0 ) of the upper sedimentary layers with respect to the underlying bedrock and a probable
lower bound of seismic ground motion amplification factor (A0 ). This conclusion is valid for
any 1D subsurface structure and not only for the simple cases of a layer overlying a bedrock.
It should be noted that using these measures, additional information can be estimated, such as
the soil vulnerability index, K g (= A20 / f 0 ), which attempts to quantify seismic strain (hence
damage) on ground and structures (Nakamura 1996, 2000).
Ambient noise data were collected with a quite uniform grid of 60 measurement sites inside
the urban area of Grevena (Fig. 4), as well as in regions where the future town expansion
is planned. The data acquisition system was composed by a Cityshark I recorder (Chatelain
et al. 2000) connected with a Lennarz 3d/5s sensor (seismometer). A GPS system provided
the geographical positions of every measurement point. The recording duration of each mea-
surement point was set to 20 min, with the sample rate fixed at 100 Hz.
The data analysis was performed using Geopsy (Geophysical signal database for noise
array processing) software included in the software package sesarray (Wathelet 2005). Three
processing steps were followed; time windows selection, calculation of spectra and calcula-
tion of spectral ratios. The window length was set to 25 s, with no overlap between successive
ones and a 5 % cosine taper was applied to all time windows. The minimum number of time
windows depends on the f 0 value of each site and was at least 15 for sites with f 0 greater
than 2 Hz, whereas it was increased to 20 for sites with f 0 less than 2 Hz. Spectra calculation
was based on FFT (fast Fourier transform) method and derived for all time windows of each

123
Author's personal copy
Bull Earthquake Eng

Fig. 4 Single station measurement sites of ambient noise in the study area. The numbers correspond to the
ID of every measurement point. Geological formations are in accordance with Fig. 1

ground motion component. Thus, the final Fourier spectrum for each ground motion compo-
nent was the arithmetic average of spectra derived from every time window, with a typical
frequency range of 0.2–20Hz. Spectral amplitudes were smoothed with a 40 % factor using
the methodology proposed by Konno and Ohmachi (1998). The final processing stage was the
calculation of spectral ratio between the horizontal to vertical component. Horizontal compo-
nents were geometrically averaged and divided with the vertical one for every time window,
with the final HVSR curve computed as the arithmetic average of all time windows. It should
be noted that the interpretation of HVSR curves was based on the available geological and
geotechnical data, in combination with the recommended quantitative criteria concerning the
reliability of the HVSR curves and the “clarity” of the observed peak(s) (SESAME 2004).
Therefore, an effort was made to minimize the subjectivity of interpretation, as well as to
estimate the robustness of HVSR method results related to site effects.
Evaluation of site response on ground motion by single station noise measurements refers
to the vicinity of the measurement point. Consequently, the application of the HVSR method
provides information on the local 1D structure, i.e. a representative “column”, with properties
(layer thickness, velocity distribution of seismic waves) varying only with depth. However, if
a satisfactory number of ambient noise measurements are available, it is possible to identify
the spatial distribution of f 0 and A0 values, especially if the measurements are performed
along a dense measurement grid.
The obtained results of the HVSR method exhibited in general two groups of resonant
peaks. The first group corresponded to relatively broad peaks, typically in the range of 1–2
Hz. The second group of peaks was identified at much higher frequencies, widely distrib-
uted between 3 and 20 Hz, often with much sharper HVSR amplitude peaks. In some cases,
both types of peaks were clearly identified on the HVSR curves. It should be noted that for

123
Author's personal copy
Bull Earthquake Eng

several sites a practically flat HVSR curve was observed since no significant peak(s) could
be identified inside the frequency range of 0.2–20 Hz. A probable explanation for this pattern
is the low impendence contrast of the two shallow sedimentary layers formations. A possible
alternative is the presence of a hard rock formation at very shallow depths, however, this
scenario must be ruled out according to available geological and geotechnical information.
In the case of low impendence contrast sites, the HVSR method is not able to describe the
site response of the subsurface structure to ground motion. Therefore, observation of flat
HVSR curves at a site does not necessarily imply negligible or even small ground motion
amplifications. In summary, four different groups of patterns of HVSR curves were observed,
which either display a single low frequency peak, a single high frequency peak, double peak
in both low and high frequencies and a flat HVSR curve, allowing the possible identification
of zones in the study area with similar response behavior.
In order to study the significance of these frequency peaks on the HVSR curves, the cor-
responding spatial distribution of the first, lower frequency group of peaks, denoted as f 0
and the second, higher frequency group of peaks, denoted as f 1 , is presented in Fig. 5. The
corresponding HVSR amplitude distributions, similarly denoted as A0 and A1 , are presented
in Fig. 6, respectively, for all ambient noise measurement sites. Notice that both Figs. 5 and
6 also contain the results of the corresponding individual HVSR curves from array noise
measurements (Fig. 2, presented later in detail), where results for the inner and middle array
circle are presented as a single point average (due to the spatial scale of the corresponding
maps of Figs. 5, 6) and results for the array outer circle are presented as individual mea-
surement points. In general the results from array measurements are in good agreement with
the neighboring corresponding single station estimates, though somewhat lower values are
observed for the DEYAG site which is the only case where f 0 is lower than 1 Hz.
From Figs. 5 and 6 it is clear that sites with higher HVSR resonant frequency, f 1 , values
(greater than 3 Hz) are mainly located on the Holocene formations to the northern part of
the Grevena town. In most cases the corresponding HVSR amplitude values (denoted as A1 )
range between 2 and 4, though higher values are also locally observed. It is quite evident
that the HVSR method is quite capable to recognize the influence of the relatively thin recent
Holocene sedimentary layer and its local contact with the underlying Pliocene–Pleistocene
formations on the site response. On the other hand, the lower HVSR resonant frequency, f 0 ,
values (less than 3 Hz) can be identified at much less observation sites, mostly located in
the western section of the city, however showing a rather complicated spatial distribution.
The corresponding HVSR curves and the low f 0 values observed suggest that the local site
response can be interpreted as the influence from the deeper geological formations (effective
bedrock consisting of Oligocene–Miocene molassic formations), almost independently from
the Holocene to Pleistocene–Pliocene contrast of the upper sedimentary layers. An alterna-
tive interpretation is to consider the possibility of strong lateral variation in the superficial
layers (e.g. velocity inversion), a pattern quite common in such a complex sedimentary basin.
Again for most sites the HVSR amplitude values (denoted as A0 ) range between 2 and 4,
though some sites exhibit higher values.
In order to examine the validity of the proposed explanations for the dual ( f 0 and f 1 )
resonant frequencies observed in the HVSR curves, it was necessary to obtain information
regarding the subsurface geophysical structure (e.g. velocity distribution). For this reason
ambient noise array measurements were performed at selected sites and combined with the
available geotechnical information, in order to create realistic velocity distributions for the
study area and perform appropriate noise modeling for testing the proposed dual resonant
frequency hypothesis.

123
Author's personal copy
Bull Earthquake Eng

Fig. 5 Spatial distribution of (a) the lower, f 0 , and (b) the higher, f 1 , resonant frequency, as revealed by the
HVSR single station ambient noise measurements (circles Single station HVSR, octagons Array measure-
ments). Geological formations are in accordance with Fig. 1

123
Author's personal copy
Bull Earthquake Eng

Fig. 6 Spatial distribution of the HVSR Amplification factor for (a) the lower, A0 , and (b) the higher, A1 ,
resonant frequency sites, as revealed by the HVSR single station ambient noise measurements (circles Single
station HVSR, octagons Array measurements). Geological formations are in accordance with Fig. 1

123
Author's personal copy
Bull Earthquake Eng

4.2 Array technique application

The application of array techniques leads to satisfactory results regarding the determination
of the local geophysical structure, under the hypothesis that ambient noise wavefield mainly
consists of surface waves which propagate through an approximately horizontally stratified
structure (Tokimatsu 1997). Inside this 1D heterogeneous medium, surface waves display
dispersion phenomenon, namely the variation of apparent phase velocity with frequency. Dis-
persion characteristics are strongly dependant on body waves’ velocity (mainly S-waves),
density and thickness of each layer which comprises the subsurface structure (Murphy and
Shah 1988; Aki and Richards 2002). Thus, dispersion (or phase velocity) curves of surface
waves are directly associated with the subsurface structure properties.
Analysis of ambient noise data using array techniques follows two main processing stages,
namely the extraction of surface waves’ dispersion curve and the application of inversion
methods for the estimation of the local 1D seismic wave velocity distribution. The extraction
of the dispersion curve from ambient noise data is usually achieved using the frequency-
wavenumber ( f –k) method (Lacoss et al. 1969; Capon 1969; Horike 1985; Kvaerna and
Ringdahl 1986; Okada 2003). The f –k method is able to detect high energy seismic waves
contained in ambient noise complex wavefield, although it is not possible to recognize the
nature (body or surface waves, fundamental or higher modes) of these waves (Okada 2003).
Hence, the extracted dispersion curve is related with surface waves, because we assume that
they are dominant in ambient noise recordings.
The f –k method considers that ambient noise sources are distant from the recording sta-
tions, so the incoming wavefiled consists from plane waves. Assuming a plane wave with
frequency, f , propagating in a homogenous medium with random velocity and direction (or
equally with wavenumbers k x and k y across the horizontal axis X and Y respectively), arrival
times can be easily calculated in every recording station of the array. Therefore, arrival times
can be shifted according to their respective time delays with respect to a reference station
(usually chosen as the center of the array geometry), stacked and divided to several segments
called time windows. For all time windows the f –k spectrum and cross spectrums are cal-
culated for the available frequency range (Okada 2003; Ohrnberger et al. 2004) and the f –k
cross spectrum values are examined on the wavenumber (k x , k y ) plane (array output) for
each frequency. The maximum value of the array output on the wavenumber plane allows
the estimation of the apparent velocity and impinging azimuth of waves propagating through
the array. Dispersion curve extraction over an adequate frequency range at each site depends
on the geometry and size (aperture) of the array and the spectral amplitudes of ambient noise
wavefield that are modified due to the filtering effect (e.g. attenuation of seismic wave energy)
of the subsurface structure (Woods and Lintz 1973; Wathelet et al. 2007). The determination
of the usable frequency range from the dispersion curve can be computed by the theoretical
array response function or the array transfer function (Woods and Lintz 1973; Asten and
Henstridge 1984; Wathelet 2005; Wathelet et al. 2007), which describes the ability of the
applied array geometry to reliably extract the corresponding part of the dispersion curve.
In the present study, the inversion of the surface waves’ dispersion curve was based on
a neighborhood algorithm which allows the incorporation of a priori information about the
subsurface structure parameters (Wathelet et al. 2004; Wathelet 2005). The neighborhood
algorithm is a stochastic direct search method for finding models of acceptable data fit inside
a multi-dimensional parameter space (Sambridge 1999). The inverted structure parameters
(parameter space) are the P and S wave velocities, density and thickness of each layer. The
algorithm generates theoretical ground models with a random seed from the parameter space
and the dispersion curves are computed (forward problem) for all these models. Probability

123
Author's personal copy
Bull Earthquake Eng

density is assumed to be uniform over the whole parameter space, whereas the space limits
are defined a priori from the user. For a satisfactory investigation of the parameter space,
the number of dispersion curve computations can be very high (a few thousands to tens
of thousands). Furthermore, due to data uncertainties and the non-linearity of the problem
itself, the solution of the dispersion curve inversion is generally non-unique. Comparison of
the computational results with the experimental dispersion curve provides misfit values that
quantify the quality of the model generated curves. If the experimental dispersion curve is
given with an uncertainty estimate, the misfit is given by (Wathelet et al. 2004):

 nF
 (xdi − xci )2
mis f it =  (1)
i=0
σi2 n F

where xdi is the velocity of the data curve at frequency f i , xci is the velocity of the calculated
curve at frequency f i , σi is the uncertainty of the frequency samples considered and n F is the
number of the frequency samples considered. If no uncertainty is provided, σi is replaced by
xdi in Eq. (1). The originality of the neighborhood algorithm is to use previous seeds from the
parameter space for guiding the search to improved models. When the computational results
do not display an appropriate data fit (high misfit values, systematical deviation from the
experimental dispersion curve, etc), then the limits of parameter space should be redefined.
Probably the best compromise is to start with the simplest model and progressively add new
layers if the data are not sufficient matched (Scherbaum et al. 2003).
Measurements of ambient noise using the array technique were performed in four selected
sites inside the urban area of Grevena town (Fig. 2). Selection of sites for array installation
followed several criteria. An effort was made to have satisfactory coverage of the study area,
keeping in mind that every site should have different response to ground motion, according
to the former HVSR method results. Moreover, the application site should have appropriate
field conditions, e.g. located far from strong topographic anomalies and artificial structures.
Finally, nearby borehole availability was considered in order to cross-check array technique
results with the observed subsurface geological and geotechnical structure. The data acquisi-
tion was performed by Ref Tek 130-01 recorders connected with Güralp CMG-40T broadband
sensors (flat response between approximately 0.033–50 Hz).
The array geometry for every site was set up following a circular arrangement of thir-
teen recording stations, separated in groups (Fig. 7). Every group of sensors inside the array
geometry had a different radius (or aperture) in order to retrieve a different part of the disper-
sion curve in a certain frequency range, while the circular pattern of each group ensures the
recording of ambient noise wavefield from all azimuths. As a small aperture array provides
information about the low frequency part of the dispersion curve and large aperture for high
frequencies, it is clear that reconstruction of the dispersion curve for a wide frequency range
implies installation of several arrays with different apertures. In the present study, the array
technique was applied with the installation of recording stations on three concentric circles.
One recording station was placed at the center, six on the periphery of the inner one and three
on the periphery of the middle and outer circle. An attempt was made to maintain equal dis-
tribution of recording stations on the periphery of every circle. Therefore, the angles between
successive stations in relation with the center were constant, approximately 60◦ for the inner
circle and 120◦ for the middle and the outer one. Deviations from the ideal positions due
to field conditions at each measuring site (installation area, topographic anomalies, artificial
obstacles, etc) where minimized whenever possible.
The maximum aperture for the circular array geometry depends on the expected subsurface
structure characteristics (bedrock depth, variation of elastic properties), penetration depth,

123
Author's personal copy
Bull Earthquake Eng

Fig. 7 Array geometries employed for every application site

fundamental resonance frequency ( f 0 ) of the site and field conditions. Deep geophysical
targets require increasing penetration depth, which can be only achieved with larger array
apertures. A typical rule of thumb is that the penetration depth is roughly one third of the
maximum array aperture, though even with large arrays ambient noise methods encounter a
difficulty to obtain the part of the dispersion curve just below the site resonance frequency
( f 0 ) due to the lack of coherency of propagating waves (Wathelet et al. 2007). For the study
area, the main target was to locate the interface depth and S-wave velocity distribution of the
recent Holocene superficial deposits and the underlying Pliocene–Pleistocene sedimentary
formations. Based on the previously described geological and geotechnical information, this
contact was expected to be in the range of 5–20 m, thus the minimum aperture of the array
geometry was set to 60 m for all application sites. Moreover, since the underlying Mesohel-
lenic Graben bedrock formations were expected to be located at depth larger than 100 m,
an array aperture greater than 200 m was required. Unfortunately, installation of such arrays
in the specific urban environment was possible only in one site (GPD in Figs. 2, 7). The
final minimum and maximum apertures, f 0 parameter and estimated penetration depths are

123
Author's personal copy
Bull Earthquake Eng

Fig. 8 Vs profile (left part) for the DEYAG site and corresponding theoretical dispersion curves (right part)
plotted together with the experimental one (black curve with error bars). Every color corresponds to a different
value of the misfit function

Table 1 Minimum and maximum aperture of array geometries, resonance frequency ( f 0 and f 1 ) of the site
and empirically estimated penetration depth for every array site

Array Minimum aperture Maximum aperture f 0 or f 1 (Hz) Empirically


of array geometry of array geometry estimated
(m) (m) penetration
depth (m)

DEYAG 7 60 7.2 20
GPD 20 260 1 85
PLT 10 80 2.5 25
TEI 15 160 1 55

listed in Table 1 for every application site. Ambient noise measurements were simultaneously
performed in all array stations for one and a half hour, with a fixed sample rate of 125 Hz.
Processing of ambient noise array data was performed with the software package ses-
array (Wathelet 2005), which implements the f –k and inversion processing on array noise
data. Before the f –k analysis, the theoretical array response function was calculated, which
correspond to the specific array geometry used at each site. This response function pro-
vides a lower (kmin ) and upper (kmax ) limit of wavenumbers for which the dispersion curve
extraction can be reliably performed. Surface waves propagating with wavenumbers smaller
than kmin /2 value cannot provide reliable information for the dispersion curve, whereas for
wavenumbers larger than kmax aliasing phenomena are introduced to the results. For the

123
Author's personal copy
Bull Earthquake Eng

previous wavenumber range, the vertical components of the ambient noise recordings were
used for further processing, since we want the extracted experimental dispersion curve to
correspond only to Rayleigh surface waves and minimize Love wave interference. The noise
signal was divided in several time windows and for each of windows the Rayleigh waves’
velocity was computed from the maximum array output value on the wavenumber (k x , k y )
plane, within the previously defined wavenumber range. The phase velocity values from all
examined time windows were then grouped and appropriate histograms were constructed for
each frequency step (Ohrnberger et al. 2004). An appropriate probability density function
was applied to each constructed histogram, in order to obtain the mean phase velocity value
and its standard deviation for every frequency, and construct the final dispersion curve.
The final processing stage corresponded to the inversion of the experimental dispersion
curve (derived from f –k analysis), in order to obtain the local 1D shear wave velocity dis-
tribution at each site. Before the inversion, a priori information on the maximum limits of
the inverted parameters was set. The starting model for all sites corresponded to a single
layer overlying a halfspace, with very broad distribution of the parameter values (Vs between
50 and 2,000 m/s and layer thickness ranging between 1 and 50 m), in an attempt to simu-
late (but not actually constrain) the Holocene layer presence over the Pliocene–Pleistocene
stiffer formations. Inversion started with random seed of the neighborhood algorithm from
the parameter space. For every set of parameters the theoretical dispersion curve was com-
puted, together with a misfit value. The inversion procedure stopped when the misfit value
could not be further reduced. In case the final data fit was not satisfactory, the parameter
space was reset with different ranges and the inversion algorithm was restarted. For every
site, the inversion procedure was repeatedly applied for several starting parameter spaces, in
order to test the stability of the results and obtain a good correlation between theoretical and
experimental dispersion curve.
The final results from the array technique are presented for every site in Fig. 8, 9, 10
and 11 and are summarized in Table 2. Velocity profiles are presented for every generated
model, while the corresponding theoretical dispersion curves are plotted together with the
experimental one (black line with standard deviation bars for every data point). A rather
good correlation of the minimum misfit theoretical dispersion curve with the experimental
one is observed for all sites, while most models (except for site DEYAG) exhibit a rather
sharp increase of the Vs velocity distribution at a relatively shallow depth (7–12 m), approx-
imately where we expect the previously mentioned contact of the sedimentary formations.
The inverted velocity distribution stops at depth of 30 m for every site, as the depth to the
Miocene bedrock formations is too large to be resolved with the available dispersion curve
frequency range.
The results presented in Fig. 8, 9, 10 and 11 and in Table 2 show that shear wave velocities
for the upper Holocene sedimentary layer (mainly consisting of clay) vary between 150 and
550 m/s, with a layer thickness of 7–12 m. The corresponding shear wave velocity of the
lower Pliocene–Pleistocene (mainly sand and gravel) formations vary between 1,100 and
1,800m/s. These velocity values are rather large for the specific type of geological formation,
despite the presence of fine grain connective material and rock pebbles in these formations,
which increases the mean velocity. Nevertheless, the derived velocity profiles from the array
technique coincide very well with the geological structure defined from geotechnical bore-
holes, as the changes in velocity distribution generally follow the observed contacts between
the geological formations (see example in Fig. 12). In addition, the SPT data indicate that
the Pliocene–Pleistocene formations are quite stiff, since the penetration of the probe is not
able to reach the suitable threshold (points with “R” in the NSPT plot of Fig. 12). Thus,

123
Author's personal copy
Bull Earthquake Eng

Fig. 9 Same as Fig. 8 for the GPD site

Fig. 10 Same as Fig. 8 for the PLT site

123
Author's personal copy
Bull Earthquake Eng

Fig. 11 Same as Fig. 8 for the TEI site

this apparently high velocity estimation of the array technique results, is consistent with the
geologically and geotechnically observed subsurface structure.

5 Numerical simulation of ambient noise

The previously described array technique provided preliminary shear wave velocity profiles
for the shallow layers of the study area, in rather good agreement with the geological and
geotechnical information. These velocity models were used for synthetic (theoretical) ambi-
ent noise recordings production via numerical simulation, which were analyzed with the
HVSR method, similar to the experimental ones. The main idea behind this approach was
to compare the synthetic and experimental HVSR curves, in order to indirectly assess: a)
the validity of the velocity models as representative for the subsurface geophysical structure
of the study area and, b) their potential usage for additional numerical simulations (e.g. for
synthetic strong motion recordings).
To generate synthetic ambient noise waveforms, we must consider that the ground motion
recorded on a seismogram is a combination of the source, the earth structure through the
waves propagated, and the seismometer (Stein and Wysession 2003), hence a seismogram,
u (t) , can be written as the time series convolution of each factor:

u (t) = S(t) ∗ G (t) ∗ I(t) (2)

where S(t) is the source time function, G (t) represent the effects of the earth’s structure, and
I(t) describes the instrument response of the seismometer.

123
Bull Earthquake Eng

Table 2 Velocity profiles derived from the array noise data processing for every site

Array Superficial layers Halfspace

H1 H2

V p1 V s1 Thickness (m) V p2 V s2 Thickness (m) V p H al f space V s H al f space


(m/s) (m/s) (m/s) (m/s) (m/s) (m/s)

DEYAG 259 164 3.5 409 272 3.5 2,302 1,184


GPD 831 531 12 – – – 2,465 1,743
PLT 912 272 11 – – – 1,825 1,158
Author's personal copy

TEI 967 272 10 – – – 1,811 1,267

123
Author's personal copy
Bull Earthquake Eng

Fig. 12 A typical example of the subsurface structure lithology, water level, SPT data and Vs profiles deter-
mined from SPT data conversion and ambient noise array technique for the DEYAG site

Modeling of ambient noise sources was implemented with the code RANSOURCE (Moczo
and Kristek 2002), whereas the media response to ground motion calculated via Green func-
tions (Hisada 1994, 1995). The program generates random space-time point sources of seis-
mic noise inside a predefined source volume. The spatial distribution of potential point sources
is assumed to be regular and controlled by the prescribed minimum distance between two
neighboring point sources, minimum distance between a point source and a receiver and
maximum distance between a point source and a receiver. The temporal distribution of point
sources is controlled by the prescribed minimum and maximum numbers of point sources
acting at the same time. Each generated point source has several characteristics, such as
its spatial position, the direction and amplitude of an acting single body force and its time
function. The time function is usually either delta-like signal or pseudo-monochromatic with
duration and predominant frequency randomly generated, in an attempt to simulate typical
man-made and natural noise sources. It should be noted that the original Green functions
computation have been adapted for a viscoelastic layered halfspace with sources and receiv-
ers at close depths. The geophysical properties of the model are considered to vary only with
depth (1D simulation), whereas a plain free surface is considered. Green functions are com-
puted for every source-receiver couple, hence the final ground motion (synthetic recording)
at each receiver position is obtained from the summation of synthetic recording produced
from all sources.
Synthetic ambient noise recordings were generated for selected geotechnical borehole
positions, array sites, as well as for a generalized “average” structure model derived from

123
Author's personal copy
Bull Earthquake Eng

the existing geological, geotechnical and geophysical data, as later explained. For the geo-
technical borehole positions for which Vs information was not available, a local Vs pro-
file was constructed by converting SPT data with appropriate relations developed for soils
in the area of Greece (Athanasopoulos 1995). For those measurements where the pene-
tration depth was smaller than the one required by the SPT test, the NSPT number was
predicted by a linear extrapolation, as is often adopted in geotechnical studies. Figure 12
illustrates a typical example of this conversion for a selected geotechnical borehole of the
study area, where information from an adjacent noise array site was also available. It is
clear that this conversion procedure results in very similar velocity estimates for the shal-
lower Holocene formations (mainly clay) but gives much smaller velocities for the under-
lying Pliocene–Pleistocene formations, dominated by mixtures of gravel and sand, suggest-
ing that the adopted linear NSPT extrapolation approach does not work properly for this
formation.
Figure 13 illustrates the comparison of synthetic and experimental HVSR curves (black
and blue curves, respectively) produced for models derived from the geotechnical borehole
data. Surprisingly, the synthetic HVSR curves are practically flat for almost all cases, with
more or less constant amplitude close to one for the frequency range of interest. On the
contrary, the experimental HVSR curves exhibit significant local maxima, especially at high
frequencies, suggesting the presence of high impendence contrast at relative shallow depths.
The observed discrepancy between synthetic and experimental HVSR curves is most prob-
ably due to the use of “equivalent” models for the noise numerical simulations. As earlier
mentioned, these models cannot properly assess the real velocity contrast between the shal-
low Holocene and deeper Pliocene–Pleistocene formation. As a result these “equivalent”
models generated from SPT information are not representative of the subsurface geophys-
ical structure, underestimating the shear wave velocity at depth and resulting in almost flat
synthetic HVSR curves.
In Fig. 14 the synthetic HVSR curves (black color) are presented for the four array sites for
which a velocity model was derived from the array data. In the same plots the experimental
HVSR curves (multi-colored) are also presented for the recording stations of the inner and
middle array circles (including the central measurement). Contrary to Fig. 13, a very good
correlation of synthetic and experimental HVSR curves is observed for high frequency peaks
( f 0 > 3 Hz), especially for sites DEYAG and TEI, and to a lesser extend for site PLT. More-
over, the absence of a high frequency peak for site GPD is confirmed from both simulated and
experimental data. Furthermore, the high frequency fundamental resonance frequency ( f 1 )
is better reconstructed from the synthetic curves in comparison to the corresponding ampli-
fication factor (A1 ). This observed agreement suggests that the Vs profile obtained from the
array technique correctly depicts the depth and the impendence contrast of the surface Holo-
cene layer to the underlying Pliocene–Pleistocene formations and adequately describes the
subsurface geophysical structure.
In order to further examine and support the correlation of the higher resonant frequencies,
f 1 , with the presence of the relatively thin Holocene sedimentary layer, we examined the
variation of the observed resonant period, T1 , values from the HVSR curves against the Holo-
cene sediment thickness available from geotechnical boreholes next to noise measurement
sites. The results are presented in Fig. 15, where a general increase of the resonant period is
observed with the layer thickness, as expected, despite the scatter of the available data. In the
same plot the corresponding quarter-wavelength lines have been plotted for various Vs values.
It is interesting to notice that the determined Vs velocity range is in good agreement with the
corresponding values derived from the array noise measurements (see Table 2) suggesting
that the hypothesized correlation is realistic.

123
Author's personal copy
Bull Earthquake Eng

Fig. 13 Synthetic (black) and experimental (blue) HVSR curves determined from velocity models reduced
from the geotechnical borehole NSPT data. Borehole D3 is located nearby the DEYAG array site

123
Author's personal copy
Bull Earthquake Eng

Fig. 13 continued

123
Author's personal copy
Bull Earthquake Eng

Fig. 14 Synthetic (black) and experimental (multi-color) HVSR curves derived from the array models. Every
experimental HVSR curve corresponds to a recording station of the inner and middle circle of the array
geometry, including the central one

All previous results can be used for a preliminary microzonation attempt of the study
area, especially regarding the shallower Holocene formation which is depicted by the higher
f 1 HVSR values in the study area. This zonation is schematically presented in Fig. 16,
superimposed on the f 1 spatial distribution (Fig. 5b). A zone of relative lower f 1 resonant
frequencies, in the range of 3–5 Hz, is observed in the city center in the northern part of
the town, showing the presence of thicker Holocene clay-dominant soft formations, with an
indicative typical thickness of 8–10 m (zone A). To the south, the situation changes drastically
(zone C), with much higher f 1 resonant frequencies (>10 Hz), corresponding to an upper
thin lower velocity layer (typically < 4 m). This thin layer may locally become insignificant,
a pattern compatible with the presence of a large number of HVSR curves with no higher
resonant HVSR peak (red circles) in this zone. The remaining western and eastern parts of
the city, as well as the west section of the city center (zone B) exhibit a gradual increase of
f 1 frequencies as we move away from zone A, with intermediate values (5–10 Hz, mostly
around 7–8 Hz), corresponding to intermediate shallow layer thicknesses (typically 6–10 m).

123
Author's personal copy
Bull Earthquake Eng

Fig. 15 Variation of the


observed HVSR resonant period,
T1 , against the corresponding
Holocene sedimentary layer
thickness, determined from
geotechnical boreholes
neighboring to noise
measurement sites. The plotted
lines correspond to the simple
quarter-wavelength fundamental
period variation for various
average Vs velocities of the
Holocene layer. Notice that the
proposed velocity range is in
good agreement with the array
results presented in Table 2 for
top-soil formations

A rather sharp transition zone (depicted with a white dashed line in Fig. 16) exists between
zone A and part of zone B with the higher frequency zone C, crossing the city center in an
almost E–W direction.
An important observation is that in three out of four array sites presented in Fig. 14
(except site PLT), the experimental HVSR curves exhibit a second broad and pronounced
low frequency peak around 1 Hz, which is not observed at the synthetic ones. Regarding the
single low frequency peak observed at station 05 of PLT array, we consider it as a probably
artificial effect (i.e. bad coupling between sensor and ground), since it was identified at only
one station from the whole array and it is not systematic. It should be noted that the synthetic
HVSR curves were generated using a model that only employed the shallow velocity con-
trast between the Holocene and underlying Pliocene–Pleistocene formations. One possible
explanation is that the observed low frequency peak is associated with the deep Oligocene–
Miocene Mesohellenic Graben molassic formations (effective bedrock) that are found in the
broader study area and continue beneath the sedimentary valley of Grevena town. Since the
expected bedrock depth is estimated to exceed 100 m, it was not included and could not be
recovered by the array noise modeling, as the expected penetration depth was much less than
85 m in all cases.
The previously proposed explanation should be considered with some caution. In general,
the low frequency peaks display some spatial inconsistency, contrary to the high frequency
peaks which show better spatial coherency. This observation, together with the low frequency
variability at array sites (Fig. 14), suggests a possible lack of 1D geometry underneath the
array sites. Moreover, it indicates that these low frequency peaks could also be explained by
considering e.g. lateral variations into the superficial layers. Furthermore, the low frequency
values presented in Fig. 14 are ranging between 0.7 and 1.5 Hz and compatible but slightly
smaller from those presented in Fig. 5a for the single station measurements, where the low
frequency values are not less than 1 Hz. However, if we consider the average low frequency
value derived from the inner and the middle circle stations of the array, a good correspondence
with neighboring single station values regarding both f 0 – f 1 and A0 –A1 values is observed,

123
Author's personal copy
Bull Earthquake Eng

Fig. 16 Preliminary microzonation of the Grevena town, on the basis of the higher, f 1 , resonant frequency, as
revealed by the HVSR analysis of ambient noise measurements and the available array and geotechnical data
in the study area. Zone A corresponds to areas with f 1 values in the range of 3–5 Hz with typical thickness
of the upper sedimentary Holocene layer 8–10 m, Zone B exhibits intermediate values of f 1 (5–10 Hz) and
layer thickness (6–10 m), whereas Zone C displays higher f 1 values (>10 Hz) related with very thin surface
Holocene layer (<4 m). The white dashed line depicts a rather sharp transition zone between Zones A–B and
Zone C. Geological formations are in accordance with Fig. 1

as depicted in Figs. 5 and 6. This is further illustrated in Fig. 17 for a typical array (TEI site),
where the f 0 and f 1 values are displayed for every station of the array as well as for nearby
single station measurement sites. Though the pattern presented in Fig. 14 does not clarify
the source of the low frequency peak, it suggests that also exhibits a spatial coherency, hence
its relation with the local velocity structure (similar to the high frequency peak) should be
considered as possible.
In order to overcome the limitation of the noise array models and explore the possibility
that both HVSR peaks are related to the local velocity structure, as already shown for the
high frequency peak, f 1 , a generalized “average” geophysical model was tested by including
a halfspace at large depths (Oligocene–Miocene bedrock) with a relative high impendence
contrast in comparison to the overlying Pliocene–Pleistocene sands. A large number of syn-
thetic HVSR curves were produced by a detailed parametric analysis of the ground model,
allowing the thickness of the Pliocene–Pleistocene sedimentary layer to vary between 100
and 200 m, whereas the Vs velocity of the halfspace was also allowed to vary between 2,000
and 3,500 m/s.
Figure 18 illustrates the synthetic HVSR curve (black color) for a typical model that
adequately describes the main locally observed patterns of the available noise data (double
resonant frequencies) and its geophysical properties are listed in Table 3. Two local maxima
are observed for this typical model, one at 2 Hz and another at 7.5 Hz, which are associated
for the model of Table 3 with the sediment/bedrock interface and the contact of the upper

123
Author's personal copy
Bull Earthquake Eng

Fig. 17 Low (a) and high (b) resonant frequency values, as revealed from HVSR analysis on ambient noise
recordings of TEI array stations (green dots). The single station HVSR values are also plotted (red dots),
showing a rather good spatial coherency, e.g. verifying the gradual decrease of f 1 values from ∼8 Hz in the
W–NW to ∼5.5–6 Hz in the E–SE part of the area

123
Author's personal copy
Bull Earthquake Eng

Fig. 18 Synthetic (black) and experimental (multi-color) HVSR curves derived from the optimal generalized
“average” 1D geophysical model. The line numbers correspond to the measurement site ID presented in Fig.
4. In a, b some typical examples of experimental HVSR curves that exhibit a rather good correlation of both
peaks with the synthetic one are presented, whereas c, d display examples of good correlation regarding only
the low and only the high frequency peak, respectively

sedimentary layers, respectively. In Fig. 18a, b we present some typical examples of experi-
mental HVSR curves that exhibit a rather good correlation of both peaks. On the other hand,
Fig. 18c illustrates examples of good correlation only regarding the low frequency peak,
whereas Fig. 18d regarding only the second high frequency peak. It should be noted that the
experimental HVSR curves shown in Fig. 18 do not follow a geographical criterion, but a
“pattern-like” criterion with measurements coming from a variety of places.
Nevertheless, it should be emphasized that although the synthetic HVSR curve derived
from the generalized “average” geophysical model displays a good “pattern” correlation
with most of the single station measurements, it is not in quantitative agreement with several
experimental HVSR curves (Fig. 14), especially in the low frequency range. The main reason
for this discrepancy is that the average geophysical model used for numerical simulations
of ambient noise recordings is an indicative 1D model, hence it is not able to describe the

123
Author's personal copy
Bull Earthquake Eng

Table 3 Geophysical parameters of the generalized “average” 1D geophysical model used for the numerical
simulation of ambient noise recordings in the study area

Layers Thickness (m) Density V p (m/s) Vs (m/s) Quality Quality


(gr/m3 ) factor factor
Qp Qs

H1 10 1,900 385 250 30 10


H2 150 2,000 1,845 1,200 90 30
Halfspace – 2,500 5,385 3,500 300 100

complexity and details of a 3D structure like the one in the study area, especially of the
strongly varying Holocene formation which significantly affects the f 1 HVSR frequency.
Therefore, the purpose of this 1D model is not to propose a single model for the study area
that can interpret the HVSR curves for all sites but simply to indicate the possibility to obtain
two independent HVSR peaks, controlled by the two main sedimentary model discontinu-
ities of the subsurface structure. Thus, this is more a suggestion of a general model feature,
rather than a strict model determination and does not provide accurate quantitative estimates
for each recording site. Given the available geological and geotechnical information it is
not possible to perform reliable individual calibrations for each site and additional effort is
needed in order to clearly define a refined 3D deep geophysical model for the study area,
including mainly additional geotechnical and geophysical surveys.
The previously presented simulation suggests that although we cannot constrain the details
of the deepest part of the geophysical model, a possible explanation for the low frequency
peak is the presence of a deep, high contrast bedrock formation, in agreement with the
available geological information. The fact that the corresponding low frequency peak is not
always observed can be due to the local variations of the sediments/bedrock interface (faults,
transition zones, layer mixing, large slope interface, etc.) and the corresponding effects on
the impendence contrast. Similarly, as earlier discussed, cases where the corresponding high
frequency peak is not observed on the HVSR curves probably correspond to sites for which
a sufficient impendence contrast does not exist but a rather gradual migration from the softer
clayish shallow Holocene formations to the underlying Pliocene–Pleistocene sands is found,
as some geotechnical boreholes also indicate (e.g. Fig. 3).

6 Discussion and conclusions

A detailed study of the geophysical structure and evaluation of site effects has been performed
for the town of Grevena (North–West Greece) using ambient noise recordings and 1D numer-
ical simulation modeling. Analysis of ambient noise measurements was implemented using
the HVSR method and the array technique.
The HVSR method was able to depict the resonant frequency of the upper sedimentary
layers on a quite uniform grid of 60 measurement sites inside the study area. The results
showed the presence of two groups of resonance peaks, one at relatively low (typically 1–2
Hz) and one at high (typically 3–20 Hz) frequencies, often with both groups identified on the
HVSR curves. An attempt was made to define zones with similar response behavior and to
associate local maxima of the HVSR curves with specific geological layers. In this context,
the observed flat HVSR curves were considered to probably indicate areas of low impen-
dence contrast at depth. Low frequency peaks (∼1 Hz) on HVSR curves exhibit medium

123
Author's personal copy
Bull Earthquake Eng

(2–4) and often relatively high amplitudes (A0 > 4), possibly controlled by the large depth
Oligocene–Miocene Mesohellenic Graben molassic formations (effective bedrock). On the
other hand, high frequency peaks (∼3–20 Hz) correspond to relatively low HVSR values
(typically A0 ∼ 3) and are clearly associated with the presence of the upper Holocene sedi-
mentary layer (mainly clays), whenever this formation exhibits a significant contrast to the
underlying Pliocene–Pleistocene formations (mainly sands).
In order to validate the proposed explanation, 1D geophysical models were obtained with
the array technique application for four selected sites inside the Grevena town. In all cases,
the shallow Holocene layer exhibited Vs velocities ranging from 150 to 530 m/s, overlying
a relative high Vs (1,100–1,800 m/s) halfspace, corresponding to the Pliocene–Pleistocene
sediments.
The geophysical structure results were further validated using numerical simulation of
ambient noise and HVSR results obtained from the synthetic noise recordings. Three groups
of models were used, derived from selected geotechnical boreholes by converting SPT data to
Vs profiles, Vs profiles from the array technique results and a generalized “average” geophys-
ical model, determined on the basis of array models and the geological information. Ground
models derived from geotechnical borehole converted NSPT data produced almost flat syn-
thetic HVSR curves, suggesting that they are not adequate, at least for the deeper formations
of the study area, to realistically approximate the real subsurface geophysical structure. On
the other hand, the array technique models exhibited a much better performance, in good
agreement with the geology of the study area showing a very good correlation between
synthetic and experimental HVSR curves for high frequencies (∼5–10 Hz).
Using this approach it was not possible to reconstruct the low frequency peaks (∼1–2
Hz) on the synthetic HVSR curves. Though the corresponding low frequency peaks show
a coherent but locally unstable spatial distribution, a possible explanation is that the deep
effective bedrock formation (Oligocene–Miocene formations at depths greater than 100 m)
was not included in the array velocity model. Assuming that the low frequency peak is
associated with this deep sediment/bedrock interface, allowed the successful construction
of a noise data compatible generalized geophysical model containing two layers (shallow
Holocene and deeper Pliocene–Pleistocene formations) and an underlying halfspace at depth
(Oligocene–Miocene formations). Synthetic HVSR curves exhibit two clear local maxima,
one at low and a second at high frequencies which can be correlated with deep and shallow
sedimentary formations, respectively. Nevertheless, this model predicts mainly qualitatively
and not in a strict quantitative manner the patterns of the majority of experimental HVSR
curves at single station and array sites, indicating that it is rather weak in the low frequency
range and cannot adequately describe the 3D geophysical model properties. Therefore, it
provides more of a suggestion that the observed low frequency HVSR peak can be controlled
by the deep geological formations and does not reflect the complex 3D subsurface structure
of the study area.
The approach presented in the present work involved a combined processing of single
station and array noise measurements and their numerical simulation, which allowed con-
straining several features of the subsurface velocity model for the Grevena town. The main
target was not only to construct local models (especially of the upper Holocene formations)
but also to validate those using experimental observations from ambient noise analysis, incor-
porating constrains that are consistent with the available geological and geotechnical data.
Future work for model improvement could incorporate a more detailed mapping of the layers
spatial distribution by application of additional geotechnical and geophysical surveys, as well
as 2D/3D numerical simulation modeling of noise and seismic motion.

123
Author's personal copy
Bull Earthquake Eng

Acknowledgments This work has been performed under the “SyNaRMa” project (Information System for
Natural Risk Management in the Mediterranean), funded by the EU Community Initiative Programme INTER-
REG III B SRCHIMED (2000-2006) and co-financed by the European Regional Development (Contract No:
SC_A.1.012). We are thankful to P. Dimitriou†, E. Zacharopoulos and A. Bantis for participating to the field
measurements and to V. Mitropoulos and I. Simou for providing useful geotechnical data.

References

Aki K (1957) Space and time spectra of stationary stochastic waves, with special reference to microtremors.
Bull Earthq Res Inst 35:415–456
Aki K, Richards GP (2002) Quantitative seismology, 2nd edn. University Science Press, USA
Asten WM (1978) Geological control on the three-component spectra of Rayleigh-wave microseism. Bull
Seismol Soc Am 68(6):1623–1636
Asten WM, Henstridge DJ (1984) Array estimators and use of microseisms for reconnaissance of sedimentary
basins. Geophys 49:1828–1837
Athanasopoulos AG (1995) Empirical correlations Empirical correlations Vso − NSPT for soils of Greece:
a comparative study of reliability. In: Cakmak A, Brebbia C (eds), Soil dynamics and earthquake engi-
neering VII, Computational Mechanics Publications, pp 19–26
Bard P-Y (1999) Microtremor measurements: a tool for site effect estimation?. In: Irikura K, Kudo K, Okada
H, Sasatani T (eds) The effects of surface geology on seismic motion. Balkema, Rotterdam pp 1251–1279
Bard P-Y, SESAME Participants (2004) The SESAME project: an overview and main results. In: Proceedings
of the 13th world conference on earthquake engineering, Vancouver, 1–6 Aug 2004, Paper no. 2207
Bindi D, Parolai S, Spallarossa D, Catteneo M (2000) Site effects by H/V ratio: comparison of two different
procedures. J Earthq Eng 4:97–113
Borcherdt DR (1970) Effects of local geology on ground motion near San Francisco bay. Bull Seismol Soc
Am 60(1):29–61
Bour M, Fouissac D, Dominique P, Martin C (1998) On the use of microtremor recordings in seismic microz-
onation. Soil Dyn Earthq Eng 17:465–474
Capon J (1969) High resolution frequency-wavenumber spectrum analysis. In: Proceedings of the IEEE, vol
57, pp 1408–1418
Chatelain J-L, Gueguen P, Guillier B, Frechet J, Bondoux F, Sarrault J, Suplice P, Neuville J-M (2000) City-
Shark: a user-friendly instrument dedicated to ambient noise (microtremor) recording for site and building
response studies. Seismol Res Lett 71:698–703
Chávez-García FJ, Pedoti G, Hatzfeld D, Bard P-Y (1990) An experimental study near Thessaloniki (Northern
Greece). Bull Seismol Soc Am 80(4):784–800
Duval A-M, Bard P-Y, Meneroud J-P, Vidal S (1994) Mapping site effects with microtremors. In: Proceedings
of 5th international conference on seismic zonation, vol 2. Nice, pp 1522–1529, 17–19 Oct 1994
Field HE, Jacob HK (1995) A comparison and test of various site-response estimation techniques, including
three that are not reference-site dependent. Bull Seismol Soc Am 85(4):1127–1143
Fyen J (1990) Diurnal and seasonal variations in the microseismic noise level observed at the NORESS array.
Phys Earth Planet Inter 63:252–268
Haghshenas E, Bard P-Y, Theodulidis NSESAME WP04 Team (2008) Empirical evaluation of microtremors
H/V spectral ratio. Bull Earthq Eng 6:75–108
Hisada Y (1994) An efficient method for computing Green’s functions for a layered half-space with sources
and receivers at close depths. Bull Seismol Soc Am 84(5):1456–1472
Hisada Y (1995) An efficient method for computing Green’s functions for a layered half-space with sources
and receivers at close depths (part 2). Bull Seismol Soc Am 85(4):1080–1093
Horike M (1985) Inversion of phase velocity of long-period microtremors to the S-Wave velocity structure
down to the basement in urbanized areas. J Phys Earth 33:59–96
Konno K, Ohmachi T (1998) Ground motion characteristics estimated from spectral ratio between horizontal
and vertical components of microtremor. Bull Seismol Soc Am 88(1):228–241
Kudo K (1995) Practical estimates of site response: State of art report. In: Society of 5th international confer-
ence on seismic zonation, pp 1878–1907
Kvaerna T, Ringdahl F (1986) Stability of various f-k estimation techniques. In: Semiannual technical sum-
mary, NORSAR Scientific report, 1-86/87, Kjeller, 1 Oct 1985–31 March 1986, pp 29–40
Lacoss RT, Kelly EJ, Toksöz MN (1969) Estimation of seismic noise structure using arrays. Geophys 34:21–38
Langston CA (1979) Structure under Mount Rainier, Washington, inferred from teleseismic body waves.
J Geophys Res 84:4749–4762

123
Author's personal copy
Bull Earthquake Eng

Lermo J, Chávez-García JF (1993) Site effect evaluation using spectral ratios with only one station. Bull
Seismol Soc Am 83(5):1574–1594
Moczo P, Kristek J (2002) FD code to generate noise synthetic. Deliverable D02.09 of the SESAME project,
p 31
Mountrakis D (1985) Geology of Greece. University Studio Press, Thessaloniki
Murphy RJ, Shah KH (1988) An analysis of the effects of site geology on the characteristics of near-field
Rayleigh waves. Bull Seismol Soc Am 78(1):64–82
Nakamura Y (1989) A method for dynamic characteristics estimation of subsurface using microtremor on the
ground surface. QR Railw Tech Res Inst 30:25–33
Nakamura Y (1996) Real-time information systems for hazards mitigation. In: Proceedings of the 11th world
conference on earthquake engineering, Acapulco, 23–28 June 1986, Paper no. 2134
Nakamura Y (2000) Clear identification of fundamental idea of Nakamura’s technique and its applications.
In: Proceedings of the 12th world conference on earthquake engineering, Auckland, 30 Jan–4 Feb, Paper
No. 2656
Nogoshi M, Igarashi T (1971) On the amplitude characteristics of microtremor (Part 2) (in Japanese with
english abstract). J Seismol Soc of Japan 24:24–40
Ohrnberger M, Schissele E, Cornou C, Bonnefoy-Claudet S, Wathelet M, Savvaidis A, Scherbaum F, Jongmans
D (2004) Frequency wavenumber and spatial autocorrelation methods for dispersion curve determina-
tion from ambient vibration recordings. In: Proceedings of the 13th world conference on earthquake
engineering, Vancouver, 1–6 Aug 2004, Paper no. 0946
Okada H (2003) The microtremor survey method. Society of Exploration Geophysics (SEG)
Papazachos CB (1990) Seismicity of the Aegean and surrounding area. Tectonophys 178:287–308
Riepl J, Bard P-Y, Hatzfeld D, Papaioannou C, Nechtschein S (1998) Detailed evaluation of site-response
estimation methods across and along the sedimentary valley of Volvi (EURO-SEISTEST). Bull Seismol
Soc Am 88(2):488–502
Sambridge M (1999) Geophysical inversion with a neighborhood algorithm I. Searching a parameter space.
Geophys J Int 103:4839–4878
Satoh T, Kawase H, Matsushima SI (2001) Differences between site characteristics obtained from microtre-
mors, S-Waves, P-Waves, and Codas. Bull Seismol Soc Am 91(2):313–334
Scherbaum F, Hinzen K-G, Ohrnberger M (2003) Determination of shallow shear wave velocity profiles in
the Cologne/Germany area using ambient vibrations. Geophys J Int 152:597–612
SESAME Project (2004) Deliverable D23.12, Guidelines for the implementation of the H/V spectral ratio
technique on ambient vibrations: measurements, processing and interpretation, available from the web
site: http://sesame-fp5.obs.ujf-grenoble.fr/Papers/HV_User_Guidelines.pdf
Steidl HJ, Tumarkin GA, Archuleta JR (1996) What is a reference site?. Bull Seismol Soc Am 86(6):1733–
1748
Stein S, Wysession M (2003) An introduction to seismology, earthquakes and earth structure. Blackwell Pub-
lishing, USA
Tokimatsu K (1997) Geotechnical site characterization using surface waves. In: Earthquake geotechnical
engineering. Balkema, Rotterdam, pp 1333–1368
Wathelet M, Jongmans D, Ohrnberger M (2004) Surface-wave inversion using a direct search algorithm and
its application to ambient vibration measurements. Near Surf Geophys 2:211–221
Wathelet M (2005) Array recordings of ambient vibrations: surface-wave inversion. PhD thesis, University of
Liège, Belgium
Wathelet M, Jongmans D, Ohrnberger M, Bonnefoy-Claudet S (2007) Array performances for ambient vibra-
tions on a shallow structure and consequences over Vs inversion. J Seismol 12:1–19
Woods WJ, Lintz RP (1973) Plane waves at small arrays. Geophys 38:1023–1104

123

You might also like