Next Article in Journal
Developments in 3D Visualisation of the Rail Tunnel Subsurface for Inspection and Monitoring
Previous Article in Journal
Virtual and Augmented Reality as a Novel Opportunity to Unleash the Power of Radiotherapy in the Digital Era: A Scoping Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Local Site Effects Investigation in Durres City (Albania) Using Ambient Noise, after the 26 November 2019 (M6.4) Destructive Earthquake

1
Institute of Engineering Seismology and Earthquake Engineering, 55535 Thessaloniki, Greece
2
Department of Seismology, Institute of Geosciences, Polytechnic University Tirana, 1000 Tirana, Albania
3
Department of Geophysics, Aristotle University Thessaloniki, 54124 Thessaloniki, Greece
4
GeoSeis-IT Consulting, London N14 6JF, UK
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(22), 11309; https://doi.org/10.3390/app122211309
Submission received: 31 August 2022 / Revised: 3 November 2022 / Accepted: 4 November 2022 / Published: 8 November 2022
(This article belongs to the Section Earth Sciences)

Abstract

:
Site characterization of metropolitan areas, especially after an earthquake, is of paramount importance for interpretation of spatial damage distribution and taking measures that assure realistic design actions to strengthen existing constructions and create new ones. Such a case is the city of Durres, Albania, that was hit by the disastrous earthquake of 26 November 2019 (M6.4). Significant differences in structural damage were observed throughout the city, despite its uniform epicentral distance (approximately 15 km); this could be either due to varying vulnerability of the affected constructions and/or to spatial variation of strong ground motion in the city, resulting from local site effects; the latter factor was investigated in this study. This was achieved by taking single station ambient noise measurements throughout the city, at approximately 80 sites. Ambient noise measurements are favorable, as acquiring ambient noise data is an easy and effective noninvasive approach within urban environments. Measurements were processed using the widely applied Horizontal-to-Vertical Spectral Ratio (HVSR) method, following the SESAME project (2004) guidelines. Their fundamental and dominant frequencies, fo and fd, respectively, were calculated and related to the iso-depth contours of the investigated area, as well as their corresponding amplitudes, Ao, and Ad. These experimental parameters and the HVSR curves were used to group all examined sites into classes with similar properties. This clustering provided a zonation map with four categories consisting of similar shapes and amplitudes, applicable to the city of Durres. This map can be utilized as a first level zonation of local site effects for the city. In addition, dynamic properties of soil profiles in selected sites were investigated and tested using 1D synthetic ambient noise data, based on the Hisada (1994, 1995) simulation method, and compared to experimental HVSRs in proximity to the selected sites. A comparison of the proposed four categories zonation map to the observed damage of the 26 November 2019, mainshock is attempted and evaluated. The four categories zonation map with similar expected local site effects proposed in this study can be used as a first level seismic microzonation of Durres. Undoubtedly, corrections for 2D/3D effects on ground shaking must be applied to sites lying in the edges of the Durres basin.

1. Introduction

After the disastrous earthquake of Durres, Albania, on 26 November 2019 (M6.4) characterizing the city in terms of soil properties related to seismic site response was deemed to be of paramount importance. Despite the recent methodological and technological developments, determining the seismic ground motion response of various geologic formations with traditional methods still proves to be a difficult and costly task [1,2]. Large-scale urban areas that need to be densely covered, or areas with moderate to low seismicity are especially difficult, due to the time period required to compile a statistically significant number of seismic records in order to produce results using traditional methods (e.g., Standard Spectral Ratio; [3]).
In the last three decades, ambient noise recordings have been extensively employed and have proven to be a useful tool in determining local site effects. Ambient noise is defined as the constant vibration of the Earth’s surface, generated either by human activities (e.g., machinery, traffic, pedestrians) in high frequency ranges (f > ~1 Hz) or by natural phenomena (e.g., wind, variations in atmospheric pressure, sea waves) in low frequency ranges (f < ~1 Hz). It is often assumed that ambient noise recordings are dominated by Rayleigh and Love waves at more than one wavelength from the source [4].
The most common technique, routinely applied today, relies on the calculation of the Horizontal-to-Vertical Spectral Ratio (HVSR) of ambient seismic noise, typically recorded at a location by a three-component velocity sensor. According to [5], this single-station ambient noise approach was initially developed in Japan by [6], based on the work by [7] for characterizing site response under seismic loading, and was later popularized and disseminated by [8,9]. In the early 2000s, the application and further development of the method exhibited mobility and broader use in Europe (e.g., [10,11]) and later in Canada ([12,13]) and South America [14]. In general, HVSR analysis is currently applied worldwide to estimate the site fundamental frequency, and in certain circumstances site amplification factors, as it is observed in weak ground motion of earthquake excitation (linear soil behavior), though this factor is still debatable among the community (among others; [1,15]).
In fact, the appropriate density of site characterization in the built environments proposed in this study is achievable by easy to perform, low cost, non-invasive geophysical methods. Such methods are based on ambient noise from both anthropogenic and natural sources. A series of investigations have been carried out over the past three decades worldwide, aimed at deriving quantitative information on site amplification through non-invasive techniques, based principally on surface wave interpretations of ambient noise measurements [2].
In urban areas affected by strong earthquakes, microzonation studies are usually performed. As a first step, ambient noise measurements are applied, and using the HVSR method. some dynamic properties of surficial geologic layers are estimated, while comparison with earthquake damage distribution is also attempted (among others [16,17,18,19]).
In summary, HVSR can provide useful information on surface geologic layers, such as their fundamental frequency (fo) and its corresponding amplitude (Ao). If used in joint inversion with dispersion curves (DC), it can constrain the shear wave velocity profile of the site under investigation. It is a low-cost and easy to apply method, constituting a first basic step in microzonation study.
After the disastrous earthquake of 26 November 2019, investigating the underlying geologic formation’s role in seismic response was deemed necessary. Basic soil dynamic properties of surficial geologic strata (e.g., fundamental frequency, amplification, shear wave velocity) must be estimated in order to be implemented in respective seismic response analyses. For this purpose, in the present work, ambient noise measurements at 80 sites were performed in the urban area of Durres, and the HVSR method was applied to estimate their fundamental and dominant frequencies, as well as their corresponding amplitudes. In addition, spatial distribution of these parameters in the city is presented and compared alongside other available geological or/and geophysical characteristics of Durres complex 3D basin. The experimental HVSRs for selected sites are compared with HVSRs from synthetic ambient noise data based on available 1D profiles, thus assessing their reliability. In addition, classification of all experimental HVSRs using their entire curve is attempted, and four discrete classes are proposed for the city of Durres. The correlation of the proposed four zones/classes with the observed damage is discussed and these zones are proposed as a satisfactory basis for any future microzonation and/or seismic risk analyses of the Durres built environment.

2. Data and Methods Used

2.1. Geographic, Demographic and Territorial Features of the Study Area

The study area predominantly covers the urban area of Durres. The city lies along the Adriatic coast of western Albania and is one of the oldest cities in the country (known as “Dyrrachium”), with a history of over 2500 years [20].
From a physical-geographical standpoint, Durres is situated in the western Lowlands, on a flat alluvial plain between the Erzeni and the Ishmi rivers, and the surrounding hills. Its highest peaks are found in the Rodon–Erzeni Hills (at 272 m) and in the Durresi Hills (at 178 m). Most of the territory is a plain shaped by the reclaim of the Durresi swamp in the last century. The area of Durres covers only 1% (332 km2) of the Albanian territory, defined by a north-south longitudinal extension of about 36 km and a width of 19 km [21]. Durres is one of Albania’s most significant urban and economic centers, representing one of the most important nodes of national and international transport, crossed by many important national and international road axes and maritime links. As the second largest metropolitan area in Albania, it has a population of approximately 202,000 [22].
According to the Institute of Statistics (INSTAT) census, up to 2011, the highest percentage of constructions in the city of Durres consisted of 1–2 story buildings [20]. It is noted that during the period of 1991–2011, the number of buildings with over 6 floors has increased by 79%, and currently constitute about 3.5% of the total number of constructions [20]. Building typologies in Durres comprise of masonry buildings, predominantly residential structures up to 1993, consisting of unreinforced load bearing masonry structures up to 5 stories, and confined masonry structures of 3–6 stories; prefabricated concrete panel buildings built between 1970–1990, of 4 and 6 stories; and concrete-frame buildings with in-fill masonry walls. The development of reinforced-concrete frame systems and infill masonry walls began after 1993, rising from 2–3 stories up to more than 10 stories [23].

2.2. Geological Setting and Seismotectonic Aspect

From a geological point of view, the Durres Bay, and the weathered hills of Tortonian age, lay on poor, thick and unconsolidated Quaternary sediments [24,25,26,27,28]. The corresponding surface geology (Figure 1a) show that the western part of the hilly chain is composed of Miocene molassic formations comprising clay, siltstone, sandstone, and gypsum, while on the central and eastern parts a composition of Pliocene sandstones and conglomerates is found [29]. The Durres basin is filled with Pliocene and Quaternary deposits. The Pliocene deposits constitute the basement of the overlying Quaternary ones, including sandstones and conglomerates found in the central and eastern flanks of the Durres hills. In the Durres basin, the Quaternary deposits include alluvial, lagoon, marshy and marine deposits, reaching an overall thickness of nearly 130 m [26,27,28,30].
From a tectonics point of view, Durres and its surroundings belong to the Peri-Adriatic Foredeep. It constitutes a northern segment along the southern convergent margin of the Eurasia Plate, where the Albanian orogenic front is thrusting over the Adria microplate, partly over the Apulian platform and the southern Adriatic basin [25,26,31]. Consequently, it is strongly influenced by the Adria-Eurasia continental collision [27,28,31,32,33]. As a result, the Durres region is affected by a compressional stress regime, acting along a series of thrust and back-thrust faults (Figure 1b). The back-thrust faults have been active since the Pliocene-Quaternary period, namely the Durres and Preza ones, while thrusts along the Durres and Tirana syncline are traced to earlier periods [27,31,33,34].
Active faults of Durres and the surrounding regions (Figure 1b) have contributed to both historical and present seismic activity [25,26,31,34,35]. This notion is reinforced by the latest strong earthquakes that struck Durres, namely the September 21 (M5.8) and November 26 (M6.4), 2019. Both events showed the same type of reverse-faulting kinematics (Figure 2), based on their focal mechanism (FM) solutions, supporting the idea of a blind reverse causative fault, NW–SE striking (140°–155°) and ENE dipping (65°–72°), rooted at the basal thrust front [32,36,37,38,39]. The results also agree with the aforementioned regional tectonics (Figure 1b), showing significant consistency between observed data and the Durres thrust fault, whose ramp dip-angle tends to decrease at the hypocenter depth of these strong earthquake [32,37,38].

2.3. Historical and Recent Seismicity

Durres and the surrounding region were hit by strong earthquakes (M > 6.0), in the past [40]. The ancient city of Durres (Dyrrachium) has been nearly destroyed by several strong historical earthquakes in 58 BC, 334 AD, 346 AD, 506 AD, 521, 1273 and 1870 (see Table 1). The historical abandonment of Durres was a consequence of a number of devastating earthquakes with epicentral intensity IMAX = VIII–IX degree (MSK-64), in 334, 346 and 506 AD (Figure 3), [40]; in 1273, with a magnitude M6.7 ([41,42]) and in 1926, with a magnitude M6.3, [40,41,42,43].
The most significant seismic events in the region (Table 1 and Figure 3), that have also impacted Durres, are the 1617 Kruja earthquake (M6.2), with an epicentral Intensity IMAX = VIII (MSK-64); the 26 August 1852 Rodoni Cape earthquake (M6.2), with an epicentral Intensity IMAX = VIII (MSK-64); the May 16, 1860 Ndroqi (Beshiri Bridge) earthquake (M6.2), with epicentral Intensity IMAX = VIII (MSK-64); the 4 February 1934 Ndroqi earthquake (M5.7); the 19 August 1970 Vrapi earthquake (M5.6), with an epicentral Intensity IMAX = VII degree (MSK-64); the 16 September 1975 Rodoni Cape earthquake (M5.4); the 15 April 1979 (M6.9) Montenegro coastal area earthquake; and the 9 January 1988 Tirana earthquake (M5.9), with an epicenter Intensity IMAX = VII-VIII degree (MSK-64). The most recent strong earthquakes, having a significant socio-economic impact for Durres and its surrounding regions, are the 21 September 2019 (M5.6), with an epicentral Intensity IMAX = VII-VIII degree (MSK-64) and the earthquake of 26 November 2019 (M6.4), with epicenter Intensity IMAX = VIII-IX degree (MSK-64), ([40,41,42,45]). (Figure 3 and Table 1).
The 16 November 2019 (M6.4) mainshock was located 16 km north of Durres city, with its epicenter on the Adriatic coast of Albania (Figure 2), about 33 km northwest of Tirana, the capital of Albania ([32,37,38,39]) It was subsequently followed by intense aftershock activity. On 21 September 2019, the same area was hit by an earthquake of moderate magnitude (M5.6), 12 km north of Durres. A posteriori it is believed to be the main foreshock of the November 26, 2019, earthquake. An absence of seismic activity until 25 November was observed. Immediate foreshock activity started on 25 November 2019, at 03:22 am (UTC), including eight earthquakes with magnitudes between 2.0–4.6 (ML). The mainshock, followed nearly one hour after the 01:47 am (UTC) foreshock (ML4.6), located approximately 4 km north of its hypocenter [45]. The seismic sequence of 26 November 2019 is apparently characterized by a high number of aftershocks [45,46]. Until 30 June 2020, approximately 2000 earthquakes occurred (Figure 2), located by the National Seismic Network of Albania (ASN).
The effects of the ground shaking of the 26 November earthquake in the city of Durres were very significant in terms of building damage. Moderately to heavily damaged buildings were observed in two broad zones. One including its central-west part, elongated in a north-south direction, and a second in its coastal area in south-to-south-eastern direction ([23,47], as well as personal communication with local structural engineers and personal autopsy during the experiment, Figure 4a).
The unique accelerometer station (DURR) was installed within the most highly affected part of the city. It showed a horizontal PGA~0.2 g and a broad response acceleration spectrum [48]. More specifically, the ground shaking at the DURR accelerometer station exhibited high spectral acceleration values (>0.3 g) for a wide frequency range (f = 1/Period), 0.8 Hz < f < 5 Hz (Figure 4b). That is, amplification of strong ground motion was intense in this frequency range.
Influence of site effects on seismic hazard and on seismic vulnerability of buildings has been documented in many cases worldwide (among others: [49,50,51,52]). Seismic ground motion is amplified according to the geophysical and geotechnical properties of the site under investigation, highlighting the need to know these properties.

2.4. Geotechnical and Geophysical Data in the Durres Region

The mainshock struck a modern urban environment, the metropolitan area of Durres, and caused extended damage downtown as well as in several villages in the epicentral area. The recent layers of surface geology (Figure 1, Figure 4 and Figure 5) are related to the highest expected seismic intensities and/or PGA values [34,53,54,55], as well as with problems concerning soil instabilities, such as induced liquefaction on poor sandy deposits of the Durres lagoon and the Durres Bay area [30]. The Quaternary deposits in the Durres basin include alluvial, lagoon, marshy and marine deposits with a total thickness of around 130 m [30].
The marshy Holocene deposits, comprised of clay, clayey loam, sand, and organic material, have played an important role in forming strong ground motion and distribution of the earthquake-induced building damage in Durres [56]. A large part of the city is founded on these marshy deposits, which reach a maximum thickness at the central part of the swamp. The geological properties and related lithologies of this area are also revealed by its name, “marshes.” The latter represents an area where expected seismic hazard can be drastically increased due to local site conditions, shallow underground water level, active faults surrounding the city and underground topography of basement rocks [30,57].
From the expected strong ground motion, the area is classified as the most hazardous among the Albanian coastal zone based on the seismic microzoning studies (Figure 5), through spectral characteristics, seismic intensities, and expected dynamic soil instabilities [30]. Based on past seismic microzonation, the expected MSK macroseismic intensity varies between VII to IX throughout the city, exhibiting an intensity increment of more than 2 units (Figure 5a). In addition, an equal depth contour map for the city of Durres (Figure 5b), shows a 3D bedrock morphology, implying a 3D basin model. The basin is built on recent Holocene marshy deposits, clays, sands, and peat, with a part of it on Pliocene clays (Figure 6), with liquefaction potential of the broader residential area [58]. The maximum thickness of sediment layers reaches up to 130 m in the central part of the basin (Figure 5).
The aforementioned characteristics are also based on bore-hole data in an East-West [IV-IV] cross-section close to the port (Figure 6) and their physical-mechanical parameters given in Table 2. The position of several boreholes, especially the deepest one, BHA, with a maximum depth of about 120 m, is shown. This cross section was developed in the framework of the Durres seismic microzonation project [30] and has been enriched with geophysical MASW measurements in selected points. One of these points refers to the accelerometer station of Durres (DURR), where the MASW method was performed to estimate the shear wave velocity profile [59]. The measured point is located ~70 m west of the DURR accelerometric station (white arrow in Figure 7). The IV-IV cross section passes through the location of the accelerometric station of Durres (DURR). Values of P and S-wave velocity and NSPT are given up to an 80 m depth profile (Figure 8).

2.5. Ambient Noise Data

For the single station ambient noise measurements two CityShark-II dataloggers (24-bits resolution) coupled with Lenartz-3D 5 sec seismometers were used. In the framework of the present work, 80 single-site measurements were performed in the city of Durres (Figure 9, red dots) during the period 18–21 February 2020. Most of the measurements were performed with a seismometer-to-ground coupling on soil conditions to avoid any high frequency contamination from an artificial surface layer (e.g., cement, asphalt, etc.). A typical ambient noise measurement installation is presented in Figure 9. All measurements were taken during the daytime, avoiding the impact of near source anthropogenic noise as much as possible. The duration of each measurement was 30 min per site.
The mesh of investigated sites was denser in the western part of the city. This choice was based on the fact that observed structural damage from the Nov. 26 mainshock, was greater at the central-to-western part of the city as well as on the steeper basement morphology of the western part, as it is evident in Figure 6. The minimum and maximum distances between the measurement points were 175 m and 750 m, respectively.
Urbanization of the Durres city has been evolved on a soft sediment basin that is elongated in a north-south direction. Soft soil deposits of the Durres basin deepen from its edges to the center, reaching a depth of ~130 m (Figure 6), indirectly implying low fundamental frequencies (fo < 1 Hz) for most of the investigated sites.

2.6. Methods Used

There are different approaches to site characterization in urban environments. Each one has its pros and cons as well as various cost levels. There are two main categories of methods for in situ site characterization; the active and passive (non-invasive) ones. The active ones (e.g., cross-hole, down-hole, CPT, etc.) usually affect the environment and thus prove challenging to implement within cities. In contrast, non-invasive methods are based on the measurement of environmental noise, and do not affect the built environment, while also being much more cost-effective to implement. Among the non-invasive methods based on ambient noise recordings, the one based on a single station measurement per site was implemented in this study. In addition, a 1D ambient noise simulation technique was used to validate the proposed soil profiles up to the bedrock for the city of Durres.
(a)
Data Processing and Analyses of Single Station HVSR
In this study, compiled ambient noise data was processed based on the SESAME project guidelines [10], and more specifically the criteria for reliability of results proposed in the Table of its 2.1 section. Their spatial distribution was investigated to divide the city in various categories of similar HVSR curves, that is, zones of similar dynamic properties (fo, Ao). In addition, for certain sites where an estimated Vs-depth profile was available (see Section 3.3). 1D synthetic ambient noise recordings were generated and their HVSRs were calculated. Comparison between experimental and synthetic HVSRs showed the degree of reliability of the latter, while any deviation from the actual HVSRs either indicated possible attenuation factor (Q) issues and/or 2D/3D imprints due to basin effects.
The data set of 80 measurements was processed in a uniform way, using the Geopsy software package (http://www.geopsy.org, accessed on 5 November 2022). Each recording of 30 min ambient noise was automatically divided into 50 s windows, using the following set of parameters [60]:
Short Time Average (STA) = 1 s, Long Time Average (LTA) = 30 s,
min[STAamplitude/LTAamplitude] = 0.20, max[STAamplitude/LTAamplitude] = 2.5
A Fourier transform was applied to each one of the selected windows, with 5% tapering, and smoothed using the approach of [61], with the b coefficient value set to 40. The mean horizontal spectra for each 50 sec time windows were derived by taking the geometric average, and then calculating the corresponding Horizontal-to-Vertical (H/V) spectral ratio.
An example of the H/V processing using Geopsy is shown in Figure 10. Certain criteria for selecting H/V peaks as fundamental or higher mode frequency (ideal and/or clear peaks) were taken into account based on the SESAME guidelines (http://sesame.geopsy.org/SES_Reports.htm, accessed on 5 November 2022). For each site the mean ± one standard deviation (H/V) spectral ratios were calculated. After the initial processing, 76 out of 80 measurements were considered of good quality and further processed in this study. These 76 HVSR curves are provided as a Supplementary Material of this study, in ascii format.
(b)
Theoretical 1D simulation of ambient noise
In modeling 1D synthetic ambient noise recordings for sites with heterogeneous subsurface structure, such as Durres, we used a two-step numerical code that has been developed in the European SESAME project. In the first step, the noise sources are represented by surface or shallow subsurface point forces randomly distributed in space, direction (vertical or horizontal), amplitude and time [62]. The source time function is either a delta-like signal (impulsive sources) or a pseudo-monochromatic signal (“machinery” sources, harmonic carriers with the Gaussian envelope). In the second step, computation of the associated wavefield is performed with different patterns in accordance with the 1D soil structure. For 1D horizontally layered media, Green’s functions are computed by using the wavenumber method proposed by [63,64], schematically presented in Figure 11.

3. Results

3.1. Experimental HVSR Results

After the 76 HVSRs were calculated, each fundamental frequency, fo, was selected along with corresponding amplitude Ao. The following four categories are proposed in terms of fundamental frequency and shown in Figure 12 (left):
Category 1: 0.20 Hz < fo < 0.60 Hz
Category 2: 0.61 Hz < fo < 0.98 Hz
Category 3: 0.99 Hz < fo < 1.70 Hz
Category 4: 1.89 Hz < fo < 20 Hz
Correspondingly, four categories are proposed in terms of amplitude and shown in Figure 12 (right):
Category 1: Ao < 2.0
Category 2: 2.01 < Ao < 2.99
Category 3: 3.00 < Ao < 3.99
Category 4: 4.00 < Ao < 5.60
In Figure 12, the bedrock depth iso-contours are also given. For Category 1 (0.20 Hz < fo < 0.60 Hz), the fundamental frequency, fo, dominates the central part of the basin with the deepest recent deposits. Category 2 (0.99 Hz < fo < 1.70 Hz) appears at the western part of the basin, while at the eastern part there is a broader transition from Category 2 to Category 3 (0.99 Hz < fo < 1.70 Hz), ending with Category 4 (1.89 Hz < fo < 20 Hz).
The amplitude, Ao, reaches high values mainly in the western part of the basin with a maximum of ~ 6, as well as in the central part, with 3.0 < Ao < 4.0. A broad zone with lower amplitudes, 2.0 < Ao < 3.0, is evident in the eastern part of the basin.
In Figure 13 the dominant frequency, fd, of the examined sites—that is, the frequency where the largest amplitude of the HVSR appears—is presented. Similarly, four categories as in the case of fo can be distinguished. The amplitude of the dominant frequency, Ad, also gives a similar pattern to that of the fundamental frequency, Ao.

3.2. Clustering of the HVSRs and Their Spatial Distribution in Durres

In Seismic Zonation maps spatial grouping of certain geophysical or/and geotechnical parameters (e.g., VS30, fo, A0, etc.) is usually attempted either in urban environment or even in wider regional scale [65,66].
To cluster spatially the 76 HVSRs, the entirety of their curve (in frequency and amplitude) was considered. The best HVSR grouping has been achieved in four categories, as shown in Figure 14. For Category 1 (0.20 Hz < fo < 0.60 Hz) an average fo ~ 0.35 Hz with average Ao ~   3 , is observed. For Category 2 (0.61 Hz < fo < 0.98 Hz) an average fo ~ 0.9 Hz with average Ao ~   3.2 while for Category 3 (0.99 Hz < fo < 1.70 Hz) fo ~ 1.2 Hz with Ao ~   2.5 , and for Category 4 (1.89 Hz < fo < 20 Hz) fo ~ 3 Hz with Ao ~   2.0 (almost flat), are observed.
Two out of seventy-six HVSR curves exhibited a double peak (Figure 15 left), while four out of seventy-six were affected by industrial noise (Figure 15 right) and were not included in the clustering procedure.
Nevertheless, fundamental frequency with depth-to-bedrock distribution shows a satisfactory correlation with large dispersion (Figure 16).
In Figure 17, all HVSRs clustered in four categories are presented for the city of Durres. It is evident that Category 1 (0.20 Hz < fo < 0.60 Hz) corresponds to the deepest part of the basin while Category 2 (0.61 Hz < fo < 0.98 Hz) corresponds to both western and eastern edge of the basin. However, in the eastern part of the basin, a gradual transition from Category 2 to Category 3 (0.99 Hz < fo < 1.70 Hz) is observed. Category 4 (1.89 Hz < fo < 20 Hz) is mainly seen in the eastern hilly area. Categories 1 and 2 exhibit an average amplitude slightly higher than that of 3, Category 3 higher than 2 and Category 4 up to that of 2.

3.3. Synthetics HVSR in Selected Sites

Synthetic data are used to test models in comparison with observations. In geophysics, 1D or 2D/3D subsoil profiles may be available after field investigation, including uncertainties according to the implemented technique. Validation testing of these models or even more their refinement can be performed after they are compared with actual observations [67,68]. Such an approach is presented in this section.
The noise sources were assumed to be randomly distributed around each station in a 1 km radius and lying at depths ranging from 0 to 10 m. The locations for which synthetic recordings of ambient noise were generated are shown in Figure 18. The minimum distance between synthetic sources was 3 m, and the maximum was 1 km. The soil profiles of each site were deduced from the output of the Durres seismic microzonation study [30] and are given into Table 3. In the absence of any information about quality factors Qp and Qs, we assumed Qp = 2Qs = Vs/5 [69,70].
Synthetic spectral ratio HVSRs were calculated using Geopsy, following the recommendations of the SESAME guidelines. In Figure 19 synthetic HVSRs (black dashed line) are plotted together with the corresponding observed ones (black solid line) using the closest experimental ambient noise measurement site.
Comparisons between synthetic HVSRs (Hisada 1D) and observed HVSRs for the examined eight sites in Figure 19 show that the fundamental frequency is generally similar for all sites, though the amplitude of synthetics is consistently higher, up to double around the fundamental frequency as in the case of P40. In two cases (1West and BHA) a double peak appears in the synthetics, which contrasts the actual data. The shape of the synthetic HVSR is different from the observed one in the cases of 1West, P38, DUR, BHA, and SH17. The underlying structure at these stations is laterally heterogeneous, indicating possible 2D/3D effects. For the sites 2West, BH15 and P40 a lower amplitude of the actual HVSRs is observed, most probably due to attenuation of surface layers (low Q-values of the uppermost layered structure). To assess this assumption, the site P40 was chosen, and the Q-values of the s layers overlying the bedrock were decreased 20-fold, thus increasing the attenuation potential. The results are presented on the bottom of Figure 19 (site P40_test) and a very good agreement is observed between synthetic and actual HVSRs. This is in agreement with the findings of [71], who showed the importance of Q-values of the surface geologic layers for generating a smooth bump in HVSRs and decreasing their amplitude around the fundamental frequency of the investigated site.

4. Discussion and Conclusions

In this study, 80 single station ambient noise measurements were performed in the city of Durres within the period of 18–21 February 2020, following the disastrous earthquake of 26 November 2019, that hit the city and inflicted heavy damages to the built environment. Seventy-six (76) out of eighty (80) measurements were processed for the purpose of this study.
The unique strong motion recording in the city’s center (DURR accelerometric station) showed PGA ~ 0.20 g and high spectral values (>0.3 g) in a broad range of periods (0.2 s ≤ T ≤ 2.0 s), implying existence of a low fundamental frequency (down to fo ~ 0.5 Hz). From the geologic and bedrock iso-depth contour maps, it is evident that the city has been developed on a basin of soft geologic layers, with an elongated north-south direction, reaching a thickness of 130 m in its center. Surface geologic layers down-town exhibit low shear-wave velocity, with VS30 ~ 200 m/s.
Data processing and analyses based on the ambient noise Horizontal-to-Vertical spectral ratio (HVSR) method showed that an extended part of the Durres city correlates to a low fundamental frequency, fo < 1.0 Hz. This is in satisfactory agreement with the characteristics of the strong motion recording in the city’s center, thus validating the findings of the applied methodology. This part is mainly in the center-west side of the city, while fundamental frequency decreases smoothly towards its eastern part (the eastern edge of the basin. The vast majority of examined sites (95%) exhibited amplitude of fundamental and dominant frequencies greater than 2, reaching a value of ~ 6. Based on the consensus that HVSR peak amplitude may represent a lower threshold of actual ground amplification than that estimated by the Standard Spectral Ratio method, one would expect equal and/or higher amplification due to basin excitation by the seismic wavefield. A second, higher mode frequency peak, predominantly found in the western edge of the basin, may either indicate a thinner surface geologic layer of high velocity contrast interface with a deeper one, or a more complex site response behavior due to edge effects of the basin. Such an observation needs further investigation using theoretical 2D/3D modeling.
A satisfactory correlation between fundamental frequency and thickness of the soft sediments is observed, though with large dispersion. This may be due either to errors of the bedrock depth estimation or to possible existence of additional superficial strata (e.g., soft colluvium, man-made deposits) which were not included in the subsoil profile. Targeted geological, geophysical and geotechnical data collection is needed to shed light on this issue and refine our understanding with respect to seisxmic risk of the city of Durres.
The four site class categories proposed in this study indirectly indicate spatial variation of amplification, in linear soil behavior. Categories 1 and 2 include zones with higher average HVSR amplitude (>3.0) and are mainly extended in the central and western part of the city. It is remarkable that moderate to heavy earthquake damage was concentrated in the central and western part of the city (Figure 4a), potentially indicating a correlation of damage to site effects. Consequently, this categorization may form the backbone of any detailed and updated microzonation study of the city of Durres. For each category, an average fundamental frequency (±1 standard deviation) can be safely assigned along with its average amplitude. The average HVSR amplitude could be converted to actual S-waves horizontal spectral amplification using appropriate correction factors of the vertical amplification (VACFs) proposed either for other regions of the world ([72,73]) or even better based on Albanian data. In summary, the ambient noise analysis implemented in this study for the city of Durres can be considered efficient and cost effective and confirms its ability to provide reliable results for seismic microzonation purposes in extended metropolitan areas. However, some assumptions made in this work (e.g., 1D consideration of subsoil profile, linear soil behavior) may constitute limitations of the research carried out and consequently additional effort is needed to improve our findings.
Finally, the 1D ambient noise synthetics HVSRs, in comparison to actual HVSRs at the same sites, can effectively validate 1D geophysical profiles for use as seismic response input to any earthquake excitation scenario. In the case of Durres, the 1D approximation seems to be satisfactory. However, seismic site response corrections due to 2D/3D effects and the attenuation factor (Q) of the upper geologic layers must be carefully investigated to realistically assess future ground shaking in Durres.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app122211309/s1.

Author Contributions

Conceptualization, N.T., E.D., L.D. and N.K.; Data curation, E.D., L.D., I.G., A.P. and A.H.; Investigation, N.T., E.D., L.D., I.G., A.H., N.K. and R.K.; Methodology, N.T., L.D.; Software, A.P.; Supervision, N.T.; Writing—original draft, N.T. and E. D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data used in this study are available upon request. Processed data can be found in the Supplementary Material attached to this study.

Acknowledgments

The Institute of Geosciences of Polytechnic University of Tirana and the Institute of Engineering Seismology & Earthquake Engineering, ITSAK-EPPO, Thessaloniki, Greece are greatly acknowledged for providing the equipment for field-data acquisition. Thalis-Panagiotis Theodoulidis substantially improved the English of the manuscript. The remarks and suggestions of two anonymous reviewers and of the Academic Editor have also significantly improved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bard, P.-Y. Microtremor Measurements: A tool for site effect estimation? In Proceedings of the the Effects of Surface Geology on Seismic Motion; Irikura, K., Kudo, K., Okada, H., Sasatani, T., Eds.; Balkema: Rotterdam, The Netherlands, 1999; pp. 1251–1279. [Google Scholar]
  2. Bard, P.-Y.H.; Cadet, B.; Endrun, M.; Hobiger, F.; Renalier, N.; Theodoulidis, M.; Ohrnberger, D.; Fäh, F.; Sabetta, P.; Teves-Costa, A.-M.; et al. From Non-invasive Site Characterization to Site Amplification: Recent Advances in the Use of Ambient Vibration Measurements. In Earthquake Engineering in Europe; Garevski, M., Ansal, A., Eds.; Springer Science & Business Media: New York, NY, USA, 2010; Chapter 6; pp. 105–123. [Google Scholar] [CrossRef]
  3. Borcherdt, R.D. Effects of Local Geology on Ground Motion near San Francisco Bay. Bull. Seismol. Soc. Am. 1970, 60, 29–61. [Google Scholar]
  4. Arai, H.; Tokimatsu, K. S-wave velocity profiling by joint inversion of microtremor dispersion curve and horizontal-to-vertical (H/V) spectrum. Bull. Seismol. Soc. Am. 2005, 95, 1766–1778. [Google Scholar] [CrossRef] [Green Version]
  5. Bonnefoy-Claudet, S.; Cotton, F.; Bard, P.-Y. The nature of noise wavefield and its applications for site effects studies: A literature review. Earth Sci. Rev. 2006, 79, 205–227. [Google Scholar] [CrossRef]
  6. Nogoshi, M.; Igarashi, T. On the Amplitude Characteristics of Microtremor (Part 2). J. Seismol. Soc. Jpn. 1971, 24, 26–40. (In Japanese) [Google Scholar]
  7. Kanai, K.; Tanaka, T. On Microtremor VIII. Bull. Earthq. Res. Inst. 1961, 39, 97–114. [Google Scholar]
  8. Nakamura, Y. A method for dynamic characteristics estimation of subsurface using microtremor on the ground surface. QR Railway Tech. Res. Inst. 1989, 30, 25–33. [Google Scholar]
  9. Nakamura, Y. Clear identification of fundamental idea of Nakamura’s technique and its applications. In Proceedings of the 12th WCEE, Auckland, New Zealand, 30 January–4 February 2000; p. 2656. [Google Scholar]
  10. SESAME Project Guidelines. 2004. Available online: http://sesame.geopsy.org/SES_Reports.htm (accessed on 1 November 2022).
  11. Bard, P.-Y.; the SESAME Participants. The SESAME project: An overview and main results. In Proceedings of the 13th WCEE, Vancouver, BC, Canada, 1–6 August 2004. [Google Scholar]
  12. Molnar, S.; Cassidy, J.F. A comparison of site response techniques using weak-motion earthquakes and microtremors. Earthq. Spectra. 2006, 22, 169–188. [Google Scholar] [CrossRef]
  13. Hunter, J.A.; Crow, H.L. (Eds.) Shear Wave Velocity Measurement Guidelines for Canadian Seismic Site Characterization in Soil and Rock; Geological Survey of Canada: Ottawa, ON, Canada, 2012; Open File 7078; p. 227. [Google Scholar] [CrossRef]
  14. Pilz, M.; Parolai, S.; Leyton, F.; Campos, J.; Zschau, J. A comparison of site response techniques using earthquake data and ambient seismic noise analysis in the large urban areas of Santiago de Chile. Geophys. J. Int. 2009, 178, 713–728. [Google Scholar] [CrossRef] [Green Version]
  15. Haghshenas, E.; Bard, P.Y.; Theodoulidis, N. Empirical evaluation of microtremor H/V spectral ratio. Bull. Earthq. Eng. 2008, 6, 75–108. [Google Scholar] [CrossRef]
  16. Panou, A.; Theodulidis, N.; Hatzidimitriou, P.; Stylianidis, K.; Papazachos, C. Ambient noise horizontal-to-vertical spectral ratio in site effects estimation and correlation with seismic damage distribution in urban environment: The case of the city of Thessaloniki (Northern Greece). Soil Dyn. Earth. Eng. 2005, 25, 261–274. [Google Scholar] [CrossRef]
  17. Albarello, D.; Cesi, C.; Eulilli, V.; Guerrini, F.; Lunedei, E.; Paolucci, E.; Pileggi, D.; Puzzilli, L.M. The contribution of ambient vibration prospecting in seismic microzoning: An example from the area damaged by the April 6, 20009 L’Aquila (Italy) earthquake. Boll. Dii Geof. Teor. Appl. 2011, 52, 513–538. [Google Scholar]
  18. Maresca, R.; Nardone, L.; Gizzi, F.T.; Potenza, M.R. Ambient noise HVSR measurements in the Avellino historical centre and surrounding area (southern Italy). Correlation with surface geology and damage caused by the 1980 Irpinia-Basilicata earthquake. Measurements V 2018, 130, 211–222. [Google Scholar] [CrossRef]
  19. Mouzakiotis, E.; Karastathis, V.; Voulgaris, N.; Papadimitriou, P. Site Amplification Assessment in the East Corinth Gulf Using 3D Finite-Difference Modeling and Local Geophysical Data. Pure Appl. Geophys. 2020, 177, 3871–3889. [Google Scholar] [CrossRef]
  20. Gjini, A.; Cullufi, H.; Deneko, E.; Xhika, P. Behavior of structure type 82/2 (RC frame), during the earthquake of 26 November 2019 in Durrës, Albania. Res. Eng. Struct. Mater. 2021, 7, 595–615. [Google Scholar] [CrossRef]
  21. Braholli, E.; Menkshi, E. Geotourism potentials of geosites in Durrës municipality, Albania. In Quaestiones Geographicae; Bogucki Wydawnictwo Naukowe: Poznań, Poland, 2021; pp. 63–73. [Google Scholar]
  22. Xhafa, S.; Hasani, B. Urban Planning Challenges in the Peripheral Areas of Durres City (Porto Romano). Mediterr. J. Soc. Sci. 2013, 4, 605–613. [Google Scholar] [CrossRef] [Green Version]
  23. Moiriat, D.; Tasellari, A.; Taylor, C. Influence of Structural and Site Factors on Earthquake Consequences in Durrës after the M 6.4 Albania Strong Motion of November 26, 2019. In Proceedings of the International Symposium on Durrës Earthquakes and Eurocodes (ISDEE-2020), Tirana, Albania, 21–22 September 2020. [Google Scholar]
  24. Shehu, R.; Shallo, M.; Kodra, A.; Vranaj, A.; Gjata, K.; Gjata, T.; Melo, V.; Yzeiri, D.; Bakiaj, H.; Xhomo, A.; et al. Geological Map of Albania in Scale 1:200.000; “Hamit Shijaku” Publishing-House: Tirana, Albania, 1983. [Google Scholar]
  25. Aliaj, S. Neotektonika dhe Sizmotektonika e Shqiperise. Master’s Thesis, Archive of Seismological Institute, Tirana, Albania, 1988. [Google Scholar]
  26. Aliaj, S.; Melo, V.; Hyseni, A.; Skrami, J.; Mehillka, L.L.; Muço, B.; Sulstarova, E.; Prifti, K.; Pasko, P.; Prillo, S. Neotectonic Structure of Albania Final Report in Archive of Seismological Institute of Academy of Sciences; Institute of Academy of Sciences: Tirana, Albania, 1996. [Google Scholar]
  27. Aliaj, S.; Baldassarre, G.; Shkupi, D. Quaternary subsidence zones in Albania: Some case studies. Bull. Eng. Geol. Environ. 2001, 59, 313–318. [Google Scholar] [CrossRef]
  28. Skrami, J. Structural and neotectonic features of the Periadriatic Depression (Albania) detected by seismic interpretation. Bull. Greek Geol. Soc. 2001, 34, 1601–1609. [Google Scholar] [CrossRef] [Green Version]
  29. Xhomo, A.; Kodra, A.; Dimo, L.; Xhafa, Z.; Nazaj, S.; Nakuçi, V.; Yzeiraj, D.; Shallo, M.; Vranaj, A.; Melo, V. Geological Map of Albania, Scale 1:200 000; Geological Survey of Albania: Tirana, Albania, 2002. [Google Scholar]
  30. Koçiu, S.; Sulstarova, E.; Aliaj, S.; Duni, L.; Peçi, V.; Konomi, N.; Dakoli, H.; Fuga, I.; Goga, K.; Zeqo, A.; et al. Seismic Microzonation of Durresi Town, Internal Report; IGEWE: Tirana, Albania, 1985. (In Albanian) [Google Scholar]
  31. Mihaljevic, J.; Zupancic, P.; Kuka, N.; Kaluderovic, N.; Koçi, R.; Markusic, S.; Salic, R.; Dushi, E.; Begu, E.; Duni, L.; et al. BSHAP Seismic Source Characterization Models for the Western Balkan Region. Bull. Earthq. Engin. 2017, 23, 3963–3985. [Google Scholar] [CrossRef]
  32. Aliaj, S.; Sulstarova, E.; Muço, B.; Koçiu, S. Seismotectonic Map of Albania in Scale 1:500.000; Seismological Institute: Tirana, Albanian, 2000. [Google Scholar]
  33. Jouanne, F.; Mugnier, J.L.; Koci, R.; Bushati, S.; Matev, K.; Kuka, N.; Shinko, I.; Kociu, S.; Duni, L. GPS constrains on current tectonic of Albania. Tectonophysics 2012, 554–557, 50–62. [Google Scholar] [CrossRef]
  34. Aliaj, S.; Koçiu, S.; Muço, B.; Sulstarova, E. Seismicity, Seismotectonics and Seismic Hazard Assessment in Albania; Academy of Sciences of Albania: Tirana, Albania, 2010. [Google Scholar]
  35. Sulstarova, E.; Kociu, S.; Aliaj, S. Seismic Zonation of Albania; Publication of Academy of Sciences of Albania and Seismological Centre of Albania: Tirane, Albania, 1980; 297p. [Google Scholar]
  36. Papadopoulos, G.A.; Agalos, A.; Carydis, P.; Lekkas, E.; Mavroulis, S.; Triantafyllou, I. The 26 November 2019 Mw 6.4 AlbaniaDestructive Earthquake. Seismol. Res. Lett. 2020, 91, 3129–3138. [Google Scholar] [CrossRef]
  37. Ganas, A.; Elias, P.; Briole, P.; Cannavo, F.; Valkaniotis, S.; Tsironi, V.; Partheniou, E.I. Ground Deformation and Seismic Fault Model of the M6.4 Durres (Albania) Nov. 26, 2019 Earthquake, Based on GNSS/INSAR Observations. Geosciences 2020, 10, 210. [Google Scholar] [CrossRef]
  38. Lekkas, E.; Mavroulis, S.; Papa, D.; Carydis, P. The November 26, 2019 M 6.4 Durrës (Albania) Earthquake. Newsletter of Environmental, Disaster and Crises Management Strategies, 15, 2019. Available online: https://edcm.edu.gr/images/docs/newsletters/Newsletter_15_2019_Albania_EQ.pdf (accessed on 1 November 2022).
  39. Moshou, A.; Dushi, E.; Argyrakis, P. A Preliminary Report on the 26 November 2019, M=6.4 Durrës, Abania Earthquake, EMSC Reports. 2019. Available online: https://www.emsc-csem.org/Files/news/Earthquakes_reports/Preliminary_Report_Albania_26112019.pdf (accessed on 1 November 2022).
  40. Sulstarova, E.; Kociaj, S. The Catalogue of the Albanian Earthquakes; Ex-Seismological Center, the Academy of Sciences of Albania: Tirana, Albania, 1975. [Google Scholar]
  41. Rovida, A.; Antonucci, A. EPICA—European PreInstrumental Earthquake Catalogue; Version 1.1; Instituto Nazionale di Geofisica e Vulcanologia (INGV): Rome, Italy, 2021. [Google Scholar] [CrossRef]
  42. Stucchi, M.; Rovida, A.; Gomez Capera, A.A.; Alexandre, P.; Camelbeeck, T.; Demircioglu, M.B.; Gasperini, P.; Kouskouna, V.; Musson, R.M.W.; Radulian, M.; et al. The SHARE European Earthquake Catalogue (SHEEC) 1000-1899. J. Seismol. 2013, 17, 523–544. [Google Scholar] [CrossRef]
  43. Storchak, D.A.; Di Giacomo, D.; Engdahl, E.R.; Harris, J.; Bondár, I.; Lee, W.H.K.; Bormann, P.; Villaseñor, A. The ISC-GEM Global Instrumental Earthquake Catalogue (1900–2009): Introduction. Phys. Earth Planet. Int. 2015, 239, 48–63. [Google Scholar] [CrossRef]
  44. Scordilis, E.; Papazachos, C.; Karakaisis, G.; Karakostas, V. Accelerating seismic crustal deformation before strong mainshocks in Adriatic and its importance for earthquake prediction. J. Seismol. 2004, 8, 57–70. [Google Scholar] [CrossRef]
  45. Anton, A.; Baballëku, M.; Baltzopoulos, G.; Blagojević, N.; Bothara, J.; Brûlé, S.; Brzev, S.; Carydis, P.; Duni, L.; Dushi, E.; et al. EERI Earthquake Reconnaissance Report-M6.4 Albania Earthquake on November 26 2019; Earthquake Engineering Research Institute: Oakland, CA, USA, 2022. [Google Scholar] [CrossRef]
  46. Van der Heiden, V.; Rietbrock, A.; Tilmann, F.; Schurr, B.; Dushi, E. The November 26, 2019 Mw(6.4) Albania Earthquake post-seismic Campaign preliminary results: Machine Learning-based Solutions. In Proceedings of the Scientific Symposium: Geosciences, Achievements and Future Challenges–2021 (SSGAFC2021), Tirana, Albania, 25 November 2021. [Google Scholar]
  47. Stefanidou, S.; Sotiriadis, D.; Klimis, N.; Margaris, B.; Sextos, A.; Theodoulidis, N. Preliminary aspects on ground motion, site characterization and structural damage of Durrës earthquake (Mw6.4, 26-11-2019). In Proceedings of the International Symposium on Durrës Earthquake and Eurocodes (ISDEE-2020), Tirana, Albania, 21–22 September 2020. [Google Scholar]
  48. Duni, L.; Theodoulidis, N. Short Note on the November 26, 2019, Durrës (Albania) M6.4 Earthquake: Strong Ground Motion with Emphasis in Durrës City. 2019. Available online: https://www.emsc-csem.org/Files/news/Earthquakes_reports/Short-Note_EMSC_31122019.pdf (accessed on 1 November 2022).
  49. Bard, P.-Y.; Campillo, M.; Chávez-Garcia, F.J.; Sánchez-Sesma, F. The Mexico Earthquake of September 19, 1985—A Theoretical Investigation of Large- and Small-scale Amplification Effects in the Mexico City Valley. Earthq. Spectra 1998, 4, 609–633. [Google Scholar] [CrossRef]
  50. Ademovic, N.; Hadzima-Nyarko, M.; Zagora, N. Influence of site effects on the seismic vulnerability of masonry and reinforced concrete buildings in Tuzla (Bosnia and Herzegovina). Bull. Earthq. Eng. 2022, 20, 2643–2681. [Google Scholar] [CrossRef]
  51. Bashir, K.; Debnath, R.; Saha, R. Estimation of local site effects and seismic vulnerability using geotechnical dataset at flyover site Agartala India. Acta Geophys. 2022, 70, 1003–1036. [Google Scholar] [CrossRef]
  52. Shakib, H.; Dardaei, S.; Farhangian, H.; Torkanbouri, N.E. Seismological Aspects and Seismic Behavior of Buildings During the M 7.3 Western Iran Earthquake in Sarpol-e-zahab Region. Iran. J. Sci. Technol. Trans. Civ. Eng. 2022, 46, 3063–3079. [Google Scholar] [CrossRef]
  53. Duni, L.; Kuka, N. Seismic hazard assessment and site-dependent response spectra parameters of the current seismic design code in Albania. Acta Geod. Geoph. Hung. 2004, 39, 161–176. [Google Scholar] [CrossRef]
  54. Kuka, N.; Duni, L. Probabilistic Assessment of Seismic Hazard of Albania, Internal Report; IGEWE: Tirana, Albania, 2007; p. 41. [Google Scholar]
  55. NATO Science for Peace program, SPS Reference 984374. Improvements in the Harmonized Seismic Hazard Maps for the Western Balkan Countries (2015). Multi-Year Project Final Report. 2015, p. 61. Available online: https://www.nato.int/science/studies_and_projects/nato_funded/pdf/983054.pdf (accessed on 1 November 2022).
  56. Kociu, S.; Kociu, L. Seismic risk reduction of big coastal cities in Albania. Coasts at the Millennium. In Proceedings of the 17th International Conference of the Coastal Society, Portland, OR, USA, 9–12 July 2012. [Google Scholar]
  57. Kociu, S.; Skrami, J. Seismic Microzonation of a New Metropolitan Area of Albania. In Proceedings of the American Geophysical Union, Fall Meeting 2006, (Abstract ID:S23E-0204), San Francisco, CA, USA, 11–15 December 2006. [Google Scholar]
  58. Kociu, S. Induced Seismic Impacts Observed in Coastal Area of Albania: Case Studies. In Proceedings of the Fifth International Conference on Case Histories in Geotechnical Engineering, New York, NY, USA, 13–17 April 2004. [Google Scholar]
  59. Duni, L.; Kuka, N.; Koçi, R.; Dushi, E. Shear Waves Velocity Using the MASW Method in Durrës City, Internal Report; IGEWE: Tirana, Albania, 2020. [Google Scholar]
  60. Theodoulidis, N.; Grendas, I.; Duni, L.; Dushi, E.; Kuka, N.; Koci, R.; Rrezart, B.; Gjuzi, O. Report on Ambient Noise Measurements and Data Analyses in Durrës City, Albania; Thessaloniki-Tirana, ITSAK Internal Report. 2020, p. 29. Available online: http://www.itsak.gr/uploads/news/earthquake_reports/Report_AmbientNoise_Durres_April2020.pdf (accessed on 1 November 2022).
  61. Konno, K.; Ohmachi, T. Ground-motion characteristics estimated from spectral ratio between horizontal and vertical components of microtremor. Bull. Seismol. Soc. Am. 1998, 88, 228–241. [Google Scholar] [CrossRef]
  62. Moczo, P.; Kristek, J.; Vavrycuk, V.; Archuleta, R.; Halada, L. 3D heterogeneous staggered-grid finite-difference modeling of seismic motion with volume harmonic and arithmetic averaging of elastic moduli and densities. Bull. Seismol. Soc. Am. 2002, 92, 3042–3066. [Google Scholar] [CrossRef]
  63. Hisada, Y. An efficient method for computing Green’s functions for a layered half-space with sources and receivers at close depths. Bull. Seismol. Soc. Am. 1994, 84, 1456–1472. [Google Scholar] [CrossRef]
  64. Hisada, Y. 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. 1995, 85, 1080–1093. [Google Scholar] [CrossRef]
  65. Maringue, J.; Mendoza, L.; Sáez, E.; Yañez, G.; Montalva, G.; Soto, V.; Ayala, F.; Perez-Estay, N.; Figueroa, R.; Sepúlveda, N.; et al. Geological and geotechnical investigation of the seismic ground response characteristics in some urban and suburban sites in Chile exposed to large seismic threats. Bull. Earthq. Eng. 2022, 20, 4895–4918. [Google Scholar] [CrossRef]
  66. Van Ginkel, J.; Ruigrok, E.; Stafleu, J.; Herber, R. Development of a seismic site-response zonation map for the Netherlands. Copernic. GmbH 2022, 22, 41–63. [Google Scholar] [CrossRef]
  67. Panou, A.; Theodoulidis, N.; Hatzidimitriou, P.; Savvaidis, A.; Papazachos, C. Reliability of ambient noise horizontal-to-vertical spectral ratio in urban environment: The case of Thessaloniki city (Northern Greece). Pageoph 2005, 162, 891–912. [Google Scholar] [CrossRef]
  68. Bustos, J.; Pastén, C.; Pavez, D.; Acevedo, M.; Ruiz, S.; Astroza, R. Two-dimensional simulation of the seismic response of the Santiago Basin, Chile. Soil Dyn. Earthq. Eng. 2023, 164, 107569. [Google Scholar] [CrossRef]
  69. Aki, K.; Richards, P. Quantitative Seismology-Theory and Methods; Freeman, W.H. & Company: New York, NY, USA, 1980; 932p. [Google Scholar]
  70. Bertil, D.; Bethoux, N.; Campillo, M.; Massinon, B. Modeling crystal phases in southeast France for focal depth determination. Earth Plan. Sci. Lett. 1989, 95, 341–358. [Google Scholar]
  71. Guiller, B.J.-L.; Chatelain, M.; Hellel, D.; Machane, N.; Mezouer, R.; Ben Salem, E.H. Oubaiche Smooth bumps in H/V curves over a broad area from single-station ambient noise recordings are meaningful and reveal the importance of Q in array processing: The Boumerdes (Algeria) case. Geoph. Res. Lett. 2005, 32, L24306. [Google Scholar] [CrossRef]
  72. Ito, E.; Nakano, K.; Nagashima, F.; Kawase, H. A Method to Directly Estimate S-Wave Site Amplification Factor from Horizontal-to-Vertical Spectral Ratio of Earthquakes (eHVSRs). Bull. Seismol. Soc. Am. 2020, 110, 2892–2911. [Google Scholar] [CrossRef]
  73. Theodoulidis, N.; Maragakis, I.; Grendas, I.; Hatzidimitriou, P.; Kawase, H.; Ito, E.; Triantafyllidis, P. Estimation of S-wave Horizontal Spectral Amplification factor (HSAF) from earthquake Horizontal-toVertical Spectral ratio (eHVSR) in Greece. In Proceedings of the 3th European Conference on Earthquake Engineering and Seismology, Bucharest, Romania, 4–9 September 2022; p. 6460. [Google Scholar]
Figure 1. (a): Geological map of Durres (section from the Geological Map of Albania 1:200,000, from [29]); the profile trace line (A—B), oriented WSW—ENS, crosses the main geological units of Durres, through Porto Romano Bay; (b): Schematic geologic section, along the A—B profile, trending NE, from the Adriatic Sea to Kruja-Tirana region, [27]. The red star indicates the location the of 26 November 2019 mainshock’s hypocenter (M6.4) [39].
Figure 1. (a): Geological map of Durres (section from the Geological Map of Albania 1:200,000, from [29]); the profile trace line (A—B), oriented WSW—ENS, crosses the main geological units of Durres, through Porto Romano Bay; (b): Schematic geologic section, along the A—B profile, trending NE, from the Adriatic Sea to Kruja-Tirana region, [27]. The red star indicates the location the of 26 November 2019 mainshock’s hypocenter (M6.4) [39].
Applsci 12 11309 g001
Figure 2. Seismic sequence of September 2019–June 2020, corresponding focal mechanism solutions (FM) of the larges shocks and the active faults influencing Durres region.
Figure 2. Seismic sequence of September 2019–June 2020, corresponding focal mechanism solutions (FM) of the larges shocks and the active faults influencing Durres region.
Applsci 12 11309 g002
Figure 3. Historical and recent earthquakes affecting Durres and the surrounding region seismicity.
Figure 3. Historical and recent earthquakes affecting Durres and the surrounding region seismicity.
Applsci 12 11309 g003
Figure 4. (a) Building damage distribution in Durres (not an exhaustive list) after the mainshock of 26 November 2019 (modified from [23]); (b) The acceleration response spectrum at the DURR accelerometer station from the 26 November 2019 mainshock.
Figure 4. (a) Building damage distribution in Durres (not an exhaustive list) after the mainshock of 26 November 2019 (modified from [23]); (b) The acceleration response spectrum at the DURR accelerometer station from the 26 November 2019 mainshock.
Applsci 12 11309 g004
Figure 5. (a) Microzoning map for the city of Durres (b) iso-depth contours of Durres down to geologic rock formations (modified from [30]).
Figure 5. (a) Microzoning map for the city of Durres (b) iso-depth contours of Durres down to geologic rock formations (modified from [30]).
Applsci 12 11309 g005
Figure 6. Geologic cross-section IV—IV [30].
Figure 6. Geologic cross-section IV—IV [30].
Applsci 12 11309 g006
Figure 7. Location of the cross section IV-IV, 6.0 km long (top part); The position of DURR station (red triangle) and several boreholes (SH-xx) are also presented (bottom part). The white arrow indicates the MASW-Vs measurement profile-P39 carried out close to the DURR accelerometric station.
Figure 7. Location of the cross section IV-IV, 6.0 km long (top part); The position of DURR station (red triangle) and several boreholes (SH-xx) are also presented (bottom part). The white arrow indicates the MASW-Vs measurement profile-P39 carried out close to the DURR accelerometric station.
Applsci 12 11309 g007
Figure 8. East-west cross section and lithology (upper); Vs profile of the DURR station for the upper 30 m (lower).
Figure 8. East-west cross section and lithology (upper); Vs profile of the DURR station for the upper 30 m (lower).
Applsci 12 11309 g008
Figure 9. Measurement points within the Durres city (top); An example of measurement site (bottom).
Figure 9. Measurement points within the Durres city (top); An example of measurement site (bottom).
Applsci 12 11309 g009
Figure 10. Snapshot example of the H/V processing using Geopsy.
Figure 10. Snapshot example of the H/V processing using Geopsy.
Applsci 12 11309 g010
Figure 11. The multi-layered half space model considered by [64]. Point sources are located at the Z axis, and receivers at (r, θ, z), with the displacement components of the Green’s functions Ur, Uθ, Uz (modified from [64]).
Figure 11. The multi-layered half space model considered by [64]. Point sources are located at the Z axis, and receivers at (r, θ, z), with the displacement components of the Green’s functions Ur, Uθ, Uz (modified from [64]).
Applsci 12 11309 g011
Figure 12. Distribution of the fundamental frequency, fo, in the city of Durres (left); Distribution of the corresponding to the fundamental frequency, Ao, amplitude (right). Bedrock depth iso-contours by [30].
Figure 12. Distribution of the fundamental frequency, fo, in the city of Durres (left); Distribution of the corresponding to the fundamental frequency, Ao, amplitude (right). Bedrock depth iso-contours by [30].
Applsci 12 11309 g012
Figure 13. Distribution of the dominant frequency, fd, in the city of Durres (left); distribution of the corresponding to the dominant frequency, Ad amplitude (right). Bedrock depth iso-contours by [30].
Figure 13. Distribution of the dominant frequency, fd, in the city of Durres (left); distribution of the corresponding to the dominant frequency, Ad amplitude (right). Bedrock depth iso-contours by [30].
Applsci 12 11309 g013
Figure 14. Four categories of HVSRs for the city of Durres (average red line ± one standard deviation blue lines).
Figure 14. Four categories of HVSRs for the city of Durres (average red line ± one standard deviation blue lines).
Applsci 12 11309 g014
Figure 15. HVSRs of two sites exhibited double peaks (left); HVSRs of four sites affected by industrial noise (right).
Figure 15. HVSRs of two sites exhibited double peaks (left); HVSRs of four sites affected by industrial noise (right).
Applsci 12 11309 g015
Figure 16. Fundamental frequency (fo) in relation to the depth-to-bedrock based on the 76 HVSRs and the underlying bedrock of the Durres basin. In red circles flat HVSRs are depicted.
Figure 16. Fundamental frequency (fo) in relation to the depth-to-bedrock based on the 76 HVSRs and the underlying bedrock of the Durres basin. In red circles flat HVSRs are depicted.
Applsci 12 11309 g016
Figure 17. Spatial distribution of the four HVSR site categories in the city of Durres, together with iso-depth contours. Bedrock depth iso-contours by [30].
Figure 17. Spatial distribution of the four HVSR site categories in the city of Durres, together with iso-depth contours. Bedrock depth iso-contours by [30].
Applsci 12 11309 g017
Figure 18. West−East geologic cross section developed in the framework of the Durres seismic microzonation project and the sites (black arrows & triangles) where synthetic ambient noise recordings were generated. Bedrock depth iso-contours by [30].
Figure 18. West−East geologic cross section developed in the framework of the Durres seismic microzonation project and the sites (black arrows & triangles) where synthetic ambient noise recordings were generated. Bedrock depth iso-contours by [30].
Applsci 12 11309 g018
Figure 19. Synthetic (dashed line) versus observed (solid line) HVSRs for the eight selected sites on the West-East cross section of Figure 18.
Figure 19. Synthetic (dashed line) versus observed (solid line) HVSRs for the eight selected sites on the West-East cross section of Figure 18.
Applsci 12 11309 g019aApplsci 12 11309 g019b
Table 1. Historical and instrumental era earthquakes (M > 6.0), 58BC-2019, in the Durres region (M implying moment magnitude and with * being the equivalent moment magnitude [44].
Table 1. Historical and instrumental era earthquakes (M > 6.0), 58BC-2019, in the Durres region (M implying moment magnitude and with * being the equivalent moment magnitude [44].
DateLatitudeLongitudeDepthMag.Location
yyyy/mm/ddN-SE-WkmM
58 BC41.219.3336.5 *Durres
33441.319.5336.2 *Durres
34641.319.3336.5 *Durres
50641.319.5336.2 *Durres
52141.319.5336.0 *Durres
127341.219.3336.7 *Durres
161719.70041.500156.2 *Kruja
185219.50041.600156.2 *Rodoni Cape
186019.7041.300156.2 *Ndroqi
187019.44641.314156.5 *Durres
192619.5041.300116.3Durres
193419.6041.250125.7Ndroqi
197019.6041.15555.6Vrapi
197519.3041.50155.4Rodoni Cape
197942.1618.87166.9Montenegro
198841.2219.8155.7Tirana
2019, Sept. 2119.50341.47623.45.6Durres
2019, Nov. 2619.46941.42120.56.4Durres
Table 2. Description of physical-mechanical parameters: ϒ (T/m3)—specific weight; Δ (T/m3)—volume weight; δ (T/m3)—skeleton volume weight; W %—humidity; n %—porosity; ϕ—porosity coefficient; Φ (°)—internal friction angle; C (kg/cm2)—cohesion; μ (m/24 h)—permeability.
Table 2. Description of physical-mechanical parameters: ϒ (T/m3)—specific weight; Δ (T/m3)—volume weight; δ (T/m3)—skeleton volume weight; W %—humidity; n %—porosity; ϕ—porosity coefficient; Φ (°)—internal friction angle; C (kg/cm2)—cohesion; μ (m/24 h)—permeability.
Thickness
(m)
LithologyGranulometryϒ (T/m3)Δ (T/m3)δ (T/m3)W %n %ϕΦ (0)C (kg/cm2)μ (m/24 h)
ClayGritSand
0–3 mMarshy-lagoon sub sand with peat4.024722.641.841.3138390.8240.05-
3–15 mPowdery water-saturated sand3.015822.661.861.4835480.89260.051.0
15–28 mSub-clay1760232.681.891.535480.92140.10.05
28–130 mClay3060102.721.91.638490.95140.25-
>130 mClayey-sand basement rock ---2.752.081.911.13400.68---
Table 3. Information on the 8 site profiles (Figure 18) [30], used in the 1D synthetic ambient noise recordings.
Table 3. Information on the 8 site profiles (Figure 18) [30], used in the 1D synthetic ambient noise recordings.
2West1WestP38DURR
H (m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)
2019015001080400101501450101201400
1021015505120140020235155016.72011500
20233158010201150030250157023.32331580
07002000552331580802651600402421600
070020000700200007002000
BHASH17BH15P40
H (m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)H(m)Vs (m/s)Vp (m/s)
101201400401201400121201400101401450
202011500102011500132011500202401550
1302381580552331580302331580203001710
07002000070020000700200007002000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Theodoulidis, N.; Dushi, E.; Duni, L.; Grendas, I.; Panou, A.; Hajrullai, A.; Kuka, N.; Koci, R. Local Site Effects Investigation in Durres City (Albania) Using Ambient Noise, after the 26 November 2019 (M6.4) Destructive Earthquake. Appl. Sci. 2022, 12, 11309. https://doi.org/10.3390/app122211309

AMA Style

Theodoulidis N, Dushi E, Duni L, Grendas I, Panou A, Hajrullai A, Kuka N, Koci R. Local Site Effects Investigation in Durres City (Albania) Using Ambient Noise, after the 26 November 2019 (M6.4) Destructive Earthquake. Applied Sciences. 2022; 12(22):11309. https://doi.org/10.3390/app122211309

Chicago/Turabian Style

Theodoulidis, Nikos, Edmond Dushi, Llambro Duni, Ioannis Grendas, Areti Panou, Ardit Hajrullai, Neki Kuka, and Rexhep Koci. 2022. "Local Site Effects Investigation in Durres City (Albania) Using Ambient Noise, after the 26 November 2019 (M6.4) Destructive Earthquake" Applied Sciences 12, no. 22: 11309. https://doi.org/10.3390/app122211309

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop