Diss. ETH Nr. 15652


EARTHQUAKE SOURCE PARAMETERS IN THE ALPINE-MEDITERRANEAN REGION FROM SURFACE WAVE ANALYSIS


A dissertation submitted to the

Swiss Federal Institute of Technology
Zürich, Switzerland


Dissertation for the degree of

Doctor of Natural Sciences


presented by
Fabrizio Bernardi


Dipl. Natw. ETH


Born on June 30, 1973
Citizen of Stabio - TI, Switzerland


Accepted on the recommendation of


Prof. Dr. Domenico Giardini, examiner
Dr. Jochen Braunmiller, co-examiner
Dr. Urs Kradolfer, co-examiner
Prof. Dr. Torsten Dahm, co-examiner


2004

_______________________________________________________________________________________

Contents

1 Introduction
2 Brief overview of seismic moment tensor
3 Automatic regional moment tensor inversion in the European-Mediterranean region
 3.1 Summary
 3.2 Introduction
 3.3 Data and method
  3.3.1 Data acquisition
  3.3.2 Algorithm
  3.3.3 Inversion
 3.4 Quality assessment of results
  3.4.1 True quality
  3.4.2 Automatic assigned quality
 3.5 Discussion
  3.5.1 Performance
  3.5.2 Frequency band and location
  3.5.3 Current improvements
 3.6 Conclusions and outlook
  Acknowledgments
4 Automatic moment tensor analysis of moderate earthquakes
 4.1 Summary
 4.2 Introduction
 4.3 Distance dependent short period cut-off
 4.4 Signal-to-noise ratio cut-off
 4.5 Application to MT data-set
 4.6 Discussion
 4.7 Conclusions and outlook
5 Seismic moment from regional surface wave amplitudes - applications to digital and analog seismograms
 5.1 Summary
 5.2 Introduction
 5.3 Procedure
 5.4 Recent earthquakes
  5.4.1 Data
  5.4.2 Reference period TD
  5.4.3 Distance dependence
  5.4.4 Seismic moment
 5.5 Historical earthquakes
  5.5.1 Data
  5.5.2 Surface wave amplitude and seismic moment
  5.5.3 Seismic moment uncertainty
 5.6 Discussion and conclusions
  Acknowledgments
6 Conclusions and outlook
A Waveform fit examples
B MT from regional data
C Early instrumental seismograms
D Instrument characteristics of early instrumental seismographs

_______________________________________________________________________________________

Abstract

In this thesis, I present two methods to retrieve earthquake source parameters from regional surface wave data for moment magnitude Mw > 4.3 earthquakes. The first method involves fast and fully automatic moment tensor (MT) computation. Automatization includes near real-time earthquake alert screening, data collection from near-real time accessible broad-band stations at regional distances (D < 20o), MT computation, solution quality assessment and dissemination. The routine is triggered by events with magnitude M > 4.7 in the European-Mediterranean region. MTs are computed using long period data with PREM synthetics. I tested various long period pass bands to evaluate the influence of location accuracy and of simple 1D synthetics on MT retrieval. The best fixed period band for the entire region is currently 50 - 100 s, because relatively few stations are near-real time accessible and often the quickly available locations are of low precision. To assess solution quality, the automatic results are compared with the independent, manually derived Swiss regional moment tensor catalog. Solutions are divided into three qualities. Near real-time application from April 2000 to April 2002 resulted in 38 quality A, with well resolved Mw, depth and focal mechanism, 21 B, with well resolved Mw and 28 unreliable quality C solutions. For Mw > 5.5 we consistently obtained A solutions. Between Mw = 4.5 - 5.5 we obtained quality A and B. In a second step, I significantly improved MT retrieval relative to the standard routine by selecting different period bands for each station and component. The band width depends on the epicentral distance and on the signal-to-noise ratio of each period within each seismogram component. For events in the eastern Mediterranean Sea and the near East, the shortest period used is 50 s. For events in Europe, the short period cut-off varies from 35 s for close stations to 50 s for distant stations. Only periods that exceed a signal-to-noise ratio Rm are actually used. The use of shorter periods and removal of noisy data leads to an improvement of solution quality for 4.3 < Mw < 4.7 earthquakes. For events between May 2002 and September 2003 the new routine provided 20% more quality A solutions than the standard routine. Successful analysis of more smaller earthquakes with the new routine suggests to lower the currently implemented trigger magnitude from M > 4.7 to about M > 4.3.

The second method retrieves seismic moment Mo directly from surface wave amplitudes recorded at regional distances. The amplitude-moment relation is derived from digital broad-band data of 18 earthquakes (3.9 < Mw < 5.1) in and near Switzerland with independent Mo values. The amplitudes were measured at empirically determined, distance varying, reference periods TD. For amplitudes measured at TD, the distance attenuation term of the surface wave magnitude relation S(D) = log(A/T)max + 1.66logD is independent of distance. Mo is then defined by logMo = S(D) + 14.90 assuming 1:1 scaling of log Mo - MS, which is true for MS <~ 7.2 in continental areas. Uncertainties of ±0.30 for the 14.90-constant correspond to a factor of 2Mo uncertainty, which was verified with independent data. This relation allows fast, direct Mo determination which can be applied to any earthquake. Re-calibration of the 14.90-constant, however, is probably required for other tectonic regions. I applied the amplitude-Mo relation to determine Mo of 25 stronger 20th century events in Switzerland from analog data. The data recorded by early instrumental seismographs were collected from several European observatories. Amplitudes were measured from scans at large magnification and corrected for differences between TD and the actual measurement period. The resulting magnitudes range from Mw = 4.6 to 5.8. The Mw = 5.8 1946 event in the Valais was the largest 20th century earthquake. Magnitude uncertainties for the early-instrumental events are on the order of 0.4 units. The Mo estimates were then used for the up-date of the Earthquake Catalog of Switzerland to calibrate an intensity-Mo relationship for pre-instrumental data. Accurate intensity-Mo relations are fundamental for improved seismic hazard evaluations in regions characterized by moderate seismicity and by long recurrence times of larger events.

_______________________________________________________________________________________

Riassunto

In questo lavoro sono presentati due metodi, i quali, utilizzando onde di superficie, permettono di determinare i parametri della sorgente di terremoti regionali con magnitudine Mw > 4.3. Il primo calcola velocemente ed in maniera completamente automatica il tensore momento. Il processo di automatizzazione include l’analisi delle localizzazioni automatiche provenienti da diverse agenzie, la richiesta dei dati in tempo reale da stazioni a banda larga situate a distanza regionale (D < 20o), il calcolo del tensore momento, la deteminazione dell’affidabilita della soluzione e la sua pubblicazione (sul web e via e-mail). La procedura analizza ogni evento dichiarato con magnitudine M > 4.7 e localizzato in Europea e nell’ area mediterranea. I tensori momento sono calcolati usando dati reali a lungo periodo e librerie di sismogrammi sintetici calcolati con modello PREM. Sono stati testati diversi intervalli di periodo, focalizzando l’analisi sulla stabilitá delle soluzioni in base alla precisione della localizzazione usata ed ai sismogrammi sintetici calcolati per un semplice modello 1D. La bassa affidabilità delle localizzazioni disponibili e l’ancora scarsa disponibilità di dati a banda larga in tempo reale evidenziano che l’intervallo compreso tra i 50 ed i 100 secondi risulta essere il più adatto per l’analisi dei terremoti nell’intera area europea e mediterranea. Il confronto dei risultati con un catalogo indipendent di soluzioni tensore momento ha permesso di stabilire le regole generali per determinare l’affidabilità delle soluzioni così ottenute. Quest’ultime sono state quindi divise in tre gruppi: soluzioni di ’tipo A’, con momento sismico Mo, profondità e meccanismo focale affidabili; soluzioni di ’tipo B’, con il solo valore di Mo affidabile ed infine, soluzioni di ’tipo C’, per cui nessuno dei tre parametri elencati può essere considerato affidabile. Durante il periodo compreso tra l’aprile 2000 e l’aprile 2002, la procedura ha ottenuto 28 soluzioni di tipo A, 21 di tipo B e 28 di tipo C. Per eventi con Mw > 5.5 generalmente si sono ottenute soluzioni di tipo A, mentre per eventi con Mw = 4.5 - 5.5 si sono ottenuti soluzioni di tipo A e B. In seguito la procedura è stata significativamente migliorata selezionando l’intervallo dei periodi in base alla distanza e al rapporto tra il segnale ed il rumore contenuto nei sismogrammi. Pertanto, eventi nel Mediterraneo orientale e nel vicino oriente sono stati analizzati con periodi compresi tra 50 e 100 secondi, mentre, quelli verificatisi in Europa, con il periodo più corto che varia tra i 35 s per le stazioni più vicine ed i 50 s per quelle più lontane. Inoltre sono stati utilizzati solo i periodi il cui segnale supera un determinato rapporto tra segnale e rumore, Rm. L’utilizzo di periodi più corti e la rimozione di sismogrammi contenenti rumore permettono un significativo miglioramento soprattutto per eventi con magnitudine 4.3 < Mw < 4.7. L’analisi di eventi occorsi tra il maggio 2002 ed il settembre 2003 con la nuova procedura ha prodotto 66 soluzioni con qualità A, migliorando del 20% le prestazioni rispetto alla prima procedura. L’efficacia della nuova procedura suggerisce di spostare la magnitudine di soglia da M > 4.7 verso M > 4.3.

Con il secondo metodo è possibile calcolare il momento sismico Mo direttamente dall’ampiezza delle onde di superficie. La relazione tra l’ampiezza ed il momento sismico è stata derivata utilizzando dati digitali di 18 terremoti (3.9 < Mw < 5.1) occorsi recentemente in Svizzera e nelle zone limitrofe, per i quali esistono stime indipendenti del momento sismico ottenute tramite inversione del tensore momento. L’ampiezza delle onde di superficie à stata misurata con periodi di riferimento TD dipendenti dalla distanza e determinati empiricamente, cosicché il termine dell’attenuazione della magnitudine delle onde di superficie S(D) = log(A/T)max + 1.66logD non varia più rispetto alla distanza tra stazione ed evento. Assumendo un rapporto di 1:1 per l’espressione log Mo - MS, considerata valida per terremoti continentali con magnitudine MS <~ 7.2, il momento sismico è di conseguenza definito dalla relazione logMo = S(D) + 14.90. L’incertezza di ±0.30 della costante 14.90 porta ad un fattore di incertezza di 2Mo, verificato con dati indipendenti. Questa relazione permette di ottenere velocemente una stima di Mo per terremoti recenti ed è applicabile in ogni regione dopo aver ricalibrato la costante. L’applicazione a sismogrammi registrati su strumenti meccanici ed elettromeccanici, raccolti in vari osservatori europei ha permesso di determinare il momento sismico di 25 terremoti verificatisi in Svizzera e nelle zone limitrofe durante il 20o secolo. Attraverso l’ingrandimento digitale delle immagini dei sismogrammi si è potuta effettuare una precisa lettura dell’ampiezza e del periodo. Questo è stato successivamente corretto rispetto il periodo di riferimento TD. Le magnitudini Mw ottenute risultano comprese tra 4.6 e 5.8, quest’ultima per il più forte terremoto avvenuto in Svizzera (1946, Vallese). L’incertezza della magnitudine determinata con questi strumenti è di circa 0.4. I valori di Mo ottenuti hanno permesso la calibrazione tra l’intensità ed il momento sismico per gli eventi pre-strumentali, utile all’aggiornamento del catalogo dei terremoti in Svizzera (ECOS). Precise relazioni tra in momento sismico e l’intensità sono fondamentali per ottenere stime affidabili del l’azzardo sismico in regioni a sismicità moderata e con lunghi tempi di ricorrenza.

Chapter 1
Introduction

In this thesis I present two methods to retrieve earthquakes source parameters (seismic moment Mo, depth and focal mechanism) for moderate to strong earthquakes from regional surface waves. The first method, described in chapters 3 and 4, performs fully automatic and near-real time waveform inversion for the seismic moment tensor. The primary motivation is to provide rapid magnitude and depth estimates after strong and potentially damaging events to assist disaster relief agencies’ response efforts. The second method, described in chapter 5, retrieves Mo from surface wave amplitudes and can be applied to digital broad band seismograms or to early-instrumental analog seismograms. Here, this method is applied to derive Mo from 20th century analog data, that are crucial for long term seismic hazard studies in areas characterized by low and moderate seismicity.

Fast intervention of disaster relief agencies needs accurate Mo and depth estimates within a short time after a potentially destructive event. In the European-Mediterranean area seismicity is mainly located in the eastern Mediterranean Sea (Figure 1.1), where several destructive earthquakes (e.g. the Mw = 7.6 August 17, 1999 Izmit event) occurred during the last years. In Europe and in the central and western Mediterranean Sea seismicity is lower, but strong destructive earthquakes are not rare (e.g. the Mw = 6.0 September 6, 2002 Molise, the Mw = 7.0 May 21, 2003 Northern Algeria and the Mw = 6.4 February 24, 2004 Morocco earthquakes). When such an earthquake occurs, tele-communications in and around the epicentral area are often interrupted. Magnitude and depth estimates, that are crucial for damage evaluation, can then be estimated only using far-field seismological data. Moment tensor (MT) analysis is particularly well suited because MT inversion prevents underestimation of the magnitude compared to the standard location techniques, since moment tensor derived seismic moment of large events does not saturate (Kanamori1977).


PIC

Figure 1.1: Seismicity and near-real time accessible broadband stations in the European - Mediterranean region. Grey dots are earthquakes with magnitude M > 4 (PDE catalog) between 1999 and 2003. Dot size is proportional to the magnitude. Diamonds are broadband stations accessible in near-real time via AutoDRM from the Virtual European Broadband Seismograph Network (van Eck et al.2004) or from national data centers.

Fast size and depth estimates are also important after moderate events that do not cause serious damages, but are widely felt. Thanks to the world-wide-web, people search for information immediately after a felt earthquake. In Europe even after a moderate event, a large number of people want to be quickly informed about its magnitude and its consequences. For this reason the press, police and civil protection agencies need fast and precise information about location, size and depth of the earthquake. An example is the recent Mw = 4.5 February 23, 2004 Rigney eastern France event, that was felt over the western part of Switzerland. Within minutes the Swiss Seismological Service got swamped with calls from concerned people and press, and the web-server was heavily accessed.

Quick moment tensor solutions are routinely provided by several groups at a global scale but only for large events (Mw > 5.5). These solutions are disseminated only after revision by a seismologist. Moment tensor solutions for moderate (Mw > 4.8) events in the European-Mediterranean region are now provided by the Swiss Seismological Service, the Italian INGV and the Spanish IGN. Such solutions are computed manually and only for the more relevant earthquakes the moment tensor is available within hours after event origin time, while other earthquakes are analyzed with a delay of weeks to months. The routine presented here computes fully automatic moment tensors from earthquake alert detection, to data collection, moment tensor inversion, solution assessment and finally solution dissemination (via e-mail and web). Computing fully automatic, near-real time moment tensors for a moderate event (4.5 < Mw < 5.0) requires near-real time access to dense networks of broadband and high dynamic range seismometer at regional distance (D < 20o). Significant technological improvements and infrastructure investments during the last few years in the European-Mediterranean area allowed the development and implementation of the procedure described here.

The seismic moment tensor provides not only size and depth estimates to alert disaster relief agencies. Earthquake focal mechanisms, that can be obtained from the seismic moment tensor, provides indispensable information for seismotectonic studies, e.g. for stress field estimation in Switzerland (Kastrup et al.2004). The strain rate tensor, that can be computed from the moment tensor, combined with GPS measurements and geological data can then be used to infer deformation fields of the crust, e.g. for deformation fields in the eastern Mediterranean See (Jenny et al.2004). Seismic hazard studies then need accurate magnitude and depth estimates. The automatic routine presented here, will over time result in a complete moment tensor catalog that includes also moderate events, which are significant for regions of low and moderate seismicity.

The Swiss Seismological Service evaluates the seismic hazard in Switzerland. Accurate seismic hazard maps are indispensable for defining building codes in densely populated and industrialized regions or nuclear power plant construction requirements. An example is the Basel area, where strong destructive earthquakes have occurred in the past (e.g. the Mw = 6.9 1356 event, which is the largest event known in Europe north of the Alps) and where numerous chemical plants are operating. To evaluate seismic hazard for such areas with long term recurrent seismicity, requires a complete earthquake catalog covering long time periods.


PIC

Figure 1.2: Seismogram of the November 16, 1911 Balingen earthquake with magnitude Mw = 5.5 recorded on smoke-paper (top) by the 1200 Kg Wiechert Seismograph in Göttingen (bottom). The original size of the record is 12.5 X 4.6 cm.

The Swiss Seismological Service recently completed a comprehensive up-date of the earthquake catalog of Switzerland and surrounding regions (Fäh et al.2003). The catalog expresses the size of all events using moment magnitude (Mw). The largest part of the catalog includes pre-instrumental events where only epicentral intensities are available. For these events, the epicentral intensities were linked to the seismic moment using a epicentral intensity - seismic moment relation that was inferred using earthquakes with both epicentral intensity and seismic moment available. To enlarge this data-set, also the major 20th century events that occurred within or near Switzerland were used. Seismic moment determination of these events constitutes the second large part of this thesis.

For these events, I collected analog seismograms from various observatories in Europe that were recorded on mechanical and electromagnetic seismographs for which the instrument response is well known and the instrument calibrations were available. Figure 1.2 shows an example of a Wiechert seismograph and one of its recordings. Because of the low quality of such records, a reliable digitization was not possible for most of the collected data-set and, thus, the moment tensor inversion technique could not be applied to derive Mo estimates for most of the events investigated. Nevertheless, accurate readings of amplitude and periods were possible. I thus developed a method to evaluate seismic moment (Mo) directly from surface wave amplitude. The method, developed using broad-band digital seismograms of recent events, was then applied to the early-instrumental seismograms of the 20th century major Swiss events.

Chapter 2
Brief overview of seismic moment tensor

The seismic moment tensor describes the earthquake source as a point source. The earthquake source parameters like focal mechanism (strike, dip, slip) and magnitude (seismic moment Mo) follow from the moment tensor. The displacement u(x,t) can be written as (Aki & Richards2002)

           integral   oo     integral   integral 
                                     -d--
un(x, t) =   - oo  dt    S[ui(q,t )]cijpqnjdqq Gnp(x, t- t;q, 0)dS
(2.1)
where Gnp(x,t - t; q, 0) is the n-th component of the earth at the station in response to a unit-force excitation in p-direction at the source (Green function). q is a general position on the fault surface S and x is the station. Using the convolution, equation 2.1 can be rewritten as

           integral   integral 
                            d
un(x,t) =      [ui]cijpqnj * dqqGnpdS
              S
(2.2)
Equation 2.2 refers to a displacement un(x,t) caused by a dislocation along a fault plane S of finite extent. Assuming that the periods considered are much longer than the source volume, the earthquake source can be approximated by a point source, reducing equation 2.2 is

un(x,t) = Mpq * Gnp,q
(2.3)
where Mpq =  integral  integral S[ui]njcijpq is the seismic moment tensor. Knowing u(x,t) (measured at seismic stations) and Gnq,p(x,t) (computed for any Earth), the MT is then computed solving

       -1
M  = G   u
(2.4)
The method used in this thesis to solve for the MT uses amplitude spectra of observed and synthetic seismograms (Giardini1992). Using amplitude spectra avoids artifacts caused by filters used to select the inverting period band, that are generally used for inversion techniques in the time domain, and allows to select different period bands for different stations and components maintaining a linear approach to solve the moment tensor. This approach allows a significant improvement of the results for small and moderate events when only a sparse seismic network is available (chapter 4).

A moment tensor can be decomposed into elementary tensors following two different approaches: a pure mathematical decomposition or a decomposition into source components with a physical meaning (e.g. double-couple, explosion, tension crack components, etc.). The decomposition into physical sources is not unique. The paper of Jost & Herrmann (1989) gives an excellent overview on the different decomposition methods that are generally applied in seismology. The MT decomposition performed in this thesis is described in the next paragraphs. Generally the MT can be decomposed uniquely into an isotropic and a deviatoric part. In order to derive a general formulation of the moment tensor decomposition, let mi be the eigenvalues corresponding to the orthogonal eigenvector ai = (aix,aiy,aiz)T of M pq. Using the orthogonality of the eigenvectors, we can write the principal axis transformation of M in reverse order as:

                    |  |
                    | T|
     |          |   ||a 1||
M  = ||a   a   a ||m  ||aT ||
       1   2   3    | 2|
                    ||aT3||
     |             ||            ||             |
     ||a1x  a2x  a3x||||m1   0    0 ||||a1x  a1y  a1z||
     ||             ||||            ||||             ||
M  = |a1y  a2y  a3y|| 0   m2   0 ||a2x  a2y  a2z|
     ||             ||||            ||||             ||
     |a1z  a2z  a3z|| 0   0   m3 ||a3x  a3y  a3z|
(2.5)
From equation 2.5, we find relations between components of the eigenvectors and the moment tensor elements:

           2       2        2
Mxx =  m1a 1x + m2a 2x + m3a 3x
M   =  m  a2 + m  a2  + m  a2
  yy     1 1y    2 2y     3 3y
Mzz =  m1a2  + m2a2   + m3a2
           1z      2z       3z
Mxy =  m1a1xa1y + m2a2xa2y  + m3a3xa3y

Mxz =  m1a1xa1z + m2a2xa2z  + m3a3xa3z

Myz =  m1a1ya1z + m2a2ya2z +  m3a3ya3z
(2.6)
where m in equation 2.5 is the diagonalized moment tensor and the elements of m are the eigenvalues of M. The general moment tensor decomposition can now be defined by rewriting m as:

       |                     |   |            |
       ||tr(M  )    0       0  ||   ||m*   0    0 ||
     1 ||                     ||   ||  1         ||
m =  --|  0    tr(M )     0  | + |0   m*2   0 |
     3 ||                     ||   ||            ||
       |  0       0    tr(M )|   |0    0   m*3|
       ||                     ||
       |tr(M  )    0       0  |
     1-||                     ||   ---
m =  3 ||  0    tr(M )     0  || + m1
       |                     |
       |  0       0    tr(M )|
(2.7)
where tr(M) = m1 + m2 + m3 is the trace of the moment tensor and mi* are purely deviatoric eigenvalues mi -13tr(M). The first term of equation 2.7 describes the isotropic part of the diagonalized moment tensor, and is important to quantify the volume change in the source. The second term describes the deviatoric part and can be further decomposed using different approaches.

In this study the deviatoric part is decomposed into a double-couple component and a compensated linear vector dipole (CLVD) (Knopoff & Randall1970). Assuming |m3*|<|m 2*|<|m 1*| the deviatoric part can be rewritten as

         |                |
         |                |
         ||- F     0      0||
m--=  m* ||                ||
  1     3| 0   (F -  1)  0|
         || 0      0      1||
(2.8)
where F = -m1*/m 3* and (F - 1) = m 2*/m 3*. Note that 0 < F < 0.5. Equation 2.8 can then be decomposed into two parts representing a double couple and a CLVD

                  ||         ||       ||           ||
                  |0  0   0 |       |- 1  0   0 |
---     *         ||         ||    *  ||           ||
m1  = m 3(1-  2F )||0  -1  0 ||+ m 3F || 0   -1  0 ||
                  |         |       |           |
                  |0  0   1 |       | 0   0   2 |
(2.9)
Such decomposition has the advantage that both double couple and CLVD components have the same orientation of their principal axis system (chapter 3). To estimate the deviation of the seismic source from the model of a pure double couple Dziewonski et al. (1981) used the parameter

    |m*min|
e = ---*---
    |m max|
(2.10)
For a pure double couple e = 0 (mmin* = 0), while for a pure CLVD mechanism e = 0.5.

The focal mechanism (strike, dip and slip) of the earthquake can be solved from the six elements of the seismic moment tensor:

M11 = - Mo(sindcoscsin2f   + sin2dsincsin2f)

M   = +M   (sindcosccos2f  + 1-sin2dsincsin2f)
  12      o                  2

M13 = - Mo(cosdcosccosf   + cos2dsincsinf)
M22 = +Mo(sindcoscsin2f    - sin2dsinccos2f)

M23 = - Mo(cosdcoscsinf   - cos2dsinccosf)

M33 = +Mosin2dsinc)
(2.11)
where d is the dip, c the slip and f the strike.

The magnitude of an earthquake can be expressed in term of scalar seismic moment Mo that can be determined from the seismic moment tensor:

      1-
Mo  = 2 (| m1 |+ |m2 |)
(2.12)
where m1 and m2 are the largest eigenvalues in absolute sense. The moment magnitude Mw is defined as (Kanamori1977):

       logMo--
Mw  =   1.5  -  10.73
(2.13)

The depth of the earthquake centroid (same as hypocenter for small earthquakes) is retrieved by multiple trials of the moment tensor inversions for different depths. We do not invert for the centroid location. For each solution we obtain the normalized variance

        V~ --------------
         (synt - obs)2
V ar =   --------2----
              obs
(2.14)
The best hypocenter depth is assumed to be at the depth with the smallest normalized variance.

A more extensive and detailed description of the seismic moment tensor can be found in Jost & Herrmann (1989), Dahlen & Tromp (1998) and in Aki & Richards (2002). For details of the MT code see Giardini (1992).

Chapter 3
Automatic regional moment tensor inversion in the European-Mediterranean region


This Chaper is published in Geophy. J. Int. (Bernardi et al.2004)

3.1 Summary

We produce fast and automatic moment tensor solutions for all moderate to strong earthquakes in the European-Mediterranean region. The procedure automatically screens near real-time earthquake alerts provided by a large number of agencies. Each event with magnitude M > 4.7 triggers an automatic request for near real-time data at several national and international data centers. Moment tensor inversion is performed using complete regional long period (50 - 100 s) waveforms. Initially the data are inverted for a fixed depth to remove traces with a low signal-to-noise ratio. The remaining data are then inverted for several trial depths to find the best-fit depth. Solutions are produced within 90 minutes after an earthquake. We analyze the results for the April 2000 to April 2002 period to evaluate the performance of the procedure. For quality assessment, we compared the results with the independent Swiss regional moment tensor catalog (SRMT) and divided the 87 moment tensor solutions into three groups: 38 quality A with well-resolved Mw, depth and focal mechanism; 21 quality B with well-resolved Mw; and 28 unreliable quality C solutions. The non-homogeneous station and event distribution, varying noise level, and inaccurate earthquake locations affected solution quality. For larger events (Mw > 5.5) we consistently obtained quality A solutions. Between Mw = 4.5 - 5.5 we obtained quality A and B solutions. Solutions that pass empirical rules mimicking the a posteriori quality for our data set are automatically disseminated.

3.2 Introduction

Quick and accurate source parameter determination (magnitude, depth and focal mechanism) provides important information for fast intervention in areas strongly damaged by large earthquakes. When compared to standard determination of earthquake location and magnitude, moment tensor inversion usually provides more accurate source parameters and particularly so for the size of large events, since their moment magnitudes Mw do not saturate (Kanamori1977). The recent increase in the number of near real-time accessible broadband stations in the European-Mediterranean region now allows automatic, near real-time moment tensor analysis monitoring strong events.

Accurate and quick moment tensor inversion is routinely performed on a global scale by the Harvard Centroid-Moment Tensor (CMT) project (Dziewonski et al.1981Dziewonski & Woodhouse1983), the United States Geological Survey (Sipkin19821986) and the Earthquake Research Institute (ERI), Japan (Kawakatsu1995). These approaches use teleseismic data, limiting analysis to stronger earthquakes (Mw > 5.5). Analysis of smaller earthquakes requires data recorded at regional distances, which is becoming possible with the growing number of broadband seismic networks. Several groups in the United States (Dreger & Helmberger1993Ritsema & Lay1993Romanowicz et al.1993Nábe  lek & Xia1995Braunmiller et al.1995Thio & Kanamori1995Dreger et al.1995Pasyanos et al.1996) and Japan (Kubo et al.2002), routinely invert for the seismic moment tensor using seismic data recorded at near regional distances (D < 10o).

In the European-Mediterranean region, event-station distances are generally larger (D > 10o) than in the western United States. However, several studies (Arvidsson & Ekstrom1998Braunmiller1998) showed that source parameters of moderate events can be determined routinely with long period regional data (T > 30 - 40s) for larger event-station distances. The Swiss Seismological Service (Braunmiller et al.20002002) and the Italian Istituto Nazionale di Geofisica e Vulcanologia (INGV) (Pondrelli et al.2002) now routinely produce moment tensor solutions in the European-Mediterranean area for moderate to strong earthquakes (Mw > 4.5), and smaller earthquakes (Mw > 3.0) in the Alpine area (Braunmiller et al.20002002). Small to moderate earthquakes in the western Mediterranean region are regularly processed by the Spanish Instituto Andaluz de Geofísica (Stich et al.2003).

Here we test whether the current near real-time data availability and location accuracy are sufficient for automatic regional moment tensor retrieval using intermediate to long period surface waves (50 - 100 s). Our goal is to develop a robust procedure that provides fully automatic, fast and reliable solutions for moderate to strong earthquakes in the European-Mediterranean region.

Our automatic procedure consists of two main parts: first, moment tensor inversion for a detected earthquake and second, automatic quality assessment of the solution. We first describe the method and show the reliability of the automatic solutions. Then we present empirical rules for automatic quality assessment. Finally, we discuss source parameters (focal mechanism, depth and Mw) and investigate factors that affect solution quality.

3.3 Data and method

Ultimately, we like to obtain a moment tensor within a few minutes after an event. This requires on-line data access to stations at close epicentral distances, currently unavailable for most of our study region. We anticipate that present waiting periods due to sparse networks and time delays for data acquisition will shorten considerably in the future.

3.3.1 Data acquisition

Automatic moment tensor inversion consists of several steps (Figure 3.1). A few minutes after an event, an automatic earthquake alert (location, magnitude) is generally available, provided by the Swiss Seismological Service (SED) and other agencies linked to our institute. Automatic information coming from many agencies may include false alarms, but guarantees that we miss no significant event. We screen incoming information and any event in the European-Mediterranean area (22o < Lat < 68o,-25o < Long < 60o) with magnitude M > 4.7, independent of magnitude type [ML,mb,MS], starts the routine automatically.


PIC

Figure 3.1: Sketch of the automatic moment tensor inversion procedure. The procedure is triggered when an earthquake alert indicates an event with magnitude M > 4.7 in the European-Mediterranean region (22o - 68oN, 25oW - 60oE). Data requests are sent 50 minutes after the origin time: 30 minutes correspond to the waveform length used for analysis, we wait for 20 additional minutes to assure data availability. After another 20 minutes, data acquisition is considered complete and inversion starts, resulting in an automatic moment tensor solution within 90 minutes after an event.

We invert complete broadband waveforms recorded at regional epicentral distances (D < 20o); therefore, our data window is 30 minutes long. We wait an additional 20 minutes to assure data availability (Figure 3.1) before sending data requests to the AutoDRM (Kradolfer1996) of several international (ORFEUS, USGS) and national data centers (in Austria, Czech Republic, Germany, Israel, Norway, Switzerland) that provide near real-time broadband seismograms. The data centers automatically process these requests and usually within 20 minutes all available seismograms are received at our server to be stored and prepared for inversion. 70 minutes after event origin time, data acquisition is considered complete and the inversion starts, resulting in an automatic moment tensor solution within 90 minutes after an earthquake.

Between April 2000 and April 2002 only stations in central and northern Europe, Israel, the Caucasus region and Russia provided near real-time data (black triangles in Figure 3.2). Data availability and the response time of data centers varied. To ensure that we received a sufficient number of seismograms, we chose a relatively long waiting period before requesting data and starting the inversion. Stations in the western and central Mediterranean Sea, Turkey and eastern Europe (white triangles in Figure 3.2), became available during mid-2002 and are only shown to illustrate the evolving and improving station coverage.


PIC

Figure 3.2: Map of near real-time accessible broadband (BB) stations and automatic moment tensor solutions obtained between April 2000 and April 2002 (event Nr. 37 in central Kazakhstan is not displayed). Solid triangles show stations available for this study. Open triangles depict stations that became available after April 2002 and were not used here. We use only some of the 28 BB stations operated by the Swiss Seismological Service. No BB stations were available from the seismically active Aegean Sea region. Circles show the events analyzed, shading indicates the quality of the moment tensor solution (see section 3.4.1): dark grey circles are true quality A, light grey circles true quality B and open circles true quality C solutions.

3.3.2 Algorithm

The algorithm uses intermediate to long period three-component regional data. The inversion code, described in Giardini (1992), has already been applied to many earthquakes (Giardini1992Giardini et al.1993a,bSicilia1999). Synthetic seismograms are generated by normal mode summation (Woodhouse1988) computed for the PREM-Earth model (Dziewonski & Anderson1981) at different source depths and stored in libraries for quick access. The moment tensor is constrained to be deviatoric. We do not invert for the source centroid but compute time corrections by re-aligning data and synthetics (Giardini1992). Depth is retrieved by minimizing the normalized variance (defined as ratio of variance over data vector norm) for different trial depths.

3.3.3 Inversion

The moment tensor inversion automatically starts 70 minutes after an event (Figure 3.1). The alert is the first available location, which is not necessarily the best epicenter estimate. 70 minutes after an event, several automatic and/or manual locations are usually available and we choose the a priori most accurate. We prefer a location from a network that surrounds the epicenter. When that is not available, we look for a manual location or one provided by an agency with a large aperture network.

The seismograms are bandpass filtered between 50 and 100 s period. A lowpass filter at 50 s applied to regional seismograms minimizes the effect of inaccurately known propagation paths and allows moment tensor retrieval of moderate sized earthquakes (Mw > 4.5) with a simple average 1D velocity model (Arvidsson & Ekstrom1998). We tested other filter parameters and report results in Section 3.5.2.

The automatic inversion consists of two main steps. First, the entire data set is inverted for a fixed depth (18 km) to remove traces with a low signal-to-noise ratio (high normalized variance > 0.8) and large re-alignment. From tests we found, that the choice of the fixed depth - 18 km or different depth - has little effect on the remaining data set and basically no effect on our results. The remaining traces are then inverted for several depths. The 50 - 100 s data have little depth resolution, because long period surface-wave excitation functions have little depth variation. Thus, we apply a limited number of depth-steps with increasing step width (10, 14, 18, 25, 31, 42, 55, 75, 100, 125, 150, 175 and 200 km) to find the best fitting depth. We use a minimum 10% variance increase to estimate depth uncertainty; our uncertainty range is simply defined by the trial depths that just exceed the 10% increase. We consider this uncertainty estimate conservative, since the waveform fit degrades visibly.

3.4 Quality assessment of results

We started our procedure for automatic moment tensor inversion in April 2000. By April 2002, we obtained 87 moment tensor solutions (Figure 3.2, Table 3.1), mainly for events in the seismically active central-eastern Mediterranean region (Jackson & McKenzie1988).

We check the quality of the automatic moment tensors because solution accuracy depends on event location, location precision, station distribution and signal strength. First, we use an independent high-quality moment tensor catalog to quantify true quality. Second, to estimate solution quality automatically, we derive rules from solution parameters which reproduce overall true quality. These rules are applied automatically before disseminating solutions.

3.4.1 True quality

The Swiss automatic moment tensor (SAMT) catalog contains many moderate events not included in global catalogs, incomplete below Mw = 5.5 and with very few Mw < 5.0 events. Therefore, we compared our automatic solutions with those of the Swiss Seismological Service’s regional moment tensor (SRMT) catalog. The SRMT covers the European-Mediterranean area and is nearly complete down to mb = 4.5 (Braunmiller et al.2002). SRMT solutions are derived with a more complete data set available weeks to months after an event than the data set used for SAMT analysis. SRMT solution quality is high based on comparison with other independent source parameter estimates available for selected events (Braunmiller et al.2002). For a few events east of SRMT coverage (> 55oE), we compared our solutions with the Harvard catalog, or when not available, with magnitudes given by the USGS.



Table 3.1: Table of the 87 automatic moment tensor solutions. From left to right: Event number, PDE-location, true and assigned quality, seismic moment Mo in Nm, Mw, depth in km, orientation of the two nodal planes in degree, number of stations and components used, quickly available location used for inversion. For completeness, we include all source parameters even for true quality B and C solutions. For B, only size is well resolved and for C, none of the parameters is reliably resolved. For further studies, use Mw, depth and focal mechanism from A and only Mw from B solutions.
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time LatLongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























1 00-04-0600:10:38.045.7226.58 A Aa .410E+175.01 125 0 20 58 213 72 101 27 50 00:10:39.045.70 26.60
2 00-04-2112:23:10.037.8329.31 B Ba .114E+185.31 31 336 49 -48 103 55 -12711 13 12:23:05.036.80 29.40
3 00-05-2405:40:37.736.0422.01 A Aa .528E+185.75 25 269 47 4 176 86 137 20 44 05:40:35.736.10 21.90
4 00-06-0602:41:49.840.6932.99 A Aa .128E+196.01 10 8 43 -31 122 68 -12826 66 02:41:56.741.30 32.90
5 00-06-1301:43:14.035.1527.13 A Aa .126E+185.34 18 227 82 1 137 88 172 28 74 01:43:18.035.30 27.20
6 00-06-1521:30:29.034.4420.18 A Ba .352E+174.97 42 40 45 65 253 49 112 10 16 21:30:34.434.50 20.00
7 00-07-2519:33:56.037.1321.99 A Aa .442E+175.03 42 333 36 -62 119 57 -10925 41 19:33:47.737.50 24.00
8 00-08-2117:14:26.044.87 8.48 A Aa .281E+174.90 10 114 34 -113 322 59 -74 30 65 17:14:26.044.80 8.60
9 00-08-2313:41:28.040.6830.72 A Aa .187E+185.45 31 251 71 -168 157 78 -18 29 77 13:41:26.040.40 30.70
10 00-11-1515:05:37.038.3542.93 A Aa .288E+185.58 18 217 20 37 92 77 106 5 10 15:05:24.738.40 44.00
11 00-12-0617:11:06.439.5754.80 A Aa .307E+206.93 42 110 44 80 302 46 98 6 18 17:11:07.939.70 54.90
12 00-12-1516:44:47.638.4531.35 A Aa .193E+196.13 18 290 30 -96 117 59 -86 25 62 16:44:45.038.60 31.10
13 01-02-2518:34:42.243.46 7.47 A Aa .997E+164.60 31 249 22 93 65 67 88 16 24 18:34:32.642.90 7.40
14 01-03-1011:20:57.835.0826.37 C Ba .640E+175.14 75 146 46 110 297 47 69 11 23 11:20:59.335.60 27.00
15 01-03-2305:24:12.232.9246.65 B Ca .274E+185.56 18 307 16 22 196 84 104 3 6 05:24:21.833.60 45.30
16 01-03-2816:34:21.829.6951.18 C Ca .205E+174.81 18 313 43 54 178 55 119 5 5 16:34:22.029.80 51.20
17 01-03-3015:30:49.038.0130.94 A Aa .115E+174.64 25 44 29 -103 239 61 -82 15 24 15:30:53.038.70 30.80
18 01-04-0317:36:34.132.4547.99 C Ca .400E+175.01 18 106 38 -139 342 66 -59 4 6 17:37:27.031.50 49.80
19 01-04-0806:12:27.838.3822.18 B Ba .152E+174.73 175 337 41 74 177 50 103 15 18 06:12:27.038.30 22.30
20 01-04-0917:38:39.240.1120.37 A Aa .307E+174.93 31 294 68 22 195 69 156 18 35 17:38:51.039.80 20.40
21 01-04-1014:00:07.634.3026.17 B Ba .214E+174.82 42 4 46 -164 263 78 -44 12 14 14:00:11.034.30 26.10
22 01-05-0106:00:54.135.6427.49 A Ba .684E+175.16 10 337 37 -126 199 60 -65 12 35 06:00:55.035.80 27.40
23 01-05-0419:52:01.934.7222.74 B Ca .272E+174.89 10 13 25 28 256 78 112 2 3 19:52:32.036.00 21.00
24 01-05-1711:43:58.239.0215.47 A Aa .359E+174.97 200 152 16 -149 33 81 -75 11 17 11:43:57.538.90 15.50
25 01-05-2417:34:01.145.7526.46 A Aa .741E+175.18 150 0 30 61 212 63 105 22 42 17:33:55.645.80 26.70


























Table 3.1: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time LatLongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























26 01-05-2904:43:56.435.3927.75 A Ba .655E+175.15 25 314 65 -179 224 89 -24 12 27 04:43:56.035.40 27.70
27 01-06-1001:52:08.039.8755.89 A Ca .132E+185.35 25 114 34 57 331 61 110 2 2 01:52:11.840.50 53.60
28 01-06-2211:54:48.539.3527.34 A Aa .983E+175.27 25 353 52 -17 94 76 -14115 28 11:54:48.039.40 27.70
29 01-06-2306:52:41.935.7128.02 B Ba .558E+185.77 31 239 53 -24 345 70 -14010 23 06:52:45.435.80 28.10
30 01-06-2513:28:46.437.2036.17 A Aa .199E+185.47 10 346 19 -131 210 75 -76 10 25 13:28:49.837.10 35.90
31 01-07-1021:42:06.439.8841.59 B Ca .994E+175.27 31 284 71 -170 191 81 -18 4 6 21:41:49.939.20 44.30
32 01-07-1715:06:15.646.7411.37 A Aa .156E+174.73 10 208 40 -5 302 86 -13010 16 15:06:16.347.00 11.50
33 01-07-2005:09:39.245.7726.78 A Aa .433E+175.03 100 281 12 20 171 85 101 10 25 05:09:10.044.00 29.00
34 01-07-2600:21:37.039.0624.34 A Aa .851E+196.56 25 231 67 -178 140 88 -22 24 70 00:21:37.539.10 24.30
35 01-07-2604:53:34.039.0224.28 A Aa .267E+174.89 31 243 58 176 335 87 31 13 21 04:53:33.139.10 24.30
36 01-07-3015:24:57.039.0424.01 A Aa .478E+175.06 18 227 64 -177 136 88 -25 23 56 15:25:06.039.70 23.90
37 01-08-2215:58:01.247.2470.04 A Aa .932E+175.25 18 149 57 172 243 83 32 3 9 15:58:10.147.70 69.30
38 01-09-1315:42:52.235.5425.97 B Ba .146E+174.71 42 227 44 -7 322 84 -13312 18 15:42:59.634.90 26.90
39 01-09-1602:00:47.337.2421.88 A Aa .282E+185.57 10 118 46 -129 348 55 -56 17 43 02:00:48.437.40 22.00
40 01-09-2511:53:32.436.0132.13 C Ca .461E+185.71 42 106 87 178 196 88 2 1 2 11:53:36.135.90 32.30
41 01-09-2604:19:56.335.0427.04 C Ba .460E+175.05 10 306 7 68 147 83 92 7 12 04:19:08.631.60 30.20
42 01-10-1508:51:09.935.9922.18 B Ba .145E+174.71 125 332 65 -163 235 75 -25 10 13 08:50:44.434.80 23.60
43 01-10-1815:50:30.936.9035.04 C Ba .206E+174.81 42 164 42 -45 291 61 -12311 15 15:50:29.038.00 37.00
44 01-10-1818:08:33.938.7014.81 C Ca .345E+174.96 175 89 63 154 191 67 28 2 2 18:07:23.334.80 18.70
45 01-10-2613:32:48.238.1023.11 B Ba .168E+174.75 18 224 41 -168 126 82 -48 11 19 13:32:38.237.80 24.20
46 01-10-2816:25:21.552.72 0.93 C Ca .415E+175.02 100 231 27 0 141 89 117 1 1 16:25:25.952.80 359.40
47 01-10-2920:34:24.738.9524.25 A Aa .321E+174.94 14 249 38 -138 124 65 -59 14 32 20:21:32.738.60 25.60
48 01-10-3021:00:05.935.9429.78 B Ba .632E+175.14 75 302 69 -165 207 76 -21 14 32 20:59:56.935.70 30.60
49 01-10-3112:33:56.437.2636.05 C Ca .149E+185.38 31 336 42 -134 209 60 -56 5 7 12:33:59.337.70 36.30
50 01-11-0222:05:30.927.1854.62 C Ca .336E+174.95 125 9 16 -46 145 78 -101 1 1 22:05:29.127.00 54.60
51 01-11-0417:23:29.834.0625.43 A Ba .203E+174.81 25 264 17 62 112 74 98 4 9 17:23:27.734.00 25.00
52 01-11-0709:40:43.541.3910.11 C Ca .296E+174.92 25 11 41 59 228 55 113 3 3 09:40:53.541.90 10.10
53 01-11-1801:01:35.835.2228.44 C Ca .232E+174.85 100 44 42 -114 255 52 -69 1 1 01:00:42.931.80 33.50
54 01-11-2600:56:57.043.8912.57 A Aa .129E+174.68 18 8 53 -43 127 56 -134 6 12 00:57:01.043.60 12.50
55 01-11-2605:03:20.934.8324.28 B Ba .615E+175.13 14 133 21 115 286 70 80 13 25 05:03:08.634.50 25.60
56 01-12-1019:50:08.637.3024.62 B Ca .120E+174.66 31 38 37 70 242 54 104 1 1 19:50:04.437.20 25.00
57 01-12-1116:34:05.139.0124.28 C Ca .130E+174.68 150 159 31 80 350 58 95 3 3 16:33:22.538.30 28.00


























Table 3.1: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time Lat LongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























58 01-12-3004:06:28.034.78 27.38 B Ba .108E+185.29 100 238 29 -33 357 74 -115 8 9 04:06:58.535.70 25.00
59 02-01-0122:15:57.737.29 21.83 C Ca .248E+174.87 175 272 12 -132 135 81 -81 2 2 22:15:06.735.50 26.00
60 02-01-0909:35:56.237.90 21.22 C Ca .414E+175.01 31 206 36 -109 50 56 -76 1 2 09:35:23.436.60 23.70
61 02-01-0914:11:08.035.99 22.91 C Ca .357E+174.97 10 122 15 -119 332 76 -82 4 4 14:10:26.534.20 25.80
62 02-01-2114:34:23.838.68 27.82 C Ca .175E+185.43 10 259 69 0 349 89 -159 2 4 14:33:10.735.20 35.60
63 02-01-2204:53:52.235.68 26.68 B Ba .257E+196.21 42 55 20 -118 265 72 -79 15 36 04:53:53.735.70 26.70
64 02-01-2620:05:35.937.16 20.98 B Ba .648E+175.14 18 314 22 25 200 80 110 8 17 20:05:11.536.30 23.00
65 02-02-0307:11:28.838.49 31.31 A Aa .797E+196.54 14 241 37 -122 100 59 -67 20 59 07:11:36.138.80 31.20
66 02-02-0309:26:43.638.63 30.81 A Aa .904E+185.91 25 231 42 -51 4 58 -11921 55 09:26:43.238.60 30.90
67 02-02-0311:54:34.638.56 31.03 A Aa .104E+185.28 18 275 42 -61 58 53 -11310 20 11:54:36.338.50 31.20
68 02-02-0420:09:32.137.14357.57 B Aa .429E+175.03 10 190 19 -11 291 85 -10917 29 20:09:32.437.30 357.30
69 02-02-1403:18:01.746.37 13.23 C Ca .358E+153.64 200 316 41 -112 165 52 -71 1 2 03:18:01.046.40 13.40
70 02-02-1713:03:52.928.12 51.75 B Ba .955E+175.26 55 344 56 -167 247 79 -34 7 9 13:03:52.728.10 51.90
71 02-03-0505:23:45.840.67 25.57 C Ca .144E+174.71 42 85 80 0 355 89 170 4 4 05:23:48.340.80 25.20
72 02-03-0720:30:40.140.98 20.77 C Ca .653E+175.15 125 195 13 63 43 78 95 1 1 20:29:01.539.60 30.10
73 02-03-1012:39:38.339.01 15.65 C Ca .509E+175.07 42 129 4 -177 37 89 -85 1 1 12:40:34.042.50 13.10
74 02-03-1120:06:37.125.23 56.13 C Ca .128E+185.34 42 172 48 -54 304 52 -123 2 3 20:06:31.824.50 57.70
75 02-03-1714:46:27.850.77 6.17 C Ca .964E+164.59 75 154 54 -136 35 55 -44 1 1 14:46:27.051.10 6.00
76 02-04-0415:44:31.927.01 55.34 C Ca .899E+164.57 55 155 45 -36 273 64 -129 1 1 15:44:31.027.00 55.30
77 02-04-0504:52:23.538.48 14.74 A Aa .561E+164.44 14 40 50 44 278 57 130 6 12 04:52:22.038.40 15.10
78 02-04-0507:55:48.637.92 21.03 C Ca .322E+185.61 200 72 20 117 223 72 80 1 2 07:55:32.937.40 22.60
79 02-04-0513:14:02.042.02 24.83 B Ba .168E+174.75 10 194 20 179 285 89 69 8 13 13:14:01.042.10 24.80
80 02-04-0818:31:05.236.56 52.01 C Ca .462E+175.05 18 97 30 55 316 65 108 3 4 18:31:0.036.50 52.20
81 02-04-1508:10:06.134.66 24.59 A Ba .207E+174.81 10 23 8 177 116 89 81 8 12 08:10:06.934.90 24.20
82 02-04-1706:42:53.139.79 16.77 B Aa .410E+175.01 14 142 13 55 357 78 97 19 39 06:42:53.039.80 16.80
83 02-04-1708:47:22.727.72 56.81 C Ca .241E+185.53 25 304 17 -82 117 72 -92 1 2 08:47:23.827.90 56.80
84 02-04-1913:46:49.636.57 49.85 B Ca .787E+175.20 18 185 26 115 337 66 77 4 7 13:46:49.036.60 49.90
85 02-04-2410:51:50.942.41 21.42 A Aa .474E+185.72 14 245 35 -84 58 54 -93 22 61 10:51:51.542.40 21.40
86 02-04-2419:48:07.034.54 47.33 A Aa .172E+185.43 14 314 65 -177 223 87 -24 6 14 19:48:07.034.50 47.30
87 02-04-2420:11:21.534.43 47.23 C Ca .888E+175.24 25 32 74 -4 124 85 -164 3 3 20:11:08.637.30 38.10


























PIC

Figure 3.3: Sketch of true quality assignment. True quality of each automatic solution is based on mean principal axes’ orientation |DAx|, depth |Dz| and magnitude |DMw| differences relative to SRMT solutions (see text for details). Dark grey shaded area corresponds to true quality A, light grey to true quality B and white to true quality C.

The true quality of a SAMT solution is estimated by comparing its focal mechanism, depth and Mw relative to the SRMT solution. We distinguish three quality levels: A has well-resolved mechanisms, depths and Mw, B has only well-resolved Mw, and C is unreliable. Figure 3.3 provides a sketch  of the quality criteria described below.

The most stable focal mechanism parameters are the double couple part and the principal axes’ orientation. Moment tensor solutions for the same earthquake included in different catalogs may show differences in the non-double couple part e (ratio of smallest to largest moment tensor eigenvalues following Dziewonski et al. (1981)). These differences are often introduced by an inaccurate source location (Zhang & Lay1990), differences in the station configuration (S  ílený & Vavryc  uk2002), inaccurate path-models (Henry et al.2002Fröhlich1994) and a poor resolution of Mrf and Mrh (Kuge & Lay1994). We therefore use mean axes’ difference |DAx|, defined as the average of the differences in principal axes’ orientation, to estimate the similarity between SAMT and SRMT focal mechanisms.

The mean difference |DAx| is zero for identical axes’ orientations. Interchanging two axes (for example changing a normal to a thrust or a left lateral to a right lateral mechanism) results in |DAx| = 60o. We thus require A quality solutions to have |DAx|< 30o; solutions with |DAx| > 30o are either B or C quality, depending on the magnitude difference (Figure 3.3). Figure 3.4 (top) shows the distribution of |DAx|. For quality A solutions the mean value of |DAx| is 18.8o ± 6.7o. Two solutions with |DAx|< 30o are not quality A, because one violates the depth (quality B) and one the magnitude criterion (quality C).


PIC

Figure 3.4: Top: Mean principal axes’ orientation difference |DAx| relative to SRMT for true quality A, B and C. Middle: Depth difference |Dz| relative to SRMT. Solutions are divided into two panels: left panel shows the number of quality A, B and C solutions where the SAMT depth uncertainty range includes the SRMT depth. Right panel shows |Dz| for SRMT depth estimates outside SAMT depth uncertainty range. Bottom: |DMw| relative to SRMT. Not all B and C solutions are shown in each panel. See text for details. Color scale as in Figure 3.2.

We also checked the focal parameter agreement between the SAMT and SRMT catalogs using the radiation pattern coefficient jP (Kuge & Kawakatsu1993), a parameter describing the radiation pattern similarity of two mechanisms. Our median jP = 0.85 for A quality solutions compares well with the median of jP = 0.88 found by Helffrich (1997) who compared ERI, Harvard and USGS catalogs for shallow earthquakes (most SAMTs are for shallow earthquakes).

Because of the low-depth resolution and the discrete set of trial depths, we require that the difference |Dz| between SAMT depth range (best-fit depth plus uncertainty) and SRMT depth is < 10 km for an A quality solution. A similar scheme was proposed by Kubo et al. (2002) when comparing the Japanese regional NIED catalog with the Harvard-CMT and the Japanese Meteorological Agency (JMA) focal mechanism catalog. Differences |Dz| > 10 km result in B or C solutions (Figure 3.3). Figure 3.4 (middle) shows |Dz| for the three quality groups.

The difference between SAMT and SRMT Mw estimates must be < 0.2 units for an A or B quality solution (Mw = logMo
-1.5- - 10.73 following Kanamori (1977)). A required difference |DMw|< 0.2 is consistent with Pasyanos et al. (1996), who found that automatic and revised regional MT solutions in northern California have Mw estimates that usually differ by less than 0.2 units even when the focal mechanism and depth estimates differ strongly. Mw depends mainly on the signal amplitude and is therefore the parameter easiest to resolve. One goal for our automatic procedure is to provide robust Mw values so that disaster relief agencies may quickly estimate possible earthquake damages. Damages are governed by the rupture process and local site effects; our Mw can only help to estimate whether no, local or widespread damages are expected. Therefore we accept an Mw estimate as accurate even when the focal mechanism and/or the depth exceed their A-quality threshold: these are our quality B solutions. |DMw| > 0.2 are quality C, irrespective of focal mechanism and depth (Figure 3.3). The distribution of |DMw| is shown in Figure 3.4 (bottom).

For our data we observe that SAMTs with |DAx|< 30o usually have |Dz|< 10 km and |DMw|< 0.2. Based on our rules, the 87 automatic moment tensors are divided into 38 quality A, 21 quality B and 28 quality C solutions.

3.4.2 Automatic assigned quality

The true quality can be verified only a posteriori. For automatic solution dissemination, we derive empirical rules, matching the a posteriori true qualities that can be implemented into the automatic procedure. The rules follow three principles: (1) applied to the whole data set, the rules should closely follow the true quality; (2) the rules should not overestimate quality: true quality B solutions should not be assigned Aa, and quality C should not be assigned Aa or Ba (we use the superscript a to denote assigned quality). We want to have high confidence that an automatic Aa solution is truly A; (3) the rules should be simple.


PIC

Figure 3.5: Empirical rules to assess assigned quality are based on the number of stations and components used to obtain the MT solutions. Dark grey circles are true quality A, light grey circles B and open circles C solutions. On and right of the black thick line is assigned quality Aa (densely stippled area); the lightly stippled area marks assigned Ba; left of the vertical thick dashed line are Ca solutions (numerical values are given in Table 3.2). Thin dashed lines (from left to right) indicate one, two or three components used for each station used. Top is for the entire European-Mediterranean area; bottom for events in the southern Aegean Sea where we generally need data from more stations to obtain true A quality, because of long average station-event distances.

A combination of number of stations and components used (i.e., with good signal-to-noise ratio) provides a simple yet reasonably accurate measure to assess solution quality (Table 3.2). The empirical rules were defined based on the April 2000 to April 2002 data set and may be modified when, for example, station availability increases. Figure 3.5 shows that it is generally sufficient to use 2 or more components for each station to obtain quality A solutions even when only a few stations can be used. Using fewer stations and components results in lower quality solutions. Many solutions of earthquakes located in the larger Iran area result in quality B or C, because of the few stations available. For the southern Aegean Sea (34o < Lat. < 38.5o; 20o < Long. < 30o) we observe an interesting difference: we need more stations and components to obtain true quality A solutions (Figure 3.5). This exception is possibly due to large event-station distances (Figure 3.2).



Table 3.2: Empirical rules applied to derive assigned quality. Number of stations (St) and components (Co) used characterize solution quality (Figure 3.5). Rules for earthquakes located in the southern Aegean Sea (34o < Lat < 38.5o; 20o < Long < 30o) are slightly different (right): there, a larger number of stations is required to obtain assigned quality A.



QualityRules Rules
(Europe-Mediterranean)(Southern Aegean Sea)



A St > 10 & Co/St > 1.5 St > 16 & Co/St > 1.5
St > 5 & Co/St > 2
St > 3 & Co/St = 3



B All other solutions All other solutions



C Co < 7 Co < 7




The empirical rules were designed to not overestimate true quality. Underestimating a few events is implicitly allowed and results in fewer assigned quality Aa and Ba solutions compared to true quality. Undesired upgrade happens only for 2 events from B to Aa (6% of 34 assigned quality Aa solutions) and 3 events from C to Ba (14% of 22 assigned quality Ba solutions). No true C event is assigned to Aa. Such upgrades are confined to smaller events (Mw < 5.3). The empirical rules actually reproduce 82% of the true quality data set; most differences (11 events) are caused by quality downgrade. Based on these observations, we have high confidence that our automatically disseminated solutions cause few (or no) false alerts for large, potentially damaging earthquakes.

3.5 Discussion

3.5.1 Performance

In this section we focus on the geographical event distribution and factors that cause low true quality solutions. We also compare our Mw, depths and focal mechanisms results with the SRMT, the Harvard (CMT) and INGV (MEDNET) moment tensor results to illustrate their high consistency.


PIC

Figure 3.6: Cumulative number of stations used to obtain all quality A (black) and B (grey) solutions versus distance from epicenter. Most BB stations are located in areas of lower seismicity (Figure 3.2) resulting in large event-station distances. Obtaining more high-quality solutions or successfully analyzing smaller events in the future critically hinges on improved availability of nearby (D < 7o) stations.

The distribution of the analyzed events (circles, Figure 3.2) reflects long-term seismicity (Jackson & McKenzie1988). The distribution of high quality solutions, however, is also affected by the station distribution (black triangles, Figure 3.2). Generally, analysis is hampered by long average event-station distances (Figure 3.6). Most data come from stations at distances D > 7o with a peak around D = 14o - 15o caused by high seismicity in the Aegean Sea and high station density in central Europe. Most high quality solutions are produced for earthquakes in the central-eastern Mediterranean region, where data from stations in central Europe, Israel, the Caucasus region and Russia provide good azimuthal coverage. In the western Mediterranean region, seismicity is relatively low and the inadequate station distribution resulted in a quality B solution for the one event analyzed. Station coverage for events in the Caspian Sea and Zagros mountain regions is also low and few of the frequent events have well-recovered source parameters. In central and northern Europe we obtained few high quality automatic solutions. Although a large number of stations is available, there, the lowpass filter at 50 s precludes moment tensor retrieval for the typically smaller (Mw < 4.5) events of this region.

Not all events triggered by our automatic procedure resulted in a moment tensor. Triggered events result in no-solution when either (1) no data are available, (2) the automatic alert is a false alarm, (3) the epicenter is strongly mislocated, (4) the true magnitude is far lower than the alert magnitude or when any of these cases combine. Case (4) caused most no-solution events due to the low trigger-threshold set for analysis (M > 4.7), that assures the processing of all stronger events. We let the procedure decide whether a solution can be produced or not. For a few smaller events, few traces containing only long-period noise were inverted, resulting in quality C solutions.


PIC

Figure 3.7: Automatic Mw estimates relative to SRMT (top), CMT (middle) and MEDNET (bottom). Mean difference (m), standard deviation (s), slope and intercept of the linear regression (y = a + b . x, dashed line) are computed for qualities A and B combined. Solid line represents a one-to-one relation. Color scale as in Figure 3.2.

Earthquake size is the most stable source parameter and a few, good signal-to-noise seismograms generally constrain the seismic moment Mo. Figure 4.12 shows the high correlation of the automatic Mw estimates relative to SRMT, CMT and MEDNET Mw’s. Mean differences are very small, m < 0.02 relative to SRMT and CMT, and lower than 0.1 unit relative to MEDNET, with standard deviations close to 0.1. The linear regressions (dashed lines in Figures 4.12) have slopes close to 1 and small intercepts, roughly consistent with a one-to-one relation between the magnitude estimates. We did not interpret small apparent differences because the data set is too small. Mean differences and regressions were determined for A and B quality solutions combined, because we did not see any significant systematic difference in the A and B quality Mw estimates (Figure 3.4 bottom).


PIC

Figure 3.8: Number of true quality A and B solutions relative to SRMT Mw estimates. Color scale as in Figure 3.2.

Quality A solutions were obtained for earthquakes from Mw = 4.5 to Mw = 7.0. In general larger events (Mw > 5.5) result in quality A solutions (Figure 3.8), and earthquakes with Mw < 5.5 result in quality A or B. We obtained 3 quality B solutions for earthquakes with Mw > 5.5; in all cases the lower quality was caused by the inaccurate quick location used for the inversion. We repeated the inversion with the PDE-location. In two cases we obtained A quality solutions. One case stayed B quality, because of |DAx| = 31o, just outside the A quality criterion (|DAx|< 30o); the depth and magnitude differences both satisfied the A quality criteria. The number of MT solutions and the ratio of A/B solutions decrease for earthquakes with Mw < 5.0 due to lower signal strength and the large event-station distances (Figure 3.6).


PIC

Figure 3.9: Correlation between normalized variance and Mw for solutions of true quality A and B. Color scale as in Figure 3.2.

Figure 3.9 illustrates the limitations of long-period (T > 50 s) analysis with respect to magnitude or signal strength. Variance increases with decreasing event size, effectively setting the lower limit for retrieving MT solutions to Mw  ~~ 4.5. At such long periods signal strength at smaller magnitudes is just slightly above the noise level. We also observe that variance for quality B is higher than it is for A solutions.


PIC

Figure 3.10: SAMT depth estimate differences for true quality A solutions relative to SRMT (top), CMT (middle) and MEDNET (bottom). Differences are small: mean differences (long dashes) are m < 4 km with standard deviations (short dashes) s < 15 km.

Long period surface waves offer only limited depth resolution (Giardini1992). However, depth of quality A solutions agree very well with those in SRMT, CMT and MEDNET (Figure 3.10). The mean difference m is always < 4 km with a standard deviation s < 15 km. Our SAMT catalog contains only 3 deep earthquakes with quality A and we observe no significantly greater depth estimate differences for these events. Events with large depth difference (|Dz|> 20) are different events in each panel, reflecting the depth differences in the catalogs used for comparison. Note that quality A solutions for shallow and deep earthquakes are always correctly distinguished.


PIC

Figure 3.11: Focal mechanisms of the 38 true quality A solutions (first column), compared with SRMT (second column), Harvard-CMT (third column) and MEDNET (fourth column) solutions. For each event, event number (Table 3.1), date and location (PDE) are given. Depth (km) and Mw are indicated on the bottom and top of each focal mechanism respectively. Focal mechanisms show a very good agreement. Automatic moment tensor inversion with complete regional seismograms in the period range of 50 - 100 s allows robust MT retrieval for earthquakes with Mw > 4.5. Large differences for 3 Harvard CMT solutions (Nr. 20, 33, 36) are discussed in Section 3.5.1.

Figure 3.11 shows the focal mechanisms of the true quality A solutions together with the available SRMT, CMT and MEDNET solutions. Focal mechanisms show excellent agreement for all earthquakes: shallow, deep, weak and strong. Larger differences exist for only 3 CMT solutions (Nr. 20, 33, 36). These earthquakes are small (Mw = 5.0 - 5.1) for CMT analysis. Their CMTs have large non-double couple parts e (0.316 < e < 0.342) and large relative moment tensor uncertainties E (0.261 < E < 0.673) compared to the average values e = 0.124 and E = 0.165 of all 19589 CMT solutions (1976 until November 2002). E is defined in Davis & Fröhlich (1995). Moment tensors with high e and E have poorly constrained focal parameters (Fröhlich et al.1997).

3.5.2 Frequency band and location

Our goal is to retrieve as many quality A solutions as possible, even for smaller earthquakes (4.5 < Mw < 5.0), while minimizing the number of quality C solutions. Solution quality depends on station distribution, location accuracy and the ability to correctly match phases in the seismograms. For the given station distribution (Figure 3.2), we performed two tests. First, we tried to find the optimum frequency band for analysis with the quickly available locations and data set used for near real-time processing. The frequency band needs to match phases - easy at longer periods - and contain good signal-to-noise seismograms - higher at shorter periods. We thus performed inversions for 5 selected period ranges (40 - 60, 45 - 80, 50 - 100, 60 - 125, and 70 - 140 s). In a second step, we repeated the same analysis with the more accurate PDE-locations to see whether location accuracy affects the choice of frequency band.


PIC

Figure 3.12: Performance of the automatic MT inversion routine for 5 different period ranges with quickly available (top) and PDE (bottom) locations. From left to right: (1) number of true quality A, B and C solutions. (2) median values of the radiation pattern correlation coefficients jP (quality A), vertical dashed line is median jP = 0.88 obtained by Helffrich (1997) comparing ERI, Harvard and USGS catalogs. (3) mean depth differences Dz with error (quality A) (4) mean Mw difference DMw with standard deviation (quality A and B combined). jP , Dz and DMw are relative to SRMT. Focal mechanism, depth and Mw estimates are well-constrained for all period ranges, but performance changes: with quickly available locations, the best period range for our inversions is 50 - 100 s, with PDE-locations 45 - 80 s. Color scale as in Figure 3.2.

The number of quality A, B and C solutions changes strongly with different frequency bands (Figure 3.12, top left). The largest number of quality A solutions (and highest ratio of A/C solutions) is obtained with the 50 - 100 s band for the quickly available locations. Using the more accurate PDE-locations, the optimal period range shifts to lower periods of 45 - 80 s (Figure 3.12, bottom left). The reason for this shift is that mislocation introduces errors in the initial phase that are then mapped onto the moment tensor. The effect is smaller at longer periods (Patton & Aki1979) and the accuracy of the quick locations sets the optimum period range to 50 - 100 s. At longer periods (60 - 125, 70 - 140 s) the number of quality A and B solutions decreases for the quick and the PDE-locations because of weak signals for the smaller events. At these period ranges, only stronger earthquakes can be analyzed.

For shorter periods (40 - 60 s), the number of quality A solutions decreases strongly even for the PDE-locations. In most cases, quality A solutions at 45 - 80 s (or 50 - 100 s) become B at 40 - 60 s (Figure 3.12, bottom left). At periods below 50 s, surface waves become more sensitive to crustal thickness and average crustal velocity variations so that significant travel-time differences relative to PREM result (Larson & Ekström2001Pasyanos et al.2001). Unresolved near-source earth structure may also cause surface wave amplitude anomalies (van der Lee1998). Both phase and amplitude differences limit reliable moment tensor retrieval below 50 s period for our uneven station distribution and our average event-station distances of about 1500 km.

The frequency band chosen has little effect on the overall quality of the focal mechanism and depth estimates for quality A solutions. This also holds for the Mw estimates for A and B solutions measured relative to SRMT (Figure 3.12). The median values of the radiation pattern coefficient jP are generally high, close to the value found by Helffrich (1997), and decrease only slightly for shorter periods. There is no significant change for the mean depth difference; all means are m < 3 km with standard deviations s < 15 km. Magnitude Mw differences for quality A and B solutions are close to 0.0 unit with standard deviation s < 0.1.

From our tests, we deduce that the 50 - 100 s period range is the best average range with the quickly available locations for the entire European-Mediterranean region. It allows routine MT analysis for earthquakes down to Mw  ~~ 4.5.

3.5.3 Current improvements

Two factors limit automatic retrieval of well-resolved MT for Mw > 4.5 earthquakes: location accuracy (Figure 3.12), usually lower for smaller events (Mw < 5.0) and the limited, inhomogeneous distribution of near real-time accessible stations (Figure 3.2).

Location accuracy is more important for shorter period analysis. Accurate quick locations are starting to become routinely available from the European-Mediterranean Seismological Center (EMSC). The EMSC merges automatic arrival-time picks from several European institutions, and the virtual network’s improved station coverage and aperture is capable of accurate automatic locations (Bossu et al.2002). The locations we have received since June 2002 are of good quality.

The geographical distribution of near real-time accessible broadband stations is limited and inhomogeneous (Figure 3.2). In the context of the European MEREDIAN project (van Eck et al.2001), the ORFEUS data center now has near real-time access to an increased (and still growing) number of broadband stations. Their distribution (white triangles, Figure 3.2) partially fills gaps like the western and central Mediterranean Sea, Turkey and eastern Europe. Northern Africa, however, lacks sufficient broadband instruments. For events since mid-2002, these data are available to us. Preliminary scanning of the results shows that the smaller epicentral distances overall and improved azimuthal coverage already increased the number of quality A relative to quality B and C solutions.

Better quick locations and a denser network (and the resulting reduced epicentral distances) also may speed up analysis and lower magnitude thresholds in the future. For closer epicentral distances, 15-minute seismograms contain the entire surface wave train. Reducing the seismogram length by 50% and assuming faster data accessibility, automatic solutions could then be obtained within 45 - 60 min after an event. At closer epicentral distances, signal strength is larger and relatively less perturbed by crustal and upper mantle heterogeneities. Combined with analysis at shorter periods, we would expect more reliable source parameter estimates for smaller events than possible now.

Although improved EMSC locations and additional data from ORFEUS started becoming available in mid-2002, we did not include these more recent events here. We wanted to have constant conditions - location and data access - for the entire period covered in this paper and therefore did not mix the two intervals.

3.6 Conclusions and outlook

We presented a fast and fully automatic procedure for moment tensor retrieval of moderate to strong (Mw > 4.5) earthquakes in the European-Mediterranean region using long-period (50 - 100 s) regional (D < 20o) seismograms. Automatic solutions are currently available within 90 minutes after an event. From April 2000 to April 2002, we obtained 87 moment tensor solutions that we grouped into three qualities based on main stress axes’ orientation, depth and Mw similarity with an independent, high quality moment tensor catalog. For 38 quality A solutions, magnitude, depth and focal mechanisms are well-resolved; 21 B solutions have well-resolved magnitude; 28 C solutions are not reliable. We derived simple empirical rules based on the number of stations and components used to predict the quality of a solution without a seismologist’s interference. Automatic quality Aa MTs are disseminated to EMSC (http://emsc-csem.org), quality Aa and Ba solutions are displayed on our web page (http://www.seismo.ethz.ch/mt), Aa and Ba solutions can be obtained via e-mail (contact first author for details).

In the near future, we foresee largely improved near real-time automatic waveform analysis capabilities and successful routine MT applications to smaller earthquakes (Mw  ~~ 4.0) for most of the European-Mediterranean region. Many national networks are transitioning to BB stations with near real-time data transmission. Connecting these national data centers via internet will create a dense European-wide virtual network that could be used, for near real-time event detection, size determination and MT inversion in the European-Mediterranean region. We can realize this scenario by increasing both the number of BB stations, particularly where no or few stations currently exist, and near real-time open access to these data for the scientific community.

Acknowledgments

We thank Daniel Stich and an anonymous reviewer for their constructive comments. We thank Suzan van der Lee and Yuan Gao for providing modified moment tensor analysis codes. We appreciate discussions with Günter Bock about the topic, we will miss him. We thank Kathleen J. Jackson for enhancing the manuscript by correcting our grammar. Most figures were produced with GMT (Wessel & Smith1995). High quality near real-time broadband data were obtained from the following institutions and networks: GEOFON, Potsdam (Germany); Gräfenberg Seismological Observatory, Erlangen (Germany); Institute of Physics of the Earth, Masaryk University, Brno (Czech Republic); MEDNET, Istituto Nazionale di Geofisica e Vulcanologia, Roma (Italy); National Data Center, Soreq (Israel); Zentralanstalt für Meteorologie und Geodynamik, Wien (Austria); and the additional members of the MEREDIAN consortium and its lead agency ORFEUS, De Bilt (The Netherlands). Quick earthquake locations are send to us by EMSC (France); Geol. Survey B-W, Freiburg (Germany); GERESS array (Germany); GSR, Obninsk (Russia); IGN, Madrid (Spain); INGV, Rome (Italy); IPRG, Tel Aviv (Israel); LDG, Paris (France); NEIC (USGS); NIEP, Bucharest (Romania); SDAC, Hannover (Germany); (see http://www.seismo.ethz.ch/redpuma).

Chapter 4
Automatic moment tensor analysis of moderate earthquakes

4.1 Summary

I present an improved automatic Frequency Adaptive Regional Moment Tensor (FARMT) routine. FARMT computes moment tensors (MT) of events with magnitude down to Mw  ~~ 4.3 in the European - Mediterranean area using complete surface wave seismograms recorded at regional distances (D < 20o). Two features lead to significant improvements of FARMT compared to the original standard routine (SR) for 4.3 < Mw < 4.7 events. First, the short-period cut-off at each station depends on station-event distance. Second, FARMT performs a signal-to-noise analysis to remove noisy traces and period-bands that do not exceed a minimum signal-to-noise ratio Rm. Tests performed to asses the short period cut-off and Rm for automatic FARMT indicate that the short period cut-off for events in the eastern Mediterranean Sea and the near East is 50 s. Events in Europe can be analyzed with a distance-variable short period cut-off ranging from 35 s for the closest to 50 s for distant stations. Values of Rm = 1 - 5 lead to well resolved mechanisms and use more seismograms for inversion than the standard method. The automatic FARMT uses Rm = 2, since it assures usage of seismograms with weak, often nodal signal, and reduces the number of iterations by about 20% to reach the final MT. For the May 2002 to September 2003 data-set, the automatic FARMT resulted in 66 solutions with well determined MT, which is a 20% increase relative to SR.

4.2 Introduction

Since April 2000, the Swiss Seismological Service (SED) computes real-time and fully automatic moment tensor (MT) solutions for Mw > 4.5 earthquakes in the European-Mediterranean area (Bernardi et al.2004). Moment tensors are computed using 50 - 100 s period surface waves recorded at regional distances (D < 20o). For stronger events (M w > 5.5) the complete moment tensor (seismic moment Mo, depth and focal mechanism) is generally well resolved. For 4.5 < Mw < 5.5 events, Mo is generally well resolved while depth and focal mechanism resolution decreases with decreasing event size. Goal of this work is to improve the results for 4.5 < Mw < 5.5 events and to extend automatic analysis to even smaller events.

Fast and reliable MT analysis for moderate events in densely populated areas like central Europe is becoming a necessity for the seismological community to provide rapid information to the public. An example is the recent Mw = 4.5 February 23, 2004 Rigney, eastern France, event, that was felt over the entire western part of Switzerland. Within minutes the press and a large number of concerned people called the SED or connected to the SED’s web-page looking for information about the event. In addition to rapid information, efficient automatic MT analysis for Mw >~126' 4.5 events will relatively quickly lead to a MT catalog, even in areas with low seismicity, like most of the European - Mediterranean region that is indispensable for seismic hazard and seismotectonic studies.

The current routine - called standard routine (SR) here - satisfactorily analyzes only few events with Mw <~~ 4.8, because analysis is performed at a fixed period band of 50 - 100 s. The 50 - 100 s interval is the best single period interval for automatic MT computation in European and Mediterranean region (Bernardi et al.2004), considering the sparse set of near-real time accessible stations, the limited accuracy of quick locations, and the complicated tectonic setting.

However, since 2002 the number of near-real time data has significantly increased in the European - Mediterranean region. Many of these stations are part of the Virtual European Broadband Seismograph Network (van Eck et al.2004) and data can be collected in near-real time from the ORFEUS data-center. Accuracy of quick automatic locations for M > 4.5 events has significantly increased in the European - Mediterranean region because more stations contribute data resulting in a relatively dense, large aperture European network. The quick, more accurate locations, provided by EMSC (Maset-Roux & Bossu2004) and by ORFEUS (van Eck et al.2004), together with the VEBSN data are key factors for automatic MT analysis of moderate events.

In this chapter, I present a modified procedure, which significantly lowers the magnitude threshold for automatic MT analysis toward Mw  ~~ 4.3. For such small events, signal energy above 50 s is generally weak and MT analysis requires inclusion of shorter periods. Although more near-real time broad-band stations exist, improving azimuthal coverage and lowering event-station distances, the distribution is still not homogeneous resulting in event data-sets that include near and far stations. Thus, I derived a distance dependent short period cut-off, which allows an optimal period band setting for MT inversion using near and far stations simultaneously (Section 4.3). An additional short and/or long period cut-off depending on the signal-to-noise ratio is applied to remove those periods without sufficient signal (Section 4.4). The notation from chapter 3 is kept, quality A solutions have well resolved Mo, depth and focal mechanism, quality B only well resolved Mo, and quality C is not reliable. Quality is assessed relative to SRMT (Braunmiller et al.2002).

4.3 Distance dependent short period cut-off


PIC

Figure 4.1: Map of the 21 calibration events (indicated by the fault plane solutions) used in Section 4.3 and listed in Table 4.1. Open dots are broad-band stations used.

The standard routine SR uses a short period cut-off at 50 s for all event-station distances, ensuring synthetics and observed seismograms are in phase even for large distances (D  ~~ 20o). At shorter distances, PREM synthetics can probably be used at shorter periods (Braunmiller et al.2002). The test data-set consists of 20 events covering the entire European - Mediterranean region . The data-set includes crustal and intermediate deep events and covers the magnitude range 4.5 < Mw < 6.7. The automatic SR resulted in quality A solutions for each event, implying that a sufficient number of seismograms (from 39 to 115) could be used for inversion in the 50 - 100 s pass band. The automatic moment tensors were then revised using the PDE-location, to exclude low quality waveform fit because of inaccurate locations. These solutions are used as reference (Table 4.1 and Figure 4.1).



Table 4.1: Reference events used to calibrate the distance dependent short period cut-off. For each earthquake the event origin time and coordinates from PDE, and depth in [km], Mw, and MT derived fault plane orientations are listed .
Event














Lat.Long. z Mw
Plane 1
Plane 1




























2000040600:0545.7226.581255.0534331 2922775 117
2000060602:3940.6932.99 18 6.0635650 -4912254-128
2000061301:4135.1527.13 25 5.3622875 413785 165
2001052417:3145.7526.461505.2035538 4822462 118
2002020307:0838.4931.31 18 6.6024326-114 9066 -78
2002090601:1838.3913.71 14 5.94 3546 6125350 116
2002103110:3041.7815.01 18 5.7926564-17717487 -25
2002110115:0641.7614.85 18 5.7626158-17516986 -31
2003022220:3848.35 6.80 18 4.7816240 -6731452-107
2003032917:3943.1515.51 10 5.5230734 124 8862 69
2003041000:3738.2226.94 14 5.7524367 17733387 23
2003041109:2444.82 9.09 18 4.8729672-16720378 -17
2003052118:4037.04 3.74 10 6.74 6335 4829063 115
2003052902:1236.82 3.36 14 4.9818880 5 9784 170
2003060115:4241.6614.83 25 4.5425083-17816088 -6
2003070619:0740.4526.07 18 5.8135057 725683 147
2003071301:4538.2739.03 25 5.6634371-17125081 -19
2003072608:3338.0329.05 18 5.4831954 -14 5878-143
2003081405:1138.7020.67 25 6.2419472-17710488 -17
2003081416:1438.8620.44 18 5.47 031 10217559 82
2003091421:3944.3911.51 25 5.2924733 77 8257 98















The test consists of observing the variance increase of each seismogram component relative to its standard solution variance for different short-period cut-offs. In practice, I recomputed the MT for each reference event changing the short-period cut-off only for one station per inversion run while keeping the 50 - 100 s band for all other stations. Re-computation was used to check whether shorter period synthetics are correctly realigned. With this approach, the MT solution remains effectively unchanged, because one station does not influence the results based on a large data-set. I tested short-period cut-offs of 30, 35, 40 and 45 seconds. The long period cut-off was kept at 100 s. I did not test shorter periods, used often for manual regional MT inversion (Braunmiller et al.2002Stich et al.2003), because of the sparsity of close-by stations and the requirements of the procedure to produce stable results in a fully automatic fashion.

A waveform fit example is shown in Figure 4.2. The waveforms are the vertical components of stations ISP (D = 2.80) and DIX (D = 16.40) for the M w = 5.8 April 10, 2003 Turkey event. For station ISP, the waveform fit remains good even for short period cut-offs of T = 30 - 35 s. For DIX, the fit decreases significantly for periods T < 40 s. The waveform fit at a given short period cut-off depends on epicentral distance. This general behavior is caused by limitations imposed by the PREM model which is used throughout the study area for synthetic seismogram calculation.


PIC

Figure 4.2: Waveform fit for five period bands for the vertical components of stations ISP (D = 2.80) and DIX (D = 16.40) for the M w = 5.8 April 10, 2003 Turkey event. Solid lines are observed data, dashed lines synthetics. Numbers on left of each fit are the normalized variance (top) and realignment time shift (bottom), respectively.


PIC

Figure 4.3: Difference between normalized variance for each component (open dots) obtained with the test surface period bands relative to the 50 - 100 s fit plotted with respect to epicentral distance. Thick lines are mean variance increase, dashed lines are the respective standard deviations.

Figure 4.3 shows the normalized variance difference for each short period cut-off as a function of epicentral distance for all components (at least 441 observations per component) relative to the 50-100 fit. The mean (solid line) and standard deviation (dashed lines) are moving window values with 1o steps using all data-points in the surrounding ±1o distance. Figure 4.3 shows that the variance difference increases for longer epicentral distances (from left to right in each panel) and for shorter period cut-offs (from bottom to top in each column). This effect is slightly stronger for horizontal components, probably because of the higher noise levels and possibly anisotropy. The components are not rotated to transversal/radial and polarization anisotropy observed in the Mediterranean region (Marone et al.2004) affects both horizontal components.

The variance difference increases smoothly over the D = 20 - 200 distance interval. I arbitrarily fixed the distance interval where the mean normalized variance difference does not decrease more than -0.1 units for one of the three components. Based on Figure 4.2, even larger variance increases appear to provide adequate fits (and thus MT solutions), however, the small 0.1 value ensures stable automatic results. All components have the same distance range for the sake of simplicity. Based on this definition, the 30 s short period cut-off could be applied only for distances D < 30, were only few data are available. I thus selected 35 seconds as the shortest cut-off period. Table 4.2 lists the resulting distance - short period cut-off setting (T2). Table 4.2 also shows the standard 50 - 100 s pass-band for all distances (T0), and two additional settings, which deviate slightly from T2: T1 is a more and T3 is a less conservative setting. For full evaluation of the Ti cut-offs, I applied all four settings to a 17 month data-set in the European - Mediterranean area (section 4.5) combined with the signal-to-noise ratio dependent period band cut-off derived in the next section.



Table 4.2: Summary of the period band - distance sets. T0 refers to the standard routine that uses a 50 s short period cut-off for all distances. T2 is derived from Figure 4.3, while T1 and T3 are derived from T2.
period ranges - epicentral distances





Sets





35 - 100 s 40 - 100 s 45 - 100 s 50 - 100 s





T0 D = 2o - 20o
T1 D = 2o - 5oD = 5o - 10oD = 10o - 15oD = 15o - 20o
T2 D = 2o - 6oD = 6o - 12oD = 12o - 17oD = 17o - 20o
T3 D = 2o - 7oD = 7o - 14oD = 14o - 19oD = 19o - 20o






4.4 Signal-to-noise ratio cut-off

The standard procedure uses a 100 s long period cut-off to compute the MT. When the seismogram contains only noise or when the signal is above the noise only for periods shorter than 100 seconds, the normalized variance of the station component is high. In the SR such traces are removed iteratively (Bernardi et al.2004).

Goal of this section is to implement a more straightforward approach to remove noisy traces. With the following signal-to-noise ratio analysis the period band where the ratio exceeds a value Rm are identified and selected for MT inversion.


PIC

Figure 4.4: Example of a long period high (station MMLI, NS-component) and a low noise (station MOA, Z-component) seismogram (left). Each noise time series is divided into 14 segments of 200 s length each, with 100 s overlap. The respective power spectral densities are shown on the right (note the different scale).

This approach has two main consequences. First, when the epicentral distance criteria, for example, indicates that the 50 - 100 s period range should be used for MT inversion and if the signal-to-noise ratio of that trace does not exceed Rm for the entire range, the seismogram will be removed from the data-set. This reduces the number of iterations and shortens the computing time. Second, using only the period range where the data exceed Rm, for instance from 50 s to 70 s, lowers the variance and the seismogram is used for inversion. The resulting data-set for automatic MT inversion contains more traces than the SR, which is important for moderate events analysis, where generally only few traces with good signal-to-noise ratio exist.

The signal-to-noise ratio is determined by simple division of signal and noise power spectral density for each sampled frequency. All seismograms have 1 second sampling interval and are zero-padded to 2048 points. A 9-point smoothing is applied to the resulting power spectral density curve.

The noise is characterized by non-stationary behavior with large variations between the low and high noise levels for 10 - 100 s period waves (Peterson1993). Examples of noise variations are shown in Figure 4.4, using a typical long period high noise (station MMLI, NS-component) and low noise seismogram (station MOA, Z-component). The seismograms are divided into fourteen 200 second long sub-windows, and for each window the power spectral density (dotted lines on the right side, Figure 4.4) is computed. In both cases, the power spectral density varies by about one order of size over the entire 10 - 100 s band within the short 1500 s noise window.

The noise power spectral density PN is computed from the 200 s window before the P-wave arrival to have a representative pre-signal noise estimate. A 200 s window, of course, does not allow to solve single frequency features between 0.01 and 0.03 Hz. The goal of this analysis is not to obtain a high single frequency resolution of the noise, where a very long time series is needed, but to obtain an overall estimate of the noise level over a finite period band. A short 6 point hanning taper is applied to each seismogram to remove edge effects.


PIC

Figure 4.5: Example of signal-to-noise ratio analysis. Right: signal Ps (thick line) and noise Pn (thin line) power spectra and the ratio R = Ps/Pn (dotted line). Pn is computed using data-points between time 107 - 307 s, Ps between 308 - 358 s of the raw data. The seismogram is the NS-component of station OBKA for the Mw = 4.4 May 10, 2003 Northwestern Balkan region event (D = 3.4o). Left (from top to bottom): raw data, data filtered between 35 and 60 seconds, and between 60 and 100 seconds.

The signal power spectral density PS is computed from a window starting at the P-wave arrival and ending with the 30 second group velocity surface waves. Since the earthquake alerts that trigger the automatic procedure may not be correctly located (Bernardi et al.2004), PS analysis requires windows wide enough to contain the signal even when severe mislocation occurred. The alert that triggers the procedure is not necessarily the location used for later inversion, but for rapid data preparation I use the initial location for signal-to-noise-analysis. Thus, I assume a maximum uncertainty of 200 km. P-wave (Kennett & Engdahl1991) and surface wave group velocities (Larson & Ekström2001Pasyanos et al.2001Cotte & Laske2002) may also vary significantly within the European-Mediterranean area. I used conservative fast V p = 8.5 km/s and slow 30 s group velocity V 30 = 2.5 km/s, to avoid truncation of the signal. At each end of the signal window, I add 5% to apply a 5% hanning taper to the seismogram without cutting into the data. The window length for PS is long for far stations because of uncertainties. The signal length, that actually contributes mostly to the power spectral value is shorter. I thus normalize PS for a 200 s window.

To find the minimum signal-to-noise ratio required, I recomputed the MT of 14 4.4 < Mw < 5.0 events (Table 4.4). I also re-analyzed the Mw = 7.0 May, 2003 northern Algeria event to verify that the signal-to-noise ratio analysis does not affect well resolved MTs of large events. All 15 test events have accurate automatic location compared to their PDE-location; 12 are A-quality and 3 are B-quality solutions, using the SR. The actual period band used for inversion is defined by the minimum signal-to-noise ratio Rm(f) = PS/PN. Only seismograms that exceed Rm for a period band at least 15 s between 50 - 100 s are selected.

Tested values of Rm are 1, 1.5, 2, 2.5, 3, 4, 5, 10, 20, 50 and 100. Figure 4.5, as an example, shows the NS-component of station OBKA (D = 3.4o) for the May 10, 2003 Northwestern Balkan Mw = 4.4 event. PS (thick black line), PN (thin black line) and the resulting signal-to-noise ratio R (dotted line) are plotted. In the time domain, the signal is visible in the 35 - 60 filtered seismogram (right side, middle), while at longer periods (right side, bottom), only noise is visible. The period-band for MT inversion for this component is 50 - 80 seconds when Rm = 1 or 50 - 57 seconds when Rm = 2. In the second case the component is not used because the window width of usable signal length is less then 15 s wide.


PIC

Figure 4.6: Figure summarizing the parameters evaluated to select Rm. Values are average obtained comparing the MT solutions of 15 test earthquakes using the signal-to-noise ratio criteria with the standard solutions. Top (from left to right): DAx in [o], DM w in units and Dz in [km](continous lines) and their standard deviations (dotted lines). Bottom (from left to right): number of used components, iterations per MT analysis and normalized variance (all in %) relative to the standard solution.

To select the ”best” value of Rm, I analyzed the solution stability, the number of seismograms used, the solution fit and the reduction of iterations as a function of Rm relative to the standard solution. Figure 4.6 summarize the average performance for the 15 events. Stability of the solution is manifested by small changes in the principal axes orientation DAx (Bernardi et al.2004), moment magnitude DMw and depth Dz. For Rm < 5, DAx and DMw are small indicating stable solutions. Depth resolution does not seem to depend strongly on Rm. For Rm > 5 the number of components used (bottom, left) is less than for the standard solution; one goal of the signal-to-noise analysis is to use more data to be able to analyze smaller event, so Rm > 5 is not acceptable. Using fewer data may explain the increase of DAx and DMw for Rm > 5. The number of iterations and the normalized variance decrease steadily. Because of the fewer components used, the resulting solutions are actually less well resolved. Using large Rm values (Rm = 20 - 100), reduces the amount of data used significantly. For Rm = 100 only 12 of 15 events still have any traces (i.e. for 3 events no traces are left for inversion).

Based on Figure 4.6, the minimum signal-to-noise ratio can be set to values of 1 < Rm < 5. I selected Rm = 2 for further analysis. With Rm = 2, the solutions are stable, more data ( ~~ 20%) are used relative to the standard solution, fewer iterations ( ~~ 20%) are required and the variance is significantly reduced ( ~~ 20%). The relatively small value of Rm = 2 ensures to use small signals from nodal, far and relatively small events. Data that are still noisy and cannot be fit are automatically removed by the variance criteria during the inversion (chapter 3).

4.5 Application to MT data-set

Between May 2002 and September 2003, the SR was triggered 246 times, 55 resulted in quality A, 44 in quality B and 34 in quality C solutions (Table 4.4). 113 alerts did not result in a MT because the earthquake was too weak, or the earthquake occurred outside our area of interest or no data were available. For the same data-set, the MTs are recomputed using the distance dependent short period cut-offs Ti (section 4.3, Ti = T0 - T3) and signal-to-noise ratio Rm = 2 (section 4.4). Results are listed in Table 4.4. No event with Mw < 4.2 resulted in an A or B quality solution, which is to be expected from the period bands Ti considered. After removing these triggers our data-set consists of 169 events. As for SR, not all events resulted in MT solution (maximum number is 139, Table 4.3).



Table 4.3: Table of the 139 automatic moment tensor solutions. From left to right: Event number, PDE-location, true and assigned quality, seismic moment Mo in Nm, Mw, depth in km, orientation of the two fault planes in degree, number of stations, components and quickly available location used for inversion. For completeness, we include all source parameters even for true quality B and C solutions. For B, only size is well resolved and for C, none of the parameters is reliably resolved. For further studies, use Mw, depth and focal mechanism from A and only Mw from B solutions. Events located in eastern Mediterranean Sea and near East are computed using the T0-setting, all others using the T1-setting. Events marked with * are used to set Rm (Section 4.4).
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time LatLongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























1 02-05-0308:13:10.385.97231.14 C c 5.23e+165.08 125 135 28 151 250 76 64 2 2 54.00 24.90 15:38:27.1
2 02-05-1611:00:12.029.67251.70 C c 1e+16 4.60 200 331 78 -168 238 78 -11 2 2 28.80 51.20 11:01:01
3 02-05-1715:52:20.929.58951.96 C c 1.15e+175.31 75 196 62 177 288 88 27 2 3 28.20 58.50 15:51:22.1
4 02-05-2120:53:29.736.62724.27 A a 8.02e+175.87 75 351 79 8 259 81 168 11 30 36.70 24.50 20:53:29.0
5 02-05-2420:42:26.844.76121.61 C a 2.1e+16 4.82 10 76 23 -132 301 72 -73 12 25 46.20 22.40 20:42:46.9
6 02-06-0116:12:36.929.56851.22 C c 1.54e+175.39 75 181 38 -165 80 81 -52 1 1 26.60 58.20 16:11:19.8
7 02-06-0622:35:41.035.64826.18 B a 6.38e+165.14 31 127 28 -18 234 81 116 24 57 35.70 26.50 22:35:36.6
8 02-06-0820:13:06.844.36010.69 A a 2.63e+154.22 31 282 33 80 113 57 96 10 18 44.50 10.80 20:13:00
9 02-06-1803:19:24.333.32645.91 B c 1.9e+16 4.79 75 9 59 -31 116 63 145 3 4 36.00 43.00 03:20:03
10 02-06-1822:23:37.744.43810.81 A c 3.56e+154.30 31 109 44 81 301 45 98 2 3 44.60 10.80 22:23:43.0
11 02-06-2202:58:21.335.62649.04 A b 7.78e+186.53 10 262 48 48 135 55 126 4 12 36.30 48.50 02:58:30.1
12 02-06-2206:45:34.635.66648.92 B c 7.25e+165.18 42 237 34 58 94 60 110 3 4 32.30 50.40 06:45:06
13 02-06-2214:27:16.035.49748.97 B c 9.04e+154.57 42 276 65 -176 184 86 -24 2 2 35.50 49.10 14:27:16.0
14 02-06-2401:20:36.735.850 9.88 A a 5.9e+16 5.12 10 352 39 73 193 52 103 16 34 35.80 9.90 01:20:35.0
15 02-07-1606:56:25.533.88324.15 A b 1.87e+164.79 31 150 41 50 18 59 119 7 15 33.00 24.00 06:56:26.0
16 02-07-2205:45:05.450.875 6.22 A a 1.11e+164.63 25 287 43 -106 128 48 -75 7 16 51.20 5.80 05:44:59.2
17 02-07-2817:16:31.137.93220.69 B a 5.55e+165.10 18 214 68 179 304 89 21 31 72 38.00 20.62 17:16:30.3
18*02-07-3012:20:23.637.69829.18 A a 2.22e+164.83 31 126 27 -75 289 63 -97 14 17 38.00 28.90 12:20:28.0
19 02-08-0606:16:18.937.980-1.89 B a 5.93e+165.12 10 183 24 -25 296 79 112 11 27 38.00358.10 06:16:21.3
20 02-08-1811:52:28.840.80842.57 B b 1.52e+164.72 18 77 48 122 213 50 58 9 10 41.10 40.60 11:52:42.4
21 02-09-0201:00:03.335.70148.83 A b 8.8e+16 5.23 14 267 47 45 143 58 127 5 9 36.80 47.10 01:01:20.1
22*02-09-0209:23:45.235.08126.52 A b 1.23e+164.66 31 22 29 51 244 67 109 8 10 35.02 26.04 09:23:49.1
23 02-09-0522:19:51.538.72024.53 A a 3.47e+164.96 25 252 51 -178 161 89 -38 38 89 38.80 24.50 22:19:50.0
24 02-09-0601:21:28.638.38113.70 A a 8.5e+17 5.89 10 30 48 56 255 50 122 4111838.39 13.71 01:21:27.9
25 02-09-0601:45:30.338.43513.73 C c 8.25e+165.21 25 107 35 -170 9 84 -55 6 6 38.60 13.60 01:45:34.0


























Table 4.3: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time Lat LongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























26 02-09-0816:14:31.834.714 23.34 B a 1.43e+164.71 18 151 47 -27 261 69 133 20 28 35.00 22.00 16:14:41.0
27 02-09-0907:56:51.029.447 51.29 C c 1.76e+186.10 125 223 27 37 99 73 112 1 2 29.40 51.30 07:56:53.0
28 02-09-1002:32:51.338.466 13.70 B c 3.91e+154.33 10 126 18 -174 31 88 -71 5 5 41.80 14.00 02:32:02
29 02-09-1419:50:16.837.810 21.06 B a 6.55e+165.15 25 101 20 31 341 79 108 17 39 36.00 26.00 19:49:35.0
30 02-09-1618:48:26.766.938-18.45 C a 5.06e+175.74 10 39 70 18 303 72 159 4 12 66.80341.50 18:48:26.5
31 02-09-1909:19:42.837.810 21.12 B a 1.86e+164.78 125 317 63 -10 52 80 153 25 47 36.70 23.10 09:19:16.3
32 02-09-2023:06:03.838.455 13.74 C c 1e+17 5.27 125 107 69 164 202 75 20 5 6 34.00 15.00 23:03:12.0
33 02-09-2223:53:14.652.520 -2.15 B a 6.69e+154.49 42 194 87 0 104 89 177 11 16 52.71358.18 23:53:11.2
34 02-09-2522:28:11.931.995 49.32 A b 3.76e+175.65 14 303 40 68 150 53 107 9 26 32.14 49.35 22:28:02
35 02-09-2620:44:07.236.667 28.02 C c 1.12e+164.63 175 289 60 -171 195 82 -30 3 3 34.00 31.00 20:43:36.0
36 02-09-2706:10:44.938.437 13.69 B a 8.29e+165.22 25 135 12 -77 302 77 -92 10 18 38.32 13.70 06:10:06
37 02-09-3006:44:48.047.832 -3.20 C c 1.37e+164.69 31 211 34 59 67 60 109 4 4 48.40 3.30 06:45:48.0
38 02-10-0222:57:25.938.456 13.72 C b 6.27e+165.13 25 138 1 -127 355 88 -89 9 10 32.00 16.00 22:56:10.0
39 02-10-1205:58:50.134.772 26.37 B b 8.47e+165.22 10 195 12 -56 341 79 -97 8 13 34.79 26.65 05:58:05
40 02-10-1700:53:02.139.443 40.27 B a 3.26e+164.95 10 103 10 -140 334 83 -82 11 20 42.00 40.00 00:53:30.0
41 02-10-1915:57:13.939.039 54.91 B b 3.9e+16 5.00 42 301 61 -172 207 83 -29 8 12 39.00 54.90 15:57:13.0
42 02-10-2215:52:13.039.423 40.17 A a 1.76e+164.77 14 315 61 177 46 87 28 12 23 39.40 40.10 15:52:16.0
43 02-10-2311:01:28.742.612 17.15 B b 7.21e+154.51 25 227 11 -7 324 88 101 6 9 43.40 17.30 11:01:33.1
44 02-10-2702:50:26.237.795 15.16 B a 2.61e+164.88 10 247 48 26 139 70 135 30 55 37.70 15.20 02:50:26.0
45 02-10-3110:32:58.841.789 14.87 A a 7.14e+175.84 25 172 77 -5 263 84 167 37 92 41.78 15.01 10:32:58.0
46 02-11-0115:09:00.841.726 14.87 B a 6.19e+175.80 18 262 53 -175 169 86 -36 32 88 41.76 14.85 15:05:05
47*02-11-0609:12:46.237.950 20.68 A a 3.9e+16 5.00 25 311 58 17 211 74 147 24 50 37.99 20.71 09:12:46.1
48 02-11-0902:18:11.545.003 37.76 A a 6.2e+16 5.13 14 162 11 95 336 78 88 12 20 45.37 37.81 02:18:12.4
49 02-11-1310:48:03.645.606 10.16 B c 4.82e+154.39 125 20 76 -178 290 88 -13 4 5 45.60 10.10 10:48:00
50 02-11-1901:25:36.438.022 38.43 B b 2.28e+164.84 31 101 33 33 342 72 119 7 11 38.70 38.70 01:25:38.9
51*02-11-3008:15:46.845.726 26.56 B a 1.4e+16 4.70 125 56 32 20 308 79 121 22 34 46.20 27.00 08:15:50.3
52 02-12-0204:58:55.237.747 21.08 A a 3.01e+175.59 18 310 49 -8 46 83 139 34 81 38.05 20.97 04:59:04
53 02-12-0907:50:58.936.628 45.18 A c 1.31e+164.68 25 34 28 -69 190 63 100 3 3 36.60 45.30 07:50:58.0
54 02-12-0909:35:04.937.869 19.97 A a 1.62e+164.74 75 60 70 -174 328 84 -19 19 34 38.00 20.20 09:35:09
55 02-12-1013:51:26.736.195 -7.47 A c 5.05e+165.07 25 226 40 87 50 49 92 3 4 36.10352.40 13:51:29.7
56 02-12-1401:02:43.537.524 36.20 C c 4.51e+165.04 125 290 24 85 115 65 92 2 3 38.00 37.00 01:01:50.0
57 02-12-2304:02:36.139.027 20.22 C c 1.3e+16 4.68 150 17 58 13 280 78 148 2 3 37.40 24.30 04:04:45.3


























Table 4.3: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time Lat LongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























58 02-12-2417:03:02.934.594 47.45 C c 2.44e+175.53 10 233 11 -142 106 83 -81 5 6 32.80 52.50 17:07:22.3
59*02-12-2422:17:14.434.572 47.41 A c 1.07e+164.62 18 288 39 71 132 52 105 2 3 34.40 47.40 22:17:13.0
60 03-01-0100:55:55.936.132 2.91 C a 1.67e+175.42 200 291 69 14 196 76 158 13 14 36.40 3.30 00:56:00
61 03-01-1117:45:30.629.590 51.47 B b 1.2e+17 5.32 31 182 46 -178 91 89 -43 4 8 29.60 51.40 17:45:29.0
62 03-01-1214:38:46.934.836 46.11 C c 6.64e+175.82 200 59 61 29 314 64 147 2 2 34.50 46.10 14:38:44.0
63*03-01-2619:57:03.243.883 11.96 A a 1.71e+164.76 10 327 23 -106 165 67 -82 22 42 43.90 11.90 19:57:09
64 03-01-2620:15:03.143.875 11.95 B a 1.05e+164.62 10 336 23 -43 107 74 107 28 50 43.80 11.90 20:15:00
65 03-01-2705:26:23.039.500 39.87 A a 2.23e+186.17 25 335 72 -177 244 87 -17 14 36 39.50 39.80 05:26:23.1
66 03-02-1813:09:36.135.836 -3.48 B a 3.73e+164.98 10 49 40 -123 270 56 -64 13 23 37.20358.80 13:03:57.4
67*03-02-2220:41:03.448.342 6.57 A a 1.89e+164.79 18 176 49 -49 303 54 127 26 58 48.35 6.80 20:41:00
68*03-02-2420:22:24.739.371 20.47 A a 8.14e+154.54 31 190 30 109 348 61 79 19 21 39.50 20.60 20:22:22.0
69 03-03-0104:06:02.134.715 23.93 B c 7.79e+154.53 55 185 44 -177 93 88 -45 2 4 36.00 21.00 04:04:30.0
70 03-03-2716:10:36.443.150 15.33 A a 3e+16 4.92 18 292 29 98 102 61 85 33 74 43.15 15.44 16:10:36.8
71 03-03-2917:42:15.743.109 15.46 A a 2.8e+17 5.57 10 304 26 119 92 67 76 36 96 43.15 15.51 17:42:14.2
72 03-03-3000:56:24.443.170 15.43 C a 9.32e+154.58 150 249 24 49 112 71 106 7 14 42.90 16.00 00:56:27.0
73 03-03-3011:09:59.943.169 15.52 B c 2.27e+164.84 42 66 46 23 319 73 133 4 7 43.30 16.50 11:10:01
74 03-04-0515:10:23.844.861-28.09 C c 1.28e+175.34 200 192 29 8 94 85 119 1 1 45.00335.00 15:10:53.0
75 03-04-1000:40:15.138.221 26.95 A a 5.86e+175.78 18 241 51 -173 147 84 -38 3810438.32 26.94 00:40:15.3
76 03-04-1109:26:58.344.792 8.89 A a 2.87e+164.91 18 297 62 -166 201 78 -27 27 65 44.82 9.09 09:26:58.9
77 03-04-1123:53:47.637.580 22.67 C a 2.36e+164.85 25 317 34 94 132 55 87 13 17 36.80 24.50 23:53:22.3
78 03-04-1722:34:24.638.158 27.00 B a 1.2e+17 5.32 18 359 21 -10 98 86 110 34 78 38.19 26.24 22:34:31.7
79 03-04-1920:55:51.135.139 27.70 B b 2.29e+164.84 10 145 35 24 34 76 122 7 8 35.00 27.00 20:56:00
80 03-04-2901:51:20.236.830 21.72 A a 6.87e+165.16 42 211 53 -3 303 87 143 26 56 36.89 21.79 01:51:18.3
81*03-05-0100:27:04.739.007 40.46 A b 6.53e+186.48 25 64 83 -2 154 87 173 5 13 39.01 40.37 00:27:00
82 03-05-0311:22:40.636.884 31.53 A b 1.62e+175.41 125 357 43 145 114 66 52 6 15 36.86 31.50 11:22:40.6
83 03-05-0822:23:10.627.538 54.47 B c 5.88e+165.12 42 103 75 -177 12 87 -14 2 3 27.60 53.30 22:23:13.6
84 03-05-1006:42:49.943.911 16.98 A a 6.72e+154.49 31 129 74 -178 39 88 -15 19 39 44.00 17.80 06:42:57.0
85 03-05-1015:44:51.239.050 40.36 A b 2.26e+164.84 18 318 57 -179 228 89 -32 8 14 39.30 41.00 15:44:50.4
86*03-05-2118:44:20.136.964 3.63 A a 1.94e+196.79 10 91 27 115 243 65 77 30 87 37.04 3.74 18:44:20.6
87 03-05-2122:04:10.237.065 3.93 C c 1.11e+175.30 150 322 65 19 223 72 154 4 5 36.90 4.50 22:02:15.8
88 03-05-2123:23:43.936.925 3.49 A a 5.22e+165.08 10 70 12 96 243 77 88 19 25 37.10 3.60 23:23:44.0
89 03-05-2203:14:03.937.080 3.63 A a 3.34e+164.95 10 80 20 107 242 70 83 26 64 37.20 3.70 03:14:03


























Table 4.3: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time Lat LongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























90 03-05-2211:11:05.337.139 3.84 B a 4.79e+154.39 10 39 64 -12 135 78 154 13 24 36.90 3.90 11:11:01
91 03-05-2213:57:20.637.030 3.93 A a 4.28e+165.02 10 70 18 77 263 71 94 24 50 36.90 4.00 13:57:20.0
92 03-05-2415:51:42.837.096 3.76 B b 3.3e+15 4.28 55 319 36 -21 66 77 124 8 8 37.10 3.70 15:51:42.1
93 03-05-2419:21:24.837.040 3.96 A a 1.52e+164.72 10 119 28 149 237 76 64 25 57 37.00 4.00 19:21:25.9
94 03-05-2710:30:52.229.598 51.31 C c 5.47e+175.76 150 225 18 178 317 89 71 1 2 31.70 46.40 10:31:30.3
95 03-05-2717:11:28.836.939 3.57 A a 4.8e+17 5.72 10 50 22 56 266 71 102 32 93 37.05 3.86 17:11:35.8
96 03-05-2806:58:37.736.877 3.27 A a 1.83e+164.78 10 100 24 118 249 68 77 25 57 37.00 3.20 06:58:41.6
97 03-05-2902:14:59.836.817 3.36 A a 3.34e+164.95 10 187 80 9 96 81 170 26 65 36.70 3.50 02:15:02
98 03-05-2911:38:09.339.551 20.38 A b 6.37e+154.47 25 310 35 101 116 55 81 9 12 39.57 20.37 11:38:10.3
99 03-05-3010:47:10.034.378 26.22 A b 1.04e+164.62 42 226 55 17 125 75 143 6 9 34.00 26.00 10:47:13.0
100*03-06-0106:09:45.641.069 47.38 A b 4.01e+165.01 31 218 55 -26 323 68 142 5 9 41.00 47.80 06:06:45.3
101*03-06-0115:45:18.241.663 14.82 A a 8.89e+154.57 25 252 80 179 342 89 9 31 55 41.30 15.00 15:45:13.0
102 03-06-0907:06:39.339.893 22.30 A a 8.8e+16 5.23 25 296 36 -92 119 53 -87 21 49 39.93 22.33 07:07:40.3
103 03-06-0917:44:03.040.242 28.02 B a 1.88e+164.79 31 171 76 6 79 83 166 10 17 40.40 27.90 17:44:07
104 03-06-1608:27:50.037.695 19.72 A a 2.18e+164.83 10 358 18 102 165 71 85 16 25 37.20 19.20 08:27:48.2
105 03-06-2109:00:20.843.072 15.30 C a 8.94e+154.57 25 10 52 11 273 80 142 16 19 43.40 15.50 09:09:25.9
106 03-06-2223:46:20.439.031 28.03 A a 1.41e+164.70 25 94 50 -126 323 51 -53 14 21 39.70 28.20 23:46:29.8
107 03-06-2413:01:32.832.927 49.47 B c 1.88e+164.79 10 203 46 -36 320 64 130 1 2 36.50 43.10 13:03:27.6
108 03-06-2613:45:57.538.610 23.65 C c 4.93e+165.07 10 39 22 -67 195 69 -98 1 1 30.80 35.80 13:43:47.8
109 03-07-0619:10:28.240.445 26.02 A a 5.51e+175.76 14 349 64 5 257 84 154 3610340.45 26.07 19:10:26.2
110 03-07-0620:10:16.540.421 26.12 A a 1.61e+175.41 18 349 41 -7 84 85 131 37 97 40.66 25.83 20:10:18.5
111 03-07-0715:08:11.235.858 14.77 B a 7.38e+154.52 10 240 47 -21 345 74 135 13 16 37.00 14.30 15:05:20.2
112*03-07-0922:31:41.240.354 25.93 A a 2.32e+164.85 18 343 55 -14 81 78 144 41 88 40.58 25.71 22:31:42.6
113 03-07-1017:06:37.728.355 54.16 B c 1.67e+186.09 100 231 27 -14 334 83 116 3 7 28.47 54.05 17:07:47.2
114 03-07-1017:40:15.928.296 54.13 A b 6.44e+175.81 31 268 28 55 126 66 106 3 8 28.39 54.00 17:40:25.2
115 03-07-1123:55:44.428.472 54.03 B c 2.17e+164.83 31 54 25 93 231 64 88 3 5 28.40 54.20 23:55:47.9
116 03-07-1301:48:21.038.288 38.96 A a 3.66e+175.65 25 342 61 -171 248 82 -29 16 45 38.27 39.03 01:48:25.1
117 03-07-1913:12:39.862.038-26.73 C c 9.04e+165.24 125 81 11 -157 329 85 -79 2 6 62.00335.00 13:14:03
118 03-07-2304:56:02.038.172 28.85 B b 2.15e+175.49 42 73 43 -161 329 77 -48 6 13 36.40 34.80 04:55:13.1
119 03-07-2601:00:57.638.111 28.88 C c 1.3e+17 5.35 31 281 33 65 130 59 105 2 2 30.80 47.80 00:58:19.4
120 03-07-2608:36:49.238.019 28.92 B a 2.44e+175.53 18 236 31 -148 119 73 -62 30 75 38.09 28.84 08:36:50.3


























Table 4.3: (continued)
Location (PDE)
























Nr.
Quality
Mo MwDepth
Plane 1
Plane 2
St.Co.
Location used for inversion
























Date Time Lat LongTrueAssigned StrikeDipRakeStrikeDipRake Time Lat Long
























121 03-07-2613:31:36.238.115 28.92 C c 8.72e+165.23 25 111 29 46 338 69 110 4 5 36.80 37.10 13:30:32.6
122 03-07-2905:31:26.735.694-10.56 A a 8.39e+165.22 55 136 53 153 242 68 39 6 16 35.72349.37 05:31:30.8
123 03-08-0210:18:38.443.015 17.69 A a 1.6e+16 4.74 25 17 25 -179 286 89 -64 38 68 43.00 17.90 10:18:37.0
124 03-08-0403:08:35.065.989 5.47 A a 4.68e+165.05 18 27 41 74 228 50 103 31 81 66.00 7.40 03:03:35.7
125 03-08-1120:12:08.038.832 44.88 A b 3.98e+165.00 31 21 67 -8 115 81 157 6 11 39.50 44.90 20:12:00
126 03-08-1405:14:54.839.160 20.60 B a 3.07e+186.26 18 103 50 -4 196 86 140 36 81 39.24 20.49 05:14:55.7
127*03-08-1408:41:39.739.000 20.54 A a 2.83e+164.90 25 290 68 -4 22 85 157 21 34 39.17 20.44 08:41:42.2
128 03-08-1412:18:14.238.896 20.61 A a 7.87e+165.20 25 358 32 56 216 63 109 21 42 39.13 20.42 12:18:17.6
129 03-08-1416:18:02.538.826 20.57 A a 1.64e+175.41 14 2 37 91 180 52 89 33 75 38.86 20.44 16:18:06
130 03-08-1420:46:52.638.918 20.55 A a 1.63e+164.74 25 13 29 116 164 63 75 15 21 38.80 20.50 20:46:50.0
131 03-08-1614:49:13.538.570 20.54 A a 9.21e+154.58 25 32 64 170 127 81 25 18 19 39.10 20.30 14:49:16.6
132 03-08-2302:00:11.963.916-22.26 B c 2.56e+164.88 10 357 69 -168 263 79 -20 2 6 64.00339.00 02:02:20.0
133 03-08-2818:31:57.228.376 54.09 C c 1.03e+175.28 125 256 12 -24 10 84 101 1 1 28.10 54.15 18:32:08
134 03-08-2906:55:51.828.391 51.52 B c 2.04e+164.81 14 193 50 160 295 75 41 2 2 24.80 53.30 06:55:21.3
135 03-09-0523:30:11.134.588 26.18 B a 1.31e+164.68 75 181 68 162 277 73 22 14 20 34.00 27.30 23:29:50.9
136 03-09-1119:31:28.028.369 54.07 C c 4.57e+175.71 200 276 71 -167 182 78 -19 1 1 30.00 49.80 19:31:59.2
137 03-09-1313:46:14.336.629 26.91 B a 3.14e+164.93 42 168 51 -30 279 66 137 26 33 36.63 27.01 13:46:13.6
138 03-09-1421:42:51.944.329 11.45 A a 1.21e+175.33 25 249 27 83 76 62 93 19 51 44.39 11.51 21:42:53.3
139*03-09-2408:13:10.339.553 38.32 A a 1.92e+164.79 25 272 81 -4 3 85 171 11 21 39.50 38.30 08:13:10.0

























4.6 Discussion

The main objective of including signal-noise analyses and periods shorter than T = 50 s to the automatic MT inversion was to increase the number of reliable MT solutions. For the entire study area (chapter 3), T0, which uses the 50 s period cut-off results in the largest number of quality A solutions (63), a 15% increase relative to SR (Table 4.4). This improvement is probably due to a similar increase in the number of components used (last column Table 4.4). The period bands with smaller cut-off periods (T1-T3), overall, did not produce more quality A solutions than SR, although even more data were used than for T0. Using shorter cut-off periods, the number of quality B solutions increases resulting in a larger number of A+B solutions. This observation implies that the shorter-period cut-offs do not adequately match the phase relations between observed and synthetic data for the study area as a whole.



Table 4.4: Number of solutions with respect to epicentral area and quality class from SR and the different short period cut-off settings Ti of FARMT (Table 4.2). Only events with Mw > 4.3 (SRMT) are considered. For the two sub-areas, the number of quality A solutions is subdivided into all/ Mw < 4.7/Mw > 4.7. The last column gives the overall number of components used for all MTs. Top portion of table refers to solutions obtained with fast locations, bottom with PDE-locations. Bold refers to the T0-T1 combined set.











Conf.
All area
Europe
Aegean Sea
Nr. comp
Eastern Turkey











ABC A B C A B C











SR 554434 28/5/23 1618 27/7/20 2816 3523
T0 633936 30/5/25 132033/11/222616 4019
T1 56523133/8/251813 23/6/17 3418 4222
T2 565231 33/8/25 1912 23/8/15 3319 4243
T3 535630 32/9/23 2211 21/6/15 3420 4246






















T0 773821 34/9/25 1514 43/14/29 23 7 4218
T1 78451342/15/2712 9 36/9/27 33 4 4454
T2 75461740/12/281311 35/9/26 32 7 4491
T3 73531337/10/2817 8 36/9/26 34 5 4498












There are several additional reasons why T1-T3 produce fewer quality A solutions than T0. One reason is location accuracy. Using PDE instead of rapidly available locations for the MT inversion (bottom part of Table 4.4) results in a significant increase in quality A solutions (73 or more) with only small variations between T0-T3. The shorter cut-offs, when considering A+B solutions combined actually surpass T0 slightly. A small decrease of A and parallel increase of B solutions from T1 to T3 probaby reflects limits of the simple 1-D PREM model to predict waveforms at shorter periods and large epicentral distances. The observation that the ’best’ period band depends on location accuracy is consistent with earlier findings (chapter 3, Figure 3.12).


PIC

Figure 4.7: Distribution of the near-real time accessible broad-band data used. Light grey refers to components used in the standard routine SR, dark grey refers to additional components used with T0 and Rm = 2. Left: Number of components with respect to epicentral distance. For all Ti-settings the mean epicentral distance is 12.2o. Right: Station component azimuthal distribution. The strong dominance of the northwest azimuth is caused by the large number of events in the eastern Mediterranean Sea and of stations in central Europe. Length of the shades is proportional to number of components in each 2o-bin.

The complicated tectonics of the European-Mediterranean region combined with the still limited azimuthal coverage and large station-event distances also affect the resolution capabilities of the T1-T3 bands. Figure 4.7 shows the data distribution as a function of epicentral distance and azimuth. Even though the number of broadband stations has increased relative to the period considered in chapter 3, the average distance (12o) remained large and most stations (in central Europe) are northwest of the events (Aegean Sea region). Our results are consistent with chapter 3, where we already observed that the SR analysis for events in the Aegean Sea and Iran was less efficient (i.e. more data were required for quality A solutions, Figure 3.5). Thus, I divided the European - Mediterranean region into two areas. The first area includes the eastern Mediterranean Sea and the near East (Lat. < 40o; Long. > 20) (called EMNE here), the second area includes all other regions (called Europe here). Results for the two areas are given separately in Table 4.4 (third and fourth main column).

For Europe, the use of the T1-T3 settings results in a larger number of quality A solutions (33-32) compared to SR (28). For Mw > 4.7, differences between SR and the Ti sets are small. For Mw < 4.7, the shorter Ti sets, as expected, provide more A solutions. The difference in the number of Mw < 4.7 A-solutions between SR and the T1-set is significant at the 75% level. In Europe the average distances are short, allowing analysis with periods shorter than 50 s. Examples of waveform fits using the SR and the T1-setting are shown for two moderate events (4.3 < Mw < 4.7) in Appendix A.

For the EMNE area, the shorter T1 - T3 cut-off sets result in fewer A-solutions than SR, independent of magnitude. Only the T0-setting leads to a slight increase of A-solutions (significant at 80% level). In eastern Turkey and Iran, the long paths in thick crust (Marone et al.2004) compared to PREM require long-period analysis. In the Aegean Sea the long average station distances probably hamper the MT analysis using shorter period Ti sets.

The difference between the two regions is not due to earthquake location accuracy. Using PDE-locations for MT analysis (Table 4.4), we found that most A-solutions in Europe are found by the T1-set and in the EMNE area for the T0-set. For both regions similar numbers of Mw > 4.7 events have quality A, while the main difference lies in resolution for Mw < 4.7 events.

Division of the automatic MTs into three pre-defined quality groups may limit the evaluation of the overall improvement. For example, the August 6, 2002 Mw = 5.1 Spain event resulted in quality A using the SR, while in quality B using the T1-set. However the solutions are actually very similar: DzSR-T1 = 0 km, DMwSR-T1 = 0.01 and DAxSR-T1 = 2.2o. Using SR resulted in quality A because of DAx = 28.9o relative to the SRMT catalog, while using the T1-set resulted in a quality B solution because of DAx = 31.1o. To bypass the limitations imposed by categorization, we looked at DAx, Dz and DMw for individual events for all Ti. Generally, we found no substantial difference relative to the results presented in terms of categories.


PIC

Figure 4.8: Location of the 139 automatic MT from the combined T0-T1-setting. Black dots are quality A, grey dots quality B and white dots quality C solutions. Events bounded by the solid line to the west and north are solved with the T0-set, the others with the T1-set.

Based on Table 4.4, I suggest to use T0 for events located in the EMNE area, and T1 for European events. Combining these two sets, FARMT resulted in 66 quality A (+20% relative to SR), 44 quality B and 29 quality C solutions (Table 4.3 and Figure 4.8). A Wilcoxon test confirms that the difference in the number between the combined T0-T1-set and the SR is significant at the 95% level. Using the PDE-locations, the combined setting resulted in 85 quality A, 35 B and 16 C solutions. The large difference between the number of quality A solutions obtained with automatic and PDE-locations shows that location accuracy plays a significant role in MT resolution. The choice of the combined T0-T1-setting is simply based on maximizing the number of quality A solutions in the 1.5 year data-set. Using a different data-set or different locations might have resulted in a different ”best choice”, because the differences between the Ti settings are gradual. The main points are: distance dependent short period cut-off and signal-to-noise based data selection increases the efficiency of regional automatic MT analysis significantly. The choice of Ti depends on crustal structure, travel path length and location accuracy, and, we do see systematic differences between Europe and the EMNE area.

To check the overall validity of the minimum signal-to-noise ratio Rm = 2 used in FARMT, the MTs of the entire data-set were recomputed using the combined T0-T1 setting for Rm = 1 - 100. Using Rm = 1 - 3, the overall performance does not vary significantly and drops for Rm > 4. Using Rm = 100, only 29 events have a quality A solution, of which only 3 events are smaller than Mw = 4.8.


PIC

Figure 4.9: Overall improvement of automatic MT using the combined T0-T1-settings for Europe (left) and eastern Mediterranean Sea and near east (right). The symbol shape refers to the quality from SR solutions: triangle quality C, square quality B and hexagon quality A. The shading refers to the quality from the T1 and T0 settings: white quality C, light grey quality B and dark grey quality A. The thick continuos lines limit the area where points can possibly lie. Points above the oblique thin continuos line are quality A solutions, the oblique dashed lines indicates a 10o axes difference excess from quality A. Dotted region contains SR quality A solutions.

Figure 4.9 shows in detail the overall gradual improvement obtained with T1 and T0 for Europe (left) and the EMNE area (right), respectively. Shown is the axes difference DAx (between the SR and SRMT solutions, horizontal axis) versus the change of difference dDAx using T1 and T0. For the previously mentioned Spanish event, dDAx = -2.2o. Positive values of dDAx imply that the solution with T i is closer to the SRMT solution than SR solution. In both cases, the solution quality improved in the majority of cases ( ~~ 60%). The solutions that apparently became significantly worse (dDAx <-40o) deserve closer inspection. In the left box, the white hexagon is event Nr. 25 (Table 4.4), an aftershock minutes after the September 6, 2002 Mw = 5.9 main Sicily event. Because of the main shock, the signal-to-noise analysis removed almost all traces and only few components could be used. The quality A solution obtained using the SR is thus fortuitous. In the right box, event Nr. 146 is deep (dDAx = -64.4o, z = 153 km in SRMT) and its variance variation for the sampled depths is very smooth, with the minimum at an incorrect shallow depth with wrong mechanism. The solutions of the other two events (Nr. 13, dDAx = -52.3o and Nr. 89, dDAx = -49.8o) were obtained using only 2 and 3 components respectively using T0, and 1 and 2 components using the SR. The two quality A solutions obtained using the SR are thus fortuitous.


PIC

Figure 4.10: Left: accuracy of the fast locations used for automatic MT. Vertical axis indicates location difference (in km) with respect to the PDE-location. Color scale follows Figure 4.8. Right: general trend of near-real time accessible broad-band data availability for each triggered event. Time scale is in julian days starting at January 1, 2002

Figure 4.10 shows the evolution of location accuracy (left) and near-real time accessible broad-band data. The locations (fewer outliers) seem to have improved starting near the second half of 2003. The solution quality indicates that location error for successful MT analysis has to be smaller than  ~~ 100 km. Data availability also clearly increased by almost a factor two. It seems that a large amount of data can partially compensate location errors smaller than 100 km. More data need to be analyzed to confirm this observation.

Figure 4.11 shows the normalized variance for quality A and B solutions using the SR (left) and the combined T0-T1-setting (right). The clear variance reduction relative to the SR seems to indicate that a significant number of Mw = 4.2 - 4.7 events could be analyzed by lowering the current trigger threshold from M > 4.7 to M > 4.3, although accuracy of quick location is still a problem for such events.


PIC

Figure 4.11: Normalized variance for quality A (dark grey dots) and B (light grey dots) for SR (left) and combined T0-T1-settings (right). Lines indicate the mean variance of quality A for Mw ranges of 4.3-4.7, 4.8-5.2, 5.3-5.7 and 5.7-7.0.

Compared to the SRMT catalog, the automatic FARMT solves a similar data-set. Figure 4.12 shows the number of events analyzed with respect to Mw and to quality. The black pattern shows the events listed in the SRMT that did not result in a MT using FARMT. Basically almost all Mw > 5.2 events were triggered by the automatic procedure and resulted in a MT solution, usually of quality A with some B and few C. For 4.7 < Mw < 5.2, the number of B and C solutions increases, but still most SRMT events were also analyzed automatically. Most of the events that did not result in a MT have magnitude Mw < 4.7 and are located in the Aegean Sea and Zagros mountains, where nearby near-real time broad-band stations are rare or absent.


PIC

Figure 4.12: Number of automatic MT solutions obtained using the combined T0-T1-setting (left) compared to SR solutions (right). Dark grey area is quality A, light grey quality B, white quality C. Black area refers to MTs included in the SRMT catalog that were not triggered or did not result in an automatic MT solution.

4.7 Conclusions and outlook

We showed that fully automatic MT analysis in the European-Mediterranean region can be significantly improved for 4.3 < Mw < 4.7 earthquakes than previously possible (Bernardi et al.2004). Improvements are obtained because shorter periods are included and data selection is based on signal-to-noise ratio analysis.

We performed independent tests to establish distance dependent short period cuf-off ranges and a minimum signal-to-noise ratio Rm. Not surprisingly, we found gradual variations of the solutions as an entity making it difficult to pick a single ”best” combination. Rm = 1 - 5 leads to well resolved mechanisms and uses more seismograms for inversion than the SR. For the Ti, we observe a significant difference between source regions. For events in EMNE, that are generally farther from seismic stations and were travel paths cross thick crust, analysis at longer periods (T0-setting) is required than in Europe. In Europe shorter Ti settings result in more well resolved, smaller events than T0. T1 seems to work slightly better than even shorter cut-offs T2 and T3. These results hold for the automatic locations, that are sometimes in error of  ~~ 100 km, and for the more precise PDE-locations.

Using Rm = 2, T0 in the eastern Mediterranean Sea and near East, and T1 in Europe, we obtained 66 quality A solutions, 44 B, and 29 C solutions out of 169 analyzed events from May 2002 to September 2003. This is a significnat improvement compared to the SR that resulted in only 55 quality A solutions (and 44 B, and 34 C).

The significant difference in the number of well resolved solutions between automatic locations, that are used for the rapid analysis, and the PDE locations high-lights one main limitation of our fully automatic, rapid analysis. Even if the number of near-real time accessible broadband stations keeps increasing, we cannot get better moment tensors as long as location accuracy does not improve significantly also for moderate events. One possible approach to avoid the location problem could be a grid search based MT inversion. However, grid searches are time consuming and contradict our general goal of shortening the time between earthquake and MT solution.

The pre-selection of the components for inversion based on Rm reduces the number of iterations to get the final solution by about 20%, which is only a slight improvement towards faster result dissemination. A further development should deal with more precise differentiation between seismic signal and noise. This would bypass inversion iterations to remove noisy traces. One problem for example, is to recognize seismic spikes due to instrument self calibration; these spikes have large amplitudes. If they occur in the signal window, Rm is large independent of the actual signal. Currently the simple Rm-analysis of FARMT cannot recognize such spikes as noise.

Additional broadband stations, particularly in northern Africa and Iran, improving the azimuthal station coverage, are necessary to improve parameter resolution further. In central Europe, with its very dense broadband networks, we already can analyse relatively small events (e.g., the August 6, 2002 Mw = 4.3 Merano, north Italy, event) reliably. Continued improvements in earthquake location and data availability will allow us to lower our trigger threshold toward M  ~~ 4, which will reduce significantly the difference between the events analyzed fully automatic and the manual SRMT catalog.

Solution improvements for Mw < 4.3 events probably require fine-scale 3-D crust-mantle models. High-quality 3-D models that have the resolution for regional waveform prediction are becoming available (Boschi et al.2004Marone et al.2003) at the same time as advanced codes for efficient 3-D synthetic seismogram calculation (Komatitsch & Tromp2002). These development promises major advances for regional moment tensor inversions, that could include, for example, finite source modeling.

Chapter 5
Seismic moment from regional surface wave amplitudes - applications to digital and analog seismograms


This Chaper has been accepted for publication by the Bull. Seism. Soc. Am.

5.1 Summary

Accurate, consistent earthquake size estimates are fundamental for seismic hazard evaluation. In central Europe, seismic activity is low and long-term seismicity, available as intensities from written historical records, has to be included for meaningful assessments. We determined seismic moments Mo of 25 stronger 20th century events in Switzerland from surface wave amplitude measurements. These Mo can be used to calibrate intensity-moment relations applicable to pre-instrumental data. We derived the amplitude-moment relation using digital data from 18 earthquakes in and near Switzerland where independent Mo estimates exist. The surface wave amplitudes were measured at empirically determined distance varying reference periods TD. For amplitudes measured at TD, the distance attenuation term of the surface wave magnitude relation S(D) = log(A/T)max + 1.66logD is independent of distance. For logMo = MS + CE, we get log Mo = S(D) + 14.90. Uncertainties of ±0.3 for the 14.90-constant correspond to a factor of 2 Mo uncertainty, which was verified with independent data. Our relation allows fast, direct Mo determination for current earthquakes, and after recalibration of the constant, the relation can be applied anywhere. We applied our relation to analog seismograms from early-instrumental earthquakes in Switzerland that were collected from several European observatories. Amplitude measurements from scans were performed at large amplifications and corrected for differences between TD and actual measurement periods. The resulting magnitudes range from Mw = 4.6 to 5.8 for the largest earthquake in Switzerland during the 20th century. Uncertainties for the early-instrumental events are on the order of 0.4 magnitude units.

5.2 Introduction


PIC

Figure 5.1: Station and event map. Black triangles: analog stations used for historical earthquakes; white triangles: digital stations used to calibrate the surface wave amplitude - Mo relation. Inset shows earthquakes analyzed; black circles are early instrumental and white circles are recent events. Recent events in northern Italy were included to increase the magnitude range.

In this paper, we present seismic moment estimates of 25 stronger, 20th century earthquakes in and near Switzerland (Figure 5.1) based on amplitude measurements of regional surface waves recorded by analog instruments. In areas of low seismicity and long historical records, like Switzerland, such direct size estimates of early instrumental earthquakes are important to derive relations between instrumental data and macroseismic observations that can be applied to pre-instrumental earthquakes (Jimenez et al., in prep.). Modern instrumental data, available since 1975 in Switzerland, exist only for a few events with macroseismic fields and cannot provide such relations. Our size estimates are part of a larger effort to update the Earthquake Catalog Of Switzerland (ECOS) (Fäh et al.2003). An important design principle of ECOS is to express earthquake size uniformly in terms of moment magnitude Mw (Braunmiller et al.2004) to allow a direct event-to-event size comparison necessary for a homogeneous seismic hazard estimation from the 2000 years of seismic records.

Our data set for the early instrumental earthquakes consists of low gain analog regional seismograms. We thus determine earthquake size from the well recorded surface waves. The original surface wave magnitude MS definition by Gutenberg (1945), and later by Soloviev (1955), Kárník et al. (1962), and Vane  k et al. (1962) is given for the maximum ground velocity (A/T)max:

         (A  )
MS  = log  --     + 1.66log D + Cp
           T  max
(5.1)
where A is ground displacement in mm and T the corresponding period in seconds, distance D is in degree and Cp equals 3.3 (Kárník et al.1962). Common current practice is to measure the maximum ground displacement at a reference period T20  ~~ 20 s. At teleseismic distances, maximum velocity occurs at the dominant signal Amax with periods near 20 seconds (Panza et al.1989) and (Aki & Richards2002, pp.611-614). For larger earthquakes most measurements contributing to MS are determined at teleseismic distances, and the difference between (Amax/T20) and (A/T)max is probably small resulting in unbiased MS. At local and regional distances (D < 20o) surface waves are dominated by shorter period signals and (A/T)max is significantly larger than (Amax/T20). Therefore, measuring Amax at T20 probably leads to systematic underestimation of MS for smaller events that are predominantly recorded at close-by stations.

Several studies modified equation 5.2 to remove the distance dependent bias (Evernden1971von Seggern1977Herak & Herak1993Rezapour & Pearce1998) while measuring amplitudes at T20. Another approach, followed here, is to test whether equation 5.2 can be made independent of distance when the measurement period T is varied with distance. Globally applicable distance varying measurement periods TD have been suggested by IASPEI (1967) and Willmore (1979). Additionally, regional MS variations due to path effects my require regional corrections (Abercrombie1994Ambraseys & Free1997Ambraseys & Douglas2000).

The earthquakes analyzed here are generally too small to be well recorded at teleseismic distances; we thus use only regional records. To avoid biased size estimates due to distance dependence and regional path effects, we calibrate the surface wave amplitudes against an independent data set. We first outline our size estimation procedure before we derive a calibration function using recent digital data. Finally, we apply the calibration to determine Mo from analog data.

5.3 Procedure

Our early instrumental earthquake data set consists of scanned paper seismograms. We measured the peak surface wave amplitude and the corresponding period T. After correcting for gain, we could determine MS from equation 5.2. However, our data are regional seismograms, where MS distance bias has been documented (Evernden1971), and all measurement periods T were shorter than 20 s.

Therefore, we investigated whether the S(D)-term in equation 5.2

S(D)  = log (A/T  )    + 1.66log (D)
                 D max
(5.2)
is distance unbiased (i.e., S(D) is constant for a given earthquake). For this purpose, we obtained broadband digital data for 18 earthquakes (Figure 5.1) with similar station and event distribution as for the early instrumental data. With the correct distance varying TD equation 5.3 is distance independent.

None of the recent events was large enough to have an independent MS based on teleseismic data. Thus, we cannot verify the validity of Cp = 3.3 in equation 5.2. However, all recent events have Mo estimates from regional moment tensor (MT) inversion. Assuming log Mo = MS + CE from theoretical considerations (Kanamori & Anderson1975) and observation (Ekström1987Ekström & Dziewonski1988), we solve for K = CP + CE:

            ( -A-)
log Mo = log  TD  max + 1.66 log (D) + K
(5.3)
where Mo, (A/TD)max and D are known. The logMo -MS relation is only valid for ”small” earthquakes that do not rupture the entire width of the seismogenic zone. Our data set complies with this condition.

Originally, we had planned to determine MS from early instrumental data. As it turns out, with digital reference data, we are able to establish a reliable and generally applicable surface wave amplitude - Mo relationship.

5.4 Recent earthquakes

5.4.1 Data

We obtained digital seismograms recorded at regional distances (2o < D < 20o) for 18 small to moderate size earthquakes (M0 < 6 . 1016 Nm) that occurred in and near Switzerland since 1992 (Figure 5.1). Mo for each event (Table 5.1) comes from regional MT inversion (Braunmiller2001Braunmiller et al.2002Braunmiller2002Deichmann et al.2002); previously unpublished solutions are listed in Table B. Three additional earthquakes, with analog, digital and MT derived Mo are used for an independent check of our methodology.












Earthquake Date LatitudeLongitude zMw Mo MS - CPMw Mo










Buchs, CH 1992/05/08 06:44 47.27 9.50 6 4.3 3.7e15 0.75 4.4 4.5e15
Genevois, F 1994/12/14 08:55 45.95 6.44 18 4.3 2.8e15 0.60 4.3 3.2e15
Valpelline, IT 1996/03/31 06:08 45.94 7.40 4 4.2 1.9e15 0.55 4.3 2.8e15
Imst, A 1997/06/05 20:22 47.34 10.69 15 4.0 1.0e15 0.21 4.0 1.3e15
Fribourg, CH 1999/14/02 05:57 46.78 7.19 4 4.0 1.0e15 0.59 4.3 3.1e15
Bormio, IT 1999/12/29 20:42 46.53 10.25 4 4.9 2.4e16 1.33 4.8 1.7e16
Bormio, IT 1999/12/31 04:55 46.52 10.46 4 4.2 2.1e15 0.45 4.2 2.2e15
Bormio, IT 2000/04/06 17:40 46.54 10.34 4 4.1 1.6e15 0.17 4.0 1.2e15
Northern Italy2000/06/18 07:42 44.76 10.80 4 4.5 6.1e15 1.09 4.6 9.8e15
Northern Italy2000/08/21 17:14 44.87 8.48 4 5.1 5.5e16 1.70 5.0 4.0e16
Merano, IT 2001/07/17 15:06 46.74 11.20 15 4.8 1.9e16 1.18 4.7 1.2e16
Asti, IT 2001/07/18 22:47 44.84 8.40 4 4.4 4.5e15 0.48 4.2 2.4e15
Udine, IT 2002/02/14 03:18 46.37 13.17 15 4.7 1.3e16 1.18 4.7 1.2e16
Austria 2002/09/30 02:48 46.42 13.73 15 3.9 9.1e14 0.14 4.0 1.1e15
Lago Iseo, IT 2002/11/13 10:48 45.59 10.15 12 4.3 3.1e15 0.62 4.3 3.3e15
France 2003/02/22 20:41 48.35 6.80 12 4.8 1.6e16 1.25 4.7 1.4e16
Albstadt, DE 2003/03/22 13:36 48.20 9.01 6 3.9 8.5e14 0.53 4.3 2.7e15
Northern Italy2003/04/11 09:26 44.82 8.83 12 4.9 2.4e16 1.36 4.8 1.8e16











Table 5.1: 18 recent earthquakes used to calibrate TD. Depth z [km], moment magnitude Mw and seismic moment Mo [Nm] listed in columns 5, 6 and 7 are from moment tensor inversion. The distance independent term MS-CP = log(A/TD)max+1.66 log D is shown in column 8. Mw and Mo in columns 9 and 10 are from surface wave amplitudes (equation 5.6).

5.4.2 Reference period TD

Surface wave magnitude MS is determined from the maximum ground velocity (A/T)max. At regional distances, the maximum occurs at periods T < 20 s. Here, we want to find a stable, distance varying reference period TD where (A/TD) and A are maxima.

We narrow band-pass filter the displacement seismograms in one second intervals from a center period Ti of 4 to 20 s for all 443 station-event pairs. The ground displacement amplitude A is the vectorial sum of the instrument corrected half peak-to-peak amplitudes of the two horizontal seismogram components. We use a 4-pole causal Butterworth band-pass filter with corners at ±15% of the center period. We do not consider periods T < 3 s to avoid S-wave contamination, which dominate the records at short periods; separation based on group velocity is possible, but was not pursued.


PIC

Figure 5.2: Dominant period selection example for the July 18, 2000 Northern Italy earthquake at station BUG (D = 7.1o) and ECH (D = 4.3o). Left panel: T i are periods sampled; ground displacement A is shown in light grey (left axis) and ground velocity in A/Ti in dark grey (right axis). Diamonds are peak values. Right panel: EW-component seismograms (top displacement, bottom velocity). Boxes show maxima in seismograms band pass filtered between 3 and 20 s. At BUG Amax and (A/T)max occur at the dominant period T = 9 s. At ECH the maxima are at different periods and no dominant period exists.

For each period Ti, we measure displacement A and velocity (A/T). Signals characterized by a dominant period T can be approximated by a monochromatic signal, thus (A/T)max and Amax occur at the same period T (Figure 5.2, top). Measuring the period where only (A/T) is a maximum, may correspond to a period, which is not representative of the signal character (Figure 5.2, bottom).


PIC

Figure 5.3: Calibration of reference period TD relative to logarithm of epicentral distance D. Grey dots are dominant periods T, measured from 4 to 20 s for each station-event pair. Black dots are medians for TD from 4 to 10 s. The dashed line is the best ln-curve fitting the medians (equation 5.4).

The dominant periods T are shown as a function of epicentral distance in Figure 5.3 (grey dots). The medians (black dots) for each period between 4 and 10 s (at longer periods we have too few data points) are selected and fitted with a simple ln(D)-curve (dashed line):

TD =  - 3.99 + 6.41 ln(D)
(5.4)
with standard deviations s = ±0.24 and ±0.30 for the first and second parameter, respectively. The logarithmic curve is compatible with the non-linear increase of the dominant period T with distance and with the 20 s dominant period at teleseismic distances. Seismograms are band-passed in 1 s period intervals, thus TD is also approximated in 1 s increments (Table 5.2).



SED Willmore TD[s]
[Dmino, Dmaxo[[Dmino, Dmaxo[



2.00, 3.75 4
3.75, 4.40 2.00, 3.00 5
4.40, 5.15 3.00, 5.00 6
5.15, 6.05 5.00, 7.00 7
6.05, 7.05 7.00, 9.00 8
7.05, 8.25 9.00, 12.50 9
8.25, 9.60 12.50, 17.50 10
9.60, 11.25 11
11.25, 13.10 17.50, 20.00 12
13.10, 15.35 13
15.35, 17.90 14
17.90, 20.00 15




Table 5.2: Epicentral distances from equation 5.4 in first and following Willmore (1979) in second column versus reference period TD. To obtain Willmore’s reference periods, we defined regular distance intervals around each reference distance and we selected the corresponding median period.

5.4.3 Distance dependence

For our reference periods TD, we test whether the S(D)-term is distance biased. For each station-event pair we determine the maximum amplitude Amax at TD and calculate MS. For each event, we then determine residuals dMS,i = MS - MS,i between average and individual station measurements that are independent of the constant Cp in equation 5.2. The TD-residuals (Figure 5.4, top) show no distance dependence when considering measurements up to D = 20o (thick dashed line) even though T D were obtained using data only up to D = 10o, which represent 90% of the data. Linear regression (dM S,i = a + b D) for data points D < 10o shows negligible distance dependence (thin dashed line, a = 0.017, b = -0.003).


PIC

Figure 5.4: Residuals dMS,i = MS - MS,i versus distance D. MS is the mean for each earthquake and MS,i are station measurements. The top right of each panel shows the linear regression (a + bD) coefficients a and b (2o < D < 20o). The dashed lines are regressions for residuals 2o < D < 20o, the dotted lines for 2o < D < 10o representing 90% of the data. Top: residuals obtained using TD listed in Table 5.2. Middle: residuals using T20 depend strongly on distance. Bottom: residuals obtained using reference periods suggested by Willmore (1979) (Table 5.2).

For comparison, the residuals for Amax measured at T20 (Figure 5.4, middle) are distance dependent with underestimation at close distances (dMS,i positive). Our distance-varying TD deviate slightly from the values suggested by Willmore (1979) (Table 5.2). The residuals dMS,i obtained using Willmore’s periods (Figure 5.4, bottom) have a small distance dependence (a = 0.063 and b = -0.010) for all residuals at D < 20o, and a more relevant dependence for residuals at D < 10o (a = 0.148 and b = -0.028). MS values measured with Willmore’s periods at D = 2o differ from the more distant measurements (at 20o, respectively, 10o distance) by about 0.2 magnitude units.

The absence of distance dependence (Figure 5.4) of the station-event residuals dMS,i for our TD seems to favor our values over those given by Willmore. Given the large scatter in the residuals, we performed two statistical tests to verify the significance of the differences.

First, we applied a Sign-test (Stahel1995). The null hypothesis Ho assumes that the probability of observing positive and negative differences (d21 = (|dM S,i1|-|dM S,i2|)) between n matched pairs of two identical distributions is equal. The matched pairs are event-station residuals obtained with our (dMS,i1) and Willmore’s (dM S,i2) periods T D. If the number of positive differences d21 exceeds [n
2 ±2.3
 2 V~ n], H o can be rejected at the 99% confidence level. For our data set of n = 443, we observe 305 positive differences d21 and reject Ho.

Second, we tested how small changes in our TD affect residual distance independence. We generated 100 random gaussian distributed values for both parameters of equation 5.4 (-3.99 ± 0.24, 6.41 ± 0.30) and calculated for each of the 10000 new reference periods TD' the residuals dMS,i' and their regressions. For each case, we then calculated the expected residual difference dD = dM S'20o - dM S'2o. The dM S'D are regression results at 20o and 2o distance, respectively. Thus, dD describes the expected magnitude difference for these two distances (Figure 5.5). The mean dD is 0.007 ± 0.029. No difference is larger than 0.1 magnitude units. Our statistically worst T'D produce significantly less distance dependence than Willmore’s reference periods.


PIC

Figure 5.5: Stability of the dMS,i-residual distribution for randomly generated T'D. As measure of stability, we use the difference dD = dM S20o - dM S2o of the regression result for each set of T'D. For the 10000 sets, we obtain a mean difference of m = 0.007 ± 0.029 (long and short dashes, respectively).

For earthquakes in and near Switzerland, we determined reference periods TD that produce distance independent surface wave magnitude MS (assuming Cp is known). Differences relative to global average values given by Willmore (1979) are statistically significant and we recommend our TD (Table 5.2) for regional MS determination in central Europe.

5.4.4 Seismic moment

From theoretical considerations (Kanamori & Anderson1975), MS is expected to be proportional to log Mo for small earthquakes. Empirical analysis shows that globally this applies to MS < 5.3 events (Ekström & Dziewonski1988). For continental earthquakes the proportionality is valid up to MS < 7.2 (Ekström1987), because of the increased width of the seismogenic zone compared to oceans. Seismicity in the Alpine area is characterized by small to moderate earthquakes, thus we can assume:

logM   = M   + C
      o     S    E
(5.5)
For a global data set Ekström & Dziewonski (1988) found CE = 12.24, but also reported regional variations of up to 0.4 magnitude units. With Mo estimates based on MT inversion available for the 18 earthquakes (Table 5.1) and combining equations 5.2 and 5.5, we obtain

            (    )
log Mo  = log  A-- +  1.66 log(D) +  Cp + CE
              TD
(5.6)
We solve for K = Cp + CE and obtain an average value of K = 14.90 ± 0.32 (Mo in N m). We find no distance or seismic moment dependence of K for our data set. The scatter of K translates to an uncertainty of moment magnitude Mw = 2
3 log Mo - 6.03 (Hanks & Kanamori1979) of 0.2 units.

Equation 5.6 allows a straight forward Mo estimation from surface wave amplitude measurements valid for historic, recent or future earthquakes. The difference of Mo obtained with amplitude measurements and MT inversion is lower than a factor of 2 (Table 5.1) and is consistent with the standard deviation s = 0.32 of K (only two smaller earthquakes show a difference of a factor of 3).

Combining the global values Cp = 3.3 and CE = 12.24, we obtain K' = 15.54, which increases earthquake’s Mo by the significant factor of 4.5 relative to K = 14.90 (or a Mw increase of almost 0.4 units). The difference K'- K is outside the s uncertainty for K determined here. Biased Mo estimates from MT inversion could also cause an artificial difference. However, these Mo estimates agree excellently with values determined by the teleseismic Harvard centroid-moment tensor technique (Braunmiller et al.2002). We think the differences between K and K' are real and caused by regional effects. From our analysis, it is not possible to resolve whether Cp = 3.3, CE = 12.24 or both values are regionally different.

5.5 Historical earthquakes

5.5.1 Data

We collected seismograms and instrument calibrations for 25 stronger earthquakes that occurred during the 20th century in or near Switzerland (Figure 5.1, Table 5.3). Earthquake selection criteria were even distribution, representing different tectonic settings (Kunze1982), a wide size range and well defined intensity fields. The largest 20th century earthquake in Switzerland, the January 25, 1946 Ayent event, is part of our list.














# Date Lat. Lon.Mo Range MwRangeMSRange # Location Area












1 1905/04/29 01:5946.09 6.90 56 __________ 5.1 ________ 5.2 ________1/0/0Massif du Mont-Blanc (VS)
2 1905/12/25 17:0546.81 9.48 13 __________ 4.7 ________ 4.5 ________1/0/0Domat/Ems (GR)
3 1911/11/16 21:2548.22 9.00190160 - 410 5.5 5.4-5.7 5.7 5.6-6.01/0/2Ebingen (Germany)
4 1915/08/25 02:1346.10 7.06 10 8.3 - 12 4.6 4.6-4.7 4.4 4.3-4.52/1/0Martigny (VS)
5 1924/04/15 12:4946.30 7.96 65 36 - 120 5.2 5.0-5.4 5.2 5.0-5.54/2/0Brig (VS)
6 1925/01/08 02:4546.74 6.39 17 12 - 21 4.8 4.7-4.9 4.6 4.5-4.72/0/0Ballaigues (VD)
7 1929/03/01 10:3246.73 6.72 39 26 - 51 5.0 4.9-5.1 5.0 4.8-5.11/1/0Bioley-Magnoux (VD)
8 1933/08/12 09:5846.66 6.809.6 6.7 - 12 4.6 4.5-4.7 4.4 4.2-4.52/2/0Moudon (VD)
9 1935/06/27 17:1948.04 9.47260 74 - 480 5.6 5.2-5.8 5.8 5.3-6.13/2/0Saulgau (Germany)
101943/05/28 01:2448.27 8.98160 71 - 550 5.4 5.2-5.8 5.6 5.3-6.15/2/0Onstmettingen (Germany)
111946/01/25 17:3246.35 7.40550320 - 680 5.8 5.6-5.9 6.1 5.9-6.22/2/0Ayent (VS)
121946/01/26 03:1546.28 7.43 15 7.8 - 21 4.7 4.6-4.9 4.5 4.3-4.71/1/0Ayent (VS)
131946/05/30 03:4146.30 7.42230100 - 270 5.5 5.3-5.6 5.7 5.4-5.83/1/0Ayent (VS)
141954/05/19 09:3546.28 7.31 87 20 - 220 5.3 4.8-5.5 5.3 4.7-5.74/3/0Mayens de My (VS)
151960/03/23 23:1046.37 8.02 37 17 - 78 5.0 4.8-5.2 5.0 4.6-5.35/0/0Brig (VS)
161961/08/09 13:1046.8110.50 25 __________ 4.9 ________ 4.8 ________1/0/0San Valentino (Italy)
171964/02/17 12:2046.88 8.27 17 8.5 - 22 4.8 4.6-4.9 4.6 4.5-4.81/1/1Flüeli (OW)
181964/03/14 02:3946.87 8.32 98 32 - 160 5.3 5.0-5.4 5.4 4.9-5.74/1/0Alpnach (OW)
191968/06/27 15:4346.29 6.739.9 3.2 - 17 4.6 4.3-4.8 4.3 3.9-4.62/0/0Chablais (France)
201968/08/19 00:3646.29 6.55 13 12 - 32 4.7 4.7-5.0 4.5 4.5-4.93/0/0Chablais (France)
211971/09/29 07:1846.90 9.01 23 12 - 34 4.9 4.7-5.0 4.8 4.5-4.93/2/0Vorstegstock (GL)
221978/09/03 05:0848.28 9.01170150 - 200 5.5 5.4-5.5 5.6 5.6-5.71/1/0Ebingen (Germany)
231980/07/15 12:1747.69 7.40 18 13 - 23 4.8 4.7-4.9 4.6 4.5-4.82/0/0Habsheim (France)
241991/11/20 01:5446.73 9.53 11 8.5 - 16 4.7 4.6-4.8 4.4 4.3-4.63/0/0Vaz (GR)
251996/07/15 00:1345.94 6.099.1 8.5 - 11 4.6 4.6-4.7 4.4 4.3-4.42/1/0Epagny (France)













Table 5.3: Selected historical earthquakes. Date and location are from ECOS. Seismic moment Mo in 1015 Nm is from equation 5.7. For each M o, we give the corresponding Mw and MS (assuming CP = 3.3). Range for each Mo,Mw,MS is the lowest and highest measurement. # gives number of measurements: first, second and third number are from MSH, MSh and MSV data.

Our data are scanned images of smoke- and photo-paper (Appendix C) recordings collected from various seismological observatories in Europe (Table 5.4). Some seismographs operate since the beginning of the 20th century. Such long continuous recording times are particularly important, since they allow a direct relative size comparison between events and thus function as a check of measurement quality. The scans, the instruments calibrations and measurements are included in ECOS and are available on-line (http://www.seismo.ethz.ch).

We used only seismograms that were recorded by mechanical or electromagnetic seismographs with well known instrument characteristics. The introduction of damping to mechanical instruments and the rigorous derivation of the instrument characteristics by Wiechert (1903) produced the first reliable observations of ground motion. For a detailed description of the instrument characteristics of all early instrumental seismographs used in this study, see Appendix D. The Wiechert instruments record true ground motion at least from 0.05 to 2 Hz (Ritter2001) and non-linear effects have been observed only for high frequencies around 4 Hz (Herak et al.1996). Figure 5.6 shows gain versus period for a Wiechert seismograph. Most of our data are from mechanical Wiechert or Mainka (Mainka1923) instruments. Electromagnetic seismographs consisting of a pendulum coupled to a galvanometer were introduced by Galitzin (1914). Compared to mechanical instruments, they record with higher gain over a wider period band (Figure 5.6). Instrument responses for electromagnetic instruments are described in Galitzin (1914), Hagiwara (1958), and Willmore (1979). The measured amplitudes, after removing the instrument response, are ground displacements (in mm).








Observatory LatitudeLongitudeStationSeismograph Events #






Bruxelles 50.80 4.36UCC Wiechert 3-5,7-10,12-14
Bruxelles 50.80 4.36UCC Galitzin 7-10,12,13
Bruxelles 49.60 6.13LUX Galitzin 21
Ebre 40.82 0.49EBRE Mainka 9,12-14,18
Fürstenfeldbruck 48.15 11.60MNH Wiechert 2,5-7
Fürstenfeldbruck 48.85 10.48NRD Wiechert 5
Göttingen 51.54 9.96GTT Wiechert 300 Kg 1,3,8-10,13,14,17,18,21-25
Jena 50.95 11.58JENA Wiechert 1200-1300 Kg9-14,18
Jena 52.38 13.06POT Wiechert 4,5,9
Jena 50.65 11.62MOXASSJI/L - HSJI/L 19-21,23-25
Ljubiana 46.04 14.53LJU Sprengnether 24-25
Wien 48.25 16.36WIEN Wiechert 3-8,13,14,17-21
Zagreb 45.83 15.99ZAG Wiechert 200-300 Kg 8-22







Table 5.4: Analog data station list. We obtained scans or copies of smoke- or photo-paper seismograms from observatories (column 1) for stations listed in column 4. Station coordinates are given in latitude and longitude. Seismograph describes the instrument type: Wiechert, Mainka are mechanical, Galitzin, Sprengnether, SSJI, HSJI are electromagnetic seismographs. The last column lists the events (Table 5.3) where the sensor was used.


PIC

Figure 5.6: Amplitude-period response examples for Wiechert (top) and Galitzin (bottom) seismographs at Uccle seismic observatory. Solid lines are for vertical, dashed and dotted lines for the two horizontal seismographs. The curves are lower and upper bounds obtained from the full range of seismograph constants given in the 1933 observatory bulletins. We used similar plots for other seismographs and time periods to estimate gain uncertainty for the two seismograph types.

5.5.2 Surface wave amplitude and seismic moment

We measure maximum surface wave amplitudes (one half of peak-to-peak values) from scanned images. We correct the amplitudes to account for gain differences between true measurement period Tw and reference period TD. For scanned images, the maximum amplitude Aw is measured at period Tw with gain V w. At period TD, the amplitude would be Aw .V
VDw- (V D is the gain at TD). The true ground displacement at TD is then Aw .VD-
Vw .-1-
VD, which replaces the expression log A in equation 5.6. The amplitude - Mo relation for early instrumental earthquakes becomes then

             (A   )
log Mo = log   -w- -  log (TD) + 1.66 .log(D) + 14.90
               Vw
(5.7)
Aw is the vector-sum of the two horizontal components. For stations where only one horizontal component was available, we added 0.15 units to logMo, which assumes that the average maximum amplitudes on both horizontal components are the same. For the 59 two-component measurements, we compared vector-sum (MSH) with one-component measurements (MSh) (Figure 5.7). The average differences are 0.17 in agreement with our average correction and we observe no difference between Wiechert and Galitzin instruments.

Correction for vertical component data (MSV ) was derived comparing two horizontal component data with vertical observations (Figure 5.7). For Wiechert instruments, we have 26 three component and 13 two component (one vertical) sets. In both cases, the average difference is 0.45 units, which we added to logMo based on vertical data only. For Galitzin instruments we found a difference of 0.20 units, however this value is based only on 6 observations and we decided not to use vertical Galitzin data.

MSH, MSh and MSV are names to indicate whether the Mo-estimate is based on amplitude measurements from two or one horizontal or a vertical component, thus they are not surface wave magnitudes. Derivation of the average differences between MSH, MSh and MSV involved only differences and is therefore independent of choice of Cp in equation 5.2.

Our combined amplitude data set then consists of 59 measurements from two horizontal components, 23 from one-horizontal and 3 from vertical component data (Table 5.3, column 10). Each station contributes only once per event with preference given to MSH (two horizontal) over MSh (one horizontal) over MSV (vertical) data. For each event we have from 1 to 7 measurements with at least one from MSH. The resulting Mo estimate (Table 5.3) is the median for each event’s measurements. We did not apply station or depth corrections. Relatively strong changes of the seismograph characteristics within a short time span have been documented (Kowalle & Thürmer1991Allegretti et al.2000). With few events and infrequently available instrument calibrations, gain uncertainties probably mask possible station terms. Depth corrections are probably unnecessary since all calibration events (Table 5.1) and other instrumental seismicity (Deichmann1992) are crustal.


PIC

Figure 5.7: Top: Difference between station MS obtained with two (MSH) and one (MSh) horizontal components for Wiechert (48 grey dots) and Galitzin (11 open dots) seismographs as function of event size. The mean difference of 0.17 ± 0.08 (thick and thin dashes) for both instrument types is compatible with our 0.15 units correction. Bottom: Difference between MSH and MSV (vertical) for Wiechert seismographs (26 diamonds). Also shown are MSH values derived from MSh data (13 squares). The mean difference is 0.45 ± 0.16 units (thick and thin dashes).

5.5.3 Seismic moment uncertainty

Uncertainty of Mo estimates (equation 5.7) is mainly due to uncertainties of gain (V w) and scatter of the constant K = 14.90 ± 0.32. Epicenter location, for historical events often not well known, probably results in a distance uncertainty of 0.2o. Amplitude readings from scans are performed at large magnifications on a computer screen, thus amplitude uncertainty is about 0.5 mm and corresponds approximately to the line width drawn by the seismograph needle. Location and amplitude uncertainties have negligible effects on Mo. The gain V w depends on time-varying seismograph constants, which were regularly calibrated by station operators. However, for many analyzed events, no calibration information existed near the event time and we extrapolated to obtain gain estimates. To estimate gain uncertainty due to this procedure, we varied the instrument constants for a station between the lower and upper values given in the station bulletin or between the calibrations just before and after an event. Figure 5.6 shows examples for the resulting amplitude responses for the Wiechert and Galitzin instruments at station UCC. Similar analyses were performed for other stations and from these tests we deduced a gain uncertainty of about 30 units for mechanical and about 50 units for electromagnetic sensors. Due to the higher gain, relative gain uncertainties are smaller for electromagnetic than for mechanical instruments.

Together, effects of gain, location and amplitude result in a mean maximum expected uncertainty for MSH data of d(log Mo) = 0.13 ± 0.03 and 0.10 ± 0.04 for mechanical and electromagnetic sensors, respectively. Combined with dK = 0.32, we obtain an uncertainty in log Mo of less than 0.5, which translates to a Mw uncertainty of about 0.3 units. Uneven azimuthal station coverage and ignoring radiation pattern effects add uncertainty. We thus consider dMw = 0.4 a reasonable uncertainty estimate for our early instrumental Mw values.

5.6 Discussion and conclusions

We presented a simple procedure to determine Mo from amplitude measurements of surface waves (equation 5.6) by combining the logMo - MS relation with the ’Prague formula’ for surface wave magnitude. Our approach is valid for ”small” earthquakes that do not rupture the entire width of the seismogenic zone, which is applicable for events up to MS < 7.2 in continental areas (Ekström1987). Validity also requires the S(D)-term to be independent of distance. Measuring amplitudes at periods that vary with epicentral distance results in distance independent S(D). We determined the reference periods TD (Table 5.2) with digital data and calibrated the constant K = 14.90 in equation 5.6 using independent Mo estimates.

Three earthquakes (Table 5.5) with Mw from MT inversion and amplitude readings from digital and analog data provide an independent check of the procedure. The Mw estimates show high consistency and the differences of < 0.3 units are consistent with our expected Mw uncertainties. We observe the largest discrepancy for the 1978 event, where we had only 3, respectively 2 amplitude measurements from digital and analog data.









earthquake
Mo[N m]
Mw







MT DigitalAnalogMTDigitalAnalog







1978/09/03 5.9e16 1.2e171.7e175.1 5.4 5.5
Harvard 7.2e16 5.2
1991/11/20 9.3e15 1.4e161.1e164.6 4.7 4.7
1996/07/15 8.7e15 1.3e169.1e154.6 4.7 4.6








Table 5.5: Seismic moment Mo and moment magnitude MW from MT inversion, digital (equation 5.6) and analog data (equation 5.7). For the 1978 earthquake a Harvard CMT solution exists. These earthquakes, not used to calibrate TD, provide an independent check of equations 5.6 and 5.7.

Our TD and K apply to earthquakes and stations in Switzerland and surroundings. For other tectonic settings or on a global scale, these values probably require re-calibration. The TD given by Willmore (1979) deviate slightly, but significantly, from our values and may represent a global average. Similarly, we expect K to show some regional variations as documented in variations of CE (Ekström1987). Assuming Europe-wide applicability of our TD, regional calibration of K is straightforward using waveform modeling based Mo data that are available for a large number of earthquakes in Europe (Stich et al.2003Pondrelli et al.2002Braunmiller et al.2002) and large events (Dziewonski et al.1981) globally.

Distance dependence of the S(D)-term has long been an issue when determining MS with T20 period data (Evernden1971). At regional distances, the amplitudes at T20 period are generally smaller than shorter period signals leading to a significant underestimation of MS (Figure 5.4). Measuring at shorter periods TD, results in distance independence and allows the analysis of smaller events that do not radiate significant energy at periods near 20 s.

Our 18 event calibration data set (Table 5.1) does not include larger events with teleseismically determined MS that could be used to fix the constant CP in equation 5.2. We thus did not explicitly determine MS. However, in engineering practice MS is often used for ground motion prediction (Ambraseys & Free1997), so we give MS estimates for the early instrumental data (Table 5.3) for CP = 3.3. Recently, Ambraseys (2003) reassessed the size of 20th century Swiss earthquakes in terms of M S using amplitude/phase measurements from station bulletins. For 23 common events, the average MS difference of m = 0.01 ± 0.25 units is remarkably small. The large scatter is mainly due to some of the oldest events, which differ by up to 0.5 magnitude units.

We observed a large discrepancy between measurements from the horizontal and vertical components of 0.45 magnitude units for Wiechert seismographs (Figure 5.7). Despite the agreement with Ambraseys (2003), we thus caution the use of bulletin data for which amplitude, period and component are not explicitly labeled. Use of such data could lead to artificial scatter or biased size estimates. Whenever possible, we suggest to use original seismic records and instrument calibrations.

A large data set of early instrumental seismic records exists in Europe, where some observatories have been in operation for an entire century. Scanning and digital archiving of many of these data is currently under way (Bono et al.2002). Simple reading of amplitudes and periods combined with our amplitude-moment relation could provide a large Mo-data base without the need of time consuming digitization. Because of calibration with recent events, accuracy of early instrumental Mo estimates depends mainly on knowledge of instrument gain and representative source radiation. Even with few measurements, accuracy is probably comparable to modern digital data.

The earthquake size estimates for the early instrumental events given here differ systematically from the estimates in the Earthquake Catalog of Switzerland (Fäh et al.2003Braunmiller et al.2004). We have two reasons that explain the difference. Here, we use a larger data set of analog seismograms that includes also data from instruments other than Wiechert seismographs. All size estimates are based on two component horizontal data, either by having both components or by correcting one component data for the missing component. Assuming CP = 3.3 for our measurements, the MS values are on average 0.1 unit smaller than in ECOS. The second reason concerns the MS - Mw conversion in ECOS. Here, we directly determine Mo from amplitude measurements (equation 5.7). Conversion in ECOS was performed with log Mo = MS + 11.9, which corresponds to KECOS = 11.9 + 3.3 = 15.2 instead of K = 14.9. This causes an additional dMw = 0.2-difference. Combined, the two small systematic shifts lead to Mw estimates in ECOS that are 0.3 units higher than here. Subtracting 0.3 magnitude units for historical and early instrumental earthquakes would alleviate most of an apparent seismicity change in ECOS. The systematic difference will be considered in future versions of the earthquake catalog.

Acknowledgments

Broadband digital data were obtained from the Geofon (Germany), Geoscope (France), Gräfenberg (Germany), IRIS (USA), Mednet (Italy), ORFEUS (The Netherlands), ReNaSS (France), and USGS (USA) data centers besides SDS-Net data. We thank our colleagues for hospitality and help during early instrumental seismogram retrieval: I. Allegretti and M. Herak (Zagreb), K. Atakan (Bergen), I. Cecic (Ljubljana), S. Gregersen (Copenhagen), D. Keiser and K. Rehm (Jena), V. Kovacevic (Belgrade), P. Labak (Bratislava), A. Lenhardt (Vienna), Y. Menechal (Bruyeres-le-Châtel), C. Meester (De Bilt), J. Ritter (Göttingen), E. Schmedes (München), A. Ugalde (Barcelona), R. Verbeiren (Bruxelles), P. Wiejacz (Warsaw). Most figures were produced with GMT (Wessel & Smith1995). Random gaussian values were generated with the Vincenzo Liberatore algorithm (Case Western Reserve University, Ohio). Support for ECOS came from internal Swiss Federal Institute of Technology (ETHZ) funds, the Swiss National Science Foundation, the Swiss Federal Nuclear Safety Inspectorate (HSK), and through an Agreement of Cooperation with PEGASOS. We thank all agencies for their support.

Chapter 6
Conclusions and outlook

In this thesis I presented two methods to derive earthquake source parameters using surface waves from regional seismograms (D < 20o).

The first method computes fast and automatic moment tensor solutions for Mw > 4.3 events in the European - Mediterranean region using 35 - 100 s period surface waves. To compute the automatic moment tensor, I developed a procedure (FARMT), that adapts the period band depending on the epicentral distance and the signal-to-noise ratio. These two features allow automatic MT analysis even for moderate events (Mw = 4.3 - 4.7). The relatively small number of analyzed events with Mw < 4.7 is caused by the actual trigger magnitude threshold Mw > 4.7. The results obtained with FARMT indicate that lowering the trigger threshold could lead to a significant increase in the number of moderate earthquakes that can be successfully analyzed.

The parameters used in FARMT for the signal-to-noise analysis and for the short period cut-off are set to maximize the number of well resolved MT solutions for the currently given circumstances (locations, stations, crustal model). Improvements will lead to parameter changes that will further reduce the magnitude threshold. The current station distribution, location accuracy and PREM synthetics force FARMT to 50 s as short period cut-off in the near East and in the eastern Mediterranean Sea. In Europe the near-real time broad-band network is denser, resulting in relatively accurate quick locations even for Mw  ~~ 4.0 - 5.0 events and shorter event - station distances. As a consequence, shorter periods, down to 35 s, can be used.

The low accuracy of some quick locations not only limits the use of shorter periods, but also hampers the signal-to-noise analysis. Signal arrival is not precisely known with inaccurate locations, I thus used time windows for signal spectral analysis that are larger than what actually is needed; the signal window thus also contains a variable amount of noise. Further developments should focus on procedures that are better able to characterize and distinguish signal and noise independently from the expected signal arrival. More sophisticated pre-processing which would select only reliable data, would eliminate iterations to remove noisy traces and cut the computation time significantly toward real-time analysis.

Besides fully automatic analysis, FARMT can also be used for a semi-automatic use, for example, to analyze a large number of older events in the European - Mediterranean region or in other parts of the world. Application of FARMT outside the European - Mediterranean region probably requires, short period cut-off re-calibration.

Further substantial improvements of MT retrieval probably require the use of 3D Earth models. High resolution 3D models and codes, able to compute 3D synthetics, started to become available during the last part of this work. Currently, the main limitation to apply 3D pre-compiled synthetic libraries for a large area like the European and Mediterranean area is computer speed. With CPU speed improvements, this problem will probably be overcome within the next few years.

The second method derives seismic moment Mo directly from surface wave amplitudes filtered between 4 s and 15 s period depending on distance. The Mo - amplitude relation is derived using digital broad-band seismograms and moment tensor derived Mo for recent events in the Alpine area. This relation does not provide surface wave magnitude MS. The relation, however, allows direct moment magnitude Mw estimation of regional, moderate events without the distance dependent bias resulting from standard MS measurements and without the need for teleseismic measurements as reference values. Measuring the amplitudes at periods shorter than 20 s, allows to compute MS of small events (Mw  ~~ 3.0) that de-facto cannot be computed using 20 s surface waves.

I applied the relation to 25 events that occurred in and near Switzerland during the entire 20th century. The earthquakes were recorded by mechanical and electromechanical seismographs on smoke- and photo-paper. With the simple amplitude measurements, I obtained size estimates that are probably superior to any other estimates for these events. In areas of low seismicity and long recurrence time for strong events, like Switzerland, my Mw estimates are of particular importance, for example for seismic hazard estimates.

Recently the so called SISMOS project started. Goal of the project is to scan and digitize a significant amount of waveform data and bulletin registrations for 20th century European events. My Mo-amplitude relations shows that a reliable and homogeneous size estimation for a large data-set can be obtained in a relatively short time. Homogeneous size estimates for a large number of significant 20th century events could provide the bases for improved and consistently derived seismic hazard estimates in Europe. My relation is very well suited to provide European - wide homogenous Mw estimates. The Mo-amplitude relation could also be applied for rapid size determination. The amplitude derived Mw depend less on location accuracy and computation is much faster than for moment tensor analysis. For example, a fast Mw estimate for small and moderate events in central Europe could be available within a few tens of seconds after the earthquake occurred.

Appendix A
Waveform fit examples

Examples of waveform fits for two moderate events in central Europe, where the denser near-real time broad-band networks allow a 35 s short period cut-off. For each event the waveform fit obtained using the SR and the new FARMT are shown. For both events, FARMT used more data than the SR, and the FARMT solutions are closer to the SRMT solutions.

The first two Figures (A.1 and A.2) refer to the June 8, 2002 northern Italy, event (Nr. 8 Table 4.3), the next two (Figures A.3 and A.4) to the January 26, 2003 central Italy event (Nr. 63 Table 4.3). The plot labels on the top left show date, origin time and event coordinates used for MT analysis, best fit depth, the orientations of the two nodal planes, the minimal and the maximal frequency used, variance and moment magnitude Mw. The top right shows double couple focal mechanism; the black triangles indicate the event-station azimuth. For each station azimuth and epicentral distance are given. Z, N, E are vertical, NS and EW components. Solid lines are observed and dashed lines synthetics seismograms. The number next to each waveform is the phase realignment shift.


PIC

Figure A.1: SR waveform fit of the northern Italy earthquake.


PIC

Figure A.2: FARMT waveform fit of the northern Italy earthquake.


PIC

Figure A.3: SR waveform fit of the central Italy earthquake.


PIC

Figure A.4: FARMT waveform fit of the central Italy earthquake.

Appendix B
MT from regional data

In this appendix, we give the results of regional moment tensor analysis for events used in this paper that have not yet been published elsewhere. The method is described in Nábe  lek & Xia (1995) and application examples for Switzerland and surrounding regions are presented in Braunmiller et al. (2002) and Deichmann et al. (2002). Additional solutions are given in Braunmiller et al. (2002), Braunmiller (2002), Baer et al. (2003) and Marone et al. (in prep). The remaining 14 solutions are presented in Table B.1.










Earthquake Date LatitudeLongitude z Mo Plane 1 Plane 2








Balingen, D 1978/09/03 05:08 48.28 9.03 65.9e16 189/90/0 99/90/180
Buchs, CH 1992/05/08 06:44 47.27 9.50 63.7e15 3/332/72 102/200/63
Genevois, F 1994/12/14 08:55 45.95 6.44 182.8e15 218/79/105 344/18/38
Valpelline, I 1996/03/31 06:08 45.94 7.40 41.9e15 15/129/82 115/13/40
Epagny, F 1996/07/15 00:13 45.94 6.09 48.7e15 246/82/228 147/43/348
Northern Italy2000/08/21 17:14 44.87 8.48 45.5e16335/78/-79/112/17/-132
Merano, IT 2001/07/17 15:06 46.74 11.20 151.9e16 116/77/163 210/74/14
Asti, IT 2001/07/18 22:47 44.84 8.40 44.5e15 345/85/-83 107/9/-148
Udine, IT 2002/02/14 03:18 46.37 13.17 151.3e16 217/83/-65 322/26/-163
Austria 2002/09/30 02:48 46.42 13.73 159.1e14152/81/-140 54/51/-12
Lago Iseo, IT 2002/11/13 10:48 45.59 10.15 123.1e15 79/85/95 214/8/46
France 2003/02/22 20:41 48.35 6.80 121.6e16 304/56/124 174/47/-50
Albstadt, DE 2003/03/22 13:36 48.20 9.01 68.5e14114/73/-156 17/67/-18
Northern Italy2003/04/11 09:26 44.82 8.83 122.4e16 207/79/-16 300/74/-169









Table B.1: Fault plane solutions from MT analysis for 14 recent earthquakes. For each event we give the location area, date, latitude, longitude, depth z [km], Mo [Nm] and the orientation of the two nodal planes (Strike/Dip/Rake). For the 1978 earthquake, Mxy = Myz = 0 are constrained. MT solutions since 2000 are available at http://www.seismo.ethz.ch/mt

Appendix C
Early instrumental seismograms

This appendix includes examples of the scans of the analog early instrumental seismograms used in chapter 5. Event number refers to table 5.3. All scans, instrument calibrations, amplitude and periods measured are included in the ECOS and are available on-line at http://www.seismo.ethz.ch.
To see the seismograms, download the pdf file.

Appendix D
Instrument characteristics of early instrumental seismographs

Mechanical seismographs

In this thesis we used two type of mechanical seismographs, the Wiechert (Wiechert1903) and the Mainka-pendulum (Mainka1923). The amplitude response of both seismographs can be described as following (Wiechert1903):

                     Vo
Vw  =  V~ ----------------------2-------
        {1 - ( TTwo)2}2 + 4 p(2+ln(ljn)j)2(TTwo)2
(D.1)
Tw is the observed period, To the natural seismometer period, V o the magnification for high-frequency excitation (Tw « To) and j the damping ratio. Equation D.1 is valid for a generic damped pendulum, thus can be applied to the Mainka-pendulum (Mainka1923) and other similar seismographs. The measured amplitude A, after removing the gain V w, is true ground displacement (in mm).

Electromagnetic seismographs

Electromagnetic seismographs consist of a pendulum coupled to a galvanometer (Galitzin1914). Electromagnetic seismographs used here belong to two groups. One group where pendulum and galvanometers have different natural periods and damping constants and a second where pendulum and galvanometer have similar natural period and critical damping. Magnification V w of the first group is (Willmore1979):

Vw =  Q .f (Ts, Tg,a,b, s,Tw)
(D.2)
Q is the instrument static magnification. The amplitude characteristic f depends on the natural periods of pendulum (Ts) and galvanometer (Tg), the damping constants of pendulum (b) and galvanometer (a), the coupling coefficient (s) and the ground motion period Tw. Following Hagiwara (1958) f, can be expressed as

                          u
f =  V~ -----------------------4----------------3--
       [1 - {a + 4abb}u2  +  un2-]2 + [- 2cu + 2d un-]2
(D.3)
where
    Tg         Tw             1        (1 - s2)            a         b
n = T--,  u =  T-,   a = 1 + n2-,  b = ----n---,   c = b + n-,   d = n-+ a
      s         s

The magnification V w of the second group is (Galitzin1914):

Vw =  kA1-.T ----u----
       pl    (1 + u2)2
(D.4)
where T = Tg = Ts, l is the length of the pendulum, A1 the distance between the galvanometer’s mirror and recording drum, k the transfer factor describing the galvanometer’s sensitivity. For both types the measured amplitude A after removing the gain V w, is true ground displacement (in mm). _______________________________________________________________________________________

Bibliography

   Abercrombie, R. E., 1994. Regional bias in estimates of earthquake MS due to surface-wave path effects, Bull. Seism. Soc. Am., 84, 377-382.

   Aki, K. & Richards, P. G., 2002. Quantitative seismology, University Science Book, Sausalito, California.

   Allegretti, I., Herak, D., Herak, M., & Duda, S. J., 2000. Consistency of magnitudes of the largest earthquakes of the first half of the 20th century determined on the basis of Göttingen and Zagreb seismograms, Studia geoph. et geod, 44, 13-25.

   Ambraseys, N., 2003. Reappraisal of magnitude of 20th century earthquakes in Switzerland, J. Earthq. Eng., 2, 149-191.

   Ambraseys, N. & Douglas, J., 2000. Reappraisal of surface wave magnitudes in the Eastern Mediterranean region and the Middle East, Geophys. J. Int., 141, 357-373.

   Ambraseys, N. & Free, M., 1997. Surface wave magnitude calibration for European region earthquakes, J. Earthq. Eng., 1, 1-22.

   Arvidsson, R. & Ekstrom, G., 1998. Global CMT Analysis of moderate earthquakes, Mw > 4.5, using intermediate-period surface waves, Bull. Seism. Soc. Am., 88, 1003-1013.

   Baer, M., Deichmann, N., Braunmiller, J., Bernardi, F., Cornou, C., Fäh, D., Giardini, D., Huber, S., Kästli, P., Kind, F., Kradolfer, U., Mai, M., Maraini, S., Oprsal, I., Schler, T., Schorlemmer, D., Sellami, S., Steimen, S., Wiemer, S., Wössner, J., & Wyss, A., 2003. Earthquakes in Switzerland and surrounding regions during 2002, Eclogae Geol. Helv, 96, 313-324.

   Bernardi, F., Braunmiller, J., Kradolfer, U., & Giardini, D., 2004. Automatic regional moment tensor inversion in the European-Mediterranean region, Geophys. J. Int., 156, 1-14.

   Bono, A., Simoni, B. D., Megna, A., Nardi, A., & Palombo, B., 2002. SISMOS project: recovering and reprocessing of historical seismic data, ESC XXVIII general assembly, Genova.

   Boschi, L., Ekstrm, G., & Kustowski, B., 2004. Multiple resolution surface wave tomography: the Mediterranean basin, Geophys. J. Int., 157, 293-304.

   Bossu, R., Mazet-Roux, G., Tome, M., Carreno, E., Guilbert, J., Menechal, Y., Schindele, F., & di Giovambattista, R., 2002. EMSC early warning system and real time seismicity, EMSC Newsletter, 18, 2-5.

   Braunmiller, J., 1998. Seismotectonics of the Explorer Region and of the Blanco Transform Fault Zone, Ph.D. thesis, Oregon State University.

   Braunmiller, J., 2001. An European-Mediterranean regional moment tensor catalog - year 2000, EOS, Trans., (82), Fall Meet. Suppl., Abstract S11B-0570.

   Braunmiller, J., 2002. Moment tensor solutions of stronger earthquakes in Germany with GRSN data, in Ten years of German regional seismic network (GRSN), pp. 227-235, Deutsche Forschungsgemeinschaft.

   Braunmiller, J., Nábe  lek, J., Leitner, B., & Qamar, A., 1995. The 1993 Klamath Falls, Oregon earthquake sequence: source mechanisms from regional data, Geophys. Res. Lett., 22, 105-108.

   Braunmiller, J., Kradolfer, U., Baer, M., & Giardini, D., 2000. Regional moment-tensor inversion in the European-Mediterranean area, Orfeus Newsletter, 2, 5.

   Braunmiller, J., Kradolfer, U., Baer, M., & Giardini, D., 2002. Regional moment tensor determination in the European-Mediterranean region - initial results, Tectonophysics, 356, 5-22.

   Braunmiller, J., Deichmann, N., Giardini, D., & Wiemer, S., 2004. Homogeneous moment magnitude calibration in Switzerland, accepted by Bull. Seism. Soc. Am..

   Cotte, N. & Laske, G., 2002. Testing group velocitiy maps for Eurasia, Geophys. J. Int., 150, 639-650.

   Dahlen, F. A. & Tromp, J., 1998. Theoretical global seismology, Princeton University Press, Priceton, New Jersey.

   Davis, S. D. & Fröhlich, C., 1995. A comparison of moment tensor solutions in the Harvard CMT and USGS catalogs, EOS, 76, F381.

   Deichmann, N., 1992. Structural and rheological implications of lower-crustal earthquakes below northern Switzerland, Phys. Earth planet. Int., 69, 270-280.

   Deichmann, N., Baer, M., Braunmiller, J., Ballarin, D., Bay, F., Bernardi, F., Delouis, B., Fäh, D., Gerstenberger, M., Giardini, D., Huber, S., Kradolfer, U., Maraini, S., Oprsal, I., Schibler, R., Schler, T., Sellami, S., Steimen, S., Wiemer, S., Wössner, J., & Wyss, A., 2002. Earthquakes in Switzerland and surrounding regions during 2001, Eclogae Geol. Helv, 95, 249-261.

   Dreger, D. S. & Helmberger, D. V., 1993. Determination of source parameters at regional distances with three-component sparse network data, J. geophys. Res., 98, 8107-8125.

   Dreger, D. S., Ritsema, J., & Pasyanos, M., 1995. Broadband analysis of the 21 September, 1993, Klamath Falls earthquake sequence, Geophys. Res. Lett., 105, 997-1000.

   Dziewonski, A. M. & Anderson, D. L., 1981. Preliminary reference Earth model, Phys. Earth planet. Int., 25, 297-356.

   Dziewonski, A. M. & Woodhouse, J. H., 1983. An experiment in systematic study of global seismicity: centroid-moment tensor solutions for 201 moderate and large earthquakes of 1981, J. geophys. Res., 88, 3247-3271.

   Dziewonski, A. M., Chou, T.-A., & Woodhouse, J. H., 1981. Determination of earthquake source parameters from waveform data for studies of global and regional seismicity, J. geophys. Res., 86, 2825-2852.

   Ekström, G., 1987. A broad band method of earthquake analysis, Ph.D. thesis, Harvard University, Cambridge, MA.

   Ekström, G. & Dziewonski, M., 1988. Evidence of bias in estimations of earthquake size, Nature, 332, 319-323.

   Evernden, J. F., 1971. Variation of Rayleigh-wave amplitude with distance, Bull. Seism. Soc. Am., 61, 231-240.

   Fäh, D., Giardini, D., Bay, F., Bernardi, F., Braunmiller, J., Deichmann, N., Furrer, M., Gantner, L., Gisler, M., Jimenez, D. I. M., Kästli, P., Koglin, R., Masciadri, V., Rutz, M., Scheidegger, C., Schibler, R., Schorlemmer, D., Schwarz-Zanetti, G., Steimen, S., Sellami, S., Wiemer, S., & Wössner, J., 2003. Earthquake Catalog Of Switzerland (ECOS) and the related macroseismic database, Eclogae geol. Helv., 96, 219-236.

   Fröhlich, C., 1994. Earthquakes with non-double-couple mechanisms, Science, 264, 804-809.

   Fröhlich, C., Millard, F. C., Massel, C., & Mann, P., 1997. Constraints on Macquarie tectonics provided by Harvard focal mechanisms and teleseismic earthquake locations, J. geophys. Res., 102, 5029-5041.

   Galitzin, F. B., 1914. Vorlesungen über Seismometrie, B.G. Teuner.

   Giardini, D., 1992. Moment tensor inversion from MEDNET data (1) large worldwide earthquakes of 1990, Geophys. Res. Lett., 19, 713-716.

   Giardini, D., Boschi, E., & Palombo, B., 1993a. Moment tensor inversion from MEDNET data (2) regional earthquakes of the Mediterranean, Geophys. Res. Lett., 20, 273-276.

   Giardini, D., Palombo, B., & Boschi, E., 1993b. The determination of earthquake size and source geometry in the Mediterranean Sea, in: Recent evolution and seismicity of the Mediterranean area, pp. 213-238, Kluwer.

   Gutenberg, B., 1945. Amplitudes of Rayleigh-waves and magnitude of shallow earthquakes, Bull. Seism. Soc. Am., 35, 3-12.

   Hagiwara, T., 1958. A note on the theory of the electromagnetic seismograph, Bull. Earth. Res. Inst., 36, 139-164.

   Hanks, T. C. & Kanamori, H., 1979. A moment magnitude scale, J. geophys. Res., 84, 2348-2350.

   Helffrich, G. R., 1997. How good are routinely determined focal mechanisms? Empirical statistics based on a comparison of Harvard, USGS and ERI moment tensors, Geophys. J. Int., 131, 741-750.

   Henry, C., Woodhouse, J. H., & Das, S., 2002. Stability of earthquake moment tensor inversions: effect of the double-couple constraint, Tectonophysics, 356, 115-124.

   Herak, M. & Herak, D., 1993. Distance dependence of MS and calibrating function for 20 second Rayleigh waves, Bull. Seism. Soc. Am., 83, 1881-1892.

   Herak, M., Allegretti, I., Herak, D., & Duda, S. J., 1996. Calibration of the Wiechert seismograph relative to a reference seismometer, Geofisika, 13, 31-59.

   IASPEI, 1967. Recommendations of the comittee on magnitudes, Comptes Rendus, 15, 65.

   Jackson, J. A. & McKenzie, D., 1988. The relationship between plate motions and seismic moment tensor, and the rates of active deformation in the Mediterranean and Middle East, Geophys. J., 93, 45-73.

   Jenny, S., Goes, S., Giardini, D., & Kahle, H.-G., 2004. Earthquake recurrence parameters from seismic and geodetic strain rates in the eastern Mediterranean, Geophys. J. Int., 157, doi:10.1111/j.1365-246X.2004.02261.x.

   Jost, M. L. & Herrmann, R. B., 1989. A student’s guide and review to moment tesnsors, Seism. Res. Lett., 60, 37-57.

   Kanamori, H., 1977. The energy release in great earthquakes, J. geophys. Res., 82, 2981-2987.

   Kanamori, H. & Anderson, D. L., 1975. Theoretical basis of some empirical relations in seismology, Bull. Seism. Soc. Am., 65, 1073-1095.

   Kárník, V., Kondorskaya, N., Riznitchenko, J., Savarensky, E., Soloviev, S., Shebalin, N., Vane  k, N., & Zátopek, A., 1962. Standardisation of the earthquake magnitude scale, Studia Geophys. Geodet., 6, 42-47.

   Kastrup, U., Zoback, M. L., Deichmann, N., Evans, K. F., Giardini, D., & Michael, A. J., 2004. Stress field variations in the Swiss Alps and northern Alpine foreland derived from inversion of fault plane solutions, J. geophys. Res., B01402, doi:10.1029/2003JB002550.

   Kawakatsu, H., 1995. Automated near-realtime CMT inversion, Geophys. Res. Lett., 22, 2569-2572.

   Kennett, B. & Engdahl, E., 1991. Traveltimes for global earthquake location and phase identification, Geophys. J. Int., 105, 429-465.

   Knopoff, L. & Randall, M. J., 1970. The compensated linear-vector dipole: A possible mechanism for deep earthquakes, J. geophys. Res., 75, 4957-4963.

   Komatitsch, D. & Tromp, J., 2002. TSpectral-element simulations of global seismic wave propagation - II. Three-dimensional models, oceans, rotation and self-gravitation, Geophys. J. Int., 150, 303-318.

   Kowalle, G. & Thürmer, C., 1991. Magnitudenbestimmung für die Wiechert-Seismogramme der Station Potsdam, Zentralinstitut für Physik der Erde, Potsdam.

   Kradolfer, U., 1996. AutoDRM - the first five years, Seis. Res. Lett., 67, 30-33.

   Kubo, A., Fukuyama, E., & Nonomura, K., 2002. NIED seismic moment tensor catalogue for regional earthquakes around Japan: quality test and application, Tectonophysics, 356, 23-48.

   Kuge, K. & Kawakatsu, H., 1993. Significance of non-double couple components of deep and intermediate-depth earthquakes: implications from moment tensor inversion of long-period seismic waves, Phys. Earth planet. Int., 75, 243-266.

   Kuge, K. & Lay, T., 1994. Data-dependent non-double-couple components of shallow earthquake source mechanisms: Effects of waveform inversion instability, Geophys. Res. Lett., 21, 9-12.

   Kunze, T., 1982. Seismotektonische Bewegungen im Alpenbereich, Ph.D. thesis, Universität Stuttgart.

   Larson, E. W. & Ekström, G., 2001. Global models of surface wave group velocity, Pure and Applied Geophysics, 158, 1377-1399.

   Mainka, C., 1923. Physik der Erdbebenwellen, Verlag Gebrüder Borntraeger, Berlin.

   Marone, F., van der Meijde, M., van der Lee, S., & Giardini, D., 2003. Joint inversion of local, regional and teleseismic data for crustal thickness in the Eurasia-Africa plate boundary region, Geophys. J. Int., 154, 499-514.

   Marone, F., van der Lee, S., & Giardini, D., 2004. Shallow anisotropy in the Mediterranean mantle from surface waves, Geophys. Res. Lett., 31, L06624, doi:10.1029/2003GL018948.

   Maset-Roux, G. & Bossu, R., 2004. Report on emsc operational activities, Tech. rep., European-Mediterranean Seismological Centre.

   Nábe  lek, J. & Xia, G., 1995. Moment-tensor analysis using regional data: application to the 25 March, 1993, Scotts Mills, Oregon, earthquake, Geophys. Res. Lett., 22, 13-16.

   Panza, G. F., Duda, S. J., Cernobori, L., & Herak, M., 1989. Gutenberg’s surface-wave magnitude calibrating function: theoretical basis from synthetic seismograms, Tectonophysics, 166, 35-43.

   Pasyanos, M. E., Dreger, D. S., & Romanowicz, B., 1996. Toward real-time estimation of regional moment tensors, Bull. Seism. Soc. Am., 86, 1255-1269.

   Pasyanos, M. E., Walter, W. R., & Hazler, S. E., 2001. A surface wave dispersion study of the Middle East and North Africa for monitoring the comprehensive Nuclear-Test-Ban Treaty, Pure and Applied Geophysics, 158, 1445-1474.

   Patton, H. & Aki, K., 1979. Bias in the estimate of seismic moment tensor by the linear inversion method, Geophys. J. R. astr. Soc., 59, 479-495.

   Peterson, J., 1993. Observations and modeling of seismic background noise., U.S. Departement of Interior, Geological Survey, pp. Open-File Report 93-322.

   Pondrelli, S., Morelli, A., Ekström, G., Mazza, S., Boschi, E., & Dziewonski, A. M., 2002. European-Mediterranean regional centroid-moment tensors: 1997-2000, Phys. Earth planet. Int., 130, 71-101.

   Rezapour, M. & Pearce, R. G., 1998. Bias in surface-wave magnitude MS due to inadequate distance corrections, Bull. Seism. Soc. Am., 88, 43-61.

   Ritsema, J. & Lay, T., 1993. Rapid source mechanism determination of large (Mw > 4.5) earthquakes in western United States, Geophys. Res. Lett., 20, 1611-1614.

   Ritter, J. R. R., 2001. On the recording characteristics of the original Wiechert seismographs at Göttingen (Germany), Journal of Seismology, 6, 477-486.

   Romanowicz, B., Dreger, D., Pasyanos, M., & Uhrhammer, R., 1993. Monitoring strain release in central and northern California using broadband data, Geophys. Res. Lett., 20, 1643-1646.

   Sicilia, D., 1999. Towards regional moment tensor determination: the moment tensor of the 1994 mb = 4.8 event near the Balearic Islands and its uncertainties, Diploma thesis, ETH Zürich.

   S  ílený, J. & Vavryc  uk, V., 2002. Can unbiased source be retrieved from anisotropic waveforms by using an isotropic model of the medium?, Tectonophysics, 356, 125-138.

   Sipkin, S. A., 1982. Estimation of earthquake source parameters by the inversion of waveform data: synthetic waveforms, Phys. Earth planet. Int., 30, 242-259.

   Sipkin, S. A., 1986. Estimation of earthquake source parameters by the inversion of waveform data: global seismicity, 1981-1983, Bull. Seism. Soc. Am., 76, 1515-1541.

   Soloviev, S. L., 1955. O klassifikatsii zemletrasenii po velichine ikh energii, Trudi Geofiz. Inst. Akad. Nauk SSSR, 30, 3-31.

   Stahel, W. A., 1995. Statistische Datenanalyse, Friedr. Vieweg & Sohn Verlagsgesellschaft mbH, Braunschweig/Wiesbaden.

   Stich, D., Ammon, C. J., & Morales, J., 2003. Moment tensor solutions for small and moderate earthquakes in the Ibero-Maghreb region, J. geophys. Res., 108(B3), 2148, doi:10.1029/2002JB002057.

   Thio, H.-K. & Kanamori, H., 1995. Moment tensor inversion for local earthquakes using surface waves recorded at TERRAscope, Bull. Seism. Soc. Am., 85, 1021-1038.

   van der Lee, S., 1998. Observation and origin of Rayleigh-wave amplitude anomalies, Geophys. J. Int., 135, 691-699.

   van Eck, T., Dost, B., & Hanka, W., 2001. European scale real-time waveform data exchange, EMSC Newsletter, 17, 11.

   van Eck, T., Trabant, C., Dost, B., Hanka, W., & Giardini, D., 2004. Setting up a virtual Broadband seismograph network acorss Europe, EOS, 85, 125,129.

   Vane  k, N., Zátopek, A., Kárník, V., Kondorskaya, N., Riznitchenko, J., Savarensky, E., Soloviev, S., & Shebalin, N., 1962. Standardisation of magnitude scales, Bull. Acad. Sci. USSR, Geophys. Ser., 2, 108-111.

   von Seggern, D., 1977. Amplitude-distance for 20-second Rayleigh waves, Bull. Seism. Soc. Am., 67, 405-411.

   Wessel, P. & Smith, W. H. F., 1995. New version of the generic mapping tool released, Trans. Am. Geophys. Union, p. 329.

   Wiechert, E., 1903. Theorie der automatischen Seismographen, Weidmannsche Buchhandlung.

   Willmore, P. L., 1979. Manual of seismological observatory practice, World Data Center A for Solid Earth Geophysics.

   Woodhouse, J. H., 1988. The calculation of eigenfrequencies and eigenfunctions of the free oscillations of the Earth and the Sun, in: Seismological algorithms, pp. 321-370, Academic Press.

   Zhang, J. & Lay, T., 1990. Effects of centroid location on determination of earthquake mechanisms using long-period surface waves, Bull. Seism. Soc. Am., 80, 1205-1231.

_______________________________________________________________________________________

Curriculum Vitae

Personal

Date of birth

30 June 1973

Place of birth

Lugano, Switzerland

Citizen of

Stabio, Switzerland

Education

2000-2004

PhD. student at institute of Geophysics at the Swiss Federal Institute of Technology in Zürich, Switzerland

1994-2000

Study of Earth Sciences with major in Geophysics at the Swiss Federal Institute of Technology in Zürich, Switzerland

1988-1992

High School in Mendrisio, Switzerland