9

Click here to load reader

Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Embed Size (px)

Citation preview

Page 1: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

2006) 137–145www.elsevier.com/locate/tecto

Tectonophysics 423 (

Earthquake activity related to seismic cycles in a modelfor a heterogeneous strike-slip fault

Gert Zöller a,⁎, Sebastian Hainzl b,1, Yehuda Ben-Zion c,2, Matthias Holschneider d,3

a Institute of Physics and Institute of Mathematics, University of Potsdam, Potsdam, Germanyb Institute of Geosciences, University of Potsdam, Potsdam, Germany

c Department of Earth Sciences, University of Southern California, Los Angeles, USAd Institute of Mathematics, University of Potsdam, Potsdam, Germany

Received 16 March 2005; received in revised form 27 September 2005; accepted 25 March 2006Available online 9 May 2006

Abstract

We investigate the evolution of seismicity within large earthquake cycles in a model of a discrete strike-slip fault in elastic solid.The model dynamics is governed by realistic boundary conditions consisting of constant velocity motion of regions around thefault, static/kinetic friction and dislocation creep along the fault, and 3D elastic stress transfer. The fault consists of brittle partswhich fail during earthquakes and undergo small creep deformation between events, and aseismic creep cells which arecharacterized by high ongoing creep motion. This mixture of brittle and creep cells is found to generate realistic aftershocksequences which follow the modified Omori law and scale with the mainshock size. Furthermore, we find that the distribution ofinterevent times of the simulated earthquakes is in good agreement with observations. The temporal occurrence, however, ismagnitude-dependent; in particular, the small events are clustered in time, whereas the largest earthquakes occur quasiperiodically.Averaging the seismicity before several large earthquakes, we observe an increase of activity and a broadening scaling range ofmagnitudes when the time of the next mainshock is approached. These results are characteristics of a critical point behavior. Thepresence of critical point dynamics is further supported by the evolution of the stress field in the model, which is compatible withthe observation of accelerating moment release in natural fault systems.© 2006 Elsevier B.V. All rights reserved.

Keywords: Earthquake dynamics; Fault models; Seismicity; Ruptures; Heterogeneities; Criticality

⁎ Corresponding author. Tel.: +49 331 977 1175; fax: +49 331 9771142.

E-mail addresses: [email protected] (G. Zöller),[email protected] (S. Hainzl), [email protected](Y. Ben-Zion), [email protected] (M. Holschneider).1 Tel.: +49 331 977 5078; fax: +49 331 977 5060.2 Tel.: +1 213 740 6734; fax: +1 213 740 8801.3 Tel.: +49 331 977 1663; fax: +49 331 977 1578.

0040-1951/$ - see front matter © 2006 Elsevier B.V. All rights reserved.doi:10.1016/j.tecto.2006.03.007

1. Introduction

The complexity of spatiotemporal earthquake occur-rence on natural fault systems has been discussed andquantified in numerous studies (for an overview seeRundle et al., 2000b). While some seismicity patternsshow an almost universal behavior, others are observedless frequently and follow no obvious law. For example,aftershocks according to the modified Omori law areobserved after almost all large earthquakes in continentallithosphere (Utsu et al., 1995; Utsu, 2002), whereas

Page 2: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

138 G. Zöller et al. / Tectonophysics 423 (2006) 137–145

foreshocks are a rare phenomenon (Wyss, 1997).Moreover, the Parkfield experiment in California (Roel-offs and Langbein, 1994) demonstrates that even if a faultsegment shows relatively regular behavior over severaldecades, a prediction of future activity can still fail. Onereason for this is that the available data sets are too limitedto provide enough information for the understanding ofthe complex relationships between the various physicalmechanisms in the earth's crust. This emphasizes theimportance of model simulations which cover manyseismic cycles and allow to study the dependences ofseismicity on the underlying parameters and evolvingstress field.

Several conceptual models have been proposed todescribe properties of observed seismicity, e.g. spring-block models and cellular automata, which reproduce thefrequency–size distribution of earthquakes. These modelsare mainly based on tectonic loading and the coseismicstress transfer (Burridge and Knopoff, 1967; Bak andTang, 1989). Others have implemented additional mecha-nisms, like viscoelastic relaxation in the fault zone(Dieterich, 1972; Hainzl et al., 1999), fault strengtheningor weakening after a block slip (Ito andMatsuzaki, 1990),pore fluid flow (Nur and Booker, 1972), rate-state friction(Dieterich, 1994) and damage rheology (Ben-Zion andLyakhovsky, 2006; Shcherbakov and Turcotte, 2004).Although most of these models reproduce certainobserved phenomena, the underlying frameworks areoften abstract and the application to an actual fault systemwith its spatiotemporal complexity of earthquake occur-rence remains questionable.

In this study, we analyze earthquake catalogs gen-erated by the model of Zöller et al. (2005a) for a largeheterogeneous strike-slip fault. The model developsfurther the framework of Ben-Zion and Rice (1993) andBen-Zion (1996) for a discrete fault in a surroundingelastic solid, and it combines computational efficiencywith realistic physical properties. The simulations cover1000s of years and allow us to reproduce brittle as wellas postseismic and interseismic creep deformation. Themodel ingredients, including laws for brittle and creepdeformation, stress transfer, and boundary conditions,are compatible with empirical knowledge. Several rela-tionships between imposed model parameters (e.g.,frictional parameters, creep velocities, spatial hetero-geneities) and observed seismicity quantities like fre-quency–size distributions and aftershock clustering,have been previously quantified (Zöller et al., 2004,2005a,b). In this paper, we will focus on the temporalevolution of seismicity during seismic cycles. Manyobservational and theoretical studies have shown thatseismicity is largely time-dependent. In particular, the

seismic moment release appears to accelerate prior tosome large earthquakes (Bufe and Varnes, 1993;Bowman et al., 1998; Jaumé and Sykes, 1999). Thisphenomenon suggests that seismicity is a process thatmay be characterized by critical point behavior in termsof upcoming long-range correlations prior to a largeearthquake (Sornette and Sornette, 1990; Zöller et al.,2001; Zöller and Hainzl, 2002). Analyzing several stressand seismicity functions over seismic cycles, Ben-Zionet al. (2003) conclude that mainshock occurrence isassociated with a period where the stress-field evolvestoward a critical level of disorder. In this state, the stressfield heterogeneities are characterized by many sizescales and a brittle failure can evolve into a large earth-quake. It is, therefore, an important question whether theevolution of stress and seismicity in our model reflectssuch critical point behavior.

2. Model

In this section, we give a brief description of the faultmodel. Further details can be found in Zöller et al.(2004, 2005a,b) and references therein.

Our model includes a single rectangular faultembedded in a 3D elastic half space (Fig. 1). A faultregion of 70 km length and 17.5 km depth is covered by acomputational grid, divided into 128×32 uniform cells,where deformational processes are calculated. Tectonicloading is imposed by a motion with constant velocityvpl=35mm/year of the regions around the computationalgrid. The space-dependent loading rate provides realisticboundary conditions. Using the static stress transferfunction for slip in elastic solid, the continuous tectonicloading for each cell on the computational grid is a linearfunction of time and plate velocity vpl. Additionalloadings on a given cell occur due to brittle and creepfailures on the fault. While the loading produces anincrease of stress on the fault, the local stress may bereduced by interseismic creep and coseismic slip. As inZöller et al. (2005a), all cells on the computational gridundergo creep deformation, while brittle deformationduring an earthquake is only possible for “brittle cells”(white cells in Fig. 1). The aseismic “creep cells” (blackcells in Fig. 1) are generated by near-vertical randomwalks from the free surface to depth accounting foroffsets and step-over regions in the fault zone.

The ongoing creep motion on the grid is implemen-ted by a local constitutive law corresponding to lab-based dislocation creep (Ben-Zion, 1996):

u:creepðx; z; tÞ ¼ cðx; zÞsðx; z; tÞ3; ð1Þ

Page 3: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Fig. 1. A sketch of a 2D strike-slip fault in a 3D elastic half space.

139G. Zöller et al. / Tectonophysics 423 (2006) 137–145

where ucreep (x, z, t) is the creep velocity of the cell withcoordinates (x, z); τ(x, z, t) is the local stress at time t,and c(x, z) are time-independent coefficients, which aredifferent for the creep cells and the brittle cells. Asshown in Zöller et al. (2005a), this design can produceaftershock sequences, where the ratio of the mean creepcoefficients in the creep cells and in the brittle cells ⟨ccr⟩/⟨cb⟩ determines the exponent p of the modified Omorilaw for the rate n˙(t) of aftershocks:

n: ðtÞ ¼ c1

ðc2 þ t−tMÞp ð2Þ

In (2), c1 and c2 are time-independent numbers, andtM is the mainshock occurrence time. For this work, wechoose creep coefficients leading to p=1: The creepparameters for the brittle cells cb are randomly dis-tributed in the interval [0.9⟨cb⟩; 1.1⟨cb⟩], where ⟨cb⟩=10−7 ms−1 MPa−3. The values for the creeping cells ccrare calculated accordingly with ⟨ccr⟩=10

− 2 ms− 1

MPa−3. These choices correspond to model A in Zölleret al. (2005a). Eq. (1) results in a system of 128×32coupled ordinary differential equations, which is solvednumerically using a Runge–Kutta scheme.

The coseismic processes are governed by static/kinetic friction. An earthquake is initiated if the localstress τ(x, z; t) exceeds the static friction τs(x, z). Thenthe stress drops in cell (x, z) to the arrest stress τa(x, z)and the strength drops to the lower dynamic value τd.The brittle failure envelope τf(x, z, t) during anearthquake is thus given by

sf ðx; z; tÞ ¼ssðx; zÞ : cell ðx; zÞ failed not during this event

sdðx; zÞ : cell ðx; zÞ already failed during this event:

(3)

If τ(x, z; t)bτf (x, z; t) for all cells on the grid, theearthquake is terminated. Then, the brittle strength

recovers back to the static level for all cells: τf (x, z; t)=τs(x, z). The dynamic friction is calculated from the staticand arrest stress levels in relation to a dynamic over-shoot coefficient D:

sd ¼ ss−ss−saD

: ð4Þ

Following Madariaga (1976), we use D=1.25. Thestatic strength is constant, τs(x, z)≡10MPa, and the arreststress is chosen randomly from the interval τa(x, z)∈[0; 1 MPa].

A quasidynamic stress transfer due to coseismic slipand creep motion is calculated by means of the three-dimensional solution K(x, z; x′, z′) of Chinnery (1963)for static dislocations on rectangular patches in anelastic Poisson solid with rigidity μ=30 GPa:

Dsðx; z; tÞ ¼X

ðx V;z VÞagrid

Kðx; z; x V; z VÞDuðx V; z V; t−r=sÞ;

ð5ÞwhereΔu is the slip, r is the spatial distance between thecells (x, z) and (x′, z′), vs is a constant shear wavevelocity, and the kernel K decays like 1 / r3. The value ofvs defines the event time scale but has no influence onthe earthquake catalogs, as long as vsN0.

3. Results

We analyze an earthquake catalog covering about5000 years, which contains 200,000 earthquakes withmoment magnitudes M between 4.0 and 6.8. The seismicpotencyP0 (momentM0 divided by rigidityμ) is calculatedas the integral of slip over the rupture area during anearthquake. The range of magnitudes depends on thesegmentation of the grid into computational cells, e.g.smaller cells would allow to simulate smaller earthquakes.

Page 4: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Fig. 2. Example of the model seismicity in a time span of 60 years.

140 G. Zöller et al. / Tectonophysics 423 (2006) 137–145

On the other hand, a larger fault would result in a highermaximum magnitude. Fig. 2 shows an earthquakesequence (magnitude vs. time) from the catalog over aperiod of 60 years.

3.1. Aftershocks

The most obvious feature in Fig. 2 is the aftershockclustering following the largest events. As mentionedabove, the aftershock decay is compatible with themodified Omori law, Eq. (2), with an exponent p=1.Furthermore, it has been demonstrated that the after-shocks are predominantly concentrated at the margins ofthe fault segments (Zöller et al., 2005a), which is goodagreement with observational studies.

For the following investigation, we define a main-shock to be the largest event within ±1 month and anaftershock to be an earthquake occurring up to onemonth after the mainshock somewhere on the compu-tational grid. An important question is the dependence

Fig. 3. The number of aftershocks in the first month after a mainshock (foraftershock productivity is found to scale with 10

23Mmain (panel a) and with the

of the aftershock sequences on the mainshock magni-tude Mmain, expressed by the value of c1 in Eq. (2). Theexponent p is found to be almost independent of Mmain

in agreement with Utsu (1962, 2002). For the rate ofaftershocks, Reasenberg (1985) assumes the relation

c1f1023dMmain : ð6Þ

The number of aftershocks as a function of themainshock magnitude is given in Fig. 3a. It is clearlyvisible that the scaling law Eq. (6) is fulfilled in ourmodel. A second scaling relation is observed, if thenumber of aftershocks is plotted as a function of themainshock rupture area, leading to NA∝A1/2 and thus toNA∝R, where R is the rupture length (Fig. 3b). Com-bining both scaling laws, results in a relation P0 (orM0)∝A9/8 between seismic potency (or moment) andrupture area A of the mainshock. Plotting P0 directlyversus A of the simulated earthquakes shows (Fig. 4)that the slope of the log (P0) versus A for the small

a definition see text) as a function of the mainshock magnitude. Therupture length R∝

ffiffiffiA

p(panel b).

Page 5: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Fig. 4. Relation between potency P0 and rupture area A. The scaling forsmall events is P0∝A and for large events P0∝A9/8.

Fig. 5. The normalized interevent time distribution of the modelsimulations (black dots) compared with the result of Corral (2004) andthe distribution of earthquakes in California (ANSS catalog of M≥3earthquakes occurred between 1970 and 2004 within 29° and 43°latitude and −113° and −123° longitude).

141G. Zöller et al. / Tectonophysics 423 (2006) 137–145

events is about 1 and that the slope increases for thelarger events (P0∝A9/8). This is compatible withsimulation results of Ben-Zion and Rice (1993) andhigh resolution observations (Ben-Zion and Zhu, 2002).A scaling relation between the log (P0) and area Awith aslope close to 1 is expected for earthquakes in a dynamicregime close to criticality (Fisher et al., 1997).

3.2. Interevent times

In recent studies, it has been suggested that the dis-tribution of interevent times can be described by a uni-versal law. In particular, the distributions from differenttectonic environments, different spatial scales (fromworldwide to local seismicity) and different magnituderanges collapse, if the time Δt is rescaled with the rateRxy of seismic occurrence in a region denoted by (x, y)(Corral, 2004):

Dx;yðDtÞ ¼ Rxyd f ðRxyDtÞ; ð7Þwhere Dxy is the probability density for the intereventtime Δt, and f can be expressed by a generalized gammadistribution

f ðhÞ ¼ C1

hg−1exp −

hd

B

!ð8Þ

with parameters C, γ, δ, and B, which have beendetermined by a fit to several observational catalogs(Corral, 2004).

In Fig. 5, we compare Dxy(Δt) from Eq. (7) with twoearthquake catalogs: (1) the ANSS catalog of California(catalog ranges are given in the caption) and (2) themodel catalog. Due to the universality of Eq. (7) withrespect to different spatial scales, the comparison of ourmodel simulating a single fault of 70 km length with aregion of hundreds of kilometers including several faults

in California is straightforward without coarse grainingthe ANSS catalog. We note that due to the numericalprocedure, especially the Runge–Kutta integration ofthe creep law (Eq. (1)), the interevent times in the modelhave a finite lower limit. However, in the region wherethe interevent times are calculated, the agreement of thethree curves is remarkable. For small values of Δt, Eq.(7) deviates from the California data; for high values themodel has a slightly better correspondence with theobservational data than Eq. (7). Thus our results gene-rally support the recent findings of Corral (2004). Adetailed analysis of the small deviations from Eq. (7),and their possible origin will be addressed by means of arefined parameter space study and additional statisticaltesting in future studies.

The degree of temporal clustering of earthquakes canbe estimated by the coefficient of variation CV of theinterevent time distribution

CV ¼ r=l; ð9Þwhere σ is the standard deviation and μ the mean valueof the interevent time distribution. High values of CVdenote clustered activity, while low values representquasiperiodic occurrence of events. The case CV=1corresponds to a random Poisson process. Ben-Zion(1996) and Zöller et al. (2005b) have found that theclustering properties of the large events depend on thedegree of quenched spatial disorder of the fault. Here weshow that CV as a function of the lower magnitudecutoff has a characteristic shape (Fig. 6). The values ofCVare higher than 1 (clustered) for small and intermediateearthquakes (M≤5.4) and smaller than 1 (quasiperiodic)

Page 6: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Fig. 6. The temporal earthquake occurrence quantified by thecoefficient of variation as a function of the lower magnitude cutoff.Values larger than 1 indicate clustering, whereas lower values point toquasiperiodic behavior.

142 G. Zöller et al. / Tectonophysics 423 (2006) 137–145

for larger earthquakes. This corresponds to the case of alow degree of disorder in Zöller et al. (2005b), because thebrittle cells which participate in an earthquake have nosignificant spatial disorder. We note that this behaviorresembles the seismicity on the Parkfield segment of theSan Andreas fault, which is characterized by a quasipe-riodic occurrence of mainshocks.

A different behavior is observed on the San Jacintofault in California, where the largest events occur lessregularly and have overall smaller magnitudes. As dis-cussed in Ben-Zion (1996) and Zöller et al. (2005b), thiscan be modeled by imposing higher degrees of disorderleading to a broader range of spatial size scales, e.g. byusing a higher number of near-vertical barriers. Whilebarriers provide a simple and physically motivated way

Fig. 7. The stacked and averaged activity prior to mainshocks (225sequences). Here, mainshocks are MN6 events which are the largestearthquakes within ±10 years.

to tune the degree of disorder, other types of hetero-geneities may work as well, as long as they are able toproduce strong enough variations of the brittle faultstrength.

3.3. Foreshocks

In Zöller et al. (2005a) and in Section 3.1, it is shownthat our model produces aftershock sequences, and thatthe parameter space of the model includes regions wherethese sequences are quantitatively in good agreementwith observed aftershock activity. Foreshock activity is,however, less present in observed data. Although it isdocumented that the increase of activity prior to amainshock follows an “inverse Omori law” n˙(t)∼ (c+tM− t)− q with an exponent q≈1 (Hainzl et al., 1999),the overall smaller number of foreshocks requires verylong data sets including many mainshocks to detect aclear foreshock signal. Stacking the activity before 225mainshocks with M≥6, shows a moderate but clearlyvisible increase of the earthquake rate before a main-shock (Fig. 7). A stable estimation of the exponent q is,however, not possible from Fig. 7 due to the smallnumber of extra events. The observation of an increa-sing event rate is encouraging for the applicability of themodel to real fault zones.

3.4. Critical point behavior

The interpretation of seismicity in terms of criticalpoint dynamics has opened new perspectives for theanalysis and the understanding of earthquake data.

Fig. 8. The averaged activity and stress evolution relative tomainshocks (MN6 events which are the largest earthquakes within ±10 years).

Page 7: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

Fig. 10. The mean potency of earthquakes as a function of the stresslevel τ. The stress level is normalized to the maximum (max) andminimum (min) observed stress. The potency is found to increaseaccording to a power law Δτ−1.5, if stress approaches the maximum.

143G. Zöller et al. / Tectonophysics 423 (2006) 137–145

These include self-organized criticality (Bak and Tang,1989; Hainzl et al., 1999), growing spatial correlationlength (Zöller et al., 2001; Zöller and Hainzl, 2002) andaccelerating moment release (Bufe and Varnes, 1993;Jaumé and Sykes, 1999). A key question is whether theapproach to the critical point can be measured by stressor seismicity functions as addressed by Ben-Zion et al.(2003). Their analysis of catalogs from three end-member models leads to the conclusion that large earth-quakes are preceded by a period of criticality charac-terized by a highly disordered stress field with a broadrange of size scales. In this section, we search for signalsof criticality in our more realistic version of the model.

Fig. 8 (bottom) shows the average earthquake ratebefore and after a large event as a function of time, basedagain on stacking of 225 large event (M≥6) sequences.The top panel of Fig. 8 is the corresponding mean stress(normalized) on the computational grid. The mainshockitself is characterized by a significant stress drop. In thefollowing period of aftershocks, lasting for about1.2 years, the mean stress still decreases before it reco-vers and increases again. This stress increase is almostlinear and lasts until the next mainshock occurs.

In previous studies (Ben-Zion, 1996; Ben-Zion et al.,2003; Zöller et al., 2005b), it was demonstrated that thedegree of disorder of the stress field is correlated withthe frequency–size distribution of the seismicity: asmooth stress field produces characteristic earthquakedistributions with a frequently occurring characteristicevent, almost no intermediate earthquakes, and smallevents following a truncated Gutenberg–Richter distri-bution; in contrast, the seismicity from a disordered stress

Fig. 9. The frequency–magnitude distribution of all earthquakes,foreshocks and aftershocks, respectively. Foreshocks and aftershocksare defined to be earthquakes occurring within one month before andafter a mainshock, where the mainshock definition is the same way asbefore. The dotted lines refer to b-values of 1 and 2.

field is characterized by a Gutenberg–Richter distribu-tion over a broad range of magnitudes. Therefore, weexpect that the frequency–size distribution before amainshock has a broad scaling region, whereas theoverall seismicity follows a characteristic earthquakedistribution. Fig. 9 shows the cumulative frequency–sizedistributions for foreshocks, aftershocks, and all earth-quakes in our model. Due to the small number of fore-shocks, the corresponding curve is less smooth than thecurves for aftershocks and all events. However, it isclearly visible that the foreshock activity shows thesmallest deviations from a scaling law (denoted as solidlines) compared to the other curves. Although the overallfrequency–size distribution of earthquakes follows acharacteristic earthquake law, the periods before main-shocks are characterized by Gutenberg–Richter type sta-tistics, with a broad scaling range pointing to a disorderedstress field with a wide range of spatial size scales.

The second indication for critical point behavior isthe acceleration of seismic moment release, which isassumed to follow the power law

ðRXÞðtÞ ¼ Aþ Bðtf−tÞm; ð10Þwhere the cumulative Benioff strain (ΣΩ)(t) is calculat-ed from the moment releases M0(t′) of earthquakes attimes t′≤ t by ðRXÞðtÞ ¼ R t0 ffiffiffiffiffiffiffiffiffiffiffiffi

M0ðt VÞp

dt V. The time tf isthe mainshock occurrence time, and A, B, and m areparameters. Analyzing eight large earthquakes in Cali-fornia, Bowman et al. (1998) find values of the exponentm in the range 0.1≤m≤0.55. The theoretical analysis ofRundle et al. (2000a) of a critical point process leads tom=0.25. The observational study of Bufe and Varnes(1993) and the theoretical damage model of Ben-Zionand Lyakhovsky (2002) suggest the value m=0.3.

Page 8: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

144 G. Zöller et al. / Tectonophysics 423 (2006) 137–145

Due to the noisy character of single earthquake se-quences, it is difficult to detect accelerating momentrelease before individual mainshocks. Instead, we use amore robust representation by plotting the release ofpotency P0 as a function of the (normalized) stress levelon the fault in Fig. 10. We note that the stress increase onthe fault is linear (see Fig. 8). Consequently, the abscissarepresents effectively the time between two mainshocks.The curve shows a clear power law dependence of the(non-cumulative) potency release on the stress levelwith an exponent of s=−1.5. This results in a power lawfor the cumulative Benioff strain according to Eq. (10)with m=0.5 · s+1=0.25, which is in good agreementwith empirical and theoretical findings (Bufe andVarnes, 1993; Bowman et al., 1998; Rundle et al.,2000a; Ben-Zion and Lyakhovsky, 2002). In sum, ourresults are compatible with the hypothesis of accelerat-ing moment release and provide further support forthe presence of critical point dynamics in the simulatedseismicity.

4. Discussion and conclusions

In the present work, we have analyzed an earthquakecatalog from a recently developed model for a largeheterogeneous strike-slip fault. As shown in a previouswork, this model reproduces various characteristics ofobserved seismicity (Ben-Zion, 1996; Zöller et al.,2005a). Using different parameters, like frictional valuesand creep rates, the model can be tuned towardsobserved cases, e.g. a fault which produces clusteredseismicity obeying Gutenberg–Richter statistics or afault with quasiperiodically occurring mainshocksfollowing the characteristic earthquake distribution.Here, we have analyzed an earthquake catalog fromthe latter situation, which resembles many features ofseismicity on the Parkfield segment of the San Andreasfault in California, including quasiperiodically occur-ring mainshocks, which are followed by aftershocksequences. In contrast to short observational data setscovering some years or decades, we consider an obser-vational period of about 5000 years of simulated seis-micity. This allows us to use a stacking procedure inorder to unveil properties, which are not visible in shortand noisy data sets.

The stacking of many aftershock sequences shows ascaling relation between the mainshock magnitude andthe number of aftershocks. This gives in combinationwiththe modified Omori law an excellent explanation ofaftershock activity. In addition, the less frequentlyoccurring foreshock activity is also visible in our model.The temporal occurrence of earthquakes is quasiperiodic

for large events and clustered for intermediate and smallevents with an overall interevent time distribution thatfollows a recently proposed universal law, and is verysimilar to the interevent time distribution of Californiaseismicity. While these results demonstrate a high cor-respondence of the model output with observed seismic-ity, we also find support for the hypothesis that seismicityis associatedwith critical point dynamics. In particular, wehave shown that the scaling range of the frequency–sizedistribution becomes broader when the time of a main-shock is approached. This indicates that the stress field hasreached a certain degree of disorder, where a single failurecan evolve into a large event. Furthermore, the momentrelease accelerates in interseismic periods in good agree-ment with observational and theoretical results.

In sum, we have demonstrated that our modelprovides a very good reproduction of natural earthquakeactivity and gives new insights to various aspects of the“critical earthquake concept.” Continuing joint study ofmodel simulations and data associated with large faultzones may improve the ability to predict large earth-quakes in these faults.

Acknowledgments

We thank the Editor Stanislaw Lasocki and twoanonymous reviewers for comments which helped us toimprove the manuscript. The ANSS catalog has beenprovided by the Northern California Earthquake DataCenter (NCEDC) and the organizations that contributeddata to the NCEDC. G.Z. acknowledges support fromthe collaborative research center “Complex NonlinearProcesses” (SFB555) of the “German Research Society”(DFG).

References

Bak, P., Tang, C., 1989. Earthquakes as a self-organized criticalphenomenon. Geophys. Res. Lett. 94, 15635–15637.

Ben-Zion, Y., 1996. Stress, slip, and earthquakes in models of complexsingle-fault systems incorporating brittle and creep deformations.J. Geophys. Res. 101, 5677–5706.

Ben-Zion, Y., Lyakhovsky, V., 2002. Accelerated seismic release andrelated aspects of seismicity patterns on earthquake faults. PureAppl. Geophys. 159, 2385–2412.

Ben-Zion, Y., Rice, J.R., 1993. Earthquake failure sequences along acellular fault zone in a three-dimensional elastic solid containingasperity and nonasperity regions. J. Geophys. Res. 98, 14109–14131.

Ben-Zion, Y., Zhu, L., 2002. Potency–magnitude scaling relations forSouthern California earthquakes with 1.0≤ML≤7.0. Geophys.J. Int. 148, F1–F5.

Ben-Zion, Y., Lyakhovsky, V., 2006. Analysis of aftershocks in alithospheric model with seismogenic zone governed by damagerheology. Geophys. J. Int. 165, 197–210. doi:10.1111/j.1365-246x.2006.02878.x.

Page 9: Earthquake activity related to seismic cycles in a model for a heterogeneous strike-slip fault

145G. Zöller et al. / Tectonophysics 423 (2006) 137–145

Ben-Zion, Y., Eneva, M., Liu, Y., 2003. Large earthquake cycles andintermittent criticality on heterogeneous faults due to evolving stress andseismicity. J. Geophys. Res. 108, 2307. doi:10.1029.2002JB002121.

Bowman, D.D., Oullion, G., Sammis, C.G., Sornette, A., Sornette, D.,1998. An observational test of the critical earthquake concept.J. Geophys. Res. 103, 24, 359–24,372.

Bufe, C.G., Varnes, D.J., 1993. Predictive modeling of the seismiccycle in the greater San Francisco Bay area. J. Geophys. Res. 98,9871–9883.

Burridge, R., Knopoff, L., 1967. Model and theoretical seismicity.Bull. Seismol. Soc. Am. 57, 341–371.

Chinnery, M., 1963. The stress changes that accompany strike-slipfaulting. Bull. Seismol. Soc. Am. 53, 921–932.

Corral, Á., 2004. Long-term clustering, scaling, and universality in thetemporal occurrence of earthquakes. Phys. Rev. Lett. 92, 108501.doi:10.1103/PhysRevLett.92.108501.

Dieterich, J.H., 1972. Time-dependent friction as a possible mecha-nism for aftershocks. J. Geophys. Res. 77, 3771–3781.

Dieterich, J.H., 1994. A constitutive law for earthquake productionand its application to earthquake clustering. J. Geophys. Res. 99,2601–2618.

Fisher, D.S., Dahmen, K., Ramanathan, S., Ben-Zion, Y., 1997.Statistics of earthquakes in simple models of heterogeneous faults.Phys. Rev. Lett. 78, 4885–4888.

Hainzl, S., Zöller, G., Kurths, J., 1999. Similar power laws for fore-and aftershock sequences in a spring-block model for earthquakes.J. Geophys. Res. 104, 7243–7253.

Ito, K., Matsuzaki, M., 1990. Earthquakes as self-organized criticalphenomena. J. Geophys. Res. 95, 6853–6860.

Jaumé, S.C., Sykes, L.R., 1999. Evolving towards a critical point: areview of accelerating moment/energy release prior to large andgreat earthquakes. Pure Appl. Geophys. 155, 279–306.

Madariaga, R., 1976. Dynamics of an expanding circular fault. Bull.Seismol. Soc. Am. 66, 639–666.

Nur, A., Booker, J.R., 1972. Aftershocks caused by pore fluid flow?Science 175, 885–887.

Reasenberg, P., 1985. Second-order moment of Central Californiaseismicity. J. Geophys. Res. 90, 5479–5495.

Roeloffs, E., Langbein, J., 1994. The earthquake predicitionexperiment at Parkfield, California. Rev. Geophys. 32, 315–336.

Rundle, J.B., Turcotte, D.L., Klein, W., 2000b. Geocomplexity and thephysics of earthquakes. Geophysical Monograph, vol. 120. Amer-ican Geophysical Union, Washington, DC.

Rundle, J.B., Klein, W., Turcotte, D.L., Malamud, B.D., 2000a.Precursory seismic activation and critical point phenomena. PureAppl. Geophys 157, 2165–2182.

Shcherbakov, R., Turcotte, D.L., 2004. A damage mechanics modelfor aftershocks. Pure Appl. Geophys. 161, 2379. doi:10.1007/s00024-004-2570-x.

Sornette, A., Sornette, D., 1990. Earthquake rupture as a criticalpoint: consequences for telluric precursors. Tectonophysics 179,327–334.

Utsu, T., 1962. On the nature of three Alaskan aftershock sequences of1957 and 1958. Bull Seismol. Soc. Am. 52, 279–297.

Utsu, T., 2002. Statistical features of seismicity. International hand-book of earthquake and engineering seismology, Part B 719–732.

Utsu, T., Ogata, Y., Matsu'ura, R.S., 1995. The centenary of the Omoriformula for a decay law of aftershock activity. J. Phys. Earth 43,1–33.

Wyss, M., 1997. Cannot earthquakes be predicted? Science 278, 487.Zöller, G., Hainzl, S., 2002. A systematic spatiotemporal test of the

critical point hypothesis for large earthquakes. Geophys. Res. Lett.29, 1558. doi:10.1029/2002GL014856.

Zöller, G., Hainzl, S., Kurths, J., 2001. Observation of growingcorrelation length as an indicator for critical point behavior prior tolarge earthquakes. J. Geophys. Res. 106, 2167–2175.

Zöller, G., Holschneider, M., Ben-Zion, Y., 2004. Quasi-static and quasi-dynamic modeling of earthquake failure at intermediate scales. PureAppl. Geophys. 161, 2103–2118. doi:10.1007/s00024-004-2551-0.

Zöller, G., Hainzl, S., Holschneider, M., Ben-Zion, Y., 2005a. After-shocks resulting from creeping sections in a heterogeneous fault.Geophys. Res. Lett. 32, L03308. doi:10.1029/2004GL021871.

Zöller, G., Holschneider, M., Ben-Zion, Y., 2005b. The role ofheterogeneities as a tuning parameter of earthquake dynamics.Pure Appl. Geophys. 162, 1027–1049. doi:10.1007/s00024-004-2660-9.