You are on page 1of 19

JOURNAL OF GEOPHYSICAL RESEARCH: SOLID EARTH, VOL. 118, 1483–1501, doi:10.

1029/2012JB009481, 2013

The seismic cycle at subduction thrusts: 1. Insights


from laboratory models
F. Corbi,1 F. Funiciello,1 M. Moroni,2 Y. van Dinther,3 P. M. Mai,4 L. A. Dalguer,3 and
C. Faccenna1
Received 23 May 2012; revised 9 October 2012; accepted 21 November 2012; published 15 April 2013.

[1] Subduction megathrust earthquakes occur at the interface between the subducting and
overriding plates. These hazardous phenomena are only partially understood because of the
absence of direct observations, the restriction of the instrumental seismic record to the past
century, and the limited resolution/completeness of historical to geological archives. To
overcome these restrictions, modeling has become a key-tool to study megathrust
earthquakes. We present a novel model to investigate the seismic cycle at subduction
thrusts using complementary analog (paper 1) and numerical (paper 2) approaches. Here
we introduce a simple scaled gelatin-on-sandpaper setup including realistic tectonic
loading, spontaneous rupture nucleation, and viscoelastic response of the lithosphere.
Particle image velocimetry allows to derive model deformation and earthquake source
parameters. Analog earthquakes are characterized by “quasi-periodic” recurrence.
Consistent with elastic theory, the interseismic stage shows rearward motion, subsidence in
the outer wedge and uplift of the “coastal area” as a response of locked plate interface at
shallow depth. The coseismic stage exhibits order of magnitude higher velocities and
reversal of the interseismic deformation pattern in the seaward direction, subsidence of the
coastal area, and uplift in the outer wedge. Like natural earthquakes, analog earthquakes
generally nucleate in the deeper portion of the rupture area and preferentially propagate
upward in a crack-like fashion. Scaled rupture width-slip proportionality and seismic
moment-duration scaling verifies dynamic similarities with earthquakes. Experimental
repeatability is statistically verified. Comparing analog results with natural observations,
we conclude that this technique is suitable for investigating the parameter space influencing
the subduction interplate seismic cycle.
Citation: Corbi, F., F. Funiciello, M. Moroni, Y. van Dinther, P. M. Mai, L. A. Dalguer, and C. Faccenna (2013), The
seismic cycle at subduction thrusts: 1. Insights from laboratory models, J. Geophys. Res. Solid Earth, 118, 1483–1501,
doi:10.1029/2012JB009481.

1. Introduction are associated with the ongoing downdip motion of the sub-
ducting lithosphere, resulting in the elastic deformation of
[2] The majority of global seismic moment occurs at the adjacent plates and the transfer of stress across the locked
convergent margins. In particular, most of M ≥ 8.0 earth- part of the subduction thrust into the upper plate. If the accu-
quakes are generated along the gently dipping fault plane mulated stress exceeds the strength of the plate interface, an
(i.e., the subduction megathrust, Figure 1) representing abrupt fault slip occurs, and part of the stored elastic energy
the frictional interface between subducting and overriding
radiates as seismic waves. The entire process, including the
plates [e.g., Scholz, 1990; Pacheco and Sykes, 1992; Dixon
stress accumulation phase, the relief by the earthquake itself
and Moore, 2007; McCaffrey, 2007, 2008]. These earthquakes
and postseismic relaxation processes, is known as the “subduc-
This article is a companion to van Dinther et al. [2013] doi:10.1029/ tion seismic cycle” [e.g., Wang, 2007]. In this context, the word
2012JB009479. “cycle” does not imply any periodicity of characteristic
1
LET-Laboratory of Experimental Tectonics, Università “Roma recurrence interval or earthquake size, but rather a repeating
Tre,”Roma, Italy.
2
DICEA, Sapienza Università di Roma, Roma, Italy. spatio-temporal pattern. Methods of studying the subduction
3
Swiss Seismological Service, ETH Zürich, Zürich, Switzerland. seismic cycle include remote sensing, geophysical, and
4
Division of Physical Sciences and Engineering, 4700 King Abdullah geological observational studies of convergent margins
University of Science and Technology, Thuwal, Saudi Arabia. [e.g., Gagnon et al., 2005; Grapenthin and Freymueller,
Corresponding author: F. Corbi, LET-Laboratory of Experimental 2011] as well as numerical [e.g., Wang, 2007; Kaneko et al.,
Tectonics, Univ. “Roma Tre”, L.S.L. Murialdo 1, 00146 Roma, Italy. 2010] and analog modeling [e.g., Rosenau et al., 2009].
(fcorbi@uniroma3.it) [3] Although the general stages of the subduction seismic
©2012. American Geophysical Union. All Rights Reserved. cycle are known, many critical points remain obscure. For
2169-9313/13/10.1029/2012JB009481 example, the maximum earthquake magnitude recorded at

1483
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

the sliding along frictional interfaces during the coseismic


stage in the presence of a free surface [Anooshehpoor and
Brune, 1994, 1999; Brune and Anooshehpoor, 1997; Brune
et al., 1993]. In these models the slow wave speed of the
material allows the process of frictional sliding to be captured
without high-speed photography. However, these models
are affected by unavoidable edge effects and nonlinearities
in the rheological response, inhibiting the scalability of the
Figure 1. Schematic cross-section of a subduction zone. experimental results to nature [Rosakis et al., 2007].
[7] The first scaled laboratory models of subduction mega-
thrust earthquakes have been realized by Rosenau et al.
different subduction zones in the instrumental and historical [2009], who used granular analog materials (i.e., a mixture
records is highly variable [e.g., McCaffrey, 2008; Heuret of rice, rubber pellets, and sugar) with frictional and elastic
et al., 2011], highlighting the possibility of a variable propen- properties that are consistently scaled from the natural proto-
sity for hosting mega-events [Marzocchi et al., 2011]. Some type. Because of the capability of the model to simulate elastic
subduction zones exhibit high seismic moment release rate and plastic deformation, this approach bridges the short-term
(e.g., South Chile or East Alaska), whereas it is very low at seismic cycle and the long-term tectonic evolution of conver-
others [e.g., East Aegean; Hyndman, 2007]. Although the gent margins [Rosenau et al., 2009] and allows to study the
subduction velocity seems to be one of the most important subduction forearc tectonic evolution over multiple seismic
controlling factors for interplate seismicity [e.g., Bird et al., cycles and inherent feedback [Rosenau and Oncken, 2009].
2009; Heuret et al., 2011], the mechanisms governing the The ability to experimentally simulate long time-series of
frequency and size of megathrust earthquakes are not yet clear self-consistent seismic cycle deformation (i.e., hundreds of
[e.g., McCaffrey, 2008]. The amount of interplate locking cycles) allowed to further explore scaling of earthquake
varies both regionally and locally [e.g., Moreno et al., 2010; effects like tsunamis in a physics-based, probabilistic manner
Freymueller et al., 2008] as well as potentially temporally, [Rosenau et al., 2010]. Because of lack of temporal resolution
and identifying the parameters controlling this variability is in monitoring the experiments, however, these models were
still a challenge. not suitable to investigate the dynamics of single ruptures.
[4] Our knowledge of the parameters that control the Thus, an analog model that combines the features of the
subduction seismic cycle and the maximum seismic energy subduction seismic cycle with experimental visualizations of
release is limited due to the following constraints: (a) the rupture propagation along frictional interfaces is still missing.
seismogenic zone behavior can only be studied using indi- This type of model would be beneficial in the light of new
rect methods, because the fault is buried and located mainly observational techniques in nature and the occurrence of sev-
offshore. This provides only a snapshot of a long-term, eral recent mega-thrust events. These events delivered prime
time-dependent process; (b) the pattern of strain accumulation examples (i.e., rupture time-series of unprecedented quality)
along the subduction thrust may not be constant throughout that need to be explained by new theories and experiments.
the seismic cycle; (c) instrumental seismic records of the past [8] In this work, we present a new experimental technique
several decades do not provide enough temporal perspective to study the behavior of multiple subduction earthquake
on the recurrence of major earthquakes [e.g., Satake and cycles. This technique is part of a larger modeling attempt that
Atwater, 2007; McCaffrey, 2008]. The problem is not solved takes advantage of complementary analog (this paper) and
by the use of historical/geological records, which lack numerical modeling approaches [van Dinther et al., 2013,
completeness and resolution; and (d) our understanding of paper 2]. The main advantage of analog modeling lies in the
the behavior of the subduction thrust fault is based on a fact that a properly constructed model evolves in response
patchwork of merged observations at different convergent to the physics of the system under the applied experimental
margins that are presently at different phases of the seismic conditions. However, laboratory experiments face technical
cycle, causing uncertainties. limitations in monitoring and repeatability, in the limited
[5] Despite unavoidable assumptions that limit the appli- availability of analog materials to reproduce complex rheolo-
cability field of final results, analog and numerical modeling gies, and in the exact control of the boundary conditions. In
may provide important insights into small-scale/short-term contrast, numerical models have the advantage of a straight-
and large-scale/long-term features of earthquake physics. forward quantitative approach and a considerable flexibility
Examples include investigations on the physical processes in selecting geometry, material and system parameters as
governing dynamic rupture nucleation, growth, and arrest well as their ability to precisely define boundary conditions.
[e.g., Ben David et al., 2010; Nielsen et al., 2010], the factors However, large computational capacities are necessary to
controlling seismic slip at a point on a fault [e.g., Xia et al., simulate realistic geological features using complex rheolo-
2004; Lykotrafitis et al., 2006], and recurrence behavior gies and/or three-dimensional settings. In particular, such
[e.g., Rubinstein et al., 2012a]. Analog models with realistic models (both numerical and analog) should be considered as
geometries (i.e., wedge-shaped) suitable for studying subduc- simplified representations of nature for the difficulty in under-
tion interplate earthquakes fall into the following categories: standing and simulating the frictional behavior. The aim of
(a) purely elastic foam-rubber experiments [e.g., Brune, our combined approach is to simulate multiple subduction
1996] and (b) granular elasto-plastic models [e.g., Rosenau interplate earthquake cycles with a simple analog setup
et al., 2009]. (e.g., neglecting temperature, off-fault plasticity, fluid pressure,
[6] Foam-rubber models of shallow-angle thrust faulting sediment metamorphism, slab elasto-plastic deformation), and
were designed for investigating the basic physics governing validating it with a numerical model.

1484
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

[9] Here we introduce the novel analog modeling approach 2.1. Rheological Properties of the Analog Materials
to reproduce the subduction interplate earthquake cycle, [11] To simulate the viscoelastic behavior of the lithosphere,
including realistic tectonic loading, spontaneous nucleation we selected Pig Skin 2.5 wt% after an extensive study of the
of the frictional instabilities (i.e., analog earthquakes), and rheological and physical properties of a wide range of materi-
viscoelastic relaxation of the lithosphere. We first describe als. Pig Skin 2.5 wt% has the right properties to scale for
the laboratory methods adopted to simulate subduction thrust length, density, stress, and viscosity of lithospheric rocks in
earthquakes, including scaling criteria, material properties, a natural gravity field [Di Giuseppe et al., 2009]. In particular,
and experimental monitoring. We then present the general this material satisfies two necessary conditions for a viscoelas-
characteristics of a reference experiment supported by a series tic solid: (1) G0 and G00 , which are standard rheological quan-
of eight additional models. We provide measurements of an- tities defining the stored and dissipated energy, respectively
alog earthquake source parameters in relation to natural [e.g., Nelson and Dealy, 1993], are approximately the same
events, aiming to demonstrate that the model is able to repro- order of magnitude (i.e., the elastic deformation counterba-
duce features of the interplate seismic cycle that are common to lances the viscous deformation); and (2) G0 > G00 (see Di
many convergent margins. Cyclicity, deformation during the Giuseppe et al. [2009] for a detailed discussion). Pig Skin
seismic cycle, faulting parameters, rupture behavior, and seis- gelatin is prepared by preliminary heating and stirring a
mic moment-duration scaling are analyzed in detail by compar- 2.5 wt% distilled water solution at 60  C and cooling it at
ing experimental results with natural data. Finally, we include 10  C for 12 h [Di Giuseppe et al., 2009] to obtain a firm tex-
details on the variability and repeatability of the experimental ture. The rheology of the material has been characterized using
results to strengthen the reliability of the novel approach. a 50 mm parallel-plate rheometer (Anton Paar MCR 301)
equipped with a Peltier element to control the temperature
2. Experimental Approach during the measurements. To avoid possible sample dehydra-
tion problems during long measurements, the Peltier element
[10] In this study, we selected a viscoelastic material, Pig was saturated using a custom water bubbler.
Skin 2.5 wt% gelatin, as a rock analog because it has the [12] Gel state (i.e., solid-like) Pig Skin 2.5 wt% gelatin has
following qualities: (a) proper rheological properties to simulate a shear modulus ranging from 103 to 104 Pa, depending on the
downscaled lithospheric rock behavior under laboratory condi- age of the material (Figure 2a). Constraining the viscosity of
tions in a natural gravity field [Di Giuseppe et al., 2009], and (b) Pig Skin gelatin in its gel state is more difficult when using
the advantage of being transparent, enabling the tracking of classical rheometric techniques. However, frequency sweep
passive tracers included in the media and allowing nonintrusive, tests conducted at low frequencies provide a lower bound of
continuous monitoring of the internal velocity field during 7  104 Pa  s for viscosity.
the evolution of the model. We use gelatin-on-sandpaper and
gelatin-on-plastic interfaces to simulate velocity-weakening
and velocity-strengthening frictional behavior of the subduc- 2.2. Frictional Properties
tion thrust fault, respectively. Monitoring of the experiment [13] The frictional properties of gelatin-on-sandpaper
is performed using the Particle Image Velocimetry (PIV) tech- (i.e., an analog of the seismogenic zone of the subduction
nique [e.g., Adrian, 1991], and the derivation of source para- thrust) and gelatin-on-plastic (i.e., an analog of the updip
meters is done using an ad hoc algorithm. and downdip aseismic parts of the subduction thrust) have

Figure 2. Rheological and frictional parameters measured with a rheometer and a spring block-like
linear device, respectively. (a) Linear viscoelasticity as expressed by G0 (i.e., storage modulus; filled
symbols) and G00 (i.e., loss modulus; open symbols) at 10  C and 5% strain for Pig Skin wt % at different
ageing. The gray shaded area highlights the possible range of the storage modulus G0 at experimental
conditions [Di Giuseppe et al., 2009]. (b) Rate effect of friction for gelatin-on-sandpaper (red squares)
and gelatin-on-plastic (circles) systems. Least square logarithmic fit and 1s standard deviation are indicated
by black lines and error bars, respectively. The a-b parameter has been derived using equation (1) from a set
of velocity stepping tests performed in the 381–991 Pa normal pressure range [Corbi et al., 2011]. Each
regression is based on 12 pairs of stress-normal stress values. The correlation coefficients are R2 = 0.73
and R2 = 0.98 for gelatin-on-sandpaper and gelatin-on-plastic regression, respectively.

1485
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

been investigated using a linear spring-block-like device in event [e.g., Ruff and Kanamori, 1983], and the latter controls
which a digital dynamometer (AFG-Mecmesin) continu- the capability to transmit the triggered deformation along the
ously measures the shear force exerted during the sliding rupture plane and is linearly related to the critical slip distance
of a gelatin sample [Corbi et al., 2011]. [Ohnaka, 2003]. Pressure and driving rate seem to play sec-
[14] Here we recall the main frictional properties, including ondary roles, although they are not negligible. Independent
the static friction coefficient, ms, and the friction rate parame- of the roughness of the contact surface, stable sliding is al-
ter, a-b, which are related to the potential seismic moment ways achieved for the lowest normal loads and the highest
release of a given interface. ms is determined by measuring sliding velocities [Corbi et al., 2011].
the amount of force that resists motion of the gelatin sample [19] Stable slip at low normal load appears to be consistent
under an applied normal pressure (i.e., from the slope of the with a conditionally stable regime according to rate and state
regression line between shear and normal stresses, normalized friction framework. The effect of sliding velocity is probably
by the area of the sample on which these forces act). linked to the finite lifetime of the adhesive bonds and their
[15] The a-b parameter is defined as follows: potential to restick after the occurrence of a slip pulse
a  b ¼ dmss =d ½lnðV Þ; (1) [Baumberger et al., 2002].
[20] In particular, the sandpaper used in this study is
where mss is the steady state friction coefficient and V is the characterized by protrusion amplitudes of 0.038  0.004 mm
sliding velocity. Negative values of a-b are indicative of and interpeak distances of 0.143  0.009 mm. The gelatin-
velocity-weakening behavior (i.e., ruptures may nucleate and on-sandpaper system has ms of 0.19  0.05. The system shows
can easily propagate along the interface), whereas positive stick-slip dynamics for a loading rate of 0.6 cm/min [Corbi
values of a-b characterize a velocity-strengthening behavior et al., 2011], and a velocity-weakening behavior (a-b = –0.028)
(i.e., no rupture nucleation, and inhibited rupture propagation). has been observed (Figure 2b) [Corbi et al., 2011].
[16] Sandpaper was selected from a range of materials [21] The friction of gelatin-on-plastic is characterized by
studied in an extensive tribological investigation focused stable sliding for all tested experimental conditions (i.e., a
on roughness, normal pressure, and loading rate dependence loading rate of 0.2 to 2 cm/min and normal pressures of
of friction dynamics [Corbi et al., 2011]. Interface rough- 381–991 Pa). ms for this system is 0.02  0.01, and a-b is
ness has been quantified in terms of amplitude, Rmh, and 0.027 (Figure 2b).
wavelength, l, of geometrical perturbations with respect to
a reference baseline. The ranges of investigated roughness
are 0.10–0.01 mm and 0.71–0.03 mm for the amplitudes 2.3. Scaling
and wavelengths of protrusions, respectively. [22] The scaling procedure is crucial for designing an
[17] Because the static friction depends on the real contact analog model that simulates geodynamical processes on a
surface, we observed the highest static friction with the smoothest convenient geometric and temporal scale. According to simi-
sandpapers [e.g., Marone and Cox, 1994], according to larity theory, the scaled model should be geometrically, kine-
the adhesion frictional model proposed by Bowden and matically, dynamically, and rheologically similar to the natural
Tabor [1950]. prototype [e.g., Hubbert, 1937; Ramberg, 1981; Weijermars
[18] Frictional behavior is primarily controlled by the and Schmeling, 1986]. We identify the most important physi-
roughness of the interface. Regular stick-slip instabilities cal quantities of the investigated system (i.e., length, velocity,
occur only for contact surfaces with Rmh > 0.018 mm. Slip and force) and provide a dimensionless number (i.e., the
can nucleate and easily propagates (i.e., a–b < 0) for Rmh scaling factor) for each parameter that represents the ratio
values of 0.03–0.100 mm; otherwise, the propagation is and tuning between the model and natural quantities (Table 1).
rapidly inhibited (i.e., a–b > 0). The dynamic regime switches [23] The wedge model is designed using a model/nature
to stable sliding and velocity-strengthening behavior for Rmh scale factor, L*, of 1.57  10–6 (i.e., 1 cm in the model corre-
0.03 mm. A nonlinearity of a-b exists because the amplitude sponds to 6.4 km in nature). For models performed in a
and wavelength of protrusions act as two competing para- natural gravity field, the stress scaling factor, s*, can be
meters. The former seems to tune the magnitude of the slip obtained according to the following relation:

Table 1. Scaling Parameters


Parameters Similarity
Scaling
Quantity Symbol Unit Quantity Model (M) Nature (N) Factor (M/N)
Model parameters gravitational acceleration g m/s2 9.8 9.8 1
length L m model thickness 0.11 70,000 1.57  10–6
m mean slip 2.30  10–4 146 1.57  10–6
time (interseismic) Ti s mean recurrence time 16.55 1.49  1011 1.11  10–10
time (coseismic) Tc s mean coseismic duration 1.15 917 1.25  10–3
velocity (interseismic) Vi m/s plate velocity 1.00  10–4 7.05  10–9 1.42  104
velocity (coseismic) Vc m/s mean slip rate 6  10–4 4.79  10–1 1.25  10–3
Rv m/s mean rupture velocity 0.17 1.36  102 1.25  10–3
stress s Pa lithostatic stress 100 1.85  108 5.42  10–7
Materials properties density r kg/m3 1000 2900 3.45  10–1
shear modulus G Pa 1  103 to 1  104 1.85  109 to 1.85  1010 5.42  10–7
viscosity  Pa s 3  105 5  1021 6.00  10–17

1486
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

s* ¼ r*L*; (2)

where r* is the nature/model density scale factor. Because


Pig Skin 2.5 wt% density is  1000 kg/m3 [Di Giuseppe
et al., 2009], we obtain a stress scale, s*, of 5.42  10–7
(i.e., 1 Pa in the model is equal to 1.85 MPa in nature) for an
average crustal wedge density of 2900 kg/m3.
[24] The physical quantity controlling the elasticity of the
analog material is the shear modulus, which has the dimen-
sion of stress and, as a result, scales with the same scaling
factor, G* = 5.42  10–7. Pig skin 2.5 wt% has a shear modu-
lus in the 103–104 Pa range, depending on the aging of the
material [see Di Giuseppe et al., 2009], and therefore scales
to a natural prototype with G = 1.85  109–1.85  1010 Pa.
[25] The subduction seismic cycle is a complex process
that includes extremely variable deformation rates. Because
the experimental interseismic/coseismic time ratio is very
small (see paragraphs 3.1 and 3.4 for details on interseismic
and rupture duration, respectively) with respect to the nature,
it is not admissible to use a single time scale that accounts
for both the coseismic and interseismic stages. To overcome Figure 3. Experimental setup. A gelatin wedge with dimen-
this limitation, we use a dual time scale (i.e., one for the sions of 60  11  34 cm3 is in contact with a flat plate, the
coseismic stage and another for the interseismic stage), analog of the subducting lithosphere. A stepping motor drives
similar to Rosenau et al. [2009]. the plate at constant velocity. The interface between the gela-
[26] The coseismic stage is characterized by an instanta- tin wedge and the plate represents the subduction thrust fault.
neous elastic response in which inertial force may play an The seismogenic area (i.e., velocity weakening frictional
important role. Therefore, the coseismic time scale, Tc*, is behavior) is bounded by two aseismic zones (i.e., velocity
set keeping the Froude number (i.e., the ratio of a body’s strengthening frictional behavior) at its updip and downdip
inertia to gravitational forces) constant, which leads to the limits. A linear UV lamp allows the visualization of passive
following relation: fluorescent tracers included in the gelatin body along a
pffiffiffiffiffiffi cross-section, highlighted by the purple dashed line. The
Tc* ¼ L* (3)
cross-section is placed along the model center line. Camera 1
(i.e., 1 s in the model is equal to approximately 800 s in nature). captures the images of the entire section, while camera 2
[27] During slow deformation inertia is negligible, but focuses exclusively on the seismogenic zone region, aiming
viscous behavior becomes dominant, and consequently, the to obtain better resolved images.
interseismic time scale, Ti*, is established using the follow-
ing relation [e.g., Weijermars and Schmeling, 1986]:
Ti* ¼ *=s*; (4) a computer-controlled stepping motor. The dip of the basal
where * is the model/nature ratio of viscosity. For a model plate is set to 10 . A rigid vertical backstop is placed at the
viscosity of 3  105 Pa  s and a natural viscosity of 5  1021 Pa  s, rear of the gelatin wedge.
we obtain an interseismic time scale of 1.11  10–10 (i.e., 1 s in [31] The analog megathrust develops at the interface
the model is equal to 286 yr in nature). between the gelatin and the basal conveyor plate. The veloc-
[28] Interseismic and coseismic stage velocities are ity-weakening seismogenic zone of the subduction thrust is
derived as the ratios between L* and Ti* (i.e., 1  10–4 m/s simulated using sandpaper at the gelatin-plate interface and
in the model is equal to 22 cm/yr in nature) and between extends from 31 to 47 cm from the backstop in horizontal
L* and Tc* (i.e., 1  10–4 m/s in the model is equal to 0.1 m/s distance. This distance corresponds to a 100 km wide (in
in nature), respectively. map view) seismogenic zone (90 to 190 km from the trench)
[29] The static friction coefficient, ms, and the friction that spans a depth range of 15 to 34 km. These dimensions
rate parameter, a-b, are dimensionless and, thus, scale with are close to the average geometry extracted from worldwide
a factor of 1. subduction zone statistics [Heuret et al., 2011]. Several
physical explanations for the updip and downdip limits of
the seismogenic zone have been proposed. A recent hypoth-
2.4. Experimental Setup esis for explaining the puzzling origin of the updip limit is
[30] The experimental apparatus (Figure 3) consists of a strain localization, in combination with precipitation of
Plexiglas box that is 60 cm long, 40 cm high and 34 cm wide, calcite and quartz [Saffer et al., 2012]. The downdip limit
in which an undeformed gelatin wedge is placed. The gelatin is likely related to the brittle-ductile transition of crustal
wedge is representative of a ~380 km long and 70 km deep material [i.e., Brace and Kohlstedt, 1980] or by the intersec-
section of forearc lithosphere. Along-strike (i.e., y-direction) tion of the slab with the forearc mantle wedge [Peacock and
variations of frictional properties are not imposed as in a Hyndman, 1999]. We simulate these velocity-strengthening
quasi-two-dimensional setup. A rigid steel basal conveyer regions using an upper and lower velocity-strengthening
plate, which is the analog of the subducting plate, is driven plastic sheet covering the conveyer plate. The plastic sheets
with a velocity of 10–4 m/s using a screw jack connected to are attached to the base of the experimental setup to ensure

1487
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

that the depth range of the analog seismogenic zone remains During the acquisition step, all factors contributing to accu-
stable during the experimental run. rate velocity measurements were carefully controlled. In
particular, we optimized the experimental procedure for
2.5. Monitoring and Resolution proper tracer density and homogeneity, low image back-
[32] A transparent analog medium seeded with highly ground, and homogeneous lighting. The camera used for
reflective, neutrally buoyant markers and a suitable illumina- image acquisition was equipped with low-aberration optics,
tion system are key components for tracking deformation in and the entire system was accurately calibrated. The accuracy
analog models. We seeded Pig Skin gelatin, which has a of the subpixel estimators of PIV algorithms is on the order of
suitable transparency in the gel state [Di Giuseppe et al., 1/10 to 1/20 of a pixel. Given the camera calibration datum
2009], with 63 to 125 mm diameter fluorescent microspheres for the reference experiment, the resulting spatial resolution
(Cospheric - UVYGPMS). Microspheres emit a green lumi- is 12 mm and accuracy is 25 mm for camera 1. A spatial
nescence when excited by 365 nm wavelength light. To resolution of 3.2 mm with a sensitivity of 7 mm is obtained
reduce possible buoyancy effects (the density difference adopting camera 2.
between microspheres and gelatin is 0.03 g/cm3), the seeding
procedure was performed just before the gelatin was cooled 2.6. Source-Parameter Derivations
to 10  C. The amount of passive tracers employed was [38] Earthquake source parameters are derived by analyzing
extremely small (0.004 wt%) to prevent any effect on the the velocity field of the entire image data set (see Figure 4 for a
rheological properties of the gelatin. The setup included a flow-chart of the required processing steps). First, the velocity
60 cm long ultraviolet (UV) lamp that lights a 2 mm thick
planar volume situated in the center of the model. This results
in a nonintrusive method for measuring model internal defor-
mation that limits possible biases due to sidewall friction.
[33] The models were monitored using two CCD cameras
(ALLIED-PIKE), both of which acquired a side view
sequence of high-resolution (1600  1200 pixels2, 8 bit, 256
gray levels) digital images of the longitudinal cross-section
illuminated (x-z plane; Figure 3). One camera imaged the
entire model with a frame rate of 7.5 fps and was used to
derive the long-term model deformation. The second camera
focuses the seismogenic zone at 15 fps, aiming to increase
both the spatial and temporal resolutions. The second camera
has been prerequisite for deriving at a high enough resolution
source-parameter and dynamic rupture imaging.
[34] Because simultaneous monitoring of the upper surface
(x-y plane; Figure 3) velocity field is not possible in a single
model due to interference with the UV lamp, additional
model runs have been performed with illumination of the
shallow horizontal plane exclusively. In such cases, we used
a single camera at 7.5 fps.
[35] Images were analyzed by means of the PIV method
using MatPIV, an open-source MATLAB code [Sveen
2004]. PIV is a widely used optical method for extracting
velocity fields in viscous and granular flows [e.g., Adrian,
1991; Adam et al., 2005; Rosenau et al., 2009].
[36] Each image was sampled by overlapping interroga-
tion windows, i.e., small image regions. The Fast Fourier
Transform was employed to compute the displacement of
each window across two consecutive frames. PIV, finally,
provides the average displacement field of tracer particles
within the interrogation window. An initial 128  128 pixel2
discretization with an overlap of 50%, subsequently refined
to 64  64 pixels2, guaranteed both robust statistics within
the interrogation windows and an adequate spatial resolution
for both velocity components. The two displacement compo-
nents in each window were converted to velocities by using
the pixel size and the time interval between two consecutive
frames, i.e., the inverse of frame rate.
[37] Particle image velocimetry is subject to sources of
error. The overall evaluation of the accuracy is affected Figure 4. Flow chart describing the main technical steps to
by the acquisition step through the techniques of velocity derive earthquake source parameters from PIV-derived veloc-
extraction, and the accuracy of the velocity measurement is ity time series. Diamond shaped boxes represent processing
strongly linked to the image quality [Raffel et al., 2007]. steps; rectangular shaped boxes contain the subsequent results.

1488
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

field was derived using the PIV method applied to the full by the necessity to exclude from the image monitoring
range of sequential images. Then, we extracted the horizontal the subducting plate motion. A new time series was succes-
and vertical components of the velocity field along a profile sively created by selecting the profile peaks for each time
1 cm above the analog subduction thrust. This choice is forced (Figure 5, red line).
[39] We used an ad hoc algorithm to identify local maxima
in slip velocity, hereinafter referred to as the picking algo-
rithm (i.e., the fpeak algorithm running under MATLAB),
was used to discriminate velocity peaks (i.e., analogous to
seismic events) from the low-background-characteristic
velocity (i.e., analogous to the interseismic stage). The
algorithm requires a characteristic minimum velocity (i.e.,
the coseismic velocity threshold) and a chosen sensitivity (i.e.,
the minimum number of frames before and after velocity
peaks). A histogram counting the number of frames with a
given maximum velocity is used as a tradeoff curve in
which the optimal value for discriminating interseismic
velocity versus coseismic velocity is given by the inflection
of the curve (Figure A1a). We used 0.015 cm/s as the velo-
city threshold. Note that the same value is used for all seven
accompanying experiments. To set the picking algorithm
sensitivity, we calculated the number of picked events
for sensitivities in the 1–30 frames range (Figure A1b)
and found that the number of identified events is stable
Figure 5. Horizontal velocity component time series for sensitivities larger than 16 frames. Slightly faster or
measured along a linear section located 1 cm above the slower velocity thresholds (0.01 and 0.02 cm/s) provide
subduction thrust fault. Maximum velocity above the seis- the same number of picked events. We conclude that a
mogenic zone is highlighted in red. Grey lines are relative sensitivity larger than 16 (20 in this case) provides a
to the entire section velocities, and provide insight in the stable sampling characterized only by large events. Due to
measurement noise. Velocity peaks (full green points) faster this decision, only peaks separated by periods of slow
than a fixed velocity threshold (black line) are interpreted as velocities longer than 2.6 s are considered. This approach,
analog earthquakes (see Appendix A1 for details on setting therefore, possibly includes aftershocks and foreshocks into
the velocity threshold). one unique event.

Figure 6. (a) Spatio-temporal slip rate distribution and source parameter derivation. Contour of coseis-
mic velocity threshold (black line) and slip rate (grey shading with red 0.05 cm/s isolines) in the space/
time plane are used for identifying the hypocenter position and spatial rupture limits highlighted by the
red point and empty red squares, respectively. (b) Spatio-temporal cumulative slip evolution. Red lines
highlight 0.05 cm step contour. Pixel size of both figures is proportional to the spatial and temporal
resolution of the image analysis technique.

1489
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

[40] Space-time plots of slip rate were produced for each [44] Stick-slip behavior at the base of the wedge is associ-
event identified by the picking algorithm (e.g., Figure 6a). ated with “saw-tooth”-shaped displacement profiles due to
The onset of the stick-slip instability and the position of alternating rearward (i.e., toward the backstop) and seaward
the rupture front as a function of time throughout the rupture motions of the particles (Figure 7a). In particular, phases of
process can be imaged with high discretization by identifying rearward quasi-continuous motion occur when the gelatin
the position of the coseismic velocity threshold in these plots. wedge “sticks” on the contact interface (see the low slope in
This procedure allows for the tracking of model velocities Figure 7a). This phase coincides with a stress increase both
during all stages of the experimental rupture (i.e., initiation, at the base and inside the wedge and corresponds to the inter-
propagation, and arrest) and represents an important advance seismic period. When shear stress along the base exceeds the
compared to earlier models. For each event, space-time plots frictional strength within the velocity weakening zone, rapid
of cumulative slip evolution were produced by summing trenchward (i.e., seaward) motion of the particles occurs.
particle displacements during the coseismic stage (Figure 6b). These slip phases can be considered analogous to the coseis-
Note that, here, “slip” is the one-sided coseismic displace- mic stage. The horizontal displacement associated with the
ment because the displacement of the subducting plate is coseismic phases of the reference model ranges from 0.005
not measured. to 0.08 cm. According to the length scaling factor, the natural
equivalent slip ranges from 32 to ~500 m.
[45] The experimental cyclicity was also investigated in
3. Results terms of recurrence behavior. Therefore, interseismic periods
[41] This section summarizes the observations and analyses have been defined as the duration of prevalence of velocities
of nine models conducted under the same experimental condi- below the coseismic threshold (Figure 7b). Accordingly, the
tions (i.e., the same materials, geometric/kinematic conditions, recurrence interval of the reference model ranges between
experimental duration of 800 s). For simplicity, we refer to 2.6 and 21.7 s, with a mean of 16.5  3.8 s. According to
model #1 as the reference model. Seven supplementary mod- the interseismic time scaling factor, the recurrence interval of
els are used to constrain the variability of the experimental the natural prototype ranges between ~750 and ~6000 years.
results and test the repeatability of our reference model. The Shorter intervals are discarded a priori by the sensitivity of
resulting catalog of analog interplate earthquakes includes the picking algorithm. A wider recurrence interval range is
260 events. The model surface deformation is acquired from obtained if all the experiments are taken into account. In
an additional model (model #9) performed using the top- this case, the longest interseismic period is 35.0 s, which
view setup. is equivalent to ~10,000 years in nature.
[46] Figure 7b shows the distribution of the recurrence
intervals for the reference model and the entire dataset (blue
3.1. Cyclicity of the Deformation Pattern bars and black line, respectively). Both distributions have
[42] Here we discuss time series of the horizontal velocity the same recurrence interval mode of 19 s, verifying statisti-
component 1 cm above the subduction thrust (Figure 7a) to cal significance of the observed model periodicity.
describe and verify the response to frictional behavior of
the models. 3.2. Wedge Deformation During the Analog
[43] After an initial 40 s period, in which the gelatin Subduction Seismic Cycle: Coseismic and
wedge adjusts to the experimental apparatus and the stresses Interseismic Stages
are build up due to convergence (note that this phase has no [47] The subduction earthquake cycle alternates between
natural equivalent and is neglected in the following analysis), stages of slow strain accumulation and sudden fast motion.
the model exhibited two spatial domains, one in which Here we use top- and side-view velocity fields to demon-
cyclicity is detectable, one in which deformation is smaller strate that the analog model shares similarities to the natural
than detection threshold. interplate earthquake cycle as observed using, e.g., geodetic

Figure 7. Stick-slip dynamics and recurrence time. (a) Space-time history of horizontal displacement
(black vertical line for reference) for markers located along the base of the model. (b) Frequency histogram
of recurrence time for the reference model (blue bars) and for the entire experimental data set (black line).

1490
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

Figure 8. (a–d) Interseismic- and (e–h) coseismic-stage model deformation. Average (over a duration
specified above each panel) velocity field (red vectors) measured at the model surface (Figures 8a and
8e) and in the model cross-section (Figures 8c and 8g). Percentage of plate locking is highlighted by black
contours (contour interval: 10%; Figures 8a and 8c). Coseismic horizontal displacement is highlighted
by the black contour with step of 0.01 cm (Figures 8e and 8g). Model topography (Figures 8b and 8f)
and basal slip/backslip (Figures 8d and 8h) have been derived by cumulating incremental vertical and hor-
izontal displacements, respectively. Red and green triangles indicate the limits of the seismogenic zone
and trench location, respectively. Green lines highlight the section along which the analysis is
performed. Diagonal texture indicates areas possibly affected by boundary effects.

observations. For the sake of clarity, we focus exclusively the seismogenic zone which is highly locked during the
on the analog interseismic and coseismic stages. We skip interseismic period (Figures 8a and 8c).
the post-seismic stage for spatial resolution limitations. [50] The side-view velocity field during the interseismic
[48] Figure 8a shows particle motion of the model surface stage confirms the rearward motion of the model (Figure 8c).
during the interseismic stage. Marker velocities are generally Rearward motion is particularly evident above the analog
directed rearward. Specifically, these velocities taper toward seismogenic zone. The maximum interseismic velocity of
zero close to the model boundaries and reach their maximum ~0.004 cm/s is located 45 cm from the backstop. A slightly
in the order of tens of micrometers per second in the frontal downward direction of the velocity vectors occurs at 45 cm
and central area (i.e., 12 to 22 cm along the trench and 37 to from the backstop. This trend changes to a slightly upward
48 cm from the backstop). The area of maximum horizontal motion at 25–35 cm from the backstop.
velocity is thus shifted a few centimeters closer to the trench [51] The side-view velocity fields were used for calculating
with respect to the analog seismogenic zone. It is important the vertical surface displacement (i.e., resulting model topog-
to note that the velocity along the x-axis tends to be zero raphy) and the amount of slip and backslip during both the
approaching the trench area. This relation is due to the interseismic and coseismic stages. In particular, topography
presence of the shallow low basal friction (i.e., due to the is calculated summing incremental vertical displacement
presence of the plastic sheet) and reflects an increase of basal at the model surface, while backslip is obtained summing
backslip dislocation (i.e., the amount of plate convergence incremental horizontal displacement at 1 cm above the
during the interseismic interval for a kinematically locked analog subduction thrust.
megathrust) at the seismogenic zone. [52] Accordingly, the interseismic stage is characterized
[49] Particle velocities at each time step were used to by subsidence of the area updip of the seismogenic zone
calculate the horizontal (i.e., trench-normal) displacement. near the trench (Figure 8b). Rearward subsidence gives
Horizontal displacement was normalized by the amount of way to uplift in a 25 cm wide zone above and landward of
plate motion to provide an estimation of “plate locking” the seismogenic zone. The inflection point from subsidence
(Figure 8a). This mirrors the amount of horizontal plate to uplift is located 15 cm from the trench. The 20 cm close
convergence transferred to the upper plate during the inter- to the backstop area remains basically flat. The amplitude of
seismic stage. The percentage of plate locking is 0% the maximum observed uplift and subsidence is 0.07 cm.
to ~40% with a spatially heterogeneous distribution. The The basal backslip varies from 0 cm to 0.06 cm (Figure 8d).
highest values of plate locking consistently occur above The maximum backslip is located 45 cm from the backstop,

1491
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

near the updip limit of the seismogenic zone. The backslip of the companion paper and Figure 6a). Consistently, the rup-
distribution tapers to zero at the trench and within 30 cm ture preferentially propagates updip. In general, 80% of the
of the backstop. This condition reflects the presence of the rupture extends updip with respect to the hypocenter position.
nonsubducting, low-friction plastic sheet in these areas. This observation is statistically consistent throughout the en-
[53] During the coseismic stage (Figures 8e–8h), the markers tire data set.
located at the surface move toward the trench (i.e., seaward) at [56] The hypocenters appear to be randomly located
least one order of magnitude faster than during the interseismic throughout the seismogenic zone and are mostly unrelated
stage (Figures 8e and 8g). The fastest velocities, located along to the hypocenter location of the previous event. However,
the model centerline at 40 to 50 cm from the backstop, slow series of 3 to 4 events are “repeaters” characterized by hypo-
to approximately zero both at the trench and near the backstop. centers located in the same position (e.g., events 31–34 and
We observe a maximum horizontal (i.e., trench-orthogonal) 41–43; Figure 9a). On other occasions, sequences of 3 to 5
displacement of 0.09 cm (Figure 8e). Consistently, the highest “triggered” events with progressively shallower hypocenters
coseismic surface displacements correspond to the areas of occur (e.g., events 6–8 and 19–23; Figure 9a).
highest interseismic plate locking. [57] Rupture area controls earthquake size and scales with
[54] Surface topography, side-view velocities, and slip earthquake magnitude [e.g., Wells and Coppersmith, 1994;
show the opposite pattern of that found during the interseis- Mai and Beroza, 2000; Blaser et al., 2010; Strasser et al.,
mic periods. In particular, the model topography is charac- 2010]. Moreover, the position of the downdip and updip
terized by an uplifted area 15 cm updip of the seismogenic limits of an earthquake is important for the proximity of seis-
zone followed by subsidence 23 to 45 cm from the backstop mic sources to inland cities and the tsunamigenic potential of
(Figure 8f). The inflection point from subsidence to uplift subduction earthquakes.
coincides with the location of highest measured slip. The [58] Here we focus exclusively on the along-dip extent of
slip distribution along the plate interface slightly asymmetri- analog ruptures (i.e., the rupture width) due to experimental
cally tapers to zero at the rupture edges (Figure 8h). In the limitations (i.e., the analog model is quasi-2D). Analog
imaged analog event, the maximum slip occurs at the updip interplate events in the reference model are characterized
limit of the seismogenic zone. by rupture widths of 0.9 cm to 23.8 cm (Figure 9a). Equiva-
lent natural rupture widths range between 5 km and 150 km.
3.3. Hypocenter Location, Rupture Width, and Figure 9a shows a simplified synoptic view of the downdip
Slip Variation extent of the reference model sequence. The downdip limit
[55] The hypocenter location, one of the key source para- of all events is located within the seismogenic zone. Few
meters, is important as a reference point to represent the events (9%) are characterized by downdip rupture limits that
spatial distribution of the slip [e.g., Chu et al., 2011]. The reach, but do not cross, the base of the seismogenic zone. For
majority (~90%) of hypocenters are located inside the the majority of the events (95%), the rupture extends to well
velocity-weakening area (Figure 9a). However, in the few within the shallower velocity-strengthening area. The rupture,
remaining cases, the hypocenter is less than 2 cm from the however, never breaks the surface. Only two events are charac-
upper limit of the seismogenic zone. In 86% of the cases, terized by updip and downdip limits confined to the seismo-
the hypocenters are located in the deeper half of the rupture genic zone.
(the downdip hypocenter position with respect to the rupture [59] The average slip of an earthquake is another parame-
width is approximately 0.3 to 0.4; Figure 9a), and 95% of them ter controlling the magnitude of a seismic event. Analog
are also deeper than the peak slip-rate location (see Figure 4d earthquakes are characterized by an average “one-sided”

Figure 9. (a) Line-source slip distributions (grey shading), rupture extent (limited by empty red squares),
and hypocenters (full red dots) for the reference model events. The black dashed lines mark the limits of
the velocity weakening seismogenic zone. (b) Mean slip versus rupture width for the reference model
events showing a clear correlation between source parameters (R2 = 0.49). The linear regression line, as
well as the 95% confidence bounds are highlighted by red and grey lines, respectively.

1492
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

slip ranging between 0.005 and 0.05 cm with a mean of Mw ¼ ða=bÞ þ ð1=bÞlog10 W ; (7)
0.03  0.01 cm. The equivalent natural slip ranges from
32 m to 320 m. The absolute maximum slip (0.08 cm) is lo- where the a and b coefficients are equal to –1.86 and 0.46, re-
cated in the shallower part of the rupture of event number 36 spectively. This results in experimental moment magnitudes
(Figure 9a). The slip distribution of analog earthquakes from 5.6 to 8.8 (with an average of 8.2  0.5).
systematically exhibits a taper with a rather pronounced [62] The seismic moment is then calculated following
asymmetry (Figure 8h). In particular, most events are char- Hanks and Kanamori [1979]:
acterized by peak slip locations within the shallower half log10 Mo ¼ 1:5 Mw þ 9:1: (8)
of the rupture (Figure 9a). Analog earthquakes are character-
ized by a positive linear relation between the mean slip For analog earthquakes in the reference model, we calculate
and rupture width consistent with crack theory (Figure 9b) seismic moments ranging from 4  1017 to 1.9  1022 Nm and
[e.g., Marone and Richardson, 2006]. from 3  1016 to 2.3  1023 Nm, considering the entire data set.
[63] A clear empirical relationship between earthquake
duration and seismic moment has been found in nature
3.4. Seismic Moment vs. Duration [e.g., Abercrombie, 1995] and is used to further validate
[60] The most important physical measure for earthquake our approach. In particular, duration scales as the cube root
size is the seismic moment, Mo, which is calculated as of seismic moment for regular earthquakes and linearly for
slow earthquakes [e.g., Ide et al., 2007; Houston, 2001].
Mo ¼ G  D  W  L; (5)
[64] We compared analog earthquake moment to duration
where G is the rigidity of the rock, D is the average slip, and W with scaling laws derived for regular and slow earthquakes
and L are rupture width and length, respectively [Hanks and [Ide et al., 2007] to relate the experimental mode of slip
Kanamori, 1979]. This formulation, which requires knowl- propagation to natural end-member models (Figure 10).
edge of both the downdip and trench-parallel rupture extent, The durations of analog earthquakes (ranging from 0.07 s
is not directly applicable to our model, because we do not to 2.2 s for the reference model) were upscaled in accordance
simulate the third dimension in the presented quasi-2D setup. with the coseismic time-scaling factor Tc*. Interestingly,
To extrapolate our data to a third spatial dimension, we apply a analog earthquakes are located between regular and slow
procedure that consists of using existing rupture width- earthquakes in the moment-duration space. In particular,
magnitude scaling relationships [Wells and Coppersmith, these analog earthquakes are located in the field of shallow
1994; Mai and Beroza, 2000; Blaser et al., 2010; Strasser subduction and tsunami events (Figure 10) [cf. Peng and
et al., 2010]. Empirical relationships of the form Gomberg, 2010; Bilek, 2004], even if a slightly longer dura-
tion than regular earthquakes is observed (upscaled analog
Mw ¼ a þ blog10 ðX Þ (6) durations range from 1–29 min). The slope of the least-squares
fit of experimental data is similar to the slope of regular
are available from regression analysis based on large numbers earthquakes. This observation suggests a mechanical simi-
of events in nature (a and b are coefficients that vary with larity between analog earthquakes and earthquakes in
rupture dimension X). Our choice is supported by the follow-
ing observations: (a) the rupture width is the most robust
observable in our experiments (taking into account both the
updip and downdip limits of the rupture, an error of 0.26 cm
is estimated based on the PIV interrogation window size;
see Figure 6a); (b) the experimental average slip appears
to be too large (scaling from 32 m to ~320 m) and would,
therefore, provides unrealistic magnitudes, larger than 10;
(c) the rupture width-magnitude scaling relationships are
robust for various kinds of faulting in nature [Wells and
Coppersmith, 1994; Mai and Beroza, 2000] and with
special focus on the subduction environment [Blaser et al.,
2010; Strasser et al., 2010].
[61] The smallest rupture width of our reference model is
0.9 cm, which corresponds to 5 km in nature. According to
Mw-W scaling relationships [Wells and Coppersmith, 1994;
Blaser et al., 2010; Strasser et al., 2010], this rupture corre-
sponds to a Mw = 5.6 earthquake. The largest rupture width
of the reference model is 23.8 cm, corresponding to 150 km Figure 10. Diagram showing the relation between seismic
in nature. This rupture width has a large moment magnitude moment and rupture duration for experimental and natural
variability corresponding to Mw = 8.3, Mw = 8.8, and Mw = 9.2, earthquakes. Scaled analog earthquakes are represented with
according to the empirical relationships of Strasser et al. black open and red filled points (entire dataset and reference
[2010], Blaser et al. [2010], and Wells and Coppersmith model, respectively). The gray line is the least square fit of the
[1994], respectively. In the following, we exclusively use the entire experimental data set. Shallow subduction earthquakes
intermediate Mw-W scaling of Blaser et al. [2010]. In particu- are highlighted by small green points (compilation by Bilek
lar, we refer to their preferred relation for “orthogonal reverse- [2004]). Blue and red diagonals mark regular and slow earth-
slip” events with a rupture width in the 2–240 km range: quakes scaling laws [Ide et al., 2007], respectively.

1493
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

nature, that is, earthquakes operate as dynamically running into a set, indicating that the data dispersion does not further
shear cracks following unique source scaling relations. increase when the events of more than four models are consid-
ered together (Figures 11a and 11c). The recurrence interval
3.5. Experimental Variability and Repeatability does not display similar behavior; rather, Cv linearly increases
with an increasing number of events (Figure 11b).
[65] Providing details on the variability and repeatability [69] The nonparametric Kolmogorov-Smirnov test can be
of the experimental results is important for demonstrating used to determine whether two data sets are drawn from
the robustness of the experimental approach and for specu- the same population. In other words, this test evaluates how
lating about possible natural variability. well two data sets match, demonstrating the repeatability of
[66] We identified the most important experimental para- experimental results. The Kolmogorov-Smirnov statistic
meters for this purpose: earthquake duration, recurrence searches for the greatest difference, k, between the empirical
interval, and rupture width. These parameters describe the cumulative distribution of the two samples (e.g., the rupture
temporal features of interseismic and coseismic stages as duration data of two distinct models). This statistic is calcu-
well as the size of analog earthquakes. lated under the null hypothesis that the samples are drawn
[67] We use the coefficient of variation, Cv, to quantify the from the same distribution [Davis, 1973]. The test rejects the
variability of experimental results, and use the two-sample null hypothesis when the significance level is equal to or
Kolmogorov-Smirnov test to investigate the repeatability exceeds the p-value, p (i.e., the minimum level of significance
using eight similar models. Cv is a dimensionless value that for which the null hypothesis is rejected). The significance
describes the dispersion of a selected variable. Cv is defined level was set to 0.05. For each of the selected parameters,
as the ratio of the standard deviation of a data distribution to we tested all the possible combinations of pairs of models
its mean. Cv is considered a reasonable measure if it contains (i.e., 28 tests for each parameter; Figures 11d, 11e, and 11f).
only positive values, as is the case here. In general, high Cv [70] The comparison of rupture durations shows that there
values are associated with a broad distribution of the selected is a relatively high consistency of performance between
parameters. We computed Cv for each of the eight side-view different pairs of models. The average greatest difference
models. We found that our experiments are characterized between distributions of a pair of data sets, k, is 0.32  0.10,
by Cvs for earthquake durations of 0.40–0.95, recurrence and 82% of tests failed to reject the null hypothesis. A similar
intervals of 0.22–1.00, and rupture widths of 0.38–0.81 level of matching was observed for the rupture width. In this
(Figures 11a, 11b, and 11c, respectively). The mean Cv is case, the average k is 0.34  0.12, and 75% of tests failed to
0.70 for the rupture duration, 0.57 for the recurrence interval, reject the null hypothesis. The recurrence interval indicates
and 0.61 for the rupture width. the weakest level of similarity between cumulative distribu-
[68] To test whether there would be a similar data disper- tions; the average k is 0.41  0.18, and 57% of tests failed
sion for a larger data set, we iteratively recalculated Cv for to reject the null hypothesis.
an increasingly large set of samples by merging n models
(with n = 1,. . .,8). The case n = 1 provides the same result
obtained when only considering model #1 (the reference
model). Following this procedure, Cv always reaches a 4. Discussion
plateau at approximately 0.6–0.7, for both the earthquake [71] The new experimental setting presented in this paper
duration and rupture width when four models are merged allows for the investigation of the coseismic and interseismic

Figure 11. Measure of dispersion and repeatability of the experimental results. (a–c) Coefficient of
variation for selected parameters of a single model data set (red dashed line) and for subsamples of the
entire data set (black lines). (d–e) Results of the two-sample Kolmogorov-Smirnov goodness-of-fit null
hypothesis testing for all possible couples of experiment combinations.

1494
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

behavior of subduction thrusts. Additionally, we identify would be characterized by a Cv > 1, Cv ~ 1, or Cv < 1, re-
four major aspects of our experimental results that deserve spectively [Kuehn et al., 2008]. The observation of a Cv in
to be analyzed in detail: (a) the recurrence behavior, (b) the the 0.2–0.6 range for the entire data set seems to support
coseismic and interseismic model responses, (c) the rupture quasi-periodic recurrence as most representative of the analog
behavior, and (d) the source parameters. To validate our model behavior. A stronger tendency toward the characteristic
model, experimental results are first compared to previous earthquake is observed when perfectly elastic systems are
analog models specifically designed for investigating the subject to both fixed loading rates and elastic energy-release
seismic cycle at subduction thrusts [Rosenau et al., 2009; conditions, as in spring-block models [e.g., Baumberger
Rosenau and Oncken, 2009; Brune, 1996] and accompa- et al., 1994], Couette systems [Galeano and Rubio, 1994],
nying numerical simulations performed in the same setting and ring shear tests [e.g., Rosenau et al., 2008].
[van Dinther et al., 2013, paper 2]. To validate the approach [75] Subduction earthquake models are characterized by
we then discuss our findings with respect to natural observa- an additional degree of freedom, because the rupture size
tions. Although our models oversimplify the natural prototype, may vary from one event to another rather than constant
the analysis of the experimental results provides interesting “system-size” ruptures as in spring-block-like models. As a
insights into the complex behavior of subduction thrusts. result, a wider distribution of recurrence intervals may occur.
Moreover, adding complexities to the system (e.g., changing
4.1. Recurrence Behavior the shape of the experiment and tuning the rheological
[72] The identification of stick-slip events with regular re- response of the media to elastoplastic) would result in a non-
currence intervals allow us to compare our physically self- characteristic recurrence behavior [e.g., Rosenau et al., 2009].
consistent model behavior to simple deterministic models The numerical simulation performed under the same condi-
of stress accumulation and release that have been proposed tions as our model confirm this tendency of less characteristic
for describing earthquake cyclicity. The simplest model is events for increased complexity (see the companion paper).
the “characteristic earthquake model” [Reid, 1910], which The sole velocity-weakening friction formulation in the seis-
postulates that a specific fault ruptures at regular intervals, mogenic zone leads to almost constant recurrence intervals
generating events of similar size and rupture area. Nonchar- and rupture width, whereas adding velocity strengthening at
acteristic models include time- and slip-predictable models the updip and downdip segments of the subduction thrust
[Shimazaki and Nakata, 1980]. The time-predictable model results in noncharacteristic behavior.
implies the presence of a constant threshold strength at [76] Even if the instrumental seismic record spans only the
which the fault ruptures. This model assumes that slip during last century, long recurrence records of subduction zone
a large earthquake reduces the shear stress on the fault more seismicity are provided by well-dated geologic indicators,
than slip during a small event. Therefore, the time to the next such as coral micro-atolls [e.g., Natawidjaja et al., 2004]
earthquake can be predicted from the size of its predecessor. and paleoseismic investigations [e.g., Cisternas et al.,
Alternatively, in the slip-predictable model, the fault does 2005]. These studies indicate the diversity in the behavior
not rupture at a fixed shear stress, but each earthquake of global subduction zones. For example, quasi-periodic
reduces the shear stress on the fault to a fixed minimum earthquakes have been inferred in Sumatra [Natawidjaja
level. This model cannot be used to predict when an earth- et al., 2004], Chile [Bookhagen et al., 2006], Alaska, and
quake will occur, but predicts the minimum magnitude of the New Zealand [Lajoie, 1986]. An upper strain threshold, con-
next earthquake. Accordingly, the size of an earthquake is sistent with the time-predictable model, has been suggested
correlated with the length of the preceding recurrence interval. for the Japan subduction zone [Shimazaki and Nakata,
[73] Approximately 40% of the experimental recurrence 1980], whereas a more complex behavior, that is neither
intervals are clustered within a 5 s time window around time-predictable nor slip-predictable, has been proposed for
the average value (Figure 7b). This first-order observation, the Cascadia subduction zone [Satake and Atwater, 2007].
focused only on recurrence intervals, seems to promote a However, the applicability of earthquake recurrence models
“quasi-periodic recurrence” model. Further investigation, for natural subduction zones, where intrinsic complexities
that includes event size, is used to test the applicability and lateral frictional heterogeneities are expected, is not
of the characteristic, slip- or time-predictable model to the straightforward. Recently, analysis of 14 repeating earthquake
experimental data. We compared analog earthquake size sequences off the east coast of the Tohoku area (Japan)
with the preceding or subsequent interseismic period. For has shown that earthquake recurrence can be better explained
our data low linear correlation coefficients (R2 < 0.06) occur by a “memory-less” earthquake model [Rubinstein et al.,
when correlating parameters that determine event size (mean 2012a]. However, the understanding of sequential rupture at
slip and rupture width) with the preceding and subsequent convergent margins remains a fundamental goal in earthquake
recurrence times. The majority of analog earthquakes are sciences, and laboratory experiments can be helpful in devel-
scattered and display no clear trend. We conclude that neither oping this understanding [e.g., Rubinstein et al., 2012b].
time-predictability nor slip-predictability can be applied to the [77] The factors (e.g., subduction velocity, plate interface
experimental behavior; nor does a characteristic model fit the locking) controlling the frequency of large events at conver-
data. In the latter case both slip- and time-predictable models gent margins remain one of the key questions in modern
should fit equally well. Rather, our models tends to follow a seismology. A multicentennial return time of major earth-
cyclicity governed by a quasi-periodic recurrence interval. quakes has been constrained by both geologic observations
[74] An additional test for the applicability of the charac- [e.g., Goldfinger et al., 2003] and statistical simulations
teristic earthquake model to analog earthquakes is performed [McCaffrey, 2008], whereas our analog models provide
by analyzing the coefficient of variation, Cv, of the recur- recurrence intervals that are roughly one order of magnitude
rence time. Clustered, random or quasi-periodic events larger (~750–10000 years). Such a discrepancy might be

1495
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

explained by (a) the uncertainties in estimating the litho- inversion of surface deformation in nature, its explanation
spheric viscosity and elasticity—mainly in nature—and the remains obscure. In active convergent margins, the base of
resulting imprecision in estimating the experimental inter- the seismogenic zone and, consequently, the location of
seismic time scale factor and slip, respectively; (b) the 2D- the maximum subsidence, roughly coincides with the coast-
like character of the analog model implies that an event line [e.g., Plafker, 1972] and it can reach a maximum ampli-
has to pass through the section that is illuminated by the tude of about –2 m as in the 1960 Mw = 9.5 Chile earthquake
UV lamp to be captured. This contributes to detect a smaller [Plafker and Savage, 1970].
number of events than those occurring in the entire naturally [83] Geologic and geodetic observations helped in identi-
three-dimensional, experimental interface. fying the three primary processes that took place in the after-
math of an earthquake: (a) long-term (decadal scale) mantle
4.2. Coseismic and Interseismic Model Response viscoelastic relaxation [e.g., Pollitz et al., 2008]; (b) short
[78] The analog model successfully simulates one of the term (days to years) relaxation of shear stresses along the
most important features of the seismic cycle at subduction plate interface, which lead to seismic and aseismic afterslip
zones: the elastic rebound associated with interplate [e.g., Perfettini et al., 2005]; (c) relocking of the subduction
thrust earthquakes as recorded in geologic and geodetic data thrust fault [e.g., Gagnon et al., 2005]. The resulting post-
[e.g., Natawidjaja et al., 2004; Wang, 2007]. Accordingly, seismic, time-dependent deformation is characterized by a
we observed periods of slow rearward motion alternating slow seaward motion of some inland areas that occurs simul-
with short phases characterized by fast “seaward” motion taneously with outer-wedge landward motion as reported
(Figure 8). from Chile and Alaska where Mw = 9.5 and Mw = 9.2 earth-
[79] During the coseismic stage, the highest deformation quakes occurred in 1960 and 1964, respectively. In particu-
rates and largest displacements occur in the seaward-most lar, this pattern reflects the interference of two processes in
part of the wedge (Figures 8e and 8g). However, small the seismic cycle: the onset of plate locking at shallow depth
displacements (i.e., 0.01 cm) are observed far inland, up to and viscoelastic relaxation of the mantle beneath the conti-
45 cm from the analog trench (~300 km in nature, which is nent [e.g., Wang et al., 2012]. At present, such evidence
about the location of the volcanic arc in subduction zones). cannot be imaged in detail in the analog viscoelastic model
This finding agrees with observations during megathrust also because of the missing upper mantle contribution.
earthquakes (e.g., the 2004 Mw 9.2 Sumatra event; Chlieh [84] The interseismic stage is characterized by rearward
et al., 2007), experimental results obtained in granular elas- velocities (Figures 8a and 8c). A decrease in horizontal
toplastic models [Rosenau et al., 2009], and the companion velocities from the locked region to inland areas is a common
numerical simulation [van Dinther et al., 2013, paper 2]. feature recognized in active convergent margins [e.g., Simons
[80] Our slip distribution is asymmetrical, with the maximum et al., 2011] and elastoplastic analog models [Rosenau et al.,
slip near the updip end of the rupture (Figures 8h and 9). The 2009], as well as numerical models like the companion
sense of skewness has been proposed to correlate to the stress numerical simulation.
drop trend in a seismogenic zone [Wang and He, 2008]. A [85] Consistently, landward motion is associated with uplift
downdip increase in stress drop is responsible for the slip above the downdip limit of the seismogenic zone (Figure 8b).
distribution to be skewed “landward” [e.g., Wang and He, These experimental observations are similar, for example,
2008; Rosenau et al., 2009] and vice versa [Rudnicki and to the vertical deformation pattern inferred through repeated
Wu, 1995; Brune, 1996; Geist and Dmowska, 1999]. In partic- leveling observations at the Nankai margin [Thatcher, 1984].
ular, the discrepancy between experimental approaches could [86] One of the major difference between viscoelastic and
be related to the capability to simulate seismic waves and their elastoplastic models is in their basal backslip distributions.
amplification (resulting in trenchward increasing slip) in a Our interseismic backslip (Figure 8d) tapers to zero at the
realistic way, as in the dynamic and fully elastic models updip end of the analog subduction thrust. This behavior
(e.g., Brune [1996] and this work). In the model of Rosenau mirrors an interseismic slip rate at the subduction thrust that
et al. [2009], seismic energy could be quantitatively progressively decreases from the plate convergence rate to
absorbed by the granular nature of the wedge medium almost zero in the locked zone [Wang, 2007]. The high inter-
(Rosenau, pers. comm. 2012). However, the exact shape seismic slip rate at the shallowest part of the subduction
of the slip distribution depends also on mechanical details thrust is caused by updip aseismic velocity-strengthening
of the rupture zone [e.g., Scholz, 2002; Bilek, 2007]. plastic sheet. Elastoplastic models [Rosenau et al., 2009]
[81] Natural cases of large subduction interplate events are instead characterized by a basal backslip smoothly
reveal variable slip distributions with maximum slip recorded increasing from the base of the seismogenic zone to the trench.
both in the shallower part (e.g., the 2011 Mw = 9.0 Tohoku- This behavior suggests that the fault is locked all the way up to
Oki earthquake; Lee et al., 2011) and in the central part of the trench [Wang, 2007] as a consequence of underthrusting /
the subduction thrust (e.g., the 2004 Mw = 9.2 Sumatra earth- accretion of material. In the Rosenau et al. [2009] model, shal-
quake; Chlieh et al. [2007]). low slip occurs, therefore, along splay faults in the overriding
[82] The observed maximum coseismic subsidence does plate instead of along the plate interface.
not coincide with where the slip tapers to zero, as predicted [87] The fraction of plate convergence not accommodated
by elastic dislocation modeling [e.g., Wang, 2007] and ob- by aseismic slip in the interseismic period, referred to as
served in the companion numerical simulation [van Dinther the interseismic coupling, reveals consistent similarities
et al., 2013, paper 2], but is shifted a few centimeters toward between viscoelastic and elastoplastic models of the subduc-
the trench (Figures 8f and 8h). While the observed shift, cap- tion interplate earthquake cycle. Figure 8a shows that the
tured also in the elastoplastic analog models by Rosenau plate convergence accumulated in the model varies from
et al. [2009], might have important implications for ~40% above the seismogenic zone to zero at the inland area.

1496
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

The variability in interseismic coupling observed at subduction analog thrust interface. As far as the bimaterial contrast is
margins remains one of the least understood elements of the concerned, the analog setup simulates an upper bound of
strain accumulation process [e.g., Scholz and Campos, 2012]. the natural convergent setting, since the overriding plate
Based on geodetic data, some subduction zones are considered is expected to be the more compliant than the subducting
fully locked, whereas others are only partially locked or are one [e.g., Ma and Beroza, 2008]. Observations [e.g., Rubin
freely slipping [e.g., Dixon and Moore, 2007], with spatio- and Gillard, 2000], numerical models [e.g., Dalguer and
temporal variability possible within the same subduction zone Day, 2009], and analog models [Xia et al., 2004] show that
[e.g., Freymueller et al., 2008; Heuret et al., 2011]. if a rupture occurs on the boundary between two solids
with different elastic properties, the rupture may propagate
4.3. Rupture Behavior in the direction of the slip of the softer rocks (e.g., gelatin
[88] Monitoring the spatial and temporal behavior of in our model).
earthquake rupture (i.e., initiation, propagation, and arrest) [92] A possible consequence of asymmetry in analog
is necessary for understanding the physics of dynamic rupture earthquakes is revealed by hypocenters that are generally
propagation, which, in turn, may help us to extrapolate from deeper than the regions of peak slip (Figure 9a) and peak slip
theory to natural settings, and to thereby better characterize rate (Figure 4d in the companion paper). A similar rupture
the resulting seismic hazard. pattern has been inferred for the 2010 Mw = 8.8 Maule earth-
[89] We process the experimental data in a purely kine- quake and the 2011 Mw = 9.0 Tohoku-Oki earthquake [Lee
matic fashion, as is commonly done in seismic or geodetic et al., 2011; Vigny et al., 2011]. In particular, for the 2010
source inversions [e.g., Ide et al., 2007]. A physics-based Mw = 8.8 Maule earthquake such relationship might result
rupture propagation model within a quasi-static framework from stress-concentration induced by a downdip gradient
is extensively described in the companion paper (sections in locking [Moreno et al., 2010]. Based on the observation
4.4 and 4.6). of earthquakes with various kinematics, Mai et al. [2005]
[90] The location of rupture initiation (i.e., hypocenter) is suggested that hypocenters are not randomly located on a
a first-order source parameter serving as a reference for the fault but are close to regions of large slip.
kinematic characterization of the spatio-temporal slip distri- [93] The analog earthquake rupture front propagates
bution during analogue earthquakes. In nature, the location with an average velocity of 13.9  7.1 cm/s upward and
of the hypocenter is generally more reliable with respect to 11.1  8.1 cm/s downward. The equivalent natural rupture
other earthquake characteristics, such as the location of the velocities are 0.11  0.05 km/s and 0.09  0.06 km/s, respec-
largest slip [Ide et al., 2007]. Our experimental results show tively. Even if experimental rupture speed is slower than veloc-
locations of hypocenters are dominantly located near the ities inferred for natural interplate events [e.g., Lee et al., 2011],
base of the analog earthquake rupture (Figures 6a and 9a). analog earthquakes are not similar to slow earthquakes
This finding is in agreement with results obtained in our (as shown in Figure 10 and described in section 4.4).
numerical simulation (Figure 4c in the companion paper), in [94] The coseismic slip rate of analog earthquakes varies
elastoplastic analog models of the subduction seismic cycle from 0.015 cm/s to 0.320 cm/s. The equivalent natural slip rate
[Rosenau et al., 2009], in numerical models of crack propaga- ranges from 0.1 m/s to 2.6 m/s, which is comparable to the gen-
tion [Das and Scholz, 1983], and by analyzing selected dip- eral range of natural slip rates [e.g., Scholz, 2002] and to the slip
slip and strike-slip natural events [Mai et al., 2005]. On the rate of the 2011 Mw = 9.0 Tohoku-Oki earthquake [Lee et al.,
basis of 12 finite-source rupture models for subduction 2011].
events, Mai et al. [2005] suggested that these earthquakes [95] Another important characteristic of earthquakes is the
prefer to nucleate in the along-dip center of the fault. amount of slip, which may reach tens of meters for very
Observations of recent megathrust earthquakes reveal that large subduction interplate events. For example, the great
both cases (i.e., deep and intermediate depth nucleation) 1960 Mw = 9.5 Chile earthquake was characterized by a
may occur. The 2004 Mw = 9.2 Sumatra earthquake was maximum slip of ~40–50 m [e.g., Barrientos and Ward,
characterized by a hypocenter located in the deepest part of 1990; Moreno et al., 2009], and the 2011 Mw = 9.0
the rupture [Rhie et al., 2007], whereas in the 2010 Mw = 8.8 Tohoku-Oki earthquake had a maximum slip of ~50–60 m
Maule and the 2011 Mw = 9.0 Tohoku-Oki earthquakes, the [Simons et al., 2011; Lee et al., 2011]. The peak and average
hypocenters were located in the along-dip center of the slip of analog earthquakes scale to roughly one-order-of-
rupture [e.g., Moreno et al., 2010; Lee et al., 2011]. Few magnitude larger values (i.e., tens to a few hundreds of
analog hypocenters (i.e., <10%) are located in the shallow meters). Possible explanations for this discrepancy include
velocity-strengthening zone likely originated by the sand- the quasi-2D setting [e.g., Bilek and Lay, 2002] and the
paper/plastic sheet frictional discontinuity. absence of off-fault plasticity for the analog model
[91] Once the rupture initiates, it mainly propagates bilat- [e.g., Dunham et al., 2011; Wang and Bilek, 2011]. Propaga-
erally with preference in the updip direction. This behavior tion in the lateral third dimension and off-fault plasticity would
is consistent with the observation of granular elastoplastic absorb part of the energy that is used for slip on the analog
[Rosenau et al., 2009], foam-rubber analog models of inter- subduction thrust. Additionally, the strong bimaterial contrast
plate seismicity [Brune, 1996], and the numerical results existing in the analog model favors a larger strength and
obtained in the companion paper (Figure 4c). One possible consequent stress drop and thereby contributes positively
explanation is that the rupture follows the lithostatic in providing large slip [Ma and Beroza, 2008].
pressure gradient that results from the subduction wedge [96] Analog earthquakes can be further characterized by
[Das and Scholz, 1983]. Another possibility is that rupture their rupture mode. Rupture growth can be described accord-
asymmetry results from a strong bimaterial contrast in ing to two competing rupture styles: crack-like versus pulse-
elastic properties for gelatin and steel on each side of the like. In the former case, the nucleation region continues to slip

1497
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

throughout the earthquake, and the slipping region expands to natural events, we identified a clear dependence of ana-
until arresting phases arrive from the borders of the fault. In log earthquake duration on seismic moment. In particular,
the latter case, a healing front propagates behind the rupture analog earthquakes are located in a region between regular
front [e.g., Marone and Richardson, 2006]. Quantitatively, and slow earthquakes, since upscaled analog earthquakes
these modes can be distinguished by analyzing the duration duration is roughly one order of magnitude larger than
of slip at a certain point on a fault with respect to the duration megathrust earthquakes in nature (i.e., average duration
of the rupture of the entire fault [Rosakis et al., 2007]. In of 1.15 s correspond to ~15 min in nature). Slow earth-
particular, the pulse mode is characterized by slip at a point quakes show a much larger characteristic duration in the
on the fault that is one order of magnitude shorter than the total range of hours to months [e.g., Ide et al., 2007].
event duration [e.g., Heaton, 1982; Hartzell and Heaton, [102] The experimental data follow a moment/duration
1983; Heaton, 1990]. Distinguishing between these two scaling law characterized by the trend of regular earthquakes
rupture modes is a central issue in earthquake source physics (Figure 10). This suggests that even if the model rupture
because they predict different hazard levels [e.g., Marone duration is slightly longer than for regular natural earth-
and Richardson, 2006]. quakes, the size/duration proportionality is retained. This
[97] Figure 6a shows how we characterized the rupture be- observation allows us to speculate that the physics control-
havior of analog earthquakes. The minimum duration of slip at ling natural and analog earthquakes is similar.
a point on the fault is larger than 1/10 of the entire rupture pro-
cess, indicating a crack-like behavior. The spatio-temporal cu- 5. Conclusions
mulative slip distribution (Figure 6b) also seems to support the
[103] A new analog modeling technique has been developed
idea of growth of the rupture as a crack. However, the patch to shed light on the seismic cycle at subduction megathrusts.
with a high slip rate is concentrated in a small region that pro- The quantitative analysis of velocity time series of the
pagates upward just behind the rupture front, suggesting pos- gelatin-on-sandpaper analog system allowed us to monitor the
sible rupture growth in a mixed pulse/crack fashion. Similar model deformation and measure earthquake source parameters.
behavior has also been observed in previous elastoplastic [104] Our experimental results show that:
models of subduction interplate seismicity [Rosenau et al., [105] (a) The recurrence interval of analog earthquakes can
2009] and in incoherent frictional interface experiments be described as “quasi-periodic” recurrence interval.
[e.g., Rosakis et al., 2007]. Although a complete characterization [106] (b) The interseismic model deformation shows rear-
of rupture styles is beyond the scope of this paper, we ward motion associated with subsidence in the outer wedge
consider this the next challenge to be investigated using this and uplift of the coastal area. The coseismic stage is charac-
experimental setting. terized by faster velocities in the seaward direction. During
[98] Detailed observations of rupture growth in large sub- this stage, the model shows a subsidence of the coastal area
duction interplate earthquakes are available only for recent and uplift in the outer wedge.
mega-events, which have been captured by modern, high- [107] (c) Analog earthquakes generally nucleate in the deeper
resolution geodetic and seismological monitoring methods. portion of the rupture and propagate preferentially seaward.
For example, along-strike crack-like rupture propagation The majority of events propagate with a crack-like behavior.
has been proposed for the 2001 Mw = 8.4 Peru earthquake [108] (d) There are clear proportionalities between rupture
[Robinson et al., 2006], the 2004 Mw = 9.2 Sumatra earth- width and slip, and between seismic moment and rupture
quake [Kruger and Ohrnberger, 2005], and during the initial duration.
75 s of the 2011 Mw = 9.0 Tohoku-Oki earthquake [Lee et al., [109] (e) The size, duration, and recurrence interval of
2011]. Slip pulse behavior has been suggested for the 2010 analog earthquakes are consistent among models performed
Mw = 8.8 Maule earthquake [Heaton et al., 2011]. under the same experimental conditions. These findings sup-
4.4. Source-Parameter Scaling port the repeatability of the experimental results.
[110] This work verifies that our analog model simulates
[99] Faulting parameters are systematically related [e.g., the basic physics governing the seismic cycle and relating
Kanamori and Anderson, 1975], and a wide range of scaling rupture processes in a simplified but robust way. Qualitative
laws has been identified depending on seismic moment and quantitative comparisons between modeling results and
range [e.g., Shimazaki, 1986], intraplate/interplate events what is known about the subduction earthquake cycle in nature
[e.g., Kanamori and Anderson, 1975; Blaser et al., 2010;
suggest that our experimental approach offers an efficient and
Strasser et al., 2010], fault kinematics [e.g., Wells and
versatile tool for investigating the seismic cycle at subduction
Coppersmith, 1994; Mai and Beroza, 2000], and shape of
thrusts. Future efforts will be directed toward increasing the
the rupture [e.g., Scholz, 2002].
level of complexity of the experimental setting to systematically
[100] We examined the average slip/rupture width rela- explore the parameter space influencing the subduction
tionship, which linearly correlates with a proportionality interplate seismic cycle.
factor of 1.7  10–3 (Figure 9b). This observation might
support the W-model [e.g., Scholz, 1982]. However, the [111] Acknowledgments. We thank Matthias Rosenau and Stephan
absence of the third dimension (i.e., L) dimension in the Dominguez for their detailed and constructive reviews. Laura Sandri and
Erika Di Giuseppe are acknowledged for helping in the statistical analysis
experimental setting does not allow for a conclusive argu- and rheological characterization of gelatin, respectively. All the members
ment for such speculation. of the EURYI team are acknowledged for stimulating discussions. This
[101] Subsequently, we investigate the seismic moment/ research was supported by the European Young Investigators (EURYI)
duration relationship to test the capability of the model to Awards Scheme (Eurohorcs/ESF including funds the National Research
Council of Italy) to F.F. and SNSF grant 200021-125274 to Y.D. Models
reproduce similarities in slip behavior with respect to natural have been performed in the “Laboratory of Experimental Tectonics”
end-members (i.e., regular and slow earthquakes). Similar Dipartimento di Scienze, Universita’ Roma TRE, Rome, Italy.

1498
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

Appendix A: Setting the Coseismic Velocity


Threshold and Picking Algorithm Sensitivity

Figure A1: Setting the coseismic velocity threshold and picking algorithm sensitivity. (a) Trade-off curve (grey line) showing
the distribution of the number of frames as a function of the maximum horizontal velocity observed at each frame of the
reference model. The red point indicates the optimal value for the coseismic velocity threshold, which is 0.015 cm/s. (b) Number
of picked events as a function of picking algorithm sensitivity (red line). The number of events picked remains stable for
sensitivities larger than 16. A similar result is obtained also for velocity thresholds of 0.01 cm/s and 0.02 cm/s (grey lines).

References Bookhagen, B., H. P. Echtler, D. Melnick, M. R. Strecker, and J. Q. G.


Spencer (2006), Using uplifted Holocene beach berms for paleoseismic
Abercrombie, R. E. (1995), Earthquake source scaling relationships from analysis on the Santa Maria Island, south-central Chile, Geophys. Res.
1to5 ml using seismograms recorded at 2.5-km depth, J. Geophys. Res., Lett., 33, L15302, doi:10.1029/2006GL026734.
100(B12), 20,036–24,015. Bowden, F. P., and D. Tabor (1950), The Friction and Lubrication of
Adam, J., J. Urai, B. Wieneke, O. Oncken, K. Pfeiffer, N. Kukowski, Solids, vol. 1, Oxford Univ. Press, New York.
J. Lohrmann, S. Hoth, W. Vanderzee, and J. Schmatz (2005), Shear Brace, W. F., and D. L. Kohlstedt (1980), Limits on lithospheric stress
localisation and strain distribution during tectonic faulting - new insights imposed by laboratory experiments, J. Geophys. Res., 85, 6248–6252,
from granular-flow experiments and high-resolution optical image correla- doi:10.1029/JB085iB11p06248.
tion techniques, J. Struct. Geol., 27(2), 283–301, doi:10.1016/j.jsg.2004. Brune, J. N. (1996), Particle motion in a physical model of shallow angle
08.008. thrust faulting, Proc. Indian Acad. Sci., 105(2), 197–206.
Adrian, R. J. (1991), Particle-imaging techniques for experimental fluid Brune, J. N., and A. Anooshehpoor (1997), Frictional resistance of a fault
mechanics, Image (Rochester, N.Y.), 261–304. zone with strong rotors, Geophys. Res. Lett., 24(16), 2071–2074.
Anooshehpoor, A., and J. N. Brune (1994), Frictional heat-generation and Brune, J. N., S. Brown, and P. A. Johnson (1993), Rupture mechanism and
seismic radiation in a foam rubber model of earthquakes, Pure Appl. interface separation in foam rubber models of earthquakes: A possible
Geophys., 142(3–4): 735–747. solution to the heat-flow paradox and the paradox of large overthrusts,
Anooshehpoor, A., and J. N. Brune (1999), Wrinkle-like Weertman pulse at Tectonophysics, 218(1–3), 59–67.
the interface between two blocks of foam rubber with different velocities, Chlieh, M., J. P. Avouac, V. Hjorleifsdottir, T. A. Song, C. Ji, K. Sieh,
Geophys. Res. Lett., 26(13): 2025–2028. A. Sladen, H. Hebert, L. Prawirodirdjo, Y. Bock, and J. Galetzka
Barrientos, S. E., and S. N. Ward (1990), The 1960 Chile earthquake: (2007), Coseismic slip and afterslip of the Great Mw 9.15 Sumatra-
inversion for slip distribution from surface deformation, Geophys. J. Int., Andaman earthquake of 2004, Bull. Seismol. Soc. Am., 97(1A), S152–S173,
103, 589–598. doi:10.1785/0120050631.
Baumberger, T., F. Heslot, and B. Perrin (1994), Crossover from creep to Chu, R., S. Wei, D. V. Helmberger, Z. Zhan, L. Zhu, and H. Kanamori
inertial motion in friction dynamics, Nature, 367, 544–546, doi:10.1038/ (2011), Initiation of the great Mw 9.0 Tohoku–Oki earthquake, Earth
367544a0. Plan. Sci. Lett., 308(3-4), 277–283, doi:10.1016/j.epsl.2011.06.031.
Baumberger, T., C. Caroli, and O. Ronsin (2002), Self-healing slip pulses Cisternas, M., B. Atwater, F. Torrejo, Y. Sawai, G. Machuca, M. Lagos,
along a gel/glass interface, Phys. Rev. Lett., 88(7), 5–8, doi:10.1103/ A. Eipert, C. Youlton, I. Salgado, T. Kamataki, M. Shishikura, C. P.
PhysRevLett.88.075509. Rajendran, J. K. Malik, Y. Rizal, and M. Husni (2005), Predecessors
Ben-David, O., G. Cohen, and J. Fineberg (2010), The dynamics of the of the giant 1960 Chile earthquake, Nature, 437(7057), 404–7,
onset of frictional slip, Science, 330(6001), 211–4, doi:10.1126/science. doi:10.1038/nature03943.
1194777. Corbi, F., F. Funiciello, C. Faccenna, G. Ranalli, and A. Heuret
Bilek, S. L. (2004), Radiated seismic energy and earthquake source duration (2011), Seismic variability of subduction thrust faults: Insights from
variations from teleseismic source time functions for shallow subduction laboratory models, J. Geophys. Res., 116(B6), 1–14, doi:10.1029/
zone thrust earthquakes, J. Geophys. Res., 109(B9), 1–14, doi:10.1029/ 2010JB007993.
2004JB003039. Dalguer, L. A., and S. M. Day (2009), Asymmetric rupture of large aspect-
Bilek, S. L. (2007), Influence of subducting topography on earthquake rup- ratio faults at bimaterial interface in 3D, Geophys. Res. Lett., 36, 1–5,
ture, in The Seismogenic Zone of Subduction Thrust Faults, MARGINS doi:10.1029/2009GL040303.
Theoretical and Experimental Earth Science Series, edited by T. H. Dixon Das, S., and C. H. Scholz (1983),Why large earthquakes do not nucleate at
and J. C. Moore, Columbia Univ. Press, New York. shallow depths, Nature, 305,621–623.
Bilek, S. L., and T. Lay (2002), Tsunami earthquakes possibly widespread Davis, J. B. (1973), Statistic and data analysis in geology. New York:
manifestations of frictional conditional stability, Geophys. Res. Lett., Wiley International, 456–473.
29.14, 1673, doi:10.1029/2002GL015215. Di Giuseppe, E., F. Funiciello, F. Corbi, G. Ranalli, and G. Mojoli (2009),
Bird, P., Y. Y. Kagan, D. D. Jackson, F. P. Schoenberg, and M. J. Werner Gelatins as rock analogs: A systematic study of their rheological and
(2009), Linear and nonlinear relations between relative plate velocity physical properties, Tectonophysics, 473(3-4), 391–403, doi:10.1016/
and seismicity, Bull. Seismol. Soc. Am., 99, 3097–3113, doi:10.1785/ j.tecto.2009.03.012.
0120090082. Dixon, T., and C. Moore (2007), The seismogenic zone of the subduction
Blaser, L., F. Krüger, M. Ohrnberger, and F. Scherbaum (2010), thrust fault: introduction, in The Seismogenic Zone of Subduction
Scaling Relations of Earthquake Source Parameter Estimates with Special Thrust Faults, MARGINS Theoretical and Experimental Earth Science
Focus on Subduction Environment, Bull. Seismol. Soc. Am., 100(6), Series, edited by T. H. Dixon and J. C. Moore, Columbia Univ. Press,
2914–2926, doi:10.1785/0120100111. New York.

1499
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

Dunham, E. M., D. Belanger, L. Cong, and J. E. Kozdon (2011), Mai, P. M., P. Spudich, and J. Boatwright (2005), Hypocenter locations in
Earthquake ruptures with strongly rate-weakening friction and off- finite-source rupture models, 1-45, Bull. Seismol. Soc. Am. 95, 965–980,
fault plasticity, Part 1: Planar faults, Bull. Seismol. Soc. Am., 101, doi: 10.1785/0120040111.
2296–2307. Marone, C., and S. J. D. Cox (1994), Scaling of rock friction constitutive
Freymueller, J. T., H. Woodard, S. C. Cohen, R. Cross, J. Elliott, C. F. Larsen, parameters: The effects of surface roughness and cumulative offset on
S. Hreinsdóttir, and C. Zweck (2008), Active deformation processes in friction of gabbro, Pure Appl. Geophys., 143, 359–385, doi:10.1007/
Alaska, based on 15 Years of GPS measurements, in Active Tectonics BF00874335.
and Seismic Potential of Alaska, AGU Geophysical Monograph, 179, J. T. Marone, C., and E. Richardson (2006), Do earthquakes rupture piece by
Freymueller, P. J. Haeussler, R. Wesson, and G. Ekstrom, eds., pp. 1–42, piece or all together? Science, 313, 1748–1749, 2006.
AGU, Washington, D.C. Marzocchi W., L. Sandri, F. Funiciello, and A. Heuret (2011), On the
Gagnon, K., C. D. Chadwell, and E. Norabuena (2005), Measuring the onset frequency-magnitude distribution of converging boundaries, abstract
of locking in the Peru-Chile trench with GPS and acoustic measurements, presented at the 2011 AGU fall meeting, San Francisco, California.
Nature, 434(7030), 205–8, doi:10.1038/nature03412. McCaffrey, R. (2007), The next great earthquake, Science (New York, N.Y.),
Galeano J., and M. Rubio (1994), Experiments on earthquakes in a contin- 315(5819), 1675–6, doi:10.1126/science.1140173.
uum elastic medium, Ann. Phys. Fr. 19 735–738. Mccaffrey, R. (2008), Global frequency of magnitude 9 earthquakes, Geology,
Geist and Dmowska (1999), Local tsunamis and distributed slip at the 7–11, doi:10.1130/G24402A.1.
source, Pure appl. geophys., 154, 485–512. Moreno, M. S., J. Bolte, J. Klotz, and D. Melnick (2009), Impact of
Goldfinger C., C. H. Nelson, and J. E. Johnson (2003), Holocene earthquake megathrust geometry on inversion of coseismic slip from geodetic data:
records from the Cascadia subduction zone and northern San Andreas Application to the 1960 Chile earthquake, Geophys. Res. Lett., 36(16),
fault based on precise dating of offshore turbidites, Annu. Rev. Earth 1–5, doi:10.1029/2009GL039276.
Planet. Sci., 31, 555–77. Moreno, M., M. Rosenau, and O. Oncken (2010), 2010 Maule earthquake
Grapenthin, R., and J. T. Freymueller (2011), The dynamics of a seismic slip correlates with pre-seismic locking of Andean subduction zone,
wave field: Animation and analysis of kinematic GPS data recorded during Nature, 467(7312), 198–202, doi:10.1038/nature09349.
the 2011 Tohoku-oki earthquake, Japan, Geophys. Res. Lett., 38(18), 1–5, Natawidjaja, D. H., K. Sieh, S. N. Ward, H. Cheng, R. L. Edwards,
doi:10.1029/2011GL048405. J. Galetzka, and B. W. Suwargadi (2004), Paleogeodetic records of seis-
Hanks, T. C., and H. Kanamori (1979), A moment magnitude scale, J. Geophys. mic and aseismic subduction from central Sumatran microatolls, Indone-
Res., 84, 2348–2350. sia, J. Geophys. Res., 109(B4), 1–34, doi:10.1029/2003JB002398.
Hartzell, S. H., and T. H. Heaton (1983), Inversion of strong ground Nelson, B. I., and J. M. Dealy (1993), Dynamic mechanical analysis using
motion and teleseismic waveform data for the fault rupture history complex waveforms, in Techniques in Rheological Measurement, edited
of the 1979 Imperial Valley, California, earthquake, Bull. Seismol. by A. A. Collyer, pp. 197–224, Chapman & Hall, Cambridge.
Soc. Am., 73 (6, Part A), 1553–1583. Nielsen, S., J. Taddeucci, and S. Vinciguerra (2010), Experimental observa-
Heaton, T. H. (1982), The 1971 San Fernando earthquake; a double event? tion of stick-slip instability fronts, Geophys. J. Int., 180(2), 697–702,
Bull. Seismol. Soc. Am., 72: 2037–2062. doi:10.1111/j.1365-246X.2009.04444.x.
Heaton, T. H. (1990), Evidence for and implications of self-healing pulses Ohnaka, M. (2003), A constitutive scaling law and a unified comprehen-
of slip in earthquake rupture, Phys. Earth Planet. Inter., 64(1), 1–20. sion for frictional slip failure, shear fracture of intact rock, and
Heaton, T. H., S. E. Minson, and M. Simons (2011), Characterization of the earthquake rupture, J. Geophys. Res., 108(B2), 2080, doi:10.1029/
Slip Pulse for the 2010 M 8.8 Maule Earthquake, presented at the 2011 2000JB000123.
AGU fall meeting, San Francisco, California. Pacheco, J. F., and L. R. Sykes (1992), Seismic moment catalog of
Heuret, A., S. Lallemand, F. Funiciello, C. Piromallo, and C. Faccenna large shallow earthquakes, 1900 to 1989, Bull. Seismol. Soc. Am., 82,
(2011), Physical characteristics of subduction interface type seismogenic 1306–1349.
zones revisited, Geochem. Geophys. Geosyst., 12(1), 1–26, doi:10.1029/ Peacock, S. M., and R. D. Hyndman (1999), Hydrous minerals in the mantle
2010GC003230. wedge and the maximum depth of subduction thrust earthquakes,
Houston, H. (2001), Influence of depth, focal mechanism, and tectonic Geophys. Res. Lett., 26, 2517–2520, doi:10.1029/1999GL900558.
setting on the shape and duration of earthquake source time functions, Peng, Z., and J. Gomberg (2010), An integrated perspective of the contin-
J. Geophys. Res., 106(B6) 137–150. uum between earthquakes and slow-slip phenomena, Nat. Geosci., 3(9),
Hubbert, M. K. (1937), Theory of scale models as applied to the study of 599–607, doi:10.1038/ngeo940.
geological structures, Geol. Soc. Am. Bull., 48, 1459–1520. Perfettini H., J.-P. Avouac, and J.-C. Ruegg (2005), Geodetic displacements
Hyndman, R. D. (2007), The seismogenic zone of the subduction thrust and aftershocks following the2001Mw58.4Peruearthquake: implications
fault: what we know and don’t know, in The Seismogenic Zone of Sub- for themechanics of the earthquake cycle along subduction zones, J. Geophys.
duction Thrust Faults, MARGINS Theoretical and Experimental Earth Res., 110, B09404.
Science Series, edited by T. H. Dixon and J. C. Moore, Columbia Univ. Plafker, G. (1972), Alaskan earthquake of 1964 and Chilean earthquake of
Press, New York. 1960: Implications for arc tectonics, J. Geophys. Res., 77(5), 901–925.
Ide, S., G. C. Beroza, D. R. Shelly, and T. Uchide (2007), A scaling law for Plafker, G., and J. Savage (1970), Mechanism of the Chilean earthquake of
slow earthquakes., Nature, 447(7140), 76–9, doi:10.1038/nature05780. May 21 and 22, 1960, Geol. Soc. Am. Bull., 81, 1001–1030.
Kanamori, H., and D. L. Anderson (1975), The theoretical basis of some Pollitz, F., P. Banerjee, K. Grijalva, B. Nagarajan, and R. Burgmann (2008),
relations in seismology, Bull. Seismol. Soc. Am., 65, 1073–1095. Effect of 3-D viscoelastic structure on post-seismic relaxation from the
Kaneko, Y., J.-P. Avouac, and N. Lapusta (2010), Towards inferring 2004 M5 9.2 Sumatra earthquake, Geophys. J. Int., 173, 189–204.
earthquake patterns from geodetic observations of interseismic coupling, Raffel, M., C. E. Willert, S. T. Wereley, and J. Kompenhans (2007),
Nat. Geosci., 3(5), 363–369, doi:10.1038/ngeo843. Particle Image Velocimetry A Practical Guide, 448 pp., Oxford Univ.
Kruger, F., and M. Ohrnberger (2005), Tracking the rupture of the Mw = 9.3 Press, New York.
Sumatra earthquake over 1150 km at teleseismic distance, Nature, 435 Ramberg, H., (1981), Gravity, Deformation and the Earth’s Crust, p. 452,
(June), 937–939, doi:10.1038/nature03696. Academic Press, New York.
Kuehn, N. M., S. Hainzl, and F. Scherbaum (2008), Non-Poissonian Reid, H. (1910), The mechanics of the earthquake: The California earth-
earthquake occurrence in coupled stress release models and its effect quake of April 18, 1906, report, vol. 2, 192 pp., State Earthquake Invest.
on seismic hazard, Geophys. J. Int., 649–658, doi:10.1111/j.1365- Comm., Carnegie Inst. of Wash., Washington, D. C.
246X.2008.03835.x.Geophys. Rhie, J., D. Dreger, R. Burgmann, and B. Romanowicz (2007), Slip of the
Lajoie, K. R. (1986), Coastal Tectonics, in Active Tectonics, Studies in 2004 Sumatra-Andaman earthquake from joint inversion of long-period
Geophysics, pp. 95–124, National Academic Press, Washington, D. C. global seismic waveforms and GPS static offsets, Bull. Seismol. Soc.
Lee, S.-J., B.-S. Huang, M. Ando, H.-C. Chiu, and J.-H. Wang (2011), Am., 97(1A), S115–S127, doi:10.1785/0120050620.
Evidence of large scale repeating slip during the 2011 Tohoku-Oki earth- Robinson, D. P., S. Das, and B. Watts (2006), Earthquake rupture stalled
quake, Geophys. Res. Lett., 38(19), 1–6, doi:10.1029/2011GL049580. by a subducting fracture zone, Science, 312(5777), 1203–1205, doi:10.1126/
Lykotrafitis, G., A. J. Rosakis, and G. Ravichandran (2006), Self-healing science.1125771.
pulse-like shear ruptures in the laboratory, Science (New York, N.Y.), Rosakis, A. J., G. Lykotrafitis, K. Xia, and H. Kanamori (2007), Dynamic
313(5794), 1765–8, doi:10.1126/science.1128359. Shear Rupture in Frictional Interfaces: Speeds, Directionality and Modes,
Ma, S., and G. C. Beroza (2008), Rupture Dynamics on a Bimaterial Inter- in Treatise in Geophysics, (G. Schubert, Editor-in-Chief) Vol. 4 - Earth-
face for Dipping Faults, Bull. Seismol. Soc. Am., 98(4), 1642–1658, quake Seismology, (H. Kanamori, Volume Editor) Elsevier, Los Angeles.
doi:10.1785/0120070201. Rosenau, M., J. Lohrmann, and O. Oncken (2009), Shocks in a box: An
Mai, P. M., and G. Beroza (2000), Source Scaling Properties from Finite- analogue model of subduction earthquake cycles with application to
Fault-Rupture Models, Bull. Seismol. Soc. Am., 90(3), 604–615, doi:10.1785/ seismotectonic forearc evolution, J. Geophys. Res., 114(B1), 1–20,
0119990126. doi:10.1029/2008JB005665.

1500
CORBI ET AL.: SUBDUCTION SEISMIC CYCLE SIMULATIONS: 1

Rosenau, M., R. Nerlich, S. Brune, and O. Oncken (2010), Experimental Scholz C. H., and J. Campos (2012), The seismic coupling of subduc-
insights into the scaling and variability of local tsunamis triggered by tion zones revisited, J. Geophys. Res., 117, B05310, doi:10.1029/
giant subduction megathrust earthquakes, J. Geophys. Res., 115(B9), 1–20, 2011JB009003
doi:10.1029/2009JB007100. Simons, M., et al. (2011), The 2011 magnitude 9.0 Tohoku-Oki earthquake:
Rosenau, M., and O. Oncken (2009), Fore-arc deformation controls frequency- mosaicking the megathrust from seconds to centuries., Science, 332
size distribution of megathrust earthquakes in subduction zones, J. Geophys. (6036), 1421–1425, doi:10.1126/science.1206731.
Res., 114(B10), 1–12, doi:10.1029/2012JB009481. Strasser, F. O., M. C. Arango, and J. J. Bommer (2010), scaling of
Rosenau, M., A. Oppelt, H. Kemnitz, and O. Oncken (2008), Analogue the source dimensions of interface and intraslab subduction-zone earth-
earthquakes: Friction experiments with rice and implications for the quakes with Moment Magnitude, Seismo. Res. Lett., 81(6), 941–950,
seismic cycle and earthquake nucleation, Boll. Geofis., 49,104–108. doi:10.1785/gssrl.
Rubin, A. M., and D. Gillard (2000), Aftershock asymmetry/rupture Sveen, J. K. (2004), An introduction to matpiv v.1.6.1 Eprint no. 2, ISSN
directivity among central San Andreas fault microearthquakes, J. Geophys. 0809-4403, Department of Mathematics, University of Oslo, http://
Res.- Solid Earth, 105, 19,095–19,109. www.math.uio.no/ jks/matpiv.
Rubinstein, J. L., W. L. Ellsworth, N. M. Beeler, B. D. Kilgore, D. A. Thatcher, W. (1984), The earthquake deformation cycle at the Nankai
Lockner, and H. M. Savage (2012a), Fixed recurrence and slip models Trough, J. Geophys. Res., 89, B5, 3087–3101.
better predict earthquake behavior than the time- and slip-predictable van Dinther, Y., T. V. Gerya, F. Corbi, L. A. Dalguer, F. Funiciello, and
models: 2. Laboratory earthquakes, J. Geophys. Res., 117(B2), 1–13, P. M. Mai (2013), The seismic cycle at subduction thrusts: 2. Benchmark-
doi:10.1029/2011JB008723. ing geodynamic numerical simulations and analogue models, J. Geophys.
Rubinstein, J. L., W. L. Ellsworth, K. H. Chen, and N. Uchida (2012b), Res., doi:10.1002/jgrb.20048.
Fixed recurrence and slip models better predict earthquake behavior than Vigny, C., et al. (2011), The 2010 Mw 8.8 Maule megathrust earthquake of
the time- and slip-predictable models: 1. Repeating earthquakes, J. Geophys. Central Chile, monitored by GPS., Science (New York, N.Y.), 332(6036),
Res., 117(B2), 1–23, doi:10.1029/2011JB008724. 1417–21, doi:10.1126/science.1204132.
Ruff, L. J., and H. Kanamori (1983), Seismic coupling and uncoupling at Wang, K. (2007), Elastic and viscoelastic models of crustal deformation in
subduction zones, Tectonophysics, 99, 99–117, doi:10.1016/0040-1951 subduction earthquake cycles, in The Seismogenic Zone of Subduction
(83)90097-5. Thrust Faults, MARGINS Theoretical and Experimental Earth Science
Rudnicki, J. W., and M. Wu (1995), Mechanics of dip-slip faulting in an Series, edited by T. H. Dixon and J. C. Moore, Columbia Univ. Press,
elastic half-space, J. Geophys. Res. 100, no. B11, 22,173–22,186. New York.
Saffer, D. M., D. A. Lockner, and A. McKiernan (2012), Effects of smectite Wang, K., and J. He (2008), Effects of Frictional Behavior and Geometry of
to illite transformation on the frictional strength and sliding stability of Subduction Fault on Coseismic Seafloor Deformation, Bull. Seismol. Soc.
intact marine mudstones, Geophys. Res. Lett., 39(11), 2–7, doi:10.1029/ Am., 98(2), 571–579, doi:10.1785/0120070097.
2012GL051761. Wang, K., and S. L. Bilek (2011), Do subducting seamounts generate or stop
Satake, K., and B. F. Atwater (2007), Long-Term Perspectives on Giant large earthquakes?, Geology, 39(9), 819–822, doi:10.1130/G31856.1.
Earthquakes and Tsunamis at Subduction Zones, Ann. Rev. Earth and Wang, K., Y. Hu, and J. He (2012), Deformation cycles of subduction earth-
Planetary Sciences, 35(1), 349–374, doi:10.1146/annurev.earth.35.031306. quakes in a viscoelastic Earth, Nature, 484(7394), 327–332, doi:10.1038/
140302. nature11032.
Shimazaki K. (1986), Small and large earthquakes: The effects of the Weijermars, R., and H. Schmeling (1986), Scaling of Newtonian and non
thickness of the seismogenic layer and free surface, Maurice Ewing Ser. Newtonian fluid dynamics without inertia for quantitative modeling of
6, 209–216. rock flow due to gravity (including the concept of rheological similarity),
Shimazaki, K., and T. Nakata (1980), Time-predictable recurrence model Phys. Earth Planet. Inter., 43, 316–330, doi:10.1016/0031-9201(86)
for large earthquakes, Geophys. Res. Lett., 7(4), 279–282. 90021-X.
Scholz, C. H. (1982), Scaling laws for large earthquakes: concequences for Wells, D. L., and K. J. Coppersmith (1994), New Empirical Relationships
physical models, Bull. Seismol. Soc. Am., 72(1), 1–14. among Magnitude, Rupture Length, Rupture Width, Rupture Area, and
Scholz, C. H. (1990), Earthquakes as Chaos, Nature, 348, 197–198, Surface Displacement, Bull. Seismol. Soc. Am., 84(4), 974–1002.
doi:10.1038/348197a0. Xia, K., A. J. Rosakis, and H. Kanamori (2004), Laboratory earthquakes:
Scholz, C. H. (2002), The Mechanics of Earthquakes and Faulting, 2nd ed., the sub-Rayleigh-to-supershear rupture transition., Science (New York,
471 pp., Cambridge Univ. Press, Cambridge, U.K. N.Y.), 303(5665), 1859–61, doi:10.1126/science.1094022.

1501

You might also like