You are on page 1of 16

Electronic Seismologist

The Bayesian Earthquake Analysis


Tool
Hannes Vasyura-Bathke*1,2, Jan Dettmer3, Andreas Steinberg4, Sebastian
Heimann5, Marius Paul Isken4, Olaf Zielke1, Paul Martin Mai1, Henriette Sudhaus4,
and Sigurjón Jónsson1

Abstract
The Bayesian earthquake analysis tool (BEAT) is an open-source Python software to con-
duct source-parameter estimation studies for crustal deformation events, such as earth-
quakes and magma intrusions, by employing a Bayesian framework with a flexible
problem definition. The software features functionality to calculate Green’s functions
for a homogeneous or a layered elastic half-space. Furthermore, algorithm(s) that
explore the solution space may be selected from a suite of implemented samplers. If
desired, BEAT’s modular architecture allows for easy implementation of additional fea-
tures, for example, alternative sampling algorithms. We demonstrate the functionality
and performance of the package using five earthquake source estimation examples:
a full moment-tensor estimation; a double-couple moment-tensor estimation; an esti-
mation for a rectangular finite source; a static finite-fault estimation with variable slip;
Cite this article as Vasyura-Bathke, H.,
and a full kinematic finite-fault estimation with variable hypocenter location, rupture
J. Dettmer, A. Steinberg, S. Heimann,
velocity, and rupture duration. This software integrates many aspects of source studies M. P. Isken, O. Zielke, P. M. Mai,
and provides an extensive framework for joint use of geodetic and seismic data for H. Sudhaus, and S. Jónsson (2020). The
Bayesian Earthquake Analysis Tool,
nonlinear source- and noise-covariance estimation within layered elastic half-spaces. Seismol. Res. Lett. XX, 1–16, doi: 10.1785/
Furthermore, the software also provides an open platform for further methodological 0220190075.

development and for reproducible source studies in the geophysical community. Supplemental Material

Introduction displacements caused by a magma intrusion in the crust, an


Crustal deformation processes, such as earthquakes, volcanic isotropic moment tensor (MT) can sometimes be a reasonable
intrusions, or due to human activities, can have severe socio- approximation to model the observations (e.g., Bathke et al.,
economic impacts. For appropriate hazard assessment of these 2011; Biggs et al., 2014). For earthquakes, a full or deviatoric
phenomena, better understanding of the physical processes MT (Jost and Herrmann, 1989; Sokos et al., 2012; Tape and
causing such deformation is required. An important step Tape, 2012, 2015; Stähler and Sigloch, 2014) in combination
toward this goal is to carefully examine the geophysical observ- with a single source time function may be sufficient to explain
ables caused by, for example, an earthquake rupture, and then seismological observations when the event is not large or is
use these data to infer the spatial and temporal evolution of the recorded at larger distances. However, the point-source
earthquake. The static surface deformation caused by such assumption of the MT is generally not valid for earthquake
events can be measured by geodetic techniques such as Global ruptures. Therefore, finite sources with a spatial extent
Navigation Satellite Systems (GNSS, e.g., Global Positioning approximating earthquake fault planes have been applied.
System, Galileo, and Glonass) and Interferometric Synthetic These include simple rectangular sources (Haskell, 1964; Okada,
Aperture Radar (InSAR) (e.g., Jónsson et al., 2002; Sudhaus 1985) and more complex finite-fault sources with hundreds of
and Jónsson, 2009; Bathke et al., 2013). Rapid ground-surface source parameters (e.g., Olsen and Apsel, 1982; Hartzell and
displacements caused by a seismic wavefield, on the other hand, Heaton, 1983; Ji et al., 2002). The variety of finite-fault
can be measured with seismometers and in some cases GNSS.
Both types of measurements can be employed to study the char-
1. King Abdullah University of Science and Technology, Thuwal, Saudi Arabia; 2. Now
acteristics of the deformation source by investigating the inverse at Institute for Geosciences, University of Potsdam, Potsdam, Germany; 3. University
problem that relates source parameters to the observed data of Calgary, Calgary, Canada; 4. Department of Geoscience, Christian-Albrecht-
University, Kiel, Germany; 5. GFZ German Research Center for Geosciences, Potsdam,
(e.g., Kikuchi and Kanamori, 1982; Yagi and Fukahata, 2011). Germany
Depending on the process being studied, the source can *Corresponding author: hvasbath@uni-potsdam.de
be parameterized in various ways. For example, to explain © Seismological Society of America

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 1

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
TABLE 1
Recent, Open-Source Software for Bayesian Estimation of Deformation Sources

Source Time GF
Name UQ, Samplers Source Types Function (STF) Data Types Support References

MTfit Bayesian, MH, Single source, linear full MT, No Wave polarity, No Pugh and
RJMCMC, and DC wave amplitude White (2018)
and MC ratios

GBIS Bayesian, AMH Multiple sources, isotropic point, No Static displacements Half-space Bagnardi and
spherical, rectangular, penny- (GNSS, InSAR) only Hooper (2018)
shaped crack, and ellipsoid

Bayes ISOLA Bayesian, grid Single source, quasi-nonlinear No Seismogram Yes Vackář et al.
search, and LSQ full MT waveforms (2017)

BEAT Bayesian, AMH, Multiple sources, nonlinear full MT, Half-sinusoidal, Seismogram Yes This work
SMC, and PT rectangular source, isotropic MT, triangular, and waveforms,
CLVD, DC, and full static or boxcar static displacements
kinematic finite fault (GNSS, InSAR)

AMH, adaptive Metropolis–Hastings; Bayes ISOLA, Bayesian isolated asperities; BEAT, Bayesian earthquake analysis tool; CLVD, compensated linear vector dipole; DC, double
couple; GBIS, geodetic Bayesian inversion software; GF, Green’s function; GNSS, Global Navigation Satellite System; InSAR, Interferometric Synthetic Aperture Radar, LSQ, least-
squares inversion; MC, Monte Carlo random sampling; MH, Metropolis–Hastings; MT, moment tensor; MTfit, moment tensor fit; PT, parallel tempering; RJMCMC, reversible
jump Markov chain Monte Carlo; SMC, sequential Monte Carlo; UQ, uncertainty quantification.

representations that have been used is large, even for the same of parameters (Monelli and Mai, 2008; Razafindrakoto and Mai,
earthquake, as illustrated in the source models database (Mai 2014) and prior information can be consistently integrated in
and Thingbaijam, 2014). the problem definition (Xu et al., 2015; Dutta et al., 2018).
For meaningful interpretation of estimation results, para- Only a few software packages are available for nonlinear
meter uncertainties should be quantified. These uncertainties Bayesian inference combining static and dynamic displace-
are mostly caused by measurement errors and theory errors ment data (Table 1). For example, moment tensor fit (MTfit)
(Tarantola and Valette, 1982). The measurement error is partly samples the parameters of a full MT while fixing the centroid
due to data being contaminated by random noise during meas- location (Pugh and White, 2018). The algorithm considers
urement. This noise at seismic stations can be reduced, for amplitude ratios and the polarity of seismic phases. Although
example, by employing better instruments and recording in it does not support seismic waveforms, the software is useful
low-noise environments. The most important error compo- for seismic events of low-moment magnitudes (M w ∼ 3).
nents in InSAR data are of atmospheric origin and may be Bayesian isolated asperities (BayesISOLA) is a program that
reduced, for example, through using additional independent supports seismic waveforms to sample the parameters of a full
data or through advanced processing strategies (Bekaert et al., MT in a mixed linear–nonlinear setting (Vackář et al., 2017).
2015). Theory errors arise from assumptions in the mathemati- The location and source time are sampled via grid search and
cal formulation relating the model parameters to the observed the moment tensor components are determined with a linear
data. For example, layered models that simplify the velocity least-squares inversion. Geodetic Bayesian inversion software
structure of the Earth are used in the computation of Green’s (GBIS) provides a Bayesian framework for sources with simple
functions (GFs, e.g., Yagi and Fukahata, 2011; Minson et al., geometrical shapes (prolate spheroid, penny-shaped crack,
2013; Duputel et al., 2014). Also, the parameterization of the etc.) (Bagnardi and Hooper, 2018) that have analytic closed
deformation source, like the discretization of the earthquake form solutions, assuming a homogeneous elastic half-space
fault plane into a fixed number of patches, introduces theory Earth structure. Because of these source parameterizations,
errors (Dettmer et al., 2014). GBIS is limited to static surface displacement data, but can
Parameterizations involving the source location, geometry, be useful in volcano geodesy (Table 1). Although the advan-
and source time function are nonlinearly related to the surface tages of using Bayesian inference in deformation source esti-
displacements caused by the source. Consequently, inferring mations are undeniable, a software package that is sufficiently
the source parameters and uncertainties from displacements general for a wide range of applications has been lacking.
requires nonlinear evaluation methods, which have high- In this work, we present the Bayesian earthquake analysis
computational costs (Dettmer et al., 2007). Bayesian inference tool (BEAT), a general suite of programs for a wide range of
is increasingly being applied to solve nonlinear estimation prob- linear and nonlinear deformation source studies (Tables 1 and
lems because it has the advantage of quantifying uncertainties 2). BEAT provides the first open-source implementation for

2 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
finite-fault estimations with variable static and/or kinematic In this study, we are mostly investigating nonlinear prob-
slip parameters (including hypocenter location) using Bayesian lems for which no analytic solution exists. Therefore, we have
inference. Our approach can include static and/or dynamic to solve the problems numerically. We assume a multivariate
deformation data, and the user can specify which seismic Gaussian distribution with error scaling to estimate the scale of
phases to include in the estimation. In addition, BEAT uses GF the unknown distribution of residuals as a hierarchical param-
databases (Heimann, 2011; Heimann et al., 2019) to speed up eter for each dataset dobs;k (Fukuda and Johnson, 2008) of K
the source parameter estimations, in which the user can also datasets. Thus, the PPD becomes
customize the 1D subsurface Earth structure. Finally, the soft-
ware supports several sampling algorithms that can be chosen EQ-TARGET;temp:intralink-;df4;308;652 pm; σ 1 ; σ 2 ; …; σ K jdobs;1 ; dobs;2 ; …; dobs;K 
depending on the problem to solve and the available comput- Y
K
ing resources. In conclusion, BEAT presents a complete open- ∝ pm × Lm; σ k ; 4
source software infrastructure that can be applied to a wide k1
range of problems and can also be extended by the user to fos-
ter future development of new methods, and to support repro- with
ducible research (Mai, Schorlemmer, et al., 2016; Mai, Shearer,
1
EQ-TARGET;temp:intralink-;df5;308;562

et al., 2016). Lm; σ k  


2πσ 2k N k =2 jCxk j1=2
 
Bayesian Inference 1 T −1
× exp − 2 dobs;k − dk m Cxk dobs;k − dk m ;
To estimate model parameters of a selected deformation source 2σ k
from geodetic and/or seismic data we apply Bayesian inference. 5
The posterior probability density (PPD) of the model param-
eters pmjdobs  can be obtained following Bayes’ theorem: in which dm is the N k predicted static or transient displace-
ments for dataset k. The covariance matrix Cxk describes the
EQ-TARGET;temp:intralink-;df1;41;455 pmjdobs  ∝ pmpdobs jm; 1 variances and covariances of residuals for each dataset. The
inverses of the covariance matrices in combination with the
in which m is a vector of the parameters to estimate, that is, the hierarchical scaling parameters σ k act as weights for the differ-
source parameters and optional hierarchical parameters (often ent datasets. Although BEAT includes significant infrastruc-
called hyperparameters), and dobs are observed data, that is, ture for addressing the covariance estimation, the details on
seismic waveforms and/or geodetic surface displacements. the inference process are described elsewhere (Dettmer et al.,
A priori information is expressed through pm for m, and 2007; Yagi and Fukahata, 2011; Duputel et al., 2014; Hallo and
pdobs jm is the probability distribution of residuals σ k (con- Gallovic, 2016).
sisting of measurement and theory errors) for a given set of m.
For a given set of observed data, pdobs jm is interpreted as the Earth structure and GF
likelihood function Lm, which quantifies the probability that Before computing data predictions dm for a set of source
a particular set of model parameters gave rise to the data. The mode parameters m, we need to make assumptions about the
observed data dobs can be regarded as the real quantity dreal in elastic subsurface structure. We assume the structure is a lay-
addition to measurement errors emeas , ered elastic half-space described by the density, thickness, seis-
mic-wave velocity, and attenuation of each layer. Once the
EQ-TARGET;temp:intralink-;df2;41;259 dobs  dreal  emeas : 2 subsurface structure has been defined, a linear combination
of 10 (eight for the far field) elementary GFs is needed to com-
The forward model dm simulates displacements generated pute time series of displacements for a general moment tensor
by the chosen deformation source dsource and includes theory source (Heimann, 2011; Heimann et al., 2019).
errors eth (Tarantola and Valette, 1982), The GFs are numerically expensive to compute, and it is
thus advantageous to precompute a database of GFs prior
EQ-TARGET;temp:intralink-;df3;41;183 dm  dsource  eth : 3 to the model parameter estimation. Here, the software library
Pyrocko-GF is employed to generate databases of GFs
The theory errors result from assumptions and simplifications (Heimann et al., 2017, 2019), which is an approach that allows
in the mathematical formulation of the forward model, for customizing the subsurface layering below each recording loca-
example, from assuming a simple subsurface model that differs tion. For example, the shallow subsurface structure from the
from the true subsurface structure. Typically, measurement CRUST2.0 model (Bassin et al., 2000) could be included for
errors and theory errors are not exactly known independently each station to complement a common global Earth structure
and probability distributions are thus commonly assumed to below the crust (Kennett et al., 1995). Estimation of the MT cent-
represent these errors. roid location is possible once the GFs have been precomputed for

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 3

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
TABLE 2
Types of Deformation Sources and Their Source Specific Parameters That Can Be Estimated in the Geometry Mode
of the Bayesian Earthquake Analysis Tool (BEAT)

Deformation Source Parameters m References

Full MT I mnn , mee , mdd , mne , mnd , med , and moment magnitude Stähler and Sigloch (2014)

Full MT II u, v, κ, σ, h, and moment magnitude Tape and Tape (2015)

Isotropic MT Moment magnitude or volume change Kumagai et al. (2014)

DC Strike, dip, rake, and moment magnitude Jost and Herrmann (1989)

CLVD Dip, azimuth (of largest dipole), and moment magnitude Jost and Herrmann (1989)

Rectangular source Strike, dip, rake, length, width, slip, optional: Haskell (1964)
hypocentral location (x, y), and rupture velocity

CLVD, compensated linear vector dipole; DC, double couple; MT, moment tensor.

an appropriate source–receiver volume, which requires that the of the same type, each with a full and independent set of source
source region is discretized with a GF grid. model parameters, allowing for custom and complex model
The GF databases are setup such that different programs setups. With this flexible setup, complex source shapes such
can use them. Currently, support is included for quick seismo- as caldera ring faults (e.g., Bathke et al., 2015), faults with var-
grams (QSEIS) (Wang, 1999), quick seismograms for spherical iable strike and dip (e.g., R. Dutta et al., unpublished manu-
earth models (QSSP) (Wang et al., 2017), and post-seismic script, 2019, see Data and Resources), and curved dike
Green’s functions and post-seismic deformation computation intrusions (e.g., Chadwick et al., 2011) can be studied.
(PSGRN/PSCMP) (Wang et al., 2006). QSEIS calculates seis-
mograms based on a layered viscoelastic half-space model and Estimation of the slip history on a finite fault
should be used if local or regional setups are of interest. QSSP For a finite-fault source, under the assumption that the source
calculates seismograms based on a layered, self-gravitating fault geometry is known, the aim is to resolve the spatiotem-
Earth and should be used for teleseismic setups or if interaction poral slip pattern on subfaults. These sources are commonly
of the crust and atmosphere is important. PSGRN/PSCMP cal- discretized more finely than can be resolved by data (overpar-
culates synthetic stress, strain, displacement, tilt, and gravita- ameterization), which requires regularization to avoid unstable
tional fields on a layered viscoelastic half-space and should be solutions due to the ill-posed inverse problem (Altman and
used for static displacement data. Gondzio, 1999). Regularization is typically applied via smooth-
Therefore, this infrastructure allows computing geodetic ness constraints based on a trade-off between slip smoothness
static displacements (at any depth) and seismic waveforms at and data fit (Jónsson et al., 2002). In Bayesian inference,
any distance and depth range with desired frequency content. smoothing regularization can be achieved via a Gaussian prior
In addition, the open-source approach of BEAT makes exten- psjα, which includes a covariance matrix with off-diagonal
sions through user contributions straightforward to include. terms, whose Cholesky decomposition is equivalent to a
Laplacian finite-difference operator L of size M × M. The
Estimation of source location and geometry trade-off parameter α scales the degree of smoothing and can
The first-order parameters of a deformation source are source be assumed to be an unknown parameter (i.e., a hierarchical
location (east, north, and depth), source time function (time scaling that is estimated based on data information). Including
and duration), and source geometry. These are determined in smoothing regularization of slip values s on M discretized fault
a nonlinear estimation (equation 5), and the number of source patches, the PPD becomes (Fukuda and Johnson, 2008)
parameters m depends on the choice of deformation source.
BEAT supports a variety of point sources: a full MT, an iso- EQ-TARGET;temp:intralink-;df6;320;184 pm; σ 1 ; σ 2 ; …; σ K jdobs;1 ; dobs;2 ; …; dobs;K 
tropic MT, a double-couple MT, and a compensated linear vec- Y
K
tor dipole (CLVD) MT (Table 2). By superposition of MTs, the ∝ pm × psjα × pdobs;k jm; σ k ; 6
GF database infrastructure permits discretizing finite sources k1
of arbitrary geometry and temporal evolution. For example, the
parameters of a rectangular source (Haskell, 1964; Heimann, with
2011) with unknown rupture nucleation and rupture velocity  
can be inferred (Table 2, Fig. S1, available in the supplemental jLT Lj1=2 1
psjα  exp − LsT
Ls : 7
2πα2 M=2 2α2
EQ-TARGET;temp:intralink-;df7;320;97

material to this article). BEAT also allows for multiple sources

4 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
The application of Bayesian methods by themselves does not directly from the PPD and allows to estimate general distribu-
remove the need for regularization; it stems from the choice of tions in which no closed-form solutions are available.
spatial discretization of the fault. As long as the model is over- However, nonlinear source estimations often result in com-
parameterized, regularization is needed. However, Bayesian plex, multimodal solutions that require advanced sampling
model selection can be applied to overcome the need for algorithms to fully sample the complex multidimensional PPD
regularization and subjective discretization choices (Dettmer and to avoid misleading results. For BEAT, we implemented
et al., 2014). three advanced sampling algorithms: adaptive Metropolis–
For static geodetic data, the slip values on the fault patches Hastings (AMH) (Roberts and Rosenthal, 2001), SMC (Jarzynski,
are linearly related to the resulting surface displacements and 1997; Neal, 2001; Moral et al., 2006; Ching and Chen, 2007;
dm reduces to Gm. Here, G is the sensitivity kernel contain- Minson et al., 2013), and parallel tempering (PT) (Geyer, 1991;
ing displacement GFs for a source with unit slip for each slip Jasra et al., 2007; Dettmer and Dosso, 2012). In the following
direction (slip parallel and slip perpendicular) on each subfault paragraphs, we give a short description of each algorithm and
(Aki and Richards, 2002). With seismic waveform data, kin- refer the reader to the original publications for details.
ematic rupture parameters can be estimated. These include The AMH algorithm tunes the step size of the proposal dis-
hypocenter location and time, rupture velocity, and the rupture tribution based on the acceptance rate of the last N samples.
duration. If rupture velocities are treated as an unknown veloc- If the acceptance rate is low, smaller steps are proposed and
ity field on the fault, the Eikonal equation can be solved in vice versa. This algorithm is useful when sampling problems
each forward operation to ensure causal rupture propagation with only a few model parameters, for example, when estimat-
(Minson et al., 2013). This leads to nonlinearity and G consists ing an isotropic MT from geodetic data to resolve the location
of two additional fields, that is, the rupture duration and rup- and volume change of a point source underneath a volcano
ture start time on each slip patch for each seismic station. (Kumagai et al., 2014).
In BEAT, G is referred to as a library and can be In higher dimensional problems, particularly with seismic
assembled given the source geometry and the desired subfault data, the AMH algorithm appears to be insufficient. Here,
discretization. This process requires a GF database to be pre- the SMC algorithm performs better due to its ability to temper
computed (see the Earth Structure and GF section). When the likelihood, apply resampling, and to take advantage of par-
working with these libraries, it is useful to precompute allel computing. In the SMC algorithm, samples are obtained
them to make the numerical sampling of the posterior compu- by simulating a sequence of intermediate distributions from
tationally tractable and more efficient. However, the libra- the prior to the posterior. Each sample starts from the prior
ries can be large for kinematic cases, leading to memory and follows a trajectory through the intermediate distributions
limitations. To overcome these limitations, we recommend to the posterior, where it is independent from other samples.
to apply as much a priori information about the source param- The transitioning between distributions is typically achieved by
eters as possible when defining the problem. For example, the scaling from the prior to the posterior via an annealing param-
rupture velocity of an earthquake is likely to be close to the eter that allows samples to initially move freely in the param-
shear-wave velocity of the subsurface medium (e.g., Das, eter space, but gradually becomes more constrained by the data
2015). Consequently, the prior on the rupture velocity on each as the sample approaches the posterior. At various stages,
patch can be adjusted accordingly, if rupture propagation with resampling can be applied to refocus the large number of sam-
super shear-wave velocity is not expected. ples into regions of high likelihood. This process also helps the
algorithm to avoid local minima. Although slowly decreasing
Sampling Algorithms the annealing parameter for each trajectory, a sample transi-
The PPD is usually numerically estimated through sampling tions through the bridging distributions until the PPD is
with Markov chain Monte Carlo (MCMC) methods (Sambridge sampled. The annealing enables sampling of complex, multi-
and Mosegaard, 2002) or sequential Monte Carlo (SMC) meth- modal, and highly peaked distributions since data information
ods (Moral et al., 2006). A commonly used algorithm in the is gradually introduced. However, the number of trajectories
family of MCMC samplers is Metropolis–Hastings sampling needs to be sufficiently large with respect to the number of
(MHS) (Metropolis et al., 1953). In MHS, the solution space parameters (Minson et al., 2013). We implemented the SMC
is explored by a random walk, in which the direction and size algorithm with Python’s multiprocessing library, and
(or length) of a step is randomly proposed by drawing from a consequently, this implementation cannot scale beyond a sin-
probability distribution (Hastings, 1970). A new state in the gle node of a computer cluster.
random walk is accepted or rejected probabilistically; if the For parallel computing architectures (e.g., computer clus-
new state has a higher or equal posterior probability than the ters), we implemented the PT algorithm with the message pass-
current state, the new state is always accepted. If the new state ing interface (MPI). This algorithm generates many MCMC
has a lower posterior probability, the new state is sometimes chains in parallel [one per central processing unit (CPU)],
accepted and sometimes rejected. This algorithm samples which draw samples from a joint posterior that includes the

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 5

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
target posterior as well as tempered posteriors in which the
likelihood function is raised to a power of less than unity.
The chains are typically split such that half of the chains sam-
ple the target posterior whereas the other half samples the tem-
pered distributions. In addition, PT includes exchange moves,
in which various chains communicate and exchange informa-
tion by swapping parameter vector m. The tempered posteriors
significantly help the algorithm to widely explore the param-
eter space and reduce the tendency of getting stuck in modes.

Seismic Covariance Matrix Structure


In BEAT, the covariance matrices Cx for the seismic wave-
forms are estimated during the sampling, based on data infor-
mation and assumptions on the noise structure. It is important
to note that the noise covariance matrix substantially influ-
ences the Bayesian estimation result (Dettmer et al., 2007;
Duputel et al., 2012). This matrix is generally not known a pri- Figure 1. Bayesian earthquake analysis tool (BEAT) modular
ori and it therefore has to be considered as a part of the esti- program structure. The main problem objects determine,
mation problem. However, as the residuals include theory together with their composites, the model that contains the
rules to calculate the posterior probability. To date, BEAT con-
errors (see equations 2 and 3), this task is not trivial. In BEAT,
tains two problem classes, the geometry problem (see the
three different forms of the covariance matrix are currently Estimation of Source Location and Geometry section) and the
available: variance, exponential (toeplitz), and non- finite-fault estimation problem (see the Estimation of the Slip
toeplitz. For the variance and exponential forms, History on a Finite Fault section). These are being used by the
the data variance is estimated from the seismic data by analyz- sampler objects to explore the solution space of their respective
posterior probability density (PPD). GNSS, Global Navigation
ing the data variability prior to the P-phase arrival. The
Satellite System; InSAR, Interferometric Synthetic Aperture Radar;
non-toeplitz matrix on the other hand is estimated with PDF, probability density function.
an iterative approach and is based on the residual waveforms
for the maximum a-posterior (MAP) model parameters (Dettmer
et al., 2007). The non-Toeplitz covariance matrix accounts for
both nonstationary and correlated noise. Because it is based on for a magnitude M w > 6:5 earthquake yields a fault plane strik-
the residuals, it includes the effects of both measurement errors ing between 130° and 140°, this information can be used in the
and theory errors, for example, due to inaccurate velocity ffi mode to constrain the fault geometry. Otherwise, a refer-
structure of the subsurface. ence fault geometry must be specified from other information.
A problem-specific YAML configuration file (e.g., config_-
Software Features geometry.yaml in Fig. 2) provides an interface for the user to
BEAT is written in the Python programming language and adjust the various parameters of the problem, including prior
provides a Bayesian sampling framework for deformation information, data processing, and sampling-algorithm tuning.
source estimation. The project website, program prerequisites, Two datatypes are supported in BEAT: seismic (wave-
platform, and license information can be found in Data and forms) and geodetic (GNSS and InSAR static displacements).
Resources. Several parts of BEAT are based on the Pyrocko The data types are implemented in terms of composites that
seismological toolbox (Heimann et al., 2017). The program- can be combined in the problem setup (Fig. 1). The problem
ming structure is object-oriented and designed around two and composites classes include the formulation of the pos-
main estimation problems, which are referred to as modes and terior likelihood (equation 5), which is then synthesized to the
illustrated in Figure 1. The first mode, geometry, addresses model (Fig. 1) object that can be used by any sampler to explore
the nonlinear problem of estimating the geometry, source time the solution space of the PPD. For numerical efficiency, the
function, location and magnitude, or slip of the deformation model is implemented via the open-source libraries theano
sources (see the Estimation of Source Location and Geometry and pymc3, which provide the option to translate Python code
section). The second mode, finite-fault inversion (ffi), is used to C code, making it computationally more efficient, or to
to estimate the spatially variable static-slip values or kinematic Compute Unified Device Architecture (CUDA) C code to make
rupture parameters (see the Estimation of the Slip History on a use of graphical processing units.
Finite Fault section). While the modes can operate independ- When using the software, the directory structure of the
ently, we recommend to apply a geometry estimation prior working directory gradually grows (Fig. 2). For example, the
to an ffi estimation. For example, if a geometry estimation mode specific directories, such as geometry, are created

6 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
complexity of the source parameterization, and the available
computational resources. In the following, we will provide appro-
ximate computation times for the five estimation examples.

Example 1: full MT
Here, we used the geometry mode to estimate parameters of
an M w 5.5 MT source from simulated seismic data at regional
distances up to 1000 km (Fig. S2). Simulated waveforms were
computed using a double-couple MT source with a sinusoidal
source time function (on 0; π), and the simulated waveforms
(Figs. S3 and S4) were contaminated with simulated, tempo-
rally, and spatially correlated noise. To obtain such correlated
noise for each waveform, we summed up 300 synthesized
waveforms containing the signals of random, full MT sources
in the epicentral region (3 km) with magnitudes between 3.0
Figure 2. An example of a directory tree for a BEAT project after all and 4.7, and including five larger magnitude sources between
steps have been executed. The config_*.yaml files are the 4.5 and 5.0 with random source time variations between −20
parameter files to configure and customize the sampling process.
and 5 s with respect to the reference event. The corresponding
The results directory (created by subcommand export)
contains sampling results, estimated covariance matrices, and signal-to-noise ratio varies between 1.2 and 3.3. We applied
synthetics. The hypers directory contains sampling results of the SMC sampling with an MTSource parameterization to sam-
initial hyperparameter estimation. The linear_gfs directory ple the parameters of a full MT, its centroid location, and its
contains the Green’s functions (GFs) for the finite-fault inversion source time function.
(created by command build_gfs). stage_* folders contain To demonstrate the influence of different parameteriza-
results of the different sampling stages of the sequential Monte
tions for the data error covariance matrix (see the Seismic
Carlo (SMC) sampler (* represents the integer number of the
corresponding stage) and stage_−1 contains the PPD of the Covariance Matrix Structure section), we compared results
model parameters of the corresponding final stage. of sampling based on a diagonal covariance matrix (uncorre-
lated noise, variance type in BEAT) and a non-Toeplitz, full
covariance matrix, which accounts for time-dependent and
nonstationary noise (Dettmer et al., 2007). The comparison
during the course of the sampling and the figures directory shows that the double-couple mechanism was not retrieved in
after the results are plotted. the estimation when neglecting noise correlations (Fig. 3a,c),
but was well recovered when using a non-toeplitz covari-
Application Examples ance matrix (Fig. 3b,d). With only data noise contamination in
To demonstrate the capabilities of BEAT, we present five different this synthetic test, the uncertainty of the solution is low for the
source estimation examples. For the first example, we generated variance covariance matrix case and the data are overfitted
seismic waveforms from an MT point source and contaminated (Figs. S5 and S6). However, this overfitting was reduced using
them with synthetic correlated noise. These synthetic data were non-Toeplitz covariance matrices, and the resulting parameter
then used to estimate the components of a full MT. In the other marginals are less biased (Figs. S7–S9) than when diagonal
four examples, we considered both geodetic and seismic data of covariance matrices were used.
the 2009 L'Aquila earthquake. First, we estimate the location For this example, we used four CPU cores on a standard
and the parameters of a double-couple point source. Then, we mobile computer. We configured the SMC sampler with 2000
employed a rectangular source to estimate the fault geometry Markov chains (MCs) over 300 steps over 40 stages. During the
of the rupture assuming uniform slip. With the MAP parameters sampling, the forward model was evaluated 24 million times,
from the rectangular fault solution we then estimate spatially which took about 4 hr.
variable static slip, while keeping the fault geometry fixed. Finally,
we estimated kinematic parameters of the rupture based on the Example 2: double-couple MT
results of the previous two steps. Here, hypocentral location, rup- We considered data from the 2009 L'Aquila earthquake to esti-
ture velocity, and rupture duration were treated as unknowns. mate the double-couple moment tensor (DCSource) using
These scenarios are fully reproducible and available as step- the geometry mode of BEAT. We included P waveforms
by-step tutorials (see Data and Resources). from 35 seismic stations at teleseismic distances between
The uncertainty quantification capability of BEAT, and its 30° and 90° (Fig. S10). The seismic data were filtered with a
flexibility can result in significant computational costs. The Butterworth band-pass filter between 0.01 and 0.2 Hz. We
overall computation time depends on the amount of data, the applied a time window of −10 to 40 s around the theoretical

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 7

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
(a) +Isotropic (b) +Isotropic PPD until 400,000 samples had
been collected, which resulted
+Crack +Crack in ∼60 million forward evalua-
+Dipole +Dipole
tions and required ∼5 hr of
runtime. The SMC sampler
+CLVD +CLVD was run with 5000 MCs and
–CLVD –CLVD 400 steps that resulted in 31
stages in which the forward
–Dipole –Dipole
model has been evaluated
–Crack –Crack
∼62 million times and required
∼6 hr of runtime.
–Isotropic –Isotropic

(c) (d) Example 3: rectangular


source fault estimation
Here, we used the seismic data
from example 2, and comple-
ment them with geodetic data
to estimate the fault dimen-
sions and geometry for the
2009 L'Aquila earthquake
using the geometry mode
of BEAT. The geodetic data
were comprised of two InSAR
surface displacement maps,
derived from Envisat satellite
images acquired before and
Figure 3. Results of example 1. (a,b) Hudson plots from ensembles of 1000 moment tensor after the earthquake from both
solutions (small black focal mechanism plots) for estimations using variance and non-Toeplitz data ascending and descending ac-
covariance matrices, respectively. The large red and gray focal mechanism plots show the quisition geometries (Table S1).
maximum a-posterior (MAP) solutions and the true moment tensor (MT), respectively.
We applied the kite software
(c,d) Fuzzy focal mechanism plots based on ensembles of 1000 MT solutions drawn from the PPD
of the two estimations using the variance and non-Toeplitz data covariance matrices, respectively. (Isken et al., 2017) to spatially
downsample the interfero-
grams and to estimate the full
data error variance–covariance
P-phase arrival time to calculate the posterior probability at matrix (Sudhaus and Jónsson, 2009; Jolivet et al., 2012). The
each station. For the seismic data, we utilized the AK135 AK135 Earth velocity structure (Kennett et al., 1995) was
Earth velocity structure (Kennett et al., 1995) to compute a applied to compute a GF database for the geodetic data with
GF database with QSSP (Wang et al., 2017). To compare PSGRN/PSCMP (Wang et al., 2006). The frequency band
the results for two sampling algorithms on the PPD, we used for the seismic data is broad compared to the previous
PT and SMC sampling with the same setup. example with 0.001–0.1 Hz, because the rectangular source
For both sampling algorithms (PT and SMC), the mode and with the rupture propagating across the fault produces
MAP of the marginals are similar (Fig. 4). The estimated broad spectra.
uncertainties in the source parameters are larger for PT com- We applied the SMC algorithm, a rectangular source
pared to SMC. Fuzzy waveform misfits (Figs. S11 and S12) are parameterization, and a non-Toeplitz seismic covariance
comparable, and uncertainties obtained from the SMC algo- matrix to estimate the location, orientation, extent, and slip
rithm might be underestimated. These differences highlight of the L'Aquila source fault (Fig. S1). In addition, we esti-
that there is neither an ultimate measure nor a guarantee mated the rupture nucleation point assuming a constant rup-
for convergence of sampling algorithms, and the user must ture velocity of 3:5 km=s as well as a source start time and
carefully evaluate posterior marginals estimated by the sam- duration of a half-sinusoidal source time function. Finally,
plers (Mosegaard, 2012). a hierarchical scaling was estimated (equation 4) for the inter-
We configured the PT sampler with 20 chains, and we ran- ferograms and seismic data, as well as a bilinear ramp func-
domly applied swaps every 10–30 steps. After each swap, we tion to account for possible inaccuracies in the satellite orbit
stored the sample of each MC that was sampled from the target information.

8 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
constrained and located near
the western edge of the fault
(nucleation_x: −1 at left edge
and 1 at right edge), whereas
its down-dip location is more
uncertain, located somewhere
near the bottom edge of the
fault (nucleation_y: −1 at the
top edge and 1 at bottom edge).
The MAP model can explain
the static surface displacement
data well (Fig. S13), and the
simulated waveforms also
show a good agreement with
the observations, considering
that only uniform fault slip
was included in this model
estimation (Fig. S14).
In this example, we again
used four mobile CPU cores to
sample 1000 MCs with 100
steps each and through 22
stages of the SMC sampler.
This resulted in ∼2:2 million
forward evaluations and over
11 hr of computation time.

Example 4: static slip


distribution estimation
with Laplacian
smoothing
In this example, we consider a
finite-fault parameterization
for the 2009 L'Aquila earth-
quake and the ffi mode of
BEAT using the static InSAR
data. We employed the previ-
ously determined parameters
from the rectangular source
estimation (see the Example
3: Rectangular Source Fault
Estimation section) to fix the
Figure 4. Results of example 2. Marginal posterior distributions for model parameters of a double-
source fault location and geo-
couple source estimated from real teleseismic data of the 2009 L'Aquila earthquake obtained from metry as well as the orbital
two different sampling algorithms, SMC (yellow) and parallel tempering (PT, blue). The MAP values ramps for the InSAR scenes.
are marked by vertical colored lines. We extended the previously
obtained MAP fault dimen-
sions in length by a factor of
1.6 and in width by a factor
The results show that the strongest parameter trade-offs of 1.4. Then, we discretized the fault into 2 × 2 km patches
are between the extent of the fault (width and length), fault resulting in 121 patches (11 × 11). For each of these patches,
slip, and the depth of the upper edge of the fault (Fig. 5). we computed GFs slip parallel and perpendicular to fault
The horizontal location of the rupture nucleation point is well rake. While assuming smooth slip on neighboring patches

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 9

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
(Laplacian smoothing; equation 7), we applied the SMC algo- Figure 5. Results of example 3. 1D and 2D marginal posteriors of a
rithm to solve for the two slip parameters on each patch as well rectangular source for the 2009 L'Aquila earthquake. The red
as estimating the smoothing parameter and hierarchical scaling vertical lines in the histograms and the red dot in the 2D cor-
relation maps mark the MAP solution. In the correlation plots,
for each interferogram (equation 4). The number of parame- blue colors are regions of high probability. North and east shifts
ters estimated in this example is 245, that is, two slip param- are relative to a given reference location.
eters for each patch, one smoothing parameter and two
hierarchical scaling parameters. To initialize each MC at a rea-
sonable starting point of the high-dimensional search space, we
drew a random smoothing weight α from its prior (uniform) fault-slip model fits the observed InSAR data better than the
distribution. With these smoothing values, we inverted for the previous constant slip model (Fig. S15).
rake-parallel slip distribution via regularized nonnegative least We carried out this model parameter estimation on four
squares (Fukuda and Johnson, 2008). The resulting slip vectors mobile CPU cores as before and included sampling of 3000
in the slip-parallel direction and zero slip in the slip- MCs for 300 steps over 43 stages of the SMC sampler.
perpendicular direction were set as the starting points for During the sampling the forward model was evaluated 39 mil-
the MCs in the first stage of the SMC algorithm. lion times, which took about 3 hr.
The fault-slip distribution results (Fig. 6) show a maximum The assumption of smooth slip between neighboring slip
slip of ∼0:7 m and significant slip confined between 2 and patches can sometimes be inappropriate (Minson et al., 2013;
18 km. The slip uncertainties increase with depth as the ability Duputel et al., 2014; Ragon et al., 2018). Thus, BEAT also sup-
of static surface displacements to resolve fault slip decreases ports estimating distributed slip without a Laplacian smooth-
with depth. This solution does not consider uncertainties in ing prior, but instead the uncertainty in Earth structure is
location and geometry as found in example 3 (Fig. 5). These propagated (Duputel et al., 2014) (see the Static Slip Distri-
would enlarge the slip vector error ellipses (Sudhaus and bution without Laplacian Smoothing section in the supple-
Jónsson, 2009, 2011). The model prediction of this variable mental material, Figs. S16–S22).

10 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
Example 5: kinematic nonlinear slip estimation
with Laplacian smoothing
In this final example, we extended the L'Aquila source estima-
tion to include rupture kinematics and used both seismic wave-
forms and static InSAR data as observations. To increase the
data information about the rupture details, we filtered the
waveforms with an upper limit of 0.5 Hz and we extended time
windows around P-wave arrivals to −10 to 50 s. For the esti-
mation, we applied the ffi mode of BEAT and used the pos-
terior slip distribution result of example 4 as a slip prior to
reduce the parameter space. The kinematic rupture parameters
that we estimated include the hypocenter time and location as
well as the rupture velocity and duration on each fault patch.
In addition, we estimated smoothing parameter α and hierar-
chical residual scaling parameters. The total parameter count
in this example is 491, that is, four kinematic slip parameters
Figure 6. Results of example 4. Static slip distribution solution for for each of the 121 patches, three hypocenter parameters, one
the 2009 L'Aquila earthquake estimated using static InSAR
smoothing parameter, and three scaling parameters for the
surface displacements. The patch colors and the black arrows
show the MAP, whereas the black ellipses around the arrow tips two InSAR datasets and the seismic data. We applied the
show the two-sigma confidence bounds. SMC sampler in the estimation, assuming a non-Toeplitz
seismic covariance matrix, with MCs initialized to the pos-
terior from the previous example, as we considered it as prior
information.
The kinematic results show that the hypocenter location is
well constrained, and that the rupture nucleated near the
western edge of the fault at a depth between 10 and 12 km
(Fig. 7a). The rupture appears
(a) (b) to have spread with near con-
stant rupture velocity of
∼3 km=s across the fault.
Compared to the static solu-
tion, the maximum slip is
reduced to ∼0:6 m, and the slip
distribution is overall some-
what shallower (by ∼2 km),
that is, it is now mostly con-
(c)
fined to depths less than
16 km. Because of the addition
of seismic data, rake and slip
uncertainties on deep slip
patches are significantly re-
duced. Marginal distributions
of selected patches (Fig. S23)
show that the kinematic slip
parameters for patches far from
Figure 7. Results of example 5. (a) Kinematic slip distribution for the L'Aquila 2009 earthquake from
the joint inference on teleseismic and InSAR static data. The patch colors and the black arrows the hypocenter (e.g., patch 15)
show the MAP, whereas the black ellipses around the arrow tips indicate the two-sigma confidence are poorly resolved, with the
boundaries, and the black star marks the hypocentral location. The black continuous lines show the 95% confidence bounds on the
MAP of the evolving rupture front, with the timing of each front in seconds annotated on the rupture duration and rupture
respective isoline. The uncertainty of the rupture onset time is shown as the fuzzy isolines with light
velocity, between 0.5 and 2.6 s
gray indicating lower probability. (b) The ensemble of moment rate function solutions, with dark
and red colors indicating high-probability moment rates. The continuous black line shows the MAP and between 2.5 and 4:3 km=s,
moment rate function. (c) Marginal posterior distribution for the Laplacian smoothing weight α, respectively. This is not sur-
with the red vertical line marking the MAP solution. prising as only teleseismic data

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 11

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
(a)

(b)

were used in the estimation. Patches with high-slip values have Figure 8. Results of example 5. Geodetic InSAR data fits for the
shorter rupture durations (e.g., patch 51) with the 95% confi- 2009 L'Aquila earthquake from (a) descending and (b) ascending
dence bounds between 0.1 and 2 s. The rupture velocities of acquisition geometry. (Data panels) Geocoded unwrapped
interferograms in radar line of sight (LoS) with negative values
patches close to the hypocenter (e.g., patch 67) are slightly bet-
indicating increasing LoS distance due to subsidence. The dis-
ter constrained (with the 95% confidence bounds between 2.45 played displacement values are derived from quadtree subsam-
and 3:8 km=s) than patches further away. This kinematic sol- pling and extrapolated to each pixel that belongs to the same
ution includes an estimation of the smoothing weight α, and quadtree square. The look-vector and heading of the satellite are
the result is based on a range of α values rather than a single- shown by the two short and long arrows, respectively. (Model
panels) Synthetic surface displacements in LoS derived from the
fixed smoothing value (Fig. 7c). The moment rate function
MAP solution. The small gray rectangle shows the location and
shows a source rupture duration of ∼9–10 s with the peak orientation of the derived fault geometry from example 2,
moment release occurring between 4 and 6 s (Fig. 7b). The whereas the red rectangle shows the extended fault geometry
fit with the geodetic data is comparable to example 4 (Fig. 8), that is used in examples 4 and 5. The solid black lines mark the
but the fit with the seismic waveforms (Fig. 9) appears to be upper edge of the fault. (Residual panels) Residual surface dis-
poorer than in examples 2 and 3. The latter comparison, how- placements, that is, the difference between the model and the
data panels. Note the different color scale for the residual plots.
ever, is unfair as the seismic data in this example include
significantly more details due to the increased upper cutoff fre-
quency of the band-pass filter.
Several features from our results generally agree with pre- indicates a slight rotation of the slip vector with increasing
vious studies. The rupture exhibits predominantly a single-slip depth (Zhang et al., 2012).
region at depths between 5 and 16 km. The hypocenter is In this example, we employed 25 CPU cores on a worksta-
located in the northwest and offset from the main slip region tion to sample 10,000 MCs for 300 steps over 35 stages of the
(Cirella et al., 2009; Zhang et al., 2012). The moment rate func- SMC sampler. This resulted in ∼100 million forward model
tion (Fig. 7) shows a small peak near 2 s. Finally, the rake angle evaluations and a runtime of ∼85 hr.

12 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
Conclusions Figure 9. Results of example 5. Waveform fits for the kinematic
The BEAT is a new software for source estimations and uncer- finite-fault solution for P-wave arrivals for 16 of the 35 stations
tainty quantification in deformation source studies. It is open used. The filtered (0.001–0.5 Hz) displacement waveform data
(dark gray solid line) and the filtered synthetic displacement
source and available under the license GPL3.0 (see Data and
waveforms (red solid line) are shown together, with the brown
Resources; Vasyura-Bathke et al., 2019). BEAT provides an shading indicating 100 random draws of the filtered synthetic
extensive open-source framework to study earthquake sources displacements from the PPD. The residual waveforms are shown
with various parameterizations, including moment tensor point below each waveform as filled red-line polygons. Each trace box
sources, rectangular static and kinematic sources, and static is annotated with the station name and component as well as the
distance and azimuth from the MAP solution of the center of the
and kinematic finite-fault sources. An important aspect of
reference fault. The arrival time and the duration of each window
BEAT’s contribution is the integration of multiple methods are shown in the lower left and right, respectively.
in a unified platform. In addition, we provide the novel ability
to consider joint inversion of geodetic and seismic data for
sources in a stratified, elastic half-space with residual covari-
ance estimation, while allowing fully nonlinear treatment of all Data and Resources
The Bayesian earthquake analysis tool (BEAT) package runs under
sources in a Bayesian framework. The intention behind provid-
Linux and MacOS on Python versions ≥ 3.5 and is available at https://
ing such a unified framework to the geophysics community
github.com/hvasbath/beat (Vasyura-Bathke et al., 2019). It is distrib-
is to make research more reproducible and to accelerate the uted under the GNU General Public Licencse, version 3.0 and depends
development of comprehensive tools for deformation source on the following main Python libraries (and dependencies within):
studies. The five examples presented here demonstrate the Pyrocko (Heimann et al., 2017), pymc3 (Salvatier et al., 2015),
main functionalities of the BEAT software and can be repro- Theano (Theano Development Team, 2016), and MPI4py (Dalcin
duced through step-by-step tutorials available on the project et al., 2011). The presented examples can be reproduced following
website (see Data and Resources). the tutorials at https://hvasbath.github.io/beat/examples/index.html.

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 13

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
Seismic waveforms were originally downloaded from Incorporated Biggs, J., S. K. Ebmeier, W. P. Aspinall, Z. Lu, M. E. Pritchard, R. S.
Research Institutions for Seismology (IRIS) at https://ds.iris.edu. Sparks, and T. A. Mather (2014). Global link between deformation
Envisat satellite radar data were provided by the European Space and volcanic eruption quantified by satellite imagery, Nat. Commun.
Agency (ESA) at https://earth.esa.int/web/guest/home. All of the figures 5, no. 3471, doi: 10.1038/ncomms4471.
have been produced using Matplotlib (Hunter, 2007) and can be gen- Chadwick, W. W., S. Jónsson, D. J. Geist, M. P. Poland, D. J. Johnson,
erated through the BEAT plot subcommand. Some maps have been B. Spencer, K. S. Harpp, and A. Ruiz (2011). The May 2005 erup-
produced using Cartopy (Met Office, 2015). The supplemental material tion of Fernandina volcano, Galápagos: The first circumferential
for this article includes a PDF file further illustrating the examples. The dike intrusion observed by GPS and InSAR, Bull. Volcanol. 73,
unpublished manuscript by R. Dutta, S. Jónsson, and H. Vasyura- no. 6, 679–697.
Bathke (2019), “Simultaneous Bayesian estimation of non-planar fault Ching, J., and Y.-C. Chen (2007). Transitional Markov Chain Monte
geometry and spatiallyvariable slip,” submitted to J. Geophys. Res. Carlo method for Bayesian model updating, model class selection,
and model averaging, J. Eng. Mech. 133, no. 7, 816–832.
Cirella, A., A. Piatanesi, M. Cocco, E. Tinti, L. Scognamiglio,
Acknowledgments
A. Michelini, A. Lomax, and E. Boschi (2009). Rupture history of
The authors thank Jiří Vackář, Romain Jolivet, one anynomous
the 2009 L'Aquila (Italy) earthquake from non-linear joint inver-
reviewer, and the editors for their comments that helped to sion of strong motion and GPS data, Geophys. Res. Lett. 36, no. 19,
improve the quality of this article. This research was supported 1–6.
by King Abdullah University of Science and Technology Dalcin, L. D., R. R. Paz, P. A. Kler, and A. Cosimo (2011). Parallel
(KAUST), under Award Numbers BAS/1/1353-01-01 and distributed computing using Python, Adv. Water Resour. 34,
BAS/1/1339-01-1. H. V.-B. was partially supported by Geo.X, no. 9, 1124–1139.
the Research Network for Geosciences in Berlin and Potsdam Das, S. (2015). Supershear Earthquake Ruptures—Theory, Methods,
under the Project Number SO_087_GeoX. Henriette Sudhaus, Laboratory Experiments and Fault Superhighways: An Update,
Andreas Steinberg, and Marius Paul Isken acknowledge found- Chapter 1, Springer International Publishing, Cham, Switzerland,
ing by the German Research Foundation (DFG) through an 1–20.
Emmy-Noether Young Researcher Grant Number 276464525. Dettmer, J., and S. E. Dosso (2012). Trans-dimensional matched-field
geoacoustic inversion with hierarchical error models and interact-
Hannes Vasyura-Bathke owes the most gratitude to his beloved
ing Markov chains, J. Acoust. Soc. Am. 132, no. 4, 2239–2250.
wife Olha for her tireless support and tolerance during many
Dettmer, J., R. Benavente, P. R. Cummins, and M. Sambridge (2014).
evenings and nights spent writing the code and this article. Trans-dimensional finite-fault inversion, Geophys. J. Int. 199,
no. 2, 735–751.
References Dettmer, J., S. E. Dosso, and C. W. Holland (2007). Uncertainty esti-
Aki, K., and P. G. Richards (2002). Quantitative Seismology, Second mation in seismo-acoustic reflection travel time inversion, J.
Ed., University Science Books, Sausalito, California. Acoust. Soc. Am. 122, no. 1, 161–176.
Altman, A., and J. Gondzio (1999). Regularized symmetric indefinite Duputel, Z., P. S. Agram, M. Simons, S. E. Minson, and J. L. Beck
systems in interior point methods for linear and quadratic optimi- (2014). Accounting for prediction uncertainty when inferring sub-
zation, Optim. Method Softw. 11, nos. 1/4, 275–302. surface fault slip, Geophys. J. Int. 197, no. 1, 464–482.
Bagnardi, M., and A. Hooper (2018). Inversion of surface defor- Duputel, Z., L. Rivera, Y. Fukahata, and H. Kanamori (2012).
mation data for rapid estimates of source parameters and uncer- Uncertainty estimations for seismic source inversions, Geophys.
tainties: A Bayesian approach, Geochem. Geophys., Geosys. 19, J. Int. 190, no. 2, 1243–1256.
1–18. Dutta, R., S. Jónsson, T. Wang, and H. Vasyura-Bathke (2018).
Bassin, C., G. Laske, and G. Masters (2000). The current limits of Bayesian estimation of source parameters and associated Coulomb
resolution for surface wave tomography in North America, EOS failure stress changes for the 2005 Fukuoka (Japan) earthquake,
Trans. AGU 81, no. F897. Geophys. J. Int. 213, 261–277.
Bathke, H., M. Nikkhoo, E. Holohan, and T. Walter (2015). Insights Fukuda, J., and K. M. Johnson (2008). A fully Bayesian inversion for
into the 3D architecture of an active caldera ring-fault at Tendürek spatial distribution of fault slip with objective smoothing, Bull.
volcano through modeling of geodetic data, Earth Planet. Sci. Lett. Seismol. Soc. Am. 98, no. 3, 1128–1146.
422, 157–168. Geyer, C. J. (1991). Markov Chain Monte Carlo maximum likelihood,
Bathke, H., M. Shirzaei, and T. R. Walter (2011). Inflation and defla- Computing Science and Statistics, Proc. of the 23rd Symposium on
tion at the steep-sided Llaima stratovolcano (Chile) detected by the Interface, 156–163 pp.
using InSAR, Geophys. Res. Lett. 38, no. 10, 1–5. Hallo, M., and F. Gallovic (2016). Fast and cheap approximation of
Bathke, H., H. Sudhaus, E. Holohan, T. R. Walter, and M. Shirzaei Green function uncertainty for waveform-based earthquake source
(2013). An active ring fault detected at Tendürek volcano by using inversions, Geophys. J. Int. 207, 1012–1029.
InSAR, J. Geophys. Res. 118, no. 8, 4488–4502. Hartzell, S. H., and T. H. Heaton (1983). Inversion of strong ground
Bekaert, D. P., R. J. Walters, T. J. Wright, A. J. Hooper, and D. J. motion and teleseismic waveform data for the fault rupture history
Parker (2015). Statistical comparison of InSAR tropospheric of the 1979 Imperial Valley, California, earthquake, Bull. Seismol.
correction techniques, Remote Sens. Environ. 170, 40–47. Soc. Am. 73, no. 6, 1553–1583.

14 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
Haskell, N. A. (1964). Total energy and energy spectral density of elas- (2016). The earthquake-source inversion validation (SIV) project,
tic wave radiation from propagating faults, Bull. Seismol. Soc. Am. Seismol. Res. Lett. 87, no. 3, 690–708.
54, no. 6, 1811–1841. Mai, P. M., P. Shearer, J.-P. Ampuero, and T. Lay (2016). Standards
Hastings, W. K. (1970). Monte Carlo sampling methods using Markov for documenting finite-fault earthquake rupture models, Seismol.
chains and their applications, Biometrika 57, no. 1, 97–109. Res. Lett. 87, no. 3, 712–718.
Heimann, S. (2011). A robust method to estimate kinematic earth- Met Office (2015). Cartopy: A Cartographic Python Library with a
quake source parameters, Ph.D. Thesis, University of Hamburg, Matplotlib Interface, Devon, Exeter, United Kingdom.
Germany. Metropolis, N., A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and
Heimann, S., M. Kriegerowski, M. Isken, S. Cesca, S. Daout, F. Grigoli, E. Teller (1953). Equation of state calculations by fast computing
C. Juretzek, T. Megies, N. Nooshiri, A. Steinberg, et al. (2017). machines, J. Chem. Phys. 1087, no. 21, 1087–1092.
Pyrocko—An open-source seismology toolbox and library, GFZ Minson, S. E., M. Simons, and J. L. Beck (2013). Bayesian inversion for
Data Services, Version 0.3, doi: 10.5880/GFZ.2.1.2017.001. finite fault earthquake source models I-Theory and algorithm,
Heimann, S., H. Vasyura-Bathke, H. Sudhaus, M. P. Isken, M. Geophys. J. Int. 194, no. 3, 1701–1726.
Kriegerowski, A. Steinberg, and T. Dahm (2019). A Python frame- Monelli, D., and P. M. Mai (2008). Bayesian inference of kinematic
work for efficient use of pre-computed Green’s functions in seis- earthquake rupture parameters through fitting of strong motion
mological and other physical forward and inverse source problems, data, Geophys. J. Int. 173, no. 1, 220–232.
Solid Earth 10, no. 6, 1921–1935. Moral, P. D., A. Doucet, and A. Jasra (2006). Sequential Monte Carlo
Hunter, J. D. (2007). Matplotlib: A 2D graphics environment, samplers, J. Roy. Stat. Soc. 68, no. 3, 411–436.
Comput. Sci. Eng. 9, no. 3, 90–95. Mosegaard, K. (2012). Limits to nonlinear inversion, in Applied
Isken, M., H. Sudhaus, S. Heimann, A. Steinberg, S. Daout, and Parallel and Scientific Computing, K. Jónasson (Editor), Springer
H. Vasyura-Bathke (2017). Kite—Software for rapid earthquake Berlin Heidelberg, Berlin, Heidelberg, 11–21.
source optimisation from InSAR surface displacement, GFZ Data Neal, R. M. (2001). Annealed importance sampling, Stat. Comput. 11,
Services, Version 0.1, doi: 10.5880/GFZ.2.1.2017.002. no. 2, 125–139.
Jarzynski, C. (1997). Equilibrium free energy differences from none- Okada, Y. (1985). Surface deformation due to shear and tensile faults
quilibrium measurements: A master equation approach, Phys. Rev. in a half-space, Bull. Seismol. Soc. Am. 75, no. 4, 1135–1154.
Lett. 78, no. 14, 2690–2693. Olsen, A. H., and R. J. Apsel (1982). Finite faults and inverse theory
Jasra, A., D. A. Stephens, and C. C. Holmes (2007). On population- with applications to the 1979 Imperial Valley earthquake, Bull.
based simulation for static inference, Stat. Comput. 17, no. 3, Seismol. Soc. Am. 72, no. 6, 1969–2001.
263–279. Pugh, D. J., and R. S. White (2018). MTfit: A Bayesian approach to
Ji, C., D. J. Wald, and D. V. Helmberger (2002). Source description of seismic moment tensor inversion, Seismol. Res. Lett. 89, no. 4,
the 1999 Hector Mine, California, earthquake, part I: Wavelet 1507–1513.
domain inversion theory and resolution analysis, Bull. Seismol. Ragon, T., A. Sladen, and M. Simons (2018). Accounting for uncertain
Soc. Am. 92, no. 4, 1192–1207. fault geometry in earthquake source inversions - I: Theory and
Jolivet, R., C. Lasserre, M.-P. Doin, S. Guillaso, G. Peltzer, R. Dailu, J. simplified application, Geophys. J. Int. 214, no. 2, 1174–1190.
Sun, Z.-K. Shen, and X. Xu (2012). Shallow creep on the Razafindrakoto, H. N. T., and M. P. Mai (2014). Uncertainty in
Haiyuan Fault (Gansu, China) revealed by SAR Inter- earthquake source imaging due to variations in source time
ferometry, J. Geophys. Res. 117, no. B6, B06401, doi: 10.1029/ function and earth structure, Bull. Seismol. Soc. Am. 104, no. 2,
2011JB008732. 855–874.
Jónsson, S., H. Zebker, P. Segall, and F. Amelung (2002). Fault slip Roberts, G. O., and J. S. Rosenthal (2001). Optimal scaling for various
distribution of the 1999 Mw 7.1 Hector Mine, California, earth- Metropolis-Hastings algorithms, Stat. Sci. 16, no. 4, 351–367.
quake, estimated from satellite radar and GPS measurments, Salvatier, J., T. Wiecki, and C. Fonnesbeck (2015). Probabilistic pro-
Bull. Seismol. Soc. Am. 92, no. 4, 1377–1389. gramming in Python using PyMC, PEERJ. Comput. Sci. 2, e55, doi:
Jost, M. L., and R. B. Herrmann (1989). A student’s guide to and 10.7717/peerj-cs.55.
review of moment tensors, Seismol. Res. Lett. 60, no. 2, 37–57. Sambridge, M., and K. Mosegaard (2002). Monte Carlo methods in
Kennett, B. L. N., E. R. Engdahl, and R. Buland (1995). Constraints on geophysical inverse problems, Rev. Geophys. 40, 3-1–3-29, doi:
seismic velocities in the Earth from traveltimes, Geophys. J. Int. 10.1029/2000RG000089.
122, 108–124. Sokos, E., J. Zahradník, A. Kiratzi, J. Janský, F. Gallovič, O. Novotny,
Kikuchi, M., and H. Kanamori (1982). Inversion of complex body J. Kostelecký, A. Serpetsidaki, and G. A. Tselentis (2012). The
waves, Bull. Seismol. Soc. Am. 72, no. 2, 491–506. January 2010 Efpalio earthquake sequence in the western Corinth
Kumagai, H., Y. Maeda, M. Ichihara, N. Kame, and T. Kusakabe Gulf (Greece), Tectonophysics 530/531, 299–309.
(2014). Seismic moment and volume change of a spherical source, Stähler, S. C., and K. Sigloch (2014). Fully probabilistic seismic
Earth Planets Space 66, no. 7, 1–10. source inversion-Part 1: Efficient parameterisation, Solid Earth
Mai, P. M., and K. K. S. Thingbaijam (2014). SRCMOD: An online 5, 1055–1069.
database of finite-fault rupture models, Seismol. Res. Lett. 85, no. 6, Sudhaus, H., and S. Jónsson (2009). Improved source modelling
1348–1357. through combined use of InSAR and GPS under consideration of
Mai, P. M., D. Schorlemmer, M. Page, K. Asano, M. Causse, S. correlated data errors: Application to the June 2000 Kleifarvatn
Custodio, G. Festa, M. Galis, F. Gallovic, W. Imperatori, et al. earthquake, Iceland, Geophys. J. Int. 176, no. 2, 389–404.

Volume XX • Number XX • – 2020 • www.srl-online.org Seismological Research Letters 15

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user
Sudhaus, H., and S. Jónsson (2011). Source model for the 1997 Zirkuh Wang, R., S. Heimann, Y. Zhang, H. Wang, and T. Dahm (2017).
earthquake (MW = 7.2) in Iran derived from JERS and ERS InSAR Complete synthetic seismograms based on a spherical self-gravitating
observations, Geophys. J. Int. 185, 676–692. Earth model with an atmosphere-ocean–mantle-core structure,
Tape, W., and C. Tape (2012). A geometric setting for moment Geophys. J. Int. 210, 1739–1764.
tensors, Geophys. J. Int. 190, no. 1, 476–498. Wang, R., F. Lorenzo-Martín, and F. Roth (2006). PSGRN/PSCMP-A
Tape, W., and C. Tape (2015). A uniform parametrization of moment new code for calculating co- and post-seismic deformation, geoid
tensors, Geophys. J. Int. 202, no. 3, 2074–2081. and gravity changes based on the viscoelastic-gravitational dislo-
Tarantola, A., and B. Valette (1982). Inverse problems = Quest for cation theory, Comput. Geosci. 32, no. 4, 527–541.
information, J. Geophys. 50, 159–170. Xu, W., R. Dutta, and S. Jonsson (2015). Identifying active faults by
Theano Development Team (2016). Theano: A Python framework improving earthquake locations with InSAR data and Bayesian
for fast computation of mathematical expressions, arXiv e-prints, estimation: The 2004 Tabuk (Saudi Arabia) earthquake sequence,
available at https://arxiv.org/abs/1605.02688v1 (last accessed Bull. Seismol. Soc. Am. 105, no. 2A, 765–775.
December 2019). Yagi, Y., and Y. Fukahata (2011). Introduction of uncertainty of
Vackář, J., J. Burjánek, F. Gallovič, J. Zahradńik, and J. Clinton (2017). Green’s function into waveform inversion for seismic source proc-
Bayesian ISOLA: New tool for automated centroid moment tensor esses, Geophys. J. Int. 186, no. 2, 711–720.
inversion, Geophys. J. Int. 210, no. 2, 693–705. Zhang, Y., W. Feng, Y. Chen, L. Xu, Z. Li, and D. Forrest (2012). The
Vasyura-Bathke, H., J. Dettmer, A. Steinberg, S. Heimann, M. Isken, 2009 L'Aquila M W 6.3 earthquake: A new technique to locate the
O. Zielke, P. M. Mai, H. Sudhaus, and S. Jónsson (2019). BEAT— hypocentre in the joint inversion of earthquake rupture process,
Bayesian earthquake analysis tool, GFZ Data Services, Version1.0, Geophys. J. Int. 191, no. 3, 1417–1426.
doi: 10.5880/fidgeo.2019.024.
Wang, R. (1999). A simple orthonormalization method for stable and
efficient computation of Green’s functions, Bull. Seismol. Soc. Am. Manuscript received 25 March 2019
89, no. 3, 733–741. Published online 22 January 2020

16 Seismological Research Letters www.srl-online.org • Volume XX • Number XX • – 2020

Downloaded from https://pubs.geoscienceworld.org/ssa/srl/article-pdf/doi/10.1785/0220190075/4928889/srl-2019075.1.pdf


by University of Cambridge user

You might also like