This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\pagerange

Smooth surface reconstruction of earthquake faults from distributed potency beachballs B

Smooth surface reconstruction of earthquake faults from distributed potency beachballs

Dye SK Sato1    Yuji Yagi2    Ryo Okuwaki2    & Yukitoshi Fukahata3
1 Research Institute for Marine Geodynamics
   Japan Agency for Marine-Earth Science and Technology   
    Yokohama
   Kanagawa 236-0001    Japan
2 Institute of Life and Environmental Sciences
   University of Tsukuba    Tsukuba    Ibaraki 305-8572    Japan
3 Disaster Prevention Research Institute
   Kyoto University    Uji    Kyoto 611-0011    Japan
(Received 20xx Xxxx xx; in original form 20xx Xxxx xx; xxxx)
keywords:
Earthquake source observations; Fractures, faults, and high strain deformation zones; Inverse theory; Theoretical seismology
volume: xxx
{summary}

Earthquake faults approximate smooth surfaces of displacement discontinuity in a coarse-grained linear continuum, and the beachball nodal planes of source inelastic strain (potency) indicate the normal vector (nn-vector) of the crack face. This theoretical relation has been used to interpret estimated potency as geometrical information of earthquake faults. We formulate an inverse analysis for reconstructing a smooth, three-dimensional fault surface from a given potency field. The analysis is grounded on the definition of the nn-vector that is orthogonality to a surface. It is shown from this definition that, whereas the estimated nn-vector field is always explained by a unique curve in two dimensions, the surface reconstruction from the estimated nn-vector field is an overdetermined problem with no solution in three dimensions. We resolve this overdetermination by adopting the original definition of a fault surface as a map that approximates a thin sheet of potency onto a displacement discontinuity across a literal face. An analytical expression for this fault reconstruction is derived and confirmed to reproduce the original three-dimensional smooth surface from a synthesized noisy nn-vector field. By incorporating this result into surface reconstruction based on potency density tensor inversion, we estimate the three-dimensional source fault geometry of the 2013 Balochistan earthquake. The estimated geometry is more consistent with the fault trace than the quasi-two-dimensional shape reconstruction as anticipated by our numerical experiments.

1 Introduction

Fault geometry predestines potential rupture modes and arrests of earthquakes (Das & Aki, 1977; Aki, 1979; Kase & Day, 2006; Bruhat et al., 2016; Ozawa et al., 2023). Therefore, fault geometry inference from actual data serves as a grasp of the source complexity that happened (Barka & Kadinsky-Cade, 1988; Kodaira et al., 2006; Ando & Kaneko, 2018; Howarth et al., 2021).

Great efforts accumulate to extract information on source fault geometry from observed data (e.g., Matsu’ura & Hasegawa, 1987; Volkov et al., 2017; Dutta et al., 2021). The simplest, highly successful example is a configuration estimation using a point or rectangular fault, which infers the center of masses, angles, and lengths of a fault (Matsu’ura & Hasegawa, 1987). Detailed slip patterns could be estimated from a distributed slip inversion on a prescribed fault surface (Yabuki & Matsu’ura, 1992).

Still, the inference of fault non-planarity is challenging. Fault subdivisions with planar subfaults have been heavily utilized in practice (e.g., Matsu’ura, 1977a, b; Loveless & Meade, 2010; Avouac et al., 2014; Jolivet et al., 2014). It, however, entails a mechanical difficulty of artificial stress infinity at subfault edges (Romanet et al., 2020, 2024; Mallick & Meade, 2025), and an issue of fault geometry inference is estimating macroscopic geometry, which is smooth, consistent with linear continuum mechanics (Matsu’ura et al., 2019). Distributed source inversions involving fault geometry inferences are proved to have solutions (Beretta et al., 2008; Volkov et al., 2017; Aspri et al., 2020; Diao et al., 2023), but how to solve them is under debate (Ragon et al., 2018; Dutta et al., 2021; Shimizu et al., 2021; Wei et al., 2023; Aspri et al., 2025).

This source geometry inference is entering a new phase owing to the potency density tensor inversion (PDTI, Kikuchi & Kanamori, 1991; Kasahara & Yagi, 2010; Shimizu et al., 2020), which reveals macroscopic fault geometry as spatially smooth rotations of estimated potency tensors. A seminal research of Kikuchi & Kanamori (1991) first infers a seismic source as inelastic strain (potency) distributed over a quasi-surface instead of as displacement discontinuity across a literal face. Nonetheless, their original PDTI is indistinct from planar fault subdivision and similar to multiple point source approaches (Tsai et al., 2005; Yue & Lay, 2020) because the prior constraint they employ is only the assumption of double-coupling potency imposed as a traceless and determinant-free constraint. A fine benefit of the PDTI comes from a prior constraint upon the fault geometry, which is, to our knowledge, first introduced by Kasahara & Yagi (2010). Smoothly varying macroscopic rotations of focal mechanisms are thus extracted by simply using a smoothness constraint on the model parameter in the PDTI. Kasahara & Yagi (2010) have further developed a stable scheme by accounting for the propagation of Green’s function errors (Yagi & Fukahata, 2010, 2011). Whereas the original intention of Kasahara & Yagi (2010) was rather to extract spatial variations in slip vectors and focal mechanisms themselves, Shimizu et al. (2020) point out that spatial patterns of the potency beachball diagram also document fault geometry as fault normal vector fields. The PDTI untangles source complexities in the perspective of fault geometry thereafter (Okuwaki et al., 2020; Tadapansawut et al., 2021; Fan et al., 2022; Ohara et al., 2023; Yagi et al., 2024; Inoue et al., 2025).

This natural extension of focal mechanism determination thus nearly establishes a methodology in which the macroscopic source geometry is immediately visualized as distributed potency beachballs. Shimizu et al. (2021) were able to estimate the fault geometry by further computing the normal vector (nn-vector) estimate, which is given by one nodal plane of estimated potency, and spatially integrating estimated fault slopes given by the 90-degree rotations of nn-vectors. However, the surface reconstruction procedure in Shimizu et al. (2021) cannot represent more than quasi-two-dimensional properties such as uniaxial bends and uniaxial twists. In fact, such a limitation stems from the seemingly harmless approach of constructing an optimum fault surface as an integral of optimum fault slopes. We show below that a fault slope field expected from the estimated nn-vector field is generally not integrable in three dimensions no matter how precise the observations are, and so that the key to inferring an earthquake fault as a macroscopic smooth surface is the stochastic inference of fault shape.

2 Problem setting of earthquake fault reconstruction from distributed potency

2.1 Forward model

Source geometry inference consists of the following three model equations.

The first model equation is the representation theorem for linear continua. Displacements uiu_{i} of data components ii are expressed by linear combinations of potency areal densities D(𝝃)D(\boldsymbol{\xi}) at coordinates 𝝃\boldsymbol{\xi} on faults Γ\Gamma (Kikuchi & Kanamori, 1991):

ui=Γ𝑑Σ(𝝃)Gi(𝝃)D(𝝃),u_{i}=\int_{\Gamma}d\Sigma(\boldsymbol{\xi})G_{i}(\boldsymbol{\xi})D(\boldsymbol{\xi}), (1)

where Gi(𝝃)G_{i}(\boldsymbol{\xi}) denotes Green’s function relating uiu_{i} and D(𝝃)D(\boldsymbol{\xi}). The same formulation of geometry inference applies to time-dependent problems, and we drop time dependence until our teleseismic applications.

The second model equation is the transformation of double couple sources to fault slips. When the potency density DD is double-coupled, it is expressed by the symmetrized tensor product of a slip vector 𝐬{\bf s} and a fault normal 𝐧{\bf n}:

D=12(𝐧𝐬T+𝐬𝐧T),D=\frac{1}{2}({\bf n}{\bf s}^{\rm T}+{\bf s}{\bf n}^{\rm T}), (2)

where the superscript T{\rm T} denotes the transpose. Equation (2) tells that an nn-vector is either of the nodal planes of potency in the beachball diagram standard in seismic moment representations. It is equivalent to one of the normalized sums of the maximum and minimum eigenvectors of DD, accounting for the sign reversals of these eigenvectors. In this paper, we suppose that the normal nn^{\circ} for a point source solution (e.g., the first-motion solution) is known and assume that the nn-vector for DD is its closest nodal plane to nn^{\circ}. That is, the tilt of the nn-vector from the reference is assumed to be within 45 degrees. Increasing the number of reference point sources may satisfy this assumption. Note that seismic moment is ill-defined in a strict sense across a stiffness discontinuity (e.g., a plate boundary), but potency is well-defined even across it.

The source inversion via eq. (1) and the nodal-plane extraction by eq. (2) have established methods of analysis. The potency density tensor inversion of eq. (1) is similar to the slip inversion (Yabuki & Matsu’ura, 1992; Yagi & Fukahata, 2011). The nodal-plane extraction of eq. (2) is the same as that for moment beachballs.

What remains is the correspondence between nn-vectors and shapes of surfaces, and it is given by the definition of the nn-vector: the normal vector 𝐧(𝝃){\bf n}(\boldsymbol{\xi}) at a coordinate 𝝃\boldsymbol{\xi} on a surface Γ\Gamma is orthogonal to the infinitesimal shift d𝝃d\boldsymbol{\xi} along Γ\Gamma at 𝝃\boldsymbol{\xi} as

𝐧(𝝃)d𝝃(𝝃)=0.{\bf n}(\boldsymbol{\xi})\cdot d\boldsymbol{\xi}(\boldsymbol{\xi})=0. (3)

In this paper, a ‘smooth surface’ denotes a surface whose shape is defined by eq. (3), known as a ‘regular surface’ in differential geometry. The orthogonality to the tangential plane defines the normal vector of the curved surface, and a relative location between two arbitrary points on a smooth surface is given by the line integral 𝑑𝝃\int d\boldsymbol{\xi} that satisfies eq. (3). Since the nn-vector field is equivalent to the shape of the smooth surface excepting its rigid-body translation, it entirely records the macroscopic smooth shape of a fault. Hereafter, one location 𝝃\boldsymbol{\xi}^{\circ} on (or at least adjacent to) the fault (e.g., the rupture initiation point) is assumed to be known, and its estimation errors are neglected.

2.2 Detachment of surface reconstruction from potency estimation

Let us suppose that the potency density DD and the associated nn-vector field are evaluated given the well-known equations (1) and (2). Our focus is on how to translate them into Γ\Gamma through eq. (3). However, the domain of definition for the observed potency density DD depends on the fault shape Γ\Gamma (eq. 1). To perform the inference of DD prior to reconstructing Γ\Gamma, the representation theorem (eq. 1) must be projected onto some fixed surface, termed the reference surface Γ\Gamma^{\circ} (Fig. 1).

Refer to caption
Figure 1: Parametrization of a fault surface Γ\Gamma using a reference surface Γ\Gamma^{\circ}. The normal vector 𝐧{\bf n} of Γ\Gamma depends on the location 𝝃\boldsymbol{\xi} on Γ\Gamma parametrized by (x,y,z)(x,y,z). The xx-yy plane describes the coordinate of 𝝃\boldsymbol{\xi} along Γ\Gamma^{\circ}, and the zz axis represents the vertical elevation of Γ\Gamma from Γ\Gamma^{\circ}. In the schematic, Γ\Gamma^{\circ} is drawn as a flat face perpendicular to a reference vector 𝐧{\bf n}^{\circ} through a point 𝝃\boldsymbol{\xi}^{\circ} on Γ\Gamma.

We now introduce Γ\Gamma^{\circ} as a flat plane perpendicular to 𝐧{\bf n}^{\circ} passing through 𝝃\boldsymbol{\xi}^{\circ} (Shimizu et al., 2020, also see Fig. 1); if necessary, multiple flat planes are incorporated. Coordinates 𝝃\boldsymbol{\xi} on Γ\Gamma are parametrized as a function of the location (x,y)(x,y) taken on Γ\Gamma^{\circ} and the elevation zz of Γ\Gamma from Γ\Gamma^{\circ}. The positive direction of zz is directed to 𝐧{\bf n}^{\circ}. Then, basis function expansions are applied to the potency density by using basis functions that depend only on (x,y)(x,y). The qq-component DqD_{q} of the potency density tensor is expressed by bases Xj(x,y)X_{j}(x,y) (j=1,,Mj=1,...,M), and the coefficients are stored in a vector 𝐚{\bf a} as aj+(q1)Ma_{j+(q-1)M}:

Dq(𝝃,τ)=j=1Maj+(q1)MXj(x,y).D_{q}(\boldsymbol{\xi},\tau)=\sum_{j=1}^{M}a_{j+(q-1)M}X_{j}(x,y). (4)

Substituting it into eq. (1), we have

ui=Hij(Γ)aj,u_{i}=H_{ij}(\Gamma)a_{j}, (5)

where Hij(Γ)=Γ𝑑ΣGi(𝝃)Xj(x,y)H_{ij}(\Gamma)=\int_{\Gamma}d\Sigma G_{i}(\boldsymbol{\xi})X_{j}(x,y). Note the domain of the surface integral of Hij(Γ)H_{ij}(\Gamma) is still the unknown Γ\Gamma. In the PDTI, Hij(Γ)H_{ij}(\Gamma) is regularly approximated by a known surface Γ\Gamma^{\prime} (e.g., Γ\Gamma^{\circ}) or an estimate Γ^\hat{\Gamma} [i.e., H(Γ)H(Γ^)H(\Gamma)\simeq H(\hat{\Gamma}), see §6.1 for details]. In the above manner, the object to simultaneously estimate (fault potency and fault geometry) is decomposed into model parameters 𝐚{\bf a} and Γ\Gamma. They are formally independent in the above formulation, and their dependencies are described by an additional constraint, which is the theme of this study.

3 How to map a potency density onto a fault shape

The translation of 𝐚{\bf a} to express a potency distribution into a fault surface Γ\Gamma requires a function ff such that Γ=f(𝐚)\Gamma=f({\bf a}). When a potency density provides an nn-vector field, the function to be sought is one that transforms the nn-vector field into a surface shape. We formulate it in this section. Such a function corresponds to the inverse of eq. (3) mapping a surface Γ\Gamma onto an nn-vector field.

When eq. (3) holds, the relative displacement on Γ\Gamma, namely the fault shape, is equivalent to the field of 𝐧{\bf n}. At the same time, the tensor shape of the potency DD dictates 𝐧{\bf n} (eq. 2). Namely, 𝐧{\bf n} is doubly characterized by Γ\Gamma and DD (Fig. 2). Those two are naively equal: articulating the nn-vector defined by Γ\Gamma as 𝐧(𝝃;Γ){\bf n}(\boldsymbol{\xi};\Gamma) and the nn-vector defined by DD by 𝐧(𝝃;𝐚){\bf n}_{*}(\boldsymbol{\xi};{\bf a}),

n(𝝃;Γ)=n(𝝃;𝐚).n(\boldsymbol{\xi};\Gamma)=n_{*}(\boldsymbol{\xi};{\bf a}). (6)

Equation (6) is an absolute requirement in the conventional slip inversion, where Γ\Gamma is defined first and n(𝝃;Γ)n(\boldsymbol{\xi};\Gamma) of Γ\Gamma constrains n(𝝃;𝐚)n_{*}(\boldsymbol{\xi};{\bf a}), the tensorial shape of the potency (e.g., Fukahata & Wright, 2008; Dutta et al., 2021). In the framework of the PDTI, which permits general inelastic strains beyond the displacement discontinuity representation, the inelastic strain is expressed as the potency DD first, and n(𝝃;𝐚)n_{*}(\boldsymbol{\xi};{\bf a}) of DD provides the information of n(𝝃;Γ)n(\boldsymbol{\xi};\Gamma) via eq. (6).

Refer to caption
Figure 2: Three normal vectors determined by different fault properties, indicated by arrows of different colors. (Red) The normal vector nn field of a fault shape Γ\Gamma. (Black) The reference normal nn^{\circ} for the reference surface Γ\Gamma^{\circ}. (Blue) The normal vector nn field corresponding to the potency DD (precisely, that after a basis function expansion along the reference surface, aa).

Based on this theoretical relationship (eq. 6), Shimizu et al. (2021) have reconstructed Γ\Gamma from the potency 𝐚{\bf a} parametrized along a reference plane. The following equation is useful to transform the nn-vector field to the surface; eq. (3) describes the infinitesimal in zz attending the infinitesimal in α=x,y\alpha=x,y,

z,α=nα/nz,z_{,\alpha}=-n_{\alpha}/n_{z}, (7)

where the subscript following a comma denotes the partial derivative. The elevation zz is a single-valued function of xx and yy since we have limited the discussion to the case where the nn-vector lies within 45 degrees from nn^{\circ} for guaranteeing the unique conversion from potency tensors to nn-vectors (§2.1).

Equation (7) states that the cotangent of the nn-vector coincides with the downward gradient of a curve in a two-dimensional space, and the slip direction perpendicular to the nn-vector is exactly the fault orientation in two dimensions. Hence, Shimizu et al. (2021) averaged the potency along the short side of the reference surface and smoothly connected the resulting nn-vector along the long side to extract two-dimensional geometrical features of a uniaxially bended fault (e.g., along-strike strike angle variations and along-dip dip angle variations). The same procedure could generate quasi-two-dimensional shapes including uniaxial twists such as along-strike dip variations and along-dip strike variations (§5.3). This paper counts such quasi-two-dimensional shape reconstruction as the previous method by Shimizu et al. (2021).

However, three-dimensional surface reconstruction is beyond the scope of the method of Shimizu et al. (2021), which is ascribed to the difference in the degrees of freedom of nn-vectors between two and three dimensions. In two dimensions, the nn-vector may be expressed by the polar angle θ\theta. The transformation of a given nn-vector field to an elevation zz field is then a one-to-one mapping θz\theta\mapsto z, which provides a unique zz field to an arbitrary smooth nn field. By contrast, in three dimensions, the nn-vector may be expressed by the polar angle θ\theta and the azimuthal angle ϕ\phi. Three-dimensional surface reconstruction from the nn-vector is therefore a two-to-one mapping (θ,ϕ)z(\theta,\phi)\mapsto z from a vector field of θ\theta and ϕ\phi to a scalar field of zz, generally having no solutions for eqs. (6) and (7) unless the given nn-vector field is error-free. Namely, the surface reconstruction from potency is an overdetermined problem in three dimensions.

Its inverse, the one-to-two mapping of zz on (θ,ϕ)(\theta,\phi), is precisely the definition of the normal vector simplified as eq. (7), and it is made bijective by the smoothness of the surface. A smooth surface with no tears, the aforementioned regular surface, sets a closed-path integral of relative elevation to zero:

𝑑𝐥z=0,\oint d{\bf l}\cdot\nabla z=0, (8)

where =(x,y)\nabla=(\partial_{x},\partial_{y}) denotes the two-dimensional partial differential operator. From Stokes’ theorem, it is expressed as

𝑑Σ×z=0,\int d\Sigma\nabla\times\nabla z=0, (9)

that is,

xyz=yxz.\partial_{x}\partial_{y}z=\partial_{y}\partial_{x}z. (10)

Under this constraint on zz by eq. (10), the one-to-two mapping of zz on (θ,ϕ)(\theta,\phi) becomes bijective. Equation (10) indicates that the requirement of the smooth surface with no tears is that z\nabla z is integrable, that is, the elevation zz is a continuous function with no jumps. Quasi-two-dimensional surface reconstruction is an approximation that neglects that the slope along the long side may vary along the short side (e.g. yz,x0\partial_{y}z_{,x}\simeq 0), so it violates eq. (10) when the slope along the short side varies along the long side (e.g. xz,y0\partial_{x}z_{,y}\neq 0).

Substituting eq. (7) into eq. (10) leads to

x(ny/nz)=y(nx/nz).\partial_{x}(n_{y}/n_{z})=\partial_{y}(n_{x}/n_{z}). (11)

Two-dimensional shapes of uniaxial bending (single-curved surfaces) treated in Shimizu et al. (2021) always set both sides of eq. (11) equally to zero. However, only the error-free nn-vector field of a surface meets eq. (11) in three dimensions. Other than such special cases, the estimated normal vectors of curved surfaces violate the integrability condition of eq. (11), and that is the reason why the (θ,ϕ)z(\theta,\phi)\mapsto z mapping of eqs. (6) and (7) is overdetermined:

x(ny/nz)y(nx/nz).\partial_{x}(n_{*y}/n_{*z})\neq\partial_{y}(n_{*x}/n_{*z}). (12)

Since the source surface reconstruction via eqs. (6) and (7) is overdetermined and eq. (7) is the identity of the normal vector of a surface, surface reconstruction from given potency must read eq. (6) as some projection:

n(𝝃;Γ)𝒫n(𝝃;𝐚),n(\boldsymbol{\xi};\Gamma)\xrightarrow{\mathcal{P}}n_{*}(\boldsymbol{\xi};{\bf a}), (13)

where 𝒫\xrightarrow{\mathcal{P}} denotes some operation to remove non-integrable components such that x(ny/nz)y(nx/nz)\partial_{x}(n_{*y}/n_{*z})\neq\partial_{y}(n_{*x}/n_{*z}). Only if that non-integrable portion of nn_{*} is removed, that is, only if illogical tears are ruled out, a fault is reconstructed as a smooth surface in a coarse-grained continuum.

Because nzn\mapsto z mapping is impossible when the theoretical relationship between potency and shape of eq. (6) strictly holds, surface reconstruction from potency must loosen this premise. There remains, however, arbitrariness in eq. (13) on how that loosening is applied. We describe this subjective postulation as prior information. Equation (13) expects that the nn-vector 𝐧(𝝃,Γ){\bf n}(\boldsymbol{\xi},\Gamma) of a surface Γ\Gamma approaches the nn-vector 𝐧(𝐚){\bf n}_{*}({\bf a}) (defined by eq. 2) pointed by the nodal planes of potency beachballs. The constraint of approximating a normalized vector to a designated value yields a non-Gaussian central-limit distribution known as the von Mises-Fisher distribution; it measures the distance of vectors by the inner product, and for a uniform isotropic case,

P(Γ|𝐚;κ)=exp[κΓ𝑑Σ(𝝃)𝐧(𝝃,𝐚)𝐧(𝝃,Γ)]/𝒵,P(\Gamma|{\bf a};\kappa)=\exp[\kappa\int_{\Gamma}d\Sigma(\boldsymbol{\xi}){\bf n}_{*}(\boldsymbol{\xi},{\bf a})\cdot{\bf n}(\boldsymbol{\xi},\Gamma)]/\mathcal{Z}, (14)

where κ\kappa denotes a scale factor, and 𝒵\mathcal{Z} is a normalization constant. The domain of integration is set to Γ\Gamma, and P(Γ|𝐚;κ)P(\Gamma|{\bf a};\kappa) is independent of the way to take the reference surface Γ\Gamma^{\circ}.

Then, eq. (13) requires the limit that places 𝐧{\bf n} in the integrable nearest neighbor of non-integrable 𝐧{\bf n}_{*}, that is, the displacement discontinuity that best approximates given inelastic strain. It is termed dislocation limit hereafter, which is, in eq. (14),

κ.\kappa\to\infty. (15)

In this limit,

P(Γ|𝐚)=δ(ΓΓ(𝐚)),P(\Gamma|{\bf a})=\delta(\Gamma-\Gamma_{*}({\bf a})), (16)

where δ()\delta(\cdot) denotes Dirac delta and

Γ:=argmaxΓ|𝐚P(Γ|𝐚;κ).\Gamma_{*}:={\rm argmax}_{\Gamma|{\bf a}}P(\Gamma|{\bf a};\kappa). (17)

Here argmax{\rm argmax} denotes the arguments of the maxima. In this paper, P(Γ|𝐚)P(\Gamma|{\bf a}) is taken as limκP(Γ|𝐚,κ)\lim_{\kappa\to\infty}P(\Gamma|{\bf a},\kappa).

When (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z expected from nn_{*} is small, nn obtained from eq. (13) is not so shifted from nn_{*}. Basically, we suppose such cases and proceed by assuming P(Γ|𝐚)P(\Gamma|{\bf a}) of eq. (14). However, the overdetermination of surface reconstruction, in which any surface cannot obey the forward model eq. (2) that yields eq. (6), leaves open the question of what priors P(Γ|𝐚)P(\Gamma|{\bf a}) are appropriate to loosen eq. (2) when (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z of nn_{*} is not small. Even before the nn-vector is disturbed by measurement errors that do not satisfy eq. (6), the fault face as an earthquake source is an approximation of a fracture zone, which accounts for the nature of brittle failure that localizes within a quasi-face (Matsu’ura et al., 2019), so the exact observance of eq. (2) is originally unphysical. Thus, if read as a forward model prior to the observation, P(Γ|𝐚)P(\Gamma|{\bf a}) for returning a surface from a given nn-vector field is a probabilistic delineation of the ambiguity of an earthquake fault, an idealization of a thin sheet of inelastic deformation. P(Γ|𝐚)P(\Gamma|{\bf a}) is different from the posterior probability of fault shape. The nn-vector-geometry mapping, and therefore the surface reconstruction of earthquake faults from potency, are never beyond Bayesian probabilistic even when potency is given deterministically, and the posterior of fault shape is the convolution between the conditional prior of fault shape given potency P(Γ|𝐚)P(\Gamma|{\bf a}) and the posterior of potency (Appendix A).

4 An analytic solution for Γ\Gamma_{*} as the optimal elevation field

It has been shown inevitably probabilistic to project two degrees of freedom in the expected normal nn_{*} onto one degree of freedom in the smooth surface elevation. Bearing P(Γ|𝐚)P(\Gamma|{\bf a}) of eq. (14) in mind, we derive an analytical expression of the most probable surface estimate Γ(𝐚)\Gamma_{*}({\bf a}) given the potency 𝐚{\bf a}. This section assumes that the non-integrable component of nn_{*} (eq. 12) is small so that the solution of surface reconstruction does not depend much on the functional form of P(Γ|𝐚)P(\Gamma|{\bf a}).

We have supposed that the potency density 𝐚{\bf a} uniquely determines the field of the normal vector nn_{*}2.1), so the goal of this section is to obtain Γ\Gamma_{*} given nn_{*}. When the xx-yy plane describes the location along the reference surface Γ\Gamma^{\circ}, Γ\Gamma is parametrized by z(,)z(\cdot,\cdot), which represents the elevation of Γ\Gamma from Γ\Gamma^{\circ} as a function of the position along the reference surface (x,y)(x,y).

4.1 Governing equation

The governing equation of the problem is now the definitional identity of the nn-vector (eq. 7). It indicates the orthogonality between the nn-vector and a curved surface, which in three dimensions means more than that a 90-degree rotation (cotangent) of the nn-vector gives the downward surface slope; introducing

n\displaystyle n_{\perp} =nx2+ny2\displaystyle=\sqrt{n_{x}^{2}+n_{y}^{2}} (18)
𝐧\displaystyle{\bf n}_{\parallel} =(nx,ny)/n,\displaystyle=(n_{x},n_{y})/n_{\perp}, (19)

we can rewrite eq. (7) as

z=nnz𝐧.\nabla z=-\frac{n_{\perp}}{n_{z}}{\bf n}_{\parallel}. (20)

Namely, the nn-vector indicates the direction of the slope by its horizontal projection 𝐧{\bf n}_{\parallel} (|𝐧|=1|{\bf n}_{\parallel}|=1) as well as the downward slope by its horizontal-vertical ratio n/nzn_{\perp}/n_{z} (Fig. 3).

Refer to caption
Figure 3: Details of the orthogonality between the nn-vector 𝐧{\bf n} and the tangent plane d𝝃d\boldsymbol{\xi} of a surface. A Cartesian coordinate is introduced such that the reference-surface normal 𝐧{\bf n}^{\circ} points to the zz axis, and xx and yy represent the location along the reference surface. The surface slope direction is indicated by the horizontal projection 𝐧{\bf n}_{\parallel} of 𝐧{\bf n}. The downward surface slope coincides with the ratio of the horizontal component nn_{\perp} to the vertical component nzn_{z} of 𝐧{\bf n}.

The remaining analysis is straightforward when the nn-vector is parametrized in a spherical coordinate. Now the polar angle θ[0,π]\theta\in[0,\pi] is set as the angle between 𝐧{\bf n}^{\circ} and 𝐧{\bf n}, and the azimuthal angle ϕ[0,2π)\phi\in[0,2\pi) is measured from the xx axis. For this setting,

nz\displaystyle n_{z} =cosθ\displaystyle=\cos\theta (21)
n\displaystyle n_{\perp} =sinθ\displaystyle=\sin\theta
nx\displaystyle n_{\parallel x} =cosϕ\displaystyle=\cos\phi
ny\displaystyle n_{\parallel y} =sinϕ.\displaystyle=\sin\phi.

Note nx=sinθcosϕn_{x}=\sin\theta\cos\phi and ny=sinθsinϕn_{y}=\sin\theta\sin\phi. Plugging eq. (21) into eq. (20),

z=tanθ(cosϕsinϕ).\nabla z=-\tan\theta\left(\begin{array}[]{c}\cos\phi\\ \sin\phi\end{array}\right). (22)

Hereafter, the basis vectors for the position (θ,ϕ)(\theta,\phi) on a unit sphere pointing toward 𝐧{\bf n} are denoted as 𝐧{\bf n} (along the nn-vector), 𝐞θ{\bf e}_{\theta} (parallel to the θ\theta axis), and 𝐞ϕ{\bf e}_{\phi} (parallel to the ϕ\phi axis).

4.2 Solving a stochastic partial differential equation for fault elevation

The shape of Γ\Gamma is equivalent to both the zz field and the nn field, but formally, eq. (7) can be read as a partial differential equation of zz with a random external force (n/nz)𝐧-(n_{\perp}/n_{z}){\bf n}_{\parallel} imposed. The probability P(Γ|𝐚;κ)P(\Gamma|{\bf a};\kappa) (eq. 14) describes fluctuations of the nn-vector from the expectation nn_{*}. We derive an analytical solution of this stochastic partial differential equation. For this derivation, we assume non-integrable components (eq. 12) of nn_{*} are small; the full-order derivation becomes complicated (Appendix B).

First, we rewrite the Cartesian forms of P(Γ|𝐚;κ)P(\Gamma|{\bf a};\kappa) (eq. 14) and the partial differential equation of the elevation (eq. 7) in the spherical coordinate. In a spherical coordinate system at a unit spherical position (θ\theta_{*}, ϕ\phi_{*}) of 𝐧{\bf n}_{*}, 𝐧{\bf n} is expressed as

𝐧=χn𝐧+χθ𝐞θ+χϕ𝐞ϕ,{\bf n}=\chi_{n}^{*}{\bf n}_{*}+\chi_{\theta}^{*}{\bf e}_{\theta}^{*}+\chi_{\phi}^{*}{\bf e}_{\phi}^{*}, (23)

with associated coefficients (χn\chi_{n}^{*}, χθ\chi_{\theta}^{*}, χϕ\chi_{\phi}^{*}). The first order of the Taylor series with respect to angle fluctuations θθ\theta-\theta_{*} and ϕϕ\phi-\phi_{*} around nn_{*} yields

χθ\displaystyle\chi_{\theta}^{*} =(θθ)+𝒪(δΩ2)\displaystyle=(\theta-\theta_{*})+\mathcal{O}(\delta\Omega^{2}) (24)
χϕ\displaystyle\chi_{\phi}^{*} =(ϕϕ)sinθ+𝒪(δΩ2),\displaystyle=(\phi-\phi_{*})\sin\theta_{*}+\mathcal{O}(\delta\Omega^{2}),

where 𝒪(δΩk)\mathcal{O}(\delta\Omega^{k}) denotes the kk-th order angle fluctuations from nn_{*}. Note 𝐧(θ,ϕ)/θ=𝐞θ\partial{\bf n}(\theta,\phi)/\partial\theta={\bf e}_{\theta} and 𝐧(θ,ϕ)/ϕ=𝐞ϕsinθ\partial{\bf n}(\theta,\phi)/\partial\phi={\bf e}_{\phi}\sin\theta. The normalization condition 𝐧𝐧=1{\bf n}\cdot{\bf n}=1 of 𝐧{\bf n} and the above expressions of χθ\chi_{\theta}^{*} and χϕ\chi_{\phi}^{*} result in

χn=1[(θθ)2+(ϕϕ)2sin2θ]/2+𝒪(δΩ3).\chi_{n}^{*}=1-[(\theta-\theta_{*})^{2}+(\phi-\phi_{*})^{2}\sin^{2}\theta_{*}]/2+\mathcal{O}(\delta\Omega^{3}). (25)

Therefore, for the von Mises-Fisher distribution of eq. (14), we have

𝐧𝐧=1[(θθ)2+(ϕϕ)2sin2θ]/2+𝒪(δΩ3){\bf n}_{*}\cdot{\bf n}=1-[(\theta-\theta_{*})^{2}+(\phi-\phi_{*})^{2}\sin^{2}\theta_{*}]/2+\mathcal{O}(\delta\Omega^{3}) (26)

and

P(Γ|𝐚;κ)=exp{κ2Γ𝑑Σ[(θθ)2+(ϕϕ)2sin2θ+𝒪(δΩ3)]}/𝒵,P(\Gamma|{\bf a};\kappa)=\exp\{-\frac{\kappa}{2}\int_{\Gamma}d\Sigma[(\theta-\theta_{*})^{2}+(\phi-\phi_{*})^{2}\sin^{2}\theta_{*}+\mathcal{O}(\delta\Omega^{3})]\}/\mathcal{Z}, (27)

where 𝒵\mathcal{Z}^{\prime} represents normalization. Only the second-order fluctuations from nn_{*} matter in the dislocation limit κ\kappa\to\infty if non-integrable components (eq. 12) of nn_{*} are small.

We next evaluate the partial differential equation of the elevation (eq. 7), which is expressed as eq. (22) in the spherical coordinate. The Taylor series of eq. (22) in angle fluctuations from nn_{*} leads to

z=zcos2θ(cosϕsinϕ)(θθ)tanθ(sinϕcosϕ)(ϕϕ)+𝒪(δΩ2),\nabla z=\nabla z_{*}-\cos^{-2}\theta_{*}\left(\begin{array}[]{c}\cos\phi_{*}\\ \sin\phi_{*}\end{array}\right)(\theta-\theta_{*})-\tan\theta_{*}\left(\begin{array}[]{c}-\sin\phi_{*}\\ \cos\phi_{*}\end{array}\right)(\phi-\phi_{*})+\mathcal{O}(\delta\Omega^{2}), (28)

where z,α:=nα/nzz_{*,\alpha}:=-n_{\alpha}/n_{z}. Using eq. (21) again and taking the inner and cross products of z\nabla z and 𝐧{\bf n}_{*\parallel}, we find

θθ\displaystyle\theta-\theta_{*} =nz2𝐧(zz)+𝒪(δΩ2)\displaystyle=-n_{*z}^{2}{\bf n}_{*\parallel}\cdot(\nabla z-{\nabla z}_{*})+\mathcal{O}(\delta\Omega^{2}) (29)
ϕϕ\displaystyle\phi-\phi_{*} =(nz/n)𝐧×(zz)+𝒪(δΩ2),\displaystyle=-(n_{*z}/n_{*\perp}){\bf n}_{*\parallel}\times(\nabla z-{\nabla z}_{*})+\mathcal{O}(\delta\Omega^{2}), (30)

or explicitly, recalling 𝐧z=n/nz{\bf n}_{\parallel}\cdot\nabla z=-n_{\perp}/n_{z} and 𝐧×z=0{\bf n}_{\parallel}\times\nabla z=0 (eq. 20),

θθ\displaystyle\theta-\theta_{*} =nz2(nxz,x+nyz,y+n/nz)+𝒪(δΩ2)\displaystyle=-n_{*z}^{2}(n_{*\parallel x}z_{,x}+n_{*\parallel y}z_{,y}+n_{*\perp}/n_{*z})+\mathcal{O}(\delta\Omega^{2}) (31)
ϕϕ\displaystyle\phi-\phi_{*} =n1nz(nxz,ynyz,y)+𝒪(δΩ2).\displaystyle=-n_{*\perp}^{-1}n_{*z}(n_{*\parallel x}z_{,y}-n_{*\parallel y}z_{,y})+\mathcal{O}(\delta\Omega^{2}). (32)

Plugging eqs. (31) and (32) into P(Γ|𝐚)P(\Gamma|{\bf a}) (eq. 27 with κ\kappa\to\infty), we arrive at

ln[P(Γ|𝐚)×𝒵]=κ2{ΓdΣ[nz2(nxz,x+nyz,y+n/nz)]2\displaystyle\ln[P(\Gamma|{\bf a})\times\mathcal{Z}^{\prime}]=-\frac{\kappa}{2}\left\{\int_{\Gamma}d\Sigma[n_{*z}^{2}(n_{*\parallel x}z_{,x}+n_{*\parallel y}z_{,y}+n_{*\perp}/n_{*z})]^{2}\right. (33)
+ΓdΣ[nz(nxz,ynyz,y)]2+𝒪(δΩ3)]}.\displaystyle\left.+\int_{\Gamma}d\Sigma[n_{*z}(n_{*\parallel x}z_{,y}-n_{*\parallel y}z_{,y})]^{2}+\mathcal{O}(\delta\Omega^{3})]\right\}.

Here sinθ=n\sin\theta=n_{\perp} (eq. 21) is also used. The area element dΣ(𝝃)d\Sigma(\boldsymbol{\xi}) along the fault is larger than the area element dxdydxdy along the reference plane because of the tilt of the fault normal by θ\theta:

dΣ(𝝃)=dxdy|cosθ(𝝃)|=|nz|1dxdy.d\Sigma(\boldsymbol{\xi})=\frac{dxdy}{|\cos\theta(\boldsymbol{\xi})|}=|n_{z}|^{-1}dxdy. (34)

Note |nz|=nz|n_{z}|=n_{z} for θ\theta below 45 degrees now considered. In the dislocation limit, with small non-integrable components (eq. 12) of nn_{*}, the higher order fluctuations from nn_{*} are dropped, and the most probable estimate Γ\Gamma_{*} for given nn_{*} is expressed as:

Γ=argmin{|nz|(nxnzz,x+nynzz,y+n)2dxdy\displaystyle\Gamma_{*}={\rm argmin}\left\{\int|n_{*z}|(n_{*\parallel x}n_{*z}z_{,x}+n_{*\parallel y}n_{*z}z_{,y}+n_{*\perp})^{2}dxdy\right. (35)
+|nz|(nxz,ynyz,y)2dxdy},\displaystyle\left.+\int|n_{*z}|(n_{*\parallel x}z_{,y}-n_{*\parallel y}z_{,y})^{2}dxdy\right\},

where argmin{\rm argmin} denotes the arguments of the minima. To avoid zero divisions at nz=0n_{*z}=0, nz2n_{*z}^{2} among the weight |nz3||n_{*z}^{3}| in the first term was placed inside the brackets. The first term from the θ\theta fluctuation constrains the 𝐧{\bf n}_{\parallel}-parallel component of the fault slope z\nabla z, and the second term from the ϕ\phi fluctuation constrains the 𝐧{\bf n}_{\parallel}-perpendicular component of z\nabla z. Their weights |nz3||n_{*z}^{3}| and |nz||n_{*z}| derive from the independence of P(Γ|𝐚)P(\Gamma|{\bf a}) from the reference plane orientation, and consequently, the constraints become weaker for steeper slopes (i.e., smaller nzn_{*z} for larger θ(<π/4)\theta(<\pi/4): nz/θ=sinθ<0\partial n_{*z}/\partial\theta_{*}=-\sin\theta_{*}<0).

Thus, equations to be solved (eqs. 7 and 14) are reduced to a least square form of eq. (35). The rest of analysis is the ordinary linear inversion, which we omit to detail. We have solved it using a basis function expansion in this paper, instead of solving it in the continuous space, where the xx and yy derivatives of the nn-vector appear, complicating implementations outside differentiable programming. The basis functions employed are the quadratic B-spline functions as in the PDTI, with intervals 1.5 times wider than that of the PDTI. The surface integral of the loss function is collocated at the xx-yy positions of the spline knots adopted in the PDTI. We have removed the rigid-body translation of the fault by specifying a point with zero elevation. This translation of a fault is unconstrained in eq. (35) and does not affect fault elevations regardless of the way of removal.

The resultant estimation remains the least square using eqs. (31) and (32) even when generalizing P(Γ|𝐚)P(\Gamma|{\bf a}) to account for the expected elevation field z¯\bar{z} such that

Q(Γ|𝐚)=exp[12𝑑Σ(𝝃)𝑑Σ(𝝃)(θ(𝝃)θ(𝝃)ϕ(𝝃)ϕ(𝝃)z(𝝃)z¯(𝝃))T𝐰(𝝃,𝝃)(θ(𝝃)θ(𝝃)ϕ(𝝃)ϕ(𝝃)z(𝝃)z¯(𝝃))]/const.Q(\Gamma|{\bf a})=\exp\left[-\frac{1}{2}\int d\Sigma(\boldsymbol{\xi})\int d\Sigma(\boldsymbol{\xi}^{\prime})\left(\begin{array}[]{cc}\theta(\boldsymbol{\xi})-\theta_{*}(\boldsymbol{\xi})\\ \phi(\boldsymbol{\xi})-\phi_{*}(\boldsymbol{\xi})\\ z(\boldsymbol{\xi})-\bar{z}(\boldsymbol{\xi})\end{array}\right)^{\rm T}{\bf w}(\boldsymbol{\xi},\boldsymbol{\xi}^{\prime})\left(\begin{array}[]{cc}\theta(\boldsymbol{\xi}^{\prime})-\theta_{*}(\boldsymbol{\xi}^{\prime})\\ \phi(\boldsymbol{\xi}^{\prime})-\phi_{*}(\boldsymbol{\xi}^{\prime})\\ z(\boldsymbol{\xi}^{\prime})-\bar{z}(\boldsymbol{\xi}^{\prime})\end{array}\right)\right]/const. (36)

Our PDTI application includes such terms for computational acceleration (§6.2). The scope of eq. (36) covers a generalization using the probabilities of source locations, such as surface reconstruction from the locations and mechanisms of distributed point sources. It is also applicable to incorporating the constraints on fault edge locations, such as accounting for fault surface traces and jointing multiple fault surfaces. Note P(Γ|𝐚)P(\Gamma|{\bf a}) other than eq. (14) is reduced to the quadratic form of eq. (36) when the higher orders are dropped. The explicit representation of those higher orders dropped in the dislocation limit is given by the non-perturbative expression of the nn-vector as a function of z\nabla z (eq. 37 in the next section).

5 Surface reconstruction from synthesized noisy normal vectors

We conduct synthetic tests of the most probable surface estimation (eq. 35) given the normal vector field. Simultaneous inference of potency density and geometry is often not significantly different from reconstructing fault geometry from the PDTI prescribing potency locations (Shimizu et al., 2021, also see §6.3). So this verification of surface reconstruction from a given noisy nn-vector field serves the core verification of the PDTI involving fault shape inference, which is performed in §6.

A correct elevation z0(x,y)z_{0}(x,y) perpendicular to the reference plane xx-yy is first synthesized, and its normal vector field n0(x,y)n_{0}(x,y) is given as n(x,y)n_{*}(x,y) disturbed by a noise δn(x,y)\delta n(x,y). From this noisy n(x,y)n_{*}(x,y), we compute an estimate of the elevation Γ\Gamma_{*} and check the accuracy compared to the correct shape. All components of the nn-vector are disturbed by Gaussian random numbers of mean 0 and standard deviation σn\sigma_{n}, and their normalized values are taken as nn_{*}. The nn_{*} values are collocated on the grid points, and we assume that the grid interval is enough large to neglect spatial correlations for simplicity. The grid is the same as the knot configuration in the basis function expansion of the PDTI appearing in our application (§6.3). It is arranged with 40 points along strike xx and 4 points along dip yy spaced by 5 km, containing 160 points in total. The length scale in the following tests is nondimensionalized by setting km at unity.

The transformation of the fault slope to the fault normal in Cartesian coordinates is given by eq. (7) and the normalization condition |𝐧|2=1|{\bf n}|^{2}=1:

nz=1/1+z,x2+z,y2\displaystyle n_{z}=1/\sqrt{1+z_{,x}^{2}+z_{,y}^{2}} (37)
nα=z,α/1+z,x2+z,y2.\displaystyle n_{\alpha}=-z_{,\alpha}/\sqrt{1+z_{,x}^{2}+z_{,y}^{2}}.

For spherical coordinates, from eq. (22),

θ=atanz,x2+z,y2\displaystyle\theta={\rm atan}\sqrt{z_{,x}^{2}+z_{,y}^{2}} (38)
ϕ=atan2(z,y,z,x),\displaystyle\phi={\rm atan}2(z_{,y},z_{,x}),

where atan2(,){\rm atan}2(\cdot,\cdot) denotes 2-argument arctangent.

Two model shapes are considered. The first shape is a uniaxial twist (§5.1). It is accompanied with biaxial slopes, providing a simplest instance of a surface that is torn (i.e. eq. 11 is violated) if a uniaxial interpolation is applied to the nn-vector field. It is also an example of the surfaces that can be expressed by spline interpolation without interpolation errors, called developable surfaces. These make this geometry useful for method comparisons with the existing spline-interpolation-based approach (Shimizu et al., 2021). The second is a bowl shape (§5.2), which is an example of non-developable surfaces that intrinsically need stochastic reconstruction. A similar geometry is estimated in our case study of the 2013 Balochistan earthquake (§6.3). In both examples, the maximum horizontal component of the normal vector is approximately n=0.45n_{\perp}=0.45, and the normal vector is tilted by approximately 40 degrees from the reference plane. These are fairly curved shapes within the range where the conversion from potency beachballs to normal vectors is straightforward, that is within tilt angles of 45 degrees. There is little difference in the maximum normal between the two, but their reconstruction accuracies differ (§5.3).

Now we suppose σn<0.25\sigma_{n}<0.25. Since fairly curved surfaces are computed, large noise results in nz<0n^{*}_{z}<0 (θ>π/2\theta>\pi/2) outside the present problem setting. The nz<0n^{*}_{z}<0 condition is formally accounted for in the loss function of eq. (34), but our consideration following eq. (7) treats only the case of a single-valued zz on the xx-yy plane, where nz0n^{*}_{z}\geq 0 holds. The nn-vector determination assuming the nn-vector as the nodal plane closer to the reference has already assumed θ<π/4\theta<\pi/4, so θπ/4\theta\geq\pi/4 containing nz<0n^{*}_{z}<0 is supposed to be avoided in this study by setting multiple reference normals (§2.1). Stochastic surface reconstruction involving stochastic nn-vector determinations is left open for future study.

5.1 A uniaxial twist

The first test treats the uniaxial twist shown in Fig. 4. Its elevation is symmetrical about the short side yy (see Fig. 4 caption). Except for the center line y=0y=0, where nx=0n_{x}=0 identically, both nxn_{x} and nyn_{y} are non-zero, and the nn-vector varies three-dimensionally. The two-dimensional vizualizaiton of the shape is a birds-eye view hereafter.

Refer to caption
Figure 4: The ground truth of our synthetic test 1: z=[(yy¯)/W]sin[(xx¯)/L]z=[(y-\bar{y})/W]\sin[(x-\bar{x})/L] with W=2W=2 and L=30L=30, where (x¯,y¯)(\bar{x},\bar{y}) denotes the reference plane center.

The estimation results with a noise standard deviation of 0.05 are indicated in Fig. 5. Despite that the input nn-vectors are noisy, the estimated elevation is consistent with the original shape. Concavities and convexities at the four corners of the original shape are reproduced while the contours shift. As mentioned in the previous section, the noise component (Δn=nn0\Delta n=n_{*}-n_{0}) almost surely does not satisfy the slope integrability condition (eq. 11). Therefore, when nn_{*} is disturbed from the nn-vector of a surface, the most probable surface Γ\Gamma_{*} (eq. 35) for given nn_{*} could selectively remove the noise in nn_{*}.

Refer to caption
Figure 5: Examples of (a) informed noisy nn-vectors and (b) reconstructed elevation and nn-vector fields in our synthetic test 1. The nn-vector randomly varies from the ground truth with a standard deviation of 0.05.

5.2 A bowl shape

The second test treats a bowl-shaped surface shown in Fig. 6, which is similar to the result of our actual data analysis concerning the 2013 Balochistan earthquake (§6.3). The elevation zz is expressed by a function that is parabolic along the long side xx and linear along the short side yy (see Fig. 6 caption). When xx is interpreted as the horizontal axis and yy as the vertical, the surface horizontally bends well at the top and increases vertical tilts at the horizontal end (larger |z,x||z_{,x}| for larger yy and larger z,yz_{,y} for larger |x||x|, roughly speaking).

Refer to caption
Figure 6: The ground truth of our synthetic test 2: z=y/Y1+(x/X)2(YY2)/Y1z=-y/Y_{1}+(x/X)^{2}(Y-Y_{2})/Y_{1} with X=50X=50, Y1=3Y_{1}=3, and Y2/Y1=25Y_{2}/Y_{1}=-25.

The results for a standard deviation of 0.05 are indicated in Fig. 7 as in the first test. The reconstructed shape seems more stable than the reconstructed uniaxial twist. This difference suggests that when the nn-vector noise is at the same level, short-wavelength features such as topography are relatively susceptible to the nn-vector noise. Considering that wider-regional elevations are controlled by longer-wavelength slopes, we ascribe this tendency to the spatial averaging that low-passes the slope error in a long-wavelength scale. The higher surface reproducibility at longer wavelengths might be reinforced in the application to the PDTI, to which the spatial potency smoothing is commonly applied.

Refer to caption
Figure 7: Examples of (a) informed noisy nn-vectors and (b) reconstructed elevation and nn-vector fields in our synthetic test 2. The nn-vector randomly varies from the ground truth with a standard deviation of 0.05.

5.3 Robustness tests of surface reconstruction

Figure 8 summarizes the errors and residuals of the above two tests with variable noise levels. The estimation errors 1n^n01-\hat{n}\cdot n_{0} and data residuals 1n^n1-\hat{n}\cdot n_{*} for normal vectors (Fig. 8a, b) are both linear with the noise amplitude 1n0n1-n_{0}\cdot n^{*}. The data residuals are visibly smaller than the noise amplitude, indicating significant variance reduction. The estimation error 1n^n01-\hat{n}\cdot n_{0} is smaller than the data residual 1n^n1-\hat{n}\cdot n_{*}, suggesting that the output denoises the given nn-vector rather than explains it. The spatial average of elevation errors |zz¯||z-\bar{z}| (Γ\Gamma_{*} in Fig. 8c, d) is more scattered than that of normal vectors (Fig. 8a, b), and its average is roughly proportional to the square root of the noise amplitude of the normal vectors 1n0n1-n_{0}\cdot n^{*} that approximates squared angle fluctuations. The elevation residuals are roughly the same between the two shapes (Fig. 8c, d), and these same-order absolute residuals indicate a significant difference in relative residuals, supporting higher reproducibility of a bowl-shaped curve compared to uniaxially twisting roughness observed in the previous subsections.

Refer to caption
Figure 8: Errors and residuals in our synthetic tests, plotted in terms of nn-vectors in tests 1 (a) and 2 (b) and elevations in tests 1 (c) and 2 (d). All values are averaged over the reference plane, and 64 noise patterns are used to evaluate the means and standard deviations shown in the panels. The horizontal axis 1nn01-n_{*}\cdot n_{0} represents the noise level of input nn-vectors nn_{*} from the ground truth n0n_{0}. The vertical axis 1n^n1-\hat{n}\cdot n in (a) and (b) represents the errors 1n^n01-\hat{n}\cdot n_{0} (red) and the residuals 1n^n1-\hat{n}\cdot n_{*} (blue) of nn-vector estimates n^\hat{n}. The vertical axis |z^z0||\hat{z}-z_{0}| in (c) and (d) represents the error of estimated elevation z^\hat{z} from the ground truth z0z_{0} (orange). Elevation errors of quasi-two-dimensional surface reconstruction (yellow) following Shimizu et al. (2021) are also indicated for comparison.

For comparison, we have also evaluated quasi-two-dimensional surface construction from Shimizu et al. (2021) (quasi 2D in Fig. 8c, d). It evaluates the elevation as z(x,y)z(x¯,y¯)+x¯x𝑑xz¯,x(x)+z¯,y(x)(yy¯)z(x,y)\simeq z(\bar{x},\bar{y})+\int_{\bar{x}}^{x}dx^{\prime}\bar{z}_{,x}(x^{\prime})+\bar{z}_{,y}(x)(y-\bar{y}) by using the mean slope z¯,α=dy(nα/nz)/𝑑y\bar{z}_{,\alpha}=-\int d_{y}(n^{*}_{\alpha}/n^{*}_{z})/\int dy integrated over the short side yy. We assume z(x¯,y¯)z(\bar{x},\bar{y}) at the reference plane center (x¯,y¯)(\bar{x},\bar{y}) is known in this quasi-two-dimensional reconstruction. The stochastic surface reconstruction Γ\Gamma_{*} works well as quasi-two-dimensional reconstruction for small noise, and for large noise, rather Γ\Gamma_{*} has smaller estimation errors. Even for a uniaxial twist, for which quasi-two-dimensional approximation is errorless, the accuracy of Γ\Gamma_{*} is higher, though the difference is less than accuracy variations (Fig. 8c). Meanwhile, for a bowl shape similar to the fault shape inferred in our application (§6.3), the accuracy of Γ\Gamma_{*} is a few times greater than the quasi-two-dimensional benchmark for the maximum noise amplitude in this test (Fig. 8d).

The fault we use here has a narrow width, so quasi-two-dimensional interpolation from Shimizu et al. (2021) provides a good approximation if no noise exists. Even for such high-aspect-ratio source faults, nonetheless, the results of this study indicate that a probabilistic approach is beneficial for surface reconstruction from noisy nn-vector fields. Rather than reducing noise through the law of large numbers associated with averaging along the short sides, denoising based on the a priori knowledge that nn-vector components that do not satisfy the slope integrability condition (eq. 11) are fictitious is likely to better reproduce the original shape.

6 Application

Finally, we apply our normal-to-elevation projection to the PDTI involving surface reconstruction (Shimizu et al., 2021). The aim of this section is three-dimensional earthquake fault inference.

6.1 Incorporating three-dimensional surface reconstruction with Bayesian PDTI

For the application, we combine the prior constraint on fault geometry P(Γ|𝐚)P(\Gamma|{\bf a}) with the posterior probability of the potency density in the orthodox PDTI. We have considered static problems so far, and hereafter we also consider dynamic problems to invert teleseismic data. The result obtained thus far is applicable as is even when the representation theorem becomes time-dependent and the basis functions are spanned spatiotemporally.

The formulation of the orthodox PDTI, both for static and dynamic problems, consists of the potency likelihood P(𝐝|𝐚,Γ;𝝈2)P({\bf d}|{\bf a},\Gamma;\boldsymbol{\sigma}^{2}), which is given by data 𝐝{\bf d} through the representation theorem, and the prior P(𝐚;𝝆2)P({\bf a};\boldsymbol{\rho}^{2}) for the potency density subject to the basis function expansions 𝐚{\bf a}. The scales of variances for the likelihood 𝝈2\boldsymbol{\sigma}^{2} and weight factors for prior loss functions 𝝆2\boldsymbol{\rho}^{2} are simultaneously estimated. Usually, 𝝈2\boldsymbol{\sigma}^{2} represents the scales of observation errors and Green’s function errors (Yagi & Fukahata, 2011), and 𝝆2\boldsymbol{\rho}^{2} represents the smoothing intensities with respect to space and time (Shimizu et al., 2020). The problem to be solved here consists of these P(𝐝|𝐚,Γ;𝝈2)P({\bf d}|{\bf a},\Gamma;\boldsymbol{\sigma}^{2}) and P(𝐚;𝝆2)P({\bf a};\boldsymbol{\rho}^{2}) and the prior for the surface P(Γ|𝐚)P(\Gamma|{\bf a}), and its formal solution is the joint posterior of 𝐚{\bf a} and Γ\Gamma:

P(𝐚,Γ|𝐝;𝝈2,𝝆2)=P(𝐝|𝐚,Γ;𝝈2)P(𝐚;𝝆2)P(Γ|𝐚)/P(𝐝;𝝈2,𝝆2),P({\bf a},\Gamma|{\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2})=P({\bf d}|{\bf a},\Gamma;\boldsymbol{\sigma}^{2})P({\bf a};\boldsymbol{\rho}^{2})P(\Gamma|{\bf a})/P({\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2}), (39)

where

P(𝐝;𝝈2,𝝆2)=𝑑𝐚𝑑ΓP(𝐝|𝐚,Γ;𝝈2)P(𝐚;𝝆2)P(Γ|𝐚).P({\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2})=\int d{\bf a}d\Gamma P({\bf d}|{\bf a},\Gamma;\boldsymbol{\sigma}^{2})P({\bf a};\boldsymbol{\rho}^{2})P(\Gamma|{\bf a}). (40)

The analysis for the dislocation limit is simplified when the joint posterior P(𝐚,Γ|𝐝;𝝈2,𝝆2)P({\bf a},\Gamma|{\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2}) is decomposed into the marginal posterior of the potency P(𝐚|𝐝;𝝈2,𝝆2)=𝑑ΓP(𝐚,Γ|𝐝;𝝈2,𝝆2)P({\bf a}|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})=\int d\Gamma P({\bf a},\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) and the conditional posterior of the surface P(Γ|𝐚,𝐝;𝝈2,𝝆2)P(\Gamma|{\bf a},{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}):

P(𝐚,Γ|𝐝;𝝈2,𝝆2)=P(𝐚|𝐝;𝝈2,𝝆2)P(Γ|𝐚,𝐝;𝝈2,𝝆2).P({\bf a},\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})=P({\bf a}|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})P(\Gamma|{\bf a},{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}). (41)

From eq. (16),

P(𝐚|𝐝;𝝈2,𝝆2)\displaystyle P({\bf a}|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) =P(𝐚,Γ(𝐚)|𝐝;𝝈2,𝝆2)\displaystyle=P({\bf a},\Gamma_{*}({\bf a})|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) (42)
P(Γ|𝐚,𝐝;𝝈2,𝝆2)\displaystyle P(\Gamma|{\bf a},{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) =δ(ΓΓ(𝐚)).\displaystyle=\delta(\Gamma-\Gamma_{*}({\bf a})). (43)

That is, the evaluation of P(𝐚,Γ|𝐝;𝝈2,𝝆2)P({\bf a},\Gamma|{\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2}) is reduced to the alternate recursion of the most probable estimation of Γ\Gamma given 𝐚{\bf a} returning Γ\Gamma_{*} (eq. 43) and the inference of 𝐚{\bf a} given Γ\Gamma_{*} (eq. 42).

We evaluate the hyperparameter optimals 𝝈^2\hat{\boldsymbol{\sigma}}^{2} and 𝝆^2\hat{\boldsymbol{\rho}}^{2} by Akaike’s Bayesian Information Criterion (ABIC; Akaike, 1980):

(𝝈^2,𝝆^2)=argmax𝝈2,𝝆2P(𝐝;𝝈2,𝝆2).(\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2})={\rm argmax}_{{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}}P({\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}). (44)

The associated optimals of the potency density and fault surface are evaluated by the posterior maximum:

(𝐚^,Γ^)=argmax𝐚,ΓP(𝐚,Γ|𝐝;𝝈^2,𝝆^2).(\hat{\bf a},\hat{\Gamma})={\rm argmax}_{{\bf a},\Gamma}P({\bf a},\Gamma|{\bf d};\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2}). (45)

Neglecting the third- and higher-order moments from the probability peaks as in Shimizu et al. (2021), eqs. (40) and (42) are simplified as

P(𝐝;𝝈2,𝝆2)𝑑𝐚P(𝐝|𝐚,Γ^;𝝈2)P(𝐚;𝝆2)P({\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})\simeq\int d{\bf a}P({\bf d}|{\bf a},\hat{\Gamma};{\boldsymbol{\sigma}}^{2})P({\bf a};{\boldsymbol{\rho}}^{2}) (46)

and

P(𝐚|𝐝;𝝈2,𝝆2)P(𝐝|𝐚,Γ^;𝝈2)P(𝐚;𝝆2)P(𝐝;𝝈2,𝝆2).P({\bf a}|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})\simeq\frac{P({\bf d}|{\bf a},\hat{\Gamma};{\boldsymbol{\sigma}}^{2})P({\bf a};{\boldsymbol{\rho}}^{2})}{P({\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})}. (47)

Besides, from eqs. (41) and (43),

Γ^=Γ(𝐚^).\hat{\Gamma}=\Gamma_{*}(\hat{\bf a}). (48)

Thus, our optimum-value estimation is the alternate recursion of solving for Γ^\hat{\Gamma} given 𝐚^\hat{\bf a} and solving for (𝐚^,𝝈^2,𝝆^2)(\hat{\bf a},\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2}) given Γ^\hat{\Gamma}. The analysis with a fixed Γ\Gamma is exactly the orthodox source inversion with fixed geometry (Yagi & Fukahata, 2011; Shimizu et al., 2020), and the analysis of Γ\Gamma given 𝐚{\bf a} is what this study has examined. Once Γ^\hat{\Gamma} and (𝝈^2,𝝆^2)(\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2}) are obtained, the uncertainty of 𝐚{\bf a} is evaluable through eq. (47) by using P(𝐚|𝐝;𝝈^2,𝝆^2)P({\bf a}|{\bf d};\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2}). Hyperparameter uncertainty is similarly evaluable using P(𝐝;𝝈2,𝝆2)P({\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) for given Γ^\hat{\Gamma}, though it is a sharply peaked well-behaved distribution and follows the law of large numbers exceptionally rapidly when 𝝈2{\boldsymbol{\sigma}}^{2} and 𝝆2{\boldsymbol{\rho}}^{2} are exponential loss weights (Sato et al., 2022).

The above procedure does not include explicit Γ\Gamma uncertainty evaluations. In the present problem setting using a delta-functional P(Γ|𝐚)P(\Gamma|{\bf a}), the estimation error of Γ\Gamma solely derives from error propagation from the potency density estimation, and its evaluation requires another joint posterior decomposition P(𝐚,Γ|𝐝;𝝈2,𝝆2)=P(Γ|𝐝;𝝈2,𝝆2)P(𝐚|Γ,𝐝;𝝈2,𝝆2)P({\bf a},\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})=P(\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})P({\bf a}|\Gamma,{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) and the evaluation of P(Γ|𝐝;𝝈2,𝝆2)=𝑑𝐚P(𝐚,Γ|𝐝;𝝈2,𝝆2)P(\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})=\int d{\bf a}P({\bf a},\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}). Due to the slope integrability (eq. 11) on a smooth surface, the posterior of the fault surface P(Γ|𝐝;𝝈2,𝝆2)P(\Gamma|{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) is deviated from the posterior of the nn-vector nn_{*} field expected from the potency density (Appendix A). The evaluation of P(𝐚|Γ,𝐝;𝝈2,𝝆2)P({\bf a}|\Gamma,{\bf d};{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) corresponds to the conventional slip inversion, where the potency density 𝐚{\bf a} is forced to be a displacement discontinuity field across given Γ\Gamma.

6.2 Recipe

The computations of 𝐚^\hat{\bf a}, Γ^\hat{\Gamma}, and (𝝈^2,𝝆^2)(\hat{\boldsymbol{\sigma}}^{2},\hat{\boldsymbol{\rho}}^{2}) (eqs. 44 and 45) as per eqs. (46)–(48) are formally the same as those of Shimizu et al. (2021) that alternately evaluate (𝐚,𝝈2,𝝆2)({\bf a},{\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2}) for given Γ\Gamma and Γ\Gamma for given 𝐚{\bf a}. Therefore, simply updating their fault reconstruction stage enables their routine to reconstruct three-dimensional fault geometry.

Figure 9 outlines our computational routine following Shimizu et al. (2021). (I: Fault model initialization) A structured grid is set along the reference plane, and the spatial knots for potency density interpolations are placed at grid centers; non-planar initializations following stage V below are optionally applied. (II: PDTI with Green’s function computations) The PDTI is performed, with Green’s function approximated by point-source representations; when the point sources are allocated to grid centers, they are weighted by the areas of gridded subsurfaces. (III: Nodal plane extraction) The nn-vector estimates of point sources are computed as nodal planes of estimated potency beachballs; the nn-vector field estimate is approximately computed by their interpolation. (IV: Surface reconstruction) The elevation field is reconstructed from the nn-vector field estimate. (V: PDTI with Green’s function updates) The aforementioned structured grid is vertically projected onto the reconstructed fault surface, and the vertical positions of sources and the areas of gridded subsurfaces are updated accordingly; the PDTI updates the potency beachballs and the loop returns to the stage III.

Refer to caption
Figure 9: A schematic diagram of simultaneous potency and fault shape estimation as per eq. (50). The loop starts with initializing the fault model. The potency tensors obtained by the PDTI are converted to the nn-vector field, which reconstructs the fault surface. The reconstructed surface updates Green’s functions, and alternate potency density and shape inferences restart. The distributed potency beachballs in the figure represent the estimate of the first iteration obtained by Shimizu et al. (2020), and the others refer to the results of Fig. 10.

In summary, once the PDTI is performed, potency beachballs set an nn-vector field, which in turn reconstructs the fault surface, the continuous information of which is subdivided as a digital elevation model to update Green’s function for the next PDTI. The digital elevation model is now computed by the Visualization Toolkit (Schroeder et al., 1998), and we call it by using a Python helper PyVista (Sullivan & Kaszynski, 2019). We execute the following two operations via PyVista: (A) the projection of the structured grid generated by numpy.meshgrid onto the reconstructed fault surface (pyvista.StructuredGrid) and (B) the computation of source locations and subdivided areas on vertically projected structured grids (compute_cell_sizes, cell_centers).

We solve this recursion by a gradient descent, where the Γ\Gamma_{*} determination (eq. 35) is modified by a penalty term on the residual elevation from a given shape Γ¯\bar{\Gamma} with elevation z¯\bar{z} as

Γ(𝐚,Γ¯;λ)=argmin{nz(nxnzz,x+nynzz,y+n)2dxdy\displaystyle\Gamma_{*}^{\prime}({\bf a},\bar{\Gamma};\lambda)={\rm argmin}\left\{\int n_{*z}(n_{*\parallel x}n_{*z}z_{,x}+n_{*\parallel y}n_{*z}z_{,y}+n_{*\perp})^{2}dxdy\right. (49)
+nz(nxz,ynyz,y)2dxdy+λ1|zz¯|2dxdy},\displaystyle\left.+\int n_{*z}(n_{*\parallel x}z_{,y}-n_{*\parallel y}z_{,y})^{2}dxdy+\lambda^{-1}\int|z-\bar{z}|^{2}dxdy\right\},

and each surface update accounts for the pre-update state:

(𝝈^2(n),𝝆^2(n))\displaystyle(\hat{\boldsymbol{\sigma}}^{2(n)},\hat{\boldsymbol{\rho}}^{2(n)}) :=argmax(𝝈2,𝝆2)𝑑𝐚P(𝐝|𝐚,Γ^(n);𝝈2)P(𝐚;𝝆2)\displaystyle:={\rm argmax}_{({\boldsymbol{\sigma}}^{2},{\boldsymbol{\rho}}^{2})}\int d{\bf a}P({\bf d}|{\bf a},\hat{\Gamma}^{(n)};\boldsymbol{\sigma}^{2})P({\bf a};{\boldsymbol{\rho}}^{2}) (50)
𝐚^(n)\displaystyle\hat{\bf a}^{(n)} :=argmax𝐚P(𝐝|𝐚,Γ^(n);𝝈^2(n))P(𝐚;𝝆^2(n))\displaystyle:={\rm argmax}_{\bf a}P({\bf d}|{\bf a},\hat{\Gamma}^{(n)};\hat{\boldsymbol{\sigma}}^{2(n)})P({\bf a};\hat{\boldsymbol{\rho}}^{2(n)})
Γ^(n+1)\displaystyle\hat{\Gamma}^{(n+1)} :=Γ(𝐚^(n),Γ^(n);λ).\displaystyle:=\Gamma_{*}^{\prime}(\hat{\bf a}^{(n)},\hat{\Gamma}^{(n)};\lambda).

The elevation z^(n)\hat{z}^{(n)} of the nn-th iteration Γ^(n)\hat{\Gamma}^{(n)} is updated to z^(n+1)=z^(n)+λ[κ1lnP(Γ|𝐚^(n))]/z\hat{z}^{(n+1)}=\hat{z}^{(n)}+\lambda\partial[\kappa^{-1}\ln P(\Gamma|\hat{\bf a}^{(n)})]/\partial z, and λ\lambda is the learning rate of this gradient descent. When iterations converge, z^(n)=z^(n+1)\hat{z}^{(n)}=\hat{z}^{(n+1)} is met, and the surface optimal Γ^=Γ^(n+1)\hat{\Gamma}=\hat{\Gamma}^{(n+1)} is the solution that satisfies lnP(Γ|𝐚^)/z=0\partial\ln P(\Gamma|\hat{\bf a})/\partial z=0 for 𝐚^=𝐚^(n)\hat{\bf a}=\hat{\bf a}^{(n)}.

6.3 Case study of the 2013 Balochistan earthquake

The 2013 Balochistan earthquake struck southwestern Pakistan on 24 September 2013. It is moment magnitude (Mw) 7.7 oblique strike-slip faulting, yet variations in the centroid moment tensor solutions and geological features (e.g. Avouac et al., 2014; Jolivet et al., 2014) may collectively suggest source complexities typified by fault nonplanarity. Its slip zone is the Hoshab fault situated in the transition between the Chaman fault system and the accretionary wedge (Avouac et al., 2014), so a wedge-slicing listric fault is likely; however, it was hard for multiple point source approaches to constrain whether the listric or non-listric fault is the preferred geometry for this earthquake (Jolivet et al., 2014). The PDTI detects major along-strike variations of potency tensors as well as minor along-dip variations, but three-dimensional geometry construction was beyond the reach of previous efforts (Shimizu et al., 2020, 2021). Enabling the PDTI to reconstruct three-dimensional geometry, we attempt to extract the listric geometry of this earthquake fault.

Data and data processing in this application follow those of Shimizu et al. (2020), which consist of observed vertical components of teleseismic P waveforms converted to velocities at 36 stations shown in Figs. 2(c) and S1 of Shimizu et al. (2021). The only update from the method and data of Shimizu et al. (2021) is the three-dimensional fault reconstruction of eq. (50).

Previous studies have shown that the solution updates could mostly finish with the first iteration (Shimizu et al., 2021), so it would be helpful to delve into the initial surface reconstruction. Figure 10 plots the surface reconstruction from the potency beachballs distributed over the reference plane (Shimizu et al., 2020), which is the same initial input as that of Shimizu et al. (2021). The reference plane is taken as a vertical plane through the hypocenter with a strike angle of 226 degrees, setting the southwest and the upward positive. Figure 10 employs λ1=0\lambda^{-1}=0 and uses eq. (35) as in the synthetic tests. Normal vectors evaluated from the PDTI provide expected fault slopes as an input set (Fig. 10a). The fault elevation field is then reconstructed as an output (Fig. 10b). Slopes of the reconstructed surface well explain the input slopes (Fig. 11). Residuals of nn-vectors are generally random, with some correlation with neighboring points due to smoothing imposed on the potency density estimate. The smooth surface reconstruction provides subfault elevations and subfault areas for horizontal fault subdivision, which convert the areal density of potency to distributed potency sources, updating Green’s function for the PDTI of the next iteration.

Refer to caption
Figure 10: Surface reconstruction from the potency distributed over a reference plane. (a) Informed fault slopes, calculated from the estimated nn-vector field via eq. (7). Collocation points of informed slopes correspond to the xx-yy locations of B-spline knots for potency densities. (b) Reconstructed fault elevation when using eq. (35).
Refer to caption
Refer to caption
Figure 11: Comparisons between reconstructed and informed slopes along strike (a) and dip (b). Reconstructed and informed slopes correspond to Fig. 10b and Fig. 10a, respectively. Their residuals are also indicated.

Iterative shape updates are shown in Fig. 12. The shape is quickly constrained at the center of large slips, and the convergence is slow at the edges of small slips, where the constraint is relatively weak. High learning rates vibrate poorly constrained regions (fault edges in Fig. 12a–d) while low learning rates slow down updates (Fig. 12i–l), as is often the case with gradient descents. The inference with adjusted learning rates converges to the optimal within a few iterations (Fig. 12e–h). The first surface reconstruction of relatively high learning rates (Fig. 12b) outlines a converged fault shape, as in a quasi-two-dimensional approach by Shimizu et al. (2021). Our estimate progressively approaches the surface rupture identified in the field survey (Zinke et al., 2014, Fig. 12 red line) in spite of not incorporating the surface trace. In Shimizu et al. (2021), the dip was fixed at 90 degrees so that their estimate (Fig. 12, blue line) shifted northward from the surface trace near 50 km southwest of the hypocenter, and it is resolved in the present inversion.

Refer to caption
Figure 12: Iterations of computed gradient descents. The first to fourth iterations are shown for the learning rate inverse λ1\lambda^{-1} of 0.01 (a–d), 0.03 (e–h), and 0.1 (i–l). The star shows the epicenter, and the background topography is from SRTMGL1 (NASA JPL, 2013). The surface rupture trace (Zinke et al., 2014, red) and two-dimensional geometry reconstructed in the previous PDTI (Shimizu et al., 2021, blue) are drawn for comparison.

We thus obtain the optimal solution as in Fig. 13 (λ1=0.03\lambda^{-1}=0.03, sixth iteration). The along-strike curve is steeper at shallower depths, and the dip angle becomes sharper at greater depths, so the reconstructed fault forms a bowl-shaped surface. It agrees with the geometry expected for a listric fault. Although the input nn-vectors corresponding to the B-spline knots of the PDTI have only four points along the dip, the reconstructed surface draws along-strike-varying along-dip variations. Such smooth strike-and-dip variations had already been captured by the beachball nodal planes without performing surface reconstruction (Fig. 11), but the three-dimensional surface reconstruction has made the interpretation more intuitive. Along-depth elevations of a few dozen kilometers are consistent with the constant-dip fault model of Jolivet et al. (2014) that is generated from distributed point source modeling. At the same time, like the other model of Jolivet et al. (2014) oriented to the geological setting of listric faults, the fault reconstructed in this study holds three-dimensional attributes.

Refer to caption
Figure 13: Our optimal source model (λ1=0.03\lambda^{-1}=0.03, the sixth iteration). Distributed potency tensors are shown in the birds-eye view with the epicenter indicated by the star (left), and the absolute potency densities are projected onto the reconstructed fault surface (right). The fault surface elevation is evaluated against a reference vertical plane with a 226-degree strike angle, the origin of which is the hypocenter. The moment rate function and the moment beachball obtained by the space-time integral of the potency-rate density tensors are also plotted.

Since the dip angle unnaturally flips near 150 km strike at depth (Fig. 9), the accuracy around the southwestern edge is doubtful. The non-slipping zone is suggested to be poorly constrained even with prior information on the smoothness of potency densities. It may be overcome by trimming source faults, for example, by using informatics to determine the minimum number of beachballs required to describe source faults (Thurin, 2024), and the same might apply to the decision of the number of faults. The elevation north of the epicenter is estimated systematically lower than the surface trace as in Shimizu et al. (2021), and it is likely due to smoothing out a near-hypocenter stepover visible in the surface trace. The use of surface rupture traces as an a priori constraint would bring fault shapes closer to the actual. The method validation is in progress as above, but anticipating that case studies ensue and reveal the workings of the fault reconstruction from potency, we submit the result of the simplest method (eq. 50) as a baseline solution.

7 Concluding remarks

We study surface reconstruction from the potency field, where a smooth surface is constructed from the estimated nn-vector field pointed by the nodal planes of estimated potency beachballs. The contribution of this work is the derivation of a formula that enables the three-dimensional surface reconstruction left behind by Shimizu et al. (2021). Our investigation reveals that the three-dimensional surface reconstruction is overdetermined, thus necessitating a stochastic approach. This transition is inevitable in the surface reconstruction from potency. The same could apply to surface reconstruction from distributed point sources, and our formulation is applicable to surface reconstruction from multiple coseismic focal mechanisms (Yue & Lay, 2020), aftershock hypocenters (Fuenzalida et al., 2013; Wang et al., 2021; Sawaki et al., 2025), and earthquake swarms (Yukutake et al., 2011; Ross et al., 2020). Given the above nature of surface reconstruction elucidated, we have formulated and solved the problem of stochastic surface reconstruction from a given nn-vector field, and our synthetic tests confirm that the solution of this problem reproduces the original surface from noisy nn-vector data. Even for narrow faults, three-dimensional probabilistic approaches can better reproduce the original shape than deterministic quasi-two-dimensional approaches. A preliminary analysis of the 2013 Balochistan earthquake captured along-dip shape variability that was not reconstructed by Shimizu et al. (2021). There remains a jump to actual applications, but at present, it works.

Acknowledgements.
The first author D.S. acknowledges valuable comments from Dr. Haruo Horikawa. This work was in part supported by JSPS KAKENHI Grant Numbers JP21H05206 and JP25K01084.

Data availability

Waveform data inverted in our application are downloaded from the Seismological Facility for the Advancement of Geoscience Data Management Center and processed in Shimizu et al. (2020). The ground topography in this paper refers to SRTMGL1 (NASA JPL, 2013).

References

  • Akaike (1980) Akaike, H., 1980. Likelihood and the Bayes procedure, in “Bayesian Statistics”, ed. by jm bernardo, mh degroot, dv lindley, and afm smith.
  • Aki (1979) Aki, K., 1979. Characterization of barriers on an earthquake fault, Journal of Geophysical Research: Solid Earth, 84(B11), 6140–6148.
  • Ando & Kaneko (2018) Ando, R. & Kaneko, Y., 2018. Dynamic rupture simulation reproduces spontaneous multifault rupture and arrest during the 2016 Mw 7.9 Kaikoura earthquake, Geophysical Research Letters, 45(23), 12–875.
  • Aspri et al. (2020) Aspri, A., Beretta, E., Mazzucato, A. L., & De Hoop, M. V., 2020. Analysis of a model of elastic dislocations in geophysics, Archive for Rational Mechanics and Analysis, 236, 71–111.
  • Aspri et al. (2025) Aspri, A., Beretta, E., Lee, A., & Mazzucato, A. L., 2025. A shape derivative algorithm for reconstructing elastic dislocations in geophysics, Research in the Mathematical Sciences, 12(2), 24.
  • Avouac et al. (2014) Avouac, J.-P., Ayoub, F., Wei, S., Ampuero, J.-P., Meng, L., Leprince, S., Jolivet, R., Duputel, Z., & Helmberger, D., 2014. The 2013, Mw 7.7 Balochistan earthquake, energetic strike-slip reactivation of a thrust fault, Earth and Planetary Science Letters, 391, 128–134.
  • Barka & Kadinsky-Cade (1988) Barka, A. & Kadinsky-Cade, K., 1988. Strike-slip fault geometry in Turkey and its influence on earthquake activity, Tectonics, 7(3), 663–684.
  • Beretta et al. (2008) Beretta, E., Francini, E., & Vessella, S., 2008. Determination of a linear crack in an elastic body from boundary measurements—lipschitz stability, SIAM journal on mathematical analysis, 40(3), 984–1002.
  • Bruhat et al. (2016) Bruhat, L., Fang, Z., & Dunham, E. M., 2016. Rupture complexity and the supershear transition on rough faults, Journal of Geophysical Research: Solid Earth, 121(1), 210–224.
  • Das & Aki (1977) Das, S. & Aki, K., 1977. Fault plane with barriers: A versatile earthquake model, Journal of geophysical research, 82(36), 5658–5670.
  • Diao et al. (2023) Diao, H., Liu, H., & Meng, Q., 2023. Dislocations with corners in an elastic body with applications to fault detection, arXiv preprint arXiv:2309.09706.
  • Dutta et al. (2021) Dutta, R., Jónsson, S., & Vasyura-Bathke, H., 2021. Simultaneous Bayesian estimation of non-planar fault geometry and spatially-variable slip, Journal of Geophysical Research: Solid Earth, 126(7), e2020JB020441.
  • Fan et al. (2022) Fan, W., Okuwaki, R., Barbour, A. J., Huang, Y., Lin, G., & Cochran, E. S., 2022. Fast rupture of the 2009 Mw 6.9 Canal de Ballenas earthquake in the Gulf of California dynamically triggers seismicity in California, Geophysical Journal International, 230(1), 528–541.
  • Fuenzalida et al. (2013) Fuenzalida, A., Schurr, B., Lancieri, M., Sobiesiak, M., & Madariaga, R., 2013. High-resolution relocation and mechanism of aftershocks of the 2007 tocopilla (chile) earthquake, Geophysical Journal International, 194(2), 1216–1228.
  • Fukahata & Wright (2008) Fukahata, Y. & Wright, T. J., 2008. A non-linear geodetic data inversion using abic for slip distribution on a fault with an unknown dip angle, Geophysical Journal International, 173(2), 353–364.
  • Howarth et al. (2021) Howarth, J. D., Barth, N. C., Fitzsimons, S. J., Richards-Dinger, K., Clark, K. J., Biasi, G. P., Cochran, U. A., Langridge, R. M., Berryman, K. R., & Sutherland, R., 2021. Spatiotemporal clustering of great earthquakes on a transform fault controlled by geometry, Nature Geoscience, 14(5), 314–320.
  • Inoue et al. (2025) Inoue, N., Yamaguchi, R., Yagi, Y., Okuwaki, R., Bogdan, E., & Tadapansawut, T., 2025. A multiple asymmetric bilateral rupture sequence derived from the peculiar tele-seismic P-waves of the 2025 Mandalay, Myanmar earthquake, Seismica, 4(1), doi: 10.26443/seismica.v4i1.1691.
  • Jolivet et al. (2014) Jolivet, R., Duputel, Z., Riel, B., Simons, M., Rivera, L., Minson, S., Zhang, H., Aivazis, M., Ayoub, F., Leprince, S., et al., 2014. The 2013 Mw 7.7 Balochistan earthquake: seismic potential of an accretionary wedge, Bulletin of the Seismological Society of America, 104(2), 1020–1030.
  • Kasahara & Yagi (2010) Kasahara, A. & Yagi, Y., 2010. Complex seismic source inversion method with the data covariance matrix: Application to the 2010 Haiti earthquake, in AGU Fall Meeting Abstracts, vol. 2010, pp. S43A–2024.
  • Kase & Day (2006) Kase, Y. & Day, S., 2006. Spontaneous rupture processes on a bending fault, Geophysical Research Letters, 33(10).
  • Kikuchi & Kanamori (1991) Kikuchi, M. & Kanamori, H., 1991. Inversion of complex body waves—III, Bulletin of the Seismological Society of America, 81(6), 2335–2350.
  • Kodaira et al. (2006) Kodaira, S., Hori, T., Ito, A., Miura, S., Fujie, G., Park, J.-O., Baba, T., Sakaguchi, H., & Kaneda, Y., 2006. A cause of rupture segmentation and synchronization in the Nankai trough revealed by seismic imaging and numerical simulation, Journal of Geophysical Research: Solid Earth, 111(B9).
  • Loveless & Meade (2010) Loveless, J. P. & Meade, B. J., 2010. Geodetic imaging of plate motions, slip rates, and partitioning of deformation in japan, Journal of Geophysical Research: Solid Earth, 115(B2).
  • Mallick & Meade (2025) Mallick, R. & Meade, B. J., 2025. Smooth slip is all you need: A singularity-free boundary element method for fault slip problems, Computers & Geosciences, 196, 105820.
  • Matsu’ura (1977a) Matsu’ura, M., 1977a. Inversion of geodetic data Part I. mathematical formulation, Journal of Physics of the Earth, 25(1), 69–90.
  • Matsu’ura (1977b) Matsu’ura, M., 1977b. Inversion of geodetic data. Part II. Optimal model of conjugate fault system for the 1927 Tango earthquake, Journal of Physics of the Earth, 25(3), 233–255.
  • Matsu’ura & Hasegawa (1987) Matsu’ura, M. & Hasegawa, Y., 1987. A maximum likelihood approach to nonlinear inversion under constraints, Physics of the Earth and planetary interiors, 47, 179–187.
  • Matsu’ura et al. (2019) Matsu’ura, M., Noda, A., & Terakawa, T., 2019. Physical interpretation of moment tensor and the energetics of shear faulting, Tectonophysics, 771, 228228.
  • NASA JPL (2013) NASA JPL, 2013. NASA Shuttle Radar Topography Mission Global 1 arc second [Data set], NASA EOSDIS Land Processes Distributed Active Archive Center, doi: 10.5067/MEaSUREs/SRTM/SRTMGL1.003.
  • Ohara et al. (2023) Ohara, K., Yagi, Y., Yamashita, S., Okuwaki, R., Hirano, S., & Fukahata, Y., 2023. Complex evolution of the 2016 Kaikoura earthquake revealed by teleseismic body waves, Progress in Earth and Planetary Science, 10(1), 35.
  • Okuwaki et al. (2020) Okuwaki, R., Hirano, S., Yagi, Y., & Shimizu, K., 2020. Inchworm-like source evolution through a geometrically complex fault fueled persistent supershear rupture during the 2018 Palu Indonesia earthquake, Earth and Planetary Science Letters, 547, 116449.
  • Ozawa et al. (2023) Ozawa, S., Ando, R., & Dunham, E. M., 2023. Quantifying the probability of rupture arrest at restraining and releasing bends using earthquake sequence simulations, Earth and Planetary Science Letters, 617, 118276.
  • Ragon et al. (2018) Ragon, T., Sladen, A., & Simons, M., 2018. Accounting for uncertain fault geometry in earthquake source inversions–i: theory and simplified application, Geophysical Journal International, 214(2), 1174–1190.
  • Romanet et al. (2020) Romanet, P., Sato, D. S., & Ando, R., 2020. Curvature, a mechanical link between the geometrical complexities of a fault: Application to bends, kinks and rough faults, Geophysical Journal International, 223(1), 211–232.
  • Romanet et al. (2024) Romanet, P., Saito, T., & Fukuyama, E., 2024. The mechanics of static non-planar faults in infinitesimal strain theory, Geophysical Journal International, 239(3), 1664–1693.
  • Ross et al. (2020) Ross, Z. E., Cochran, E. S., Trugman, D. T., & Smith, J. D., 2020. 3d fault architecture controls the dynamism of earthquake swarms, Science, 368(6497), 1357–1361.
  • Sato et al. (2022) Sato, D. S., Fukahata, Y., & Nozue, Y., 2022. Appropriate reduction of the posterior distribution in fully Bayesian inversions, Geophysical Journal International, 231(2), 950–981.
  • Sawaki et al. (2025) Sawaki, Y., Shiina, T., Sagae, K., Sato, Y., Horikawa, H., Miyakawa, A., Imanishi, K., & Uchide, T., 2025. Fault geometries of the 2024 mw 7.5 noto peninsula earthquake from hypocenter-based hierarchical clustering of point-cloud normal vectors, Journal of Geophysical Research: Solid Earth, 130(4), e2024JB030233.
  • Schroeder et al. (1998) Schroeder, W., Martin, K. M., & Lorensen, W. E., 1998. The visualization toolkit an object-oriented approach to 3D graphics, Prentice-Hall, Inc.
  • Shimizu et al. (2020) Shimizu, K., Yagi, Y., Okuwaki, R., & Fukahata, Y., 2020. Development of an inversion method to extract information on fault geometry from teleseismic data, Geophysical Journal International, 220(2), 1055–1065.
  • Shimizu et al. (2021) Shimizu, K., Yagi, Y., Okuwaki, R., & Fukahata, Y., 2021. Construction of fault geometry by finite-fault inversion of teleseismic data, Geophysical Journal International, 224(2), 1003–1014.
  • Sullivan & Kaszynski (2019) Sullivan, C. & Kaszynski, A., 2019. Pyvista: 3D plotting and mesh analysis through a streamlined interface for the Visualization Toolkit (VTK), Journal of Open Source Software, 4(37), 1450.
  • Tadapansawut et al. (2021) Tadapansawut, T., Okuwaki, R., Yagi, Y., & Yamashita, S., 2021. Rupture process of the 2020 Caribbean earthquake along the Oriente transform fault, involving supershear rupture and geometric complexity of fault, Geophysical Research Letters, 48(1), e2020GL090899.
  • Thurin (2024) Thurin, J., 2024. Sparse fault representation based on moment tensor interpolation, Seismica, 3(2).
  • Tsai et al. (2005) Tsai, V. C., Nettles, M., Ekström, G., & Dziewonski, A. M., 2005. Multiple cmt source analysis of the 2004 sumatra earthquake, Geophysical Research Letters, 32(17).
  • Volkov et al. (2017) Volkov, D., Voisin, C., & Ionescu, I. R., 2017. Reconstruction of faults in elastic half space from surface measurements, Inverse Problems, 33(5), 055018.
  • Wang et al. (2021) Wang, W., Fang, L., Wu, J., Tu, H., Chen, L., Lai, G., & Zhang, L., 2021. Aftershock sequence relocation of the 2021 ms 7.4 maduo earthquake, qinghai, china, Science China Earth Sciences, 64, 1371–1380.
  • Wei et al. (2023) Wei, G., Chen, K., & Meng, H., 2023. Bayesian inversion of finite-fault earthquake slip model using geodetic data, solving for non-planar fault geometry, variable slip, and data weighting, Journal of Geophysical Research: Solid Earth, 128(2), e2022JB025225.
  • Yabuki & Matsu’ura (1992) Yabuki, T. & Matsu’ura, M., 1992. Geodetic data inversion using a Bayesian information criterion for spatial distribution of fault slip, Geophysical Journal International, 109(2), 363–375.
  • Yagi & Fukahata (2010) Yagi, Y. & Fukahata, Y., 2010. Waveform inversion for seismic source processes with uncertainty of Green’s function, in AGU Fall Meeting Abstracts, vol. 2010, pp. S53C–1990.
  • Yagi & Fukahata (2011) Yagi, Y. & Fukahata, Y., 2011. Introduction of uncertainty of Green’s function into waveform inversion for seismic source processes, Geophysical Journal International, 186(2), 711–720.
  • Yagi et al. (2024) Yagi, Y., Okuwaki, R., Hirano, S., Enescu, B., Chikamori, M., & Yamaguchi, R., 2024. Barrier-induced rupture front disturbances during the 2023 Morocco earthquake, Seismological Research Letters, 95(3), 1591–1598.
  • Yue & Lay (2020) Yue, H. & Lay, T., 2020. Resolving complicated faulting process using multi-point-source representation: Iterative inversion algorithm improvement and application to recent complex earthquakes, Journal of Geophysical Research: Solid Earth, 125(2), e2019JB018601.
  • Yukutake et al. (2011) Yukutake, Y., Ito, H., Honda, R., Harada, M., Tanada, T., & Yoshida, A., 2011. Fluid-induced swarm earthquake sequence revealed by precisely determined hypocenters and focal mechanisms in the 2009 activity at hakone volcano, japan, Journal of Geophysical Research: Solid Earth, 116(B4).
  • Zinke et al. (2014) Zinke, R., Hollingsworth, J., & Dolan, J. F., 2014. Surface slip and off-fault deformation patterns in the 2013 Mw 7.7 Balochistan, P akistan earthquake: Implications for controls on the distribution of near-surface coseismic slip, Geochemistry, Geophysics, Geosystems, 15(12), 5034–5050.

Appendix A Estimation error propagation from potency to fault shape

The prior of shape P(Γ|𝐚)P(\Gamma|{\bf a}) shown in §3 and the posterior of potency P(𝐚|𝐝)P({\bf a}|{\bf d}) [P(𝐚|𝐝;𝝈2,𝝆2)P({\bf a}|{\bf d};\boldsymbol{\sigma}^{2},\boldsymbol{\rho}^{2}) in the text] set the marginal posterior of fault shape:

P(Γ|𝐝)=𝑑𝐚P(Γ|𝐚)P(𝐚|𝐝).P(\Gamma|{\bf d})=\int d{\bf a}P(\Gamma|{\bf a})P({\bf a}|{\bf d}). (A.1)

Here we show the procedure to translate P(𝐚|𝐝)P({\bf a}|{\bf d}) into P(Γ|𝐝)P(\Gamma|{\bf d}) when P(Γ|𝐚)P(\Gamma|{\bf a}) satisfies n𝒫nn\xrightarrow{\mathcal{P}}n_{*} (eq. 13).

For brevity, we introduce the (marginal) posterior of the nn-vector field {𝐧}\{{\bf n_{*}}\} associated with the potency density:

P({𝐧}|𝐝)=𝑑𝐚P(𝐚|𝐝)𝑑𝝃δ(𝐧(𝝃)𝐧(𝝃;𝐚)).P(\{{\bf n_{*}}\}|{\bf d})=\int d{\bf a}P({\bf a}|{\bf d})\int d\boldsymbol{\xi}\delta({\bf n}_{*}(\boldsymbol{\xi})-{\bf n}_{*}(\boldsymbol{\xi};{\bf a})). (A.2)

P({𝐧}|𝐝)P(\{{\bf n_{*}}\}|{\bf d}) and P(Γ|𝐚)P(\Gamma|{\bf a}) of eq. (14) rewrite the marginal posterior of the fault shape as

P(Γ|𝐝)=d{𝐧}P({𝐧}|𝐝)exp[κ𝑑𝝃𝐧𝐧]/𝒵,P(\Gamma|{\bf d})=\int d\{{\bf n_{*}}\}P(\{{\bf n_{*}}\}|{\bf d})\exp[\kappa\int d\boldsymbol{\xi}{\bf n}\cdot{\bf n_{*}}]/\mathcal{Z}, (A.3)

the right-hand side of which describes the transformation from the posterior P({𝐧}|𝐝)P(\{{\bf n_{*}}\}|{\bf d}) of the fault normal {𝐧}\{{\bf n}_{*}\} anticipated from potency into the posterior P(Γ|𝐝)P(\Gamma|{\bf d}) of fault shape. Especially when κ\kappa\to\infty, exp[κ𝑑𝝃𝐧𝐧]\exp[\kappa\int d\boldsymbol{\xi}{\bf n}\cdot{\bf n_{*}}] is extremely sharper than P({𝐧}|𝐝)P(\{{\bf n_{*}}\}|{\bf d}) and satisfies n𝒫nn\xrightarrow{\mathcal{P}}n_{*} (eq. 13), leading to

P(Γ|𝐝)=cd{𝐧}P({𝐧}|𝐝)𝑑𝝃δ(𝐧(𝝃)𝐧(𝝃)),P(\Gamma|{\bf d})=c\int d\{{\bf n_{*}}\}P(\{{\bf n_{*}}\}|{\bf d})\int d\boldsymbol{\xi}\delta({\bf n}(\boldsymbol{\xi})-{\bf n}_{*}(\boldsymbol{\xi})), (A.4)

where cc denotes normalization. Contrary to its appearance, eq. (A.4) does not conclude that the posterior P(Γ|𝐝)P(\Gamma|{\bf d}) of fault surface Γ\Gamma equals the posterior P({𝐧}|𝐝)P(\{{\bf n_{*}}\}|{\bf d}) of the potency-conforming nn-vector nn_{*}: P(Γ|𝐝)P({𝐧}|𝐝)|𝐧=𝐧.P(\Gamma|{\bf d})\neq P(\{{\bf n_{*}}\}|{\bf d})|_{{\bf n}_{*}={\bf n}}. The nn_{*}-vector field determined from potency can exceed the definition domain of the nn-vector field on a smooth surface Γ\Gamma bounded by the surface slope integrability (eq. 11) as indicated in §3, and therefore eq. (A.4) becomes

P(Γ|𝐝)=cP({𝐧}|𝐝)|𝐧=𝐧x(ny/nz)=y(nx/nz).P(\Gamma|{\bf d})=cP(\{{\bf n_{*}}\}|{\bf d})|_{{\bf n}_{*}={\bf n}\cap\partial_{x}(n_{y}/n_{z})=\partial_{y}(n_{x}/n_{z})}. (A.5)

Thus, searching for the optimum nn-vector field of a smooth surface from the posterior of potency is a conditional optimization constrained by x(ny/nz)=y(nx/nz)\partial_{x}(n_{y}/n_{z})=\partial_{y}(n_{x}/n_{z}). The inference of the optimum elevation automatically satisfies x(ny/nz)=y(nx/nz)\partial_{x}(n_{y}/n_{z})=\partial_{y}(n_{x}/n_{z}) and evades this complication as in §3. The surface reconstruction from this P(Γ|𝐝)P(\Gamma|{\bf d}) is generally a non-linear analysis based on eq. (37) relating surface normal 𝐧{\bf n} and surface slope z\nabla z. The evaluation of eq. (37) passes through the similar procedure to that shown in §4.2. Equation (A.5) also applies to P(Γ|𝐚)P(\Gamma|{\bf a}) other than eq. (14) when n𝒫nn\xrightarrow{\mathcal{P}}n_{*} (eq. 13) is met.

We must remark that the non-uniqueness of functional forms of P(Γ|𝐚)P(\Gamma|{\bf a}), which was an issue in surface reconstruction from given potency, does not matter in surface reconstruction from the posterior probability of potency P(Γ|𝐝)P(\Gamma|{\bf d}) (eq. A.5). It could answer why the overdetermination of surface reconstruction from potency has not posed an issue in slip inversions on a movable fault surface. After all, we must embrace either the subjectivity inherent in projecting observed inelastic strains onto displacement discontinuities or the subjectivity that forbids all but displacement discontinuities: P(𝐚,Γ|𝐝)=P(𝐚|𝐝)P(Γ|𝐚,𝐝)P(𝐚|𝐝)P(Γ|𝐚)P({\bf a},\Gamma|{\bf d})=P({\bf a}|{\bf d})P(\Gamma|{\bf a},{\bf d})\to P({\bf a}|{\bf d})P(\Gamma|{\bf a}) in the text or P(𝐚,Γ|𝐝)=P(𝐚|Γ,𝐝)P(Γ|𝐝)P({\bf a},\Gamma|{\bf d})=P({\bf a}|\Gamma,{\bf d})P(\Gamma|{\bf d}) equivalent to the slip inversions with a movable fault surface (§6.1).

Appendix B The exact solution of Γ\Gamma_{*} for finite (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}

The evaluation of Γ\Gamma_{*} in the text (§4.2) has assumed that the non-integrable component (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*} expected from the nn-vector nn_{*} of potency is small. Its motive is to avoid an issue of P(Γ|𝐚)P(\Gamma|{\bf a}) selection arising from finite (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}3, the last paragraph). For deeper considerations, we show below the way to evaluate Γ\Gamma_{*} for given P(Γ|𝐚)P(\Gamma|{\bf a}) for finite (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}. We treat eq. (14) with κ\kappa\to\infty as a concrete example, and similar results follow other forms of P(Γ|𝐚)P(\Gamma|{\bf a}).

Even in the dislocation limit n𝒫nn\xrightarrow{\mathcal{P}}n_{*} (eq. 13, κ\kappa\to\infty for eq. 14), the nn-vector of a surface is necessarily deviated from the expectation nn_{*} of potency due to the non-integrable components of (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}. What is expected in this limit is the smallness of the fluctuations from nn_{*} with non-integrable components removed, that is, the nn-vector of Γ\Gamma_{*} to be sought itself. This nn-vector of Γ\Gamma_{*}, the solution to be sought, is here denoted by nn_{*}^{\prime}:

limκ𝐧(𝝃)=𝐧(𝝃;Γ).\lim_{\kappa\to\infty}{\bf n}(\boldsymbol{\xi})={\bf n}_{*}^{\prime}(\boldsymbol{\xi};\Gamma). (B.1)

Therefore, if (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*} is finite, the fast convergent series in the dislocation limit is the Taylor series of 𝐧{\bf n} around 𝐧{\bf n}_{*}^{\prime}, not the Taylor series of 𝐧{\bf n} around 𝐧{\bf n}_{*} conducted in the text.

We now perform this corrected series expansion to account for (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*} components dropped in §4.2. By utilizing

(θϕ)(𝐧𝐞θ𝐞ϕ)=(𝐞θ𝐧𝟎𝐞ϕsinθ𝐞ϕcosθ𝐧sinθ𝐞θcosθ),\left(\begin{array}[]{c}\frac{\partial}{\partial\theta}\\ \frac{\partial}{\partial\phi}\end{array}\right)\left({\bf n}\,\,{\bf e}_{\theta}\,\,{\bf e}_{\phi}\right)=\left(\begin{array}[]{ccc}{\bf e}_{\theta}&-{\bf n}&{\bf 0}\\ {\bf e}_{\phi}\sin\theta&{\bf e}_{\phi}\cos\theta&-{\bf n}\sin\theta-{\bf e}_{\theta}\cos\theta\end{array}\right), (B.2)

we obtain the second-order representation of 𝐧{\bf n} around 𝐧{\bf n}_{*}^{\prime} with respect to the angle fluctuations δΩ\delta\Omega^{\prime} from 𝐧{\bf n}_{*}^{\prime}:

𝐧=[112(θθ)2sin2θ2(ϕϕ)2]𝐧\displaystyle{\bf n}=\left[1-\frac{1}{2}(\theta-\theta_{*}^{\prime})^{2}-\frac{\sin^{2}\theta_{*}^{\prime}}{2}(\phi-\phi_{*}^{\prime})^{2}\right]{\bf n}_{*}^{\prime} (B.3)
+[(θθ)12sinθcosθ(ϕϕ)2]𝐞θ\displaystyle+\left[(\theta-\theta_{*}^{\prime})-\frac{1}{2}\sin\theta_{*}^{\prime}\cos\theta_{*}^{\prime}(\phi-\phi_{*}^{\prime})^{2}\right]{\bf e}_{\theta}^{*\prime}
+[sinθ(ϕϕ)+cosθ(θθ)(ϕϕ)]𝐞ϕ,\displaystyle+\left[\sin\theta_{*}^{\prime}(\phi-\phi*^{\prime})+\cos\theta_{*}^{\prime}(\theta-\theta_{*}^{\prime})(\phi-\phi_{*}^{\prime})\right]{\bf e}_{\phi}^{*\prime},

where (θ,ϕ)(\theta_{*}^{\prime},\phi_{*}^{\prime}) denotes the position of 𝐧{\bf n}_{*}^{\prime} on a unit sphere, and 𝐞θ{\bf e}_{\theta}^{*\prime} and 𝐞ϕ{\bf e}_{\phi}^{*\prime} represent associated basis vectors; higher orders become infinitesimally smaller in the dislocation limit. Especially for P(Γ|𝐚)P(\Gamma|{\bf a}) of eq. (14) with κ\kappa\to\infty,

𝐧𝐧=\displaystyle{\bf n}_{*}\cdot{\bf n}= 𝐧𝐧(𝐧𝐧)𝐧\displaystyle{\bf n}_{*}^{\prime}\cdot{\bf n}-({\bf n}_{*}^{\prime}-{\bf n}_{*})\cdot{\bf n} (B.4)
=\displaystyle= [112(θθ)2sin2θ2(ϕϕ)2](𝐧𝐧)𝐧.\displaystyle\left[1-\frac{1}{2}(\theta-\theta_{*}^{\prime})^{2}-\frac{\sin^{2}\theta_{*}^{\prime}}{2}(\phi-\phi_{*}^{\prime})^{2}\right]-({\bf n}_{*}^{\prime}-{\bf n}_{*})\cdot{\bf n}. (B.5)

The first term replicates the right-hand side of eq. (26) with (θ,ϕ)(\theta_{*},\phi_{*}) substituted with (θ,ϕ)(\theta_{*}^{\prime},\phi_{*}^{\prime}) and thus functions similarly to those seen in the text. The second term is the correction arising from the deviation of 𝐧{\bf n}_{*}^{\prime} from 𝐧{\bf n}_{*}, explicitly written as

(𝐧𝐧)𝐧=[112(θθ)2sin2θ2(ϕϕ)2](1𝐧𝐧)\displaystyle({\bf n}_{*}^{\prime}-{\bf n}_{*})\cdot{\bf n}=\left[1-\frac{1}{2}(\theta-\theta_{*}^{\prime})^{2}-\frac{\sin^{2}\theta_{*}^{\prime}}{2}(\phi-\phi_{*}^{\prime})^{2}\right](1-{\bf n}_{*}\cdot{\bf n}_{*}^{\prime}) (B.6)
[(θθ)12sinθcosθ(ϕϕ)2](𝐧𝐞θ)\displaystyle-\left[(\theta-\theta_{*}^{\prime})-\frac{1}{2}\sin\theta_{*}^{\prime}\cos\theta_{*}^{\prime}(\phi-\phi_{*}^{\prime})^{2}\right]({\bf n}_{*}\cdot{\bf e}_{\theta}^{*\prime})
[sinθ(ϕϕ)+cosθ(θθ)(ϕϕ)](𝐧𝐞ϕ).\displaystyle-\left[\sin\theta_{*}^{\prime}(\phi-\phi*^{\prime})+\cos\theta_{*}^{\prime}(\theta-\theta_{*}^{\prime})(\phi-\phi_{*}^{\prime})\right]({\bf n}_{*}\cdot{\bf e}_{\phi}^{*\prime}).

Angle fluctuations θθ\theta-\theta_{*}^{\prime} and ϕϕ\phi-\phi_{*}^{\prime} are expressed by eqs. (31) and (32) in the text with a substitution (θ,ϕ)(θ,ϕ)(\theta_{*},\phi_{*})\to(\theta_{*}^{\prime},\phi_{*}^{\prime}):

θθ\displaystyle\theta-\theta_{*}^{\prime} =nz2(nxz,x+nyz,y+n/nz)\displaystyle=-n_{*z}^{\prime 2}(n_{*\parallel x}^{\prime}z_{,x}+n_{*\parallel y}^{\prime}z_{,y}+n_{*\perp}^{\prime}/n_{*z}^{\prime}) (B.7)
ϕϕ\displaystyle\phi-\phi_{*}^{\prime} =n1nz(nxz,ynyz,y).\displaystyle=-n_{*\perp}^{\prime-1}n_{*z}^{\prime}(n_{*\parallel x}^{\prime}z_{,y}-n_{*\parallel y}^{\prime}z_{,y}). (B.8)

The rest is unaltered from the text, and the final result is:

Γ=argmindΣ{[(θθ)2+sin2θ(ϕϕ)2](𝐧𝐧)\displaystyle\Gamma_{*}={\rm argmin}\int d\Sigma\{[(\theta-\theta_{*}^{\prime})^{2}+\sin^{2}\theta_{*}^{\prime}(\phi-\phi_{*}^{\prime})^{2}]({\bf n}_{*}\cdot{\bf n}_{*}^{\prime}) (B.9)
+[2(θθ)+sinθcosθ(ϕϕ)2](𝐧𝐞θ)\displaystyle+[-2(\theta-\theta_{*}^{\prime})+\sin\theta_{*}^{\prime}\cos\theta_{*}^{\prime}(\phi-\phi_{*}^{\prime})^{2}]({\bf n}_{*}\cdot{\bf e}_{\theta}^{*\prime})
2[sinθ(ϕϕ)+cosθ(θθ)(ϕϕ)](𝐧𝐞ϕ)}.\displaystyle-2[\sin\theta_{*}^{\prime}(\phi-\phi*^{\prime})+\cos\theta_{*}^{\prime}(\theta-\theta_{*}^{\prime})(\phi-\phi_{*}^{\prime})]({\bf n}_{*}\cdot{\bf e}_{\phi}^{*\prime})\}.

Equation (B.9) is a self-consistent equation, where the condition to determine Γ\Gamma_{*} contains the nn-vector of Γ\Gamma_{*} (𝐧{\bf n}_{*}^{\prime}) and the basis vectors (𝐞θ{\bf e}_{\theta}^{*\prime} and 𝐞ϕ{\bf e}_{\phi}^{*\prime}) at the spherical position (θ,ϕ)(\theta_{*}^{\prime},\phi_{*}^{\prime}) of that nn-vector. For small (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}, 𝐧𝐧{\bf n}_{*}^{\prime}\simeq{\bf n}_{*} holds, and the correction becomes negligible, resulting in eq. (35) of the main text. Although the above correction is not accounted for in the main text supposing small (xyyx)z(\partial_{x}\partial_{y}-\partial_{y}\partial_{x})z_{*}, the correction may be iteratively computed with the initialization of 𝐧𝐧{\bf n}_{*}^{\prime}\simeq{\bf n}_{*}. The correction for P(Γ|𝐚)P(\Gamma|{\bf a}) other than eq. (14) is similarly evaluated using eqs. (B.3), (B.7), and (B.8).