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

Poroelastic near-field inverse scattering

Fatemeh Pourahmadian1,2 Kevish Napal1 1 Department of Civil, Environmental & Architectural Engineering, University of Colorado Boulder, USA 2 Department of Applied Mathematics, University of Colorado Boulder, USA
Abstract

A multiphysics data analytic platform is established for imaging poroelastic interfaces of finite permeability (e.g., hydraulic fractures) from elastic waveforms and/or acoustic pore pressure measurements. This is accomplished through recent advances in design of non-iterative sampling methods to inverse scattering. The direct problem is formulated via the Biot equations in the frequency domain where a network of discontinuities is illuminated by a set of total body forces and fluid volumetric sources, while monitoring the induced (acoustic and elastic) scattered waves in an arbitrary near-field configuration. A thin-layer approximation is deployed to capture the rough and multiphase nature of interfaces whose spatially varying hydro-mechanical properties are a priori unknown. In this setting, the well-posedness analysis of the forward problem yields the admissibility conditions for the contact parameters. In light of which, the poroelastic scattering operator and its first and second factorizations are introduced and their mathematical properties are carefully examined. It is shown that the non-selfadjoint nature of the Biot system leads to an intrinsically asymmetric factorization of the scattering operator which may be symmetrized at certain limits. These results furnish a robust framework for systematic design of regularized and convex cost functionals whose minimizers underpin the multiphysics imaging indicators. The proposed solution is synthetically implemented with application to spatiotemporal reconstruction of hydraulic fracture networks via deep-well measurements.

keywords:
poroelastic waves, waveform tomography, hydraulic fractures, multiphysics sensing

1 Introduction

Coupled physics processes driven by engineered stimulation of the subsurface underlie many emerging technologies germane to advanced energy infrastructure such as unconventional hydrocarbon and geothermal reservoirs [1, 2]. Optimal design and closed-loop control of such systems require real-time feedback on the nature of progressive variations in a target region [3, 4]. 3D in-situ tracking of multiphasic developments however is exceptionally challenging [5, 6]. Engineered treatments (such as fluid or gas injection) often occur in a complex background whose structure and hydro-mechanical properties are uncertain [7, 8]. Nevertheless, state-of-the-art solutions to in-situ characterization mostly rely on elastic waves [9, 10, 11, 4]. In particular, micro-seismic waves generated during shear propagation of fractures are commonly used to monitor evolving discontinuities in rock [12, 13]. Such passive schemes are nonetheless agnostic to aseismic evolution [14], and involve significant assumptions on the nature of wave motion in the subsurface which may lead to critical errors [15, 16, 17]. On the other hand, existing approaches to full waveform inversion [18, 19, 11], while offering more comprehensive solutions, are by and large computationally expensive and thus inapplicable for real-time sensing. Therefore, there exists a critical need for the next generation of imaging technologies that transcend some of these limitations.

In this vein, recent advances in rigorous design of sampling-based waveform tomography solutions [20, 21, 22, 23, 24] seem particularly relevant owing to their robustness and newfound capabilities pertinent to imaging in complex and unknown domains [21, 23, 25, 26]. So far, these developments mostly reside in the context of self-adjoint and energy preserving systems such as Helmholtz and Navier equations. Wave attenuation, however, is a defining feature of the fractured rock systems with crucial impact on subsurface exploration [27, 28, 29]. In this context, a systematic analysis of sound absorption – e.g., due to the fluid-solid interaction, and its implications on sampling-based data inversion is still lacking. This, in part, may be due to the sparse and single-physics nature of traditional data. Fast-paced progress in sensing instruments such as fiber optic (FBG) sensors – capturing high-resolution multiple physical measurements in time-space [12, 30], may bridge this gap and enable a holistic approach to subsurface characterization. This invites new data analytic tools capable of (a) expedited processing of large datasets, and (b) simultaneous inversion of multiphysics observations. The latter may particularly assist a high-fidelity reconstruction of the sought-for subterranean events.

In this paper, the theoretical foundation of poroelastic inverse scattering is established with application to spatiotemporal tracking of hydraulic fracture networks in unconventional energy systems. This method is nested within the framework of active sensing where the process zone is sequentially illuminated (during stimulation) via deep-well excitations prompting poroelastic scattering whose signature is recorded in the form of seismic and pore pressure waveforms. Thus-obtained multiphase sensory data are deployed for geometric reconstruction of treatment-induced evolution in the target region. Rooted in rigorous foundations of the inverse scattering theory [24], the proposed imaging solution carries a high spatial resolution with carefully controlled sensitivity to noise and illumination frequency.

The forward problem is posed in a suitable dimensional platform catering for simultaneous elastic and acoustic data inversion. A system of arbitrary-shaped hydraulic fractures with non-trivial (i.e., heterogeneous, dissipative, and finitely permeable) interfaces is illuminated by a plausible combination of total body forces and fluid volumetric sources in a medium governed by the coupled Biot equations [31, 32]. Thus-induced multiphase scattered fields are used to define the poroelastic scattering operator mapping the source densities to near-field measurements. In this setting, the reconstruction scheme is based on (i) a custom factorization of the near-field operator, and (b) a sequence of approximate solutions to the scattering equation, seeking (fluid and total-force) source densities whose affiliated waveform pattern matches that of a designated poroelastodynamics solution emanating from a sampling point. The latter is obtained via a sequence of carefully constructed cost functionals whose minimizers can be computed without iterations. Such minimizing solutions are then used to construct a robust multiphysics imaging indicator whose performance is examined through a set of numerical experiments.

In what follows, Section 2 introduces the direct scattering problem and the necessary tools for the inverse analysis. In particular, Section 2.4 investigates the well-posedness of the forward problem which leads to finding the admissibility conditions for the interfacial contact parameters including the elastic stiffness matrix and the permeability, effective stress and Skempton coefficients. This criteria plays a key role later in Section 3 where the essential properties of the poroelastic scattering operator and its components, in the first and second factorizations, are established. Section 4 presents the regularized cost functionals for solving the scattering equations, main theorems, and imaging functionals. Section 5 illustrates a set of numerical experiments implementing the reconstruction of (stationary and evolving) hydraulic fracture networks via the near-field measurements by way of the proposed imaging indicators.

2 Problem statement

Consider the ball \mathscr{B}^{\prime} of radius R>0R\!>\!0, centered at the origin in the subsurface Ω3\Omega\subset\mathbb{R}^{3}, encompassing the stimulation zone and sensing grid 𝒢\mathscr{G} as illustrated in Fig. 1. Let RR be sufficiently large so that the probing waves may be assumed negligible in Ω\Omega\setminus\mathscr{B}^{\prime} due to poroelastic attenuation. It is also assumed that \mathscr{B}^{\prime} does not meet Ω\partial\Omega, which in hydraulic fracturing is a reasonable premise, in particular, since a treatment region is typically four to ten kilometers below the surface while R=O(102)R=O(10^{2})[33]. The domain Ω\mathscr{B}^{\prime}\subset\Omega is described by drained Lamé coefficients λ\lambda^{\prime} and μ\mu^{\prime}, Biot modulus MM^{\prime}, total density ρ\rho^{\prime}, fluid density ρf\rho_{f}^{\prime}, apparent mass density ρa\rho_{a}^{\prime}, permeability coefficient κ\kappa^{\prime}, porosity ϕ\phi, and Biot effective stress coefficient α\alpha. A system of pre-existing and stimulation-induced fractures and flow networks Γ\Gamma^{\prime} is embedded in \mathscr{B}^{\prime} whose rough and multiphasic interfaces are characterized by a set of heterogeneous parameters, namely: the stiffness matrix 𝑲(𝝃)\boldsymbol{K}^{\prime}(\boldsymbol{\xi}^{\prime}), permeability coefficient ϰf(𝝃)\varkappa_{\textrm{\tiny f}}^{\prime}(\boldsymbol{\xi}^{\prime}), effective stress coefficient αf(𝝃)\alpha_{\textrm{\tiny f}}(\boldsymbol{\xi}^{\prime}), Skempton coefficient βf(𝝃)\beta_{\textrm{\tiny f}}(\boldsymbol{\xi}^{\prime}), and fluid pressure dissipation factor Π(𝝃)\Pi(\boldsymbol{\xi}^{\prime}) on 𝝃Γ\boldsymbol{\xi}^{\prime}\in\Gamma^{\prime}. The domain is excited by time harmonic total body forces of density 𝒈s(𝝃,ω){\boldsymbol{g}}^{\prime}_{s}(\boldsymbol{\xi}^{\prime},\omega^{\prime}) and fluid volumetric sources of density gf(𝝃,ω)g^{\prime}_{f}(\boldsymbol{\xi}^{\prime},\omega^{\prime}) on 𝝃𝒢\boldsymbol{\xi}^{\prime}\in\mathscr{G} at frequency ω\omega^{\prime}. This gives rise to poroelastic scattering observed on the sensing grid 𝒢\mathscr{G} in terms of solid displacements 𝒖(𝝃,ω)\boldsymbol{u}^{\prime}(\boldsymbol{\xi}^{\prime},\omega^{\prime}) and pore pressure p(𝝃,ω)p^{\prime}(\boldsymbol{\xi}^{\prime},\omega^{\prime}).

Refer to caption
Figure 1: Near-field illumination of a fracture network in the stimulation zone Ω\mathscr{B}^{\prime}\subset\Omega featuring disjoint multiphasic discontinuities Γ\Gamma^{\prime}\! characterized by heterogeneous elasticity 𝑲(𝝃)\boldsymbol{K}^{\prime}(\boldsymbol{\xi}^{\prime}) and permeability ϰf(𝝃)\varkappa_{\textrm{\tiny f}}^{\prime}(\boldsymbol{\xi}^{\prime}). Poroelastic incidents are induced via total body forces and fluid volumetric sources of respective densities 𝒈s(𝝃,ω){\boldsymbol{g}}^{\prime}_{s}(\boldsymbol{\xi}^{\prime},\omega^{\prime}) and gf(𝝃,ω)g^{\prime}_{f}(\boldsymbol{\xi}^{\prime},\omega^{\prime}) on a designated grid 𝝃𝒢\boldsymbol{\xi}^{\prime}\in\mathscr{G} at frequency ω\omega^{\prime}, which give rise to the scattered fields captured on 𝒢\mathscr{G} in the form of pore pressure pp^{\prime} and solid displacements 𝒖\boldsymbol{u}^{\prime}\!.

2.1 Dimensional platform

To enable multiphasic data inversion, the reference scales

μr=λ,ρr=ρ,r= 2πμ(1ϕ)ρsω2,\mu_{r}\,=\,\lambda^{\prime},\quad\rho_{r}\,=\,\rho^{\prime},\quad\ell_{r}\,=\,2\pi\sqrt{\dfrac{\mu^{\prime}}{(1-\phi){\rho}\hskip 0.56905pt^{\prime}_{s}{\omega^{\prime}}^{2}}},\vspace{-0mm} (1)

are respectively defined for stress, mass density, and length. Here, ρs\rho_{s}^{\prime} denotes the solid-phase density. Note that r\ell_{r} represents the drained shear wavelength. In this setting, all physical quantities are rendered dimensionless [34] as follows

(𝝃,ω)(𝝃,ω):=(1r𝝃,ρrμrrω),\displaystyle(\boldsymbol{\xi}^{\prime},\omega^{\prime})~{}\rightarrow~{}(\boldsymbol{\xi},\omega):=\big{(}\dfrac{1}{\ell_{r}}\boldsymbol{\xi}^{\prime},\sqrt{\dfrac{\rho_{r}}{\mu_{r}}}\ell_{r}\omega^{\prime}\big{)},
(λ,μ,M,ρ,ρf,ρa,κ,ϕ,α)\displaystyle\mathscr{B}^{\prime}(\lambda^{\prime},\mu^{\prime},M^{\prime},\rho^{\prime},\rho_{f}^{\prime},\rho_{a}^{\prime},\kappa^{\prime},\phi,\alpha)~{}\rightarrow~{} (2)
(λ,μ,M,ρ,ρf,ρa,κ,ϕ,α):=(λμr,μμr,Mμr,ρρr,ρfρr,ρaρr,μrρrrκ,ϕ,α),\displaystyle\mathscr{B}(\lambda,\mu,M,\rho,\rho_{f},\rho_{a},\kappa,\phi,\alpha):=\mathscr{B}(\dfrac{\lambda^{\prime}}{\mu_{r}},\dfrac{\mu^{\prime}}{\mu_{r}},\dfrac{M^{\prime}}{\mu_{r}},\dfrac{\rho^{\prime}}{\rho_{r}},\dfrac{\rho_{f}^{\prime}}{\rho_{r}},\dfrac{\rho_{a}^{\prime}}{\rho_{r}},\dfrac{\sqrt{\mu_{r}\rho_{r}}}{\ell_{r}}\kappa^{\prime},\phi,\alpha),
Γ(𝑲,ϰf,αf,βf,Π)Γ(𝑲,ϰf,αf,βf,Π):=Γ(rμr𝑲,μrρrϰf,αf,βf,Π),\displaystyle{\Gamma}^{\prime}(\boldsymbol{K}^{\prime},\varkappa_{\textrm{\tiny f}}^{\prime},\alpha_{\textrm{\tiny f}},\beta_{\textrm{\tiny f}},\Pi)~{}\rightarrow~{}{\Gamma}(\boldsymbol{K},\varkappa_{\textrm{\tiny f}},\alpha_{\textrm{\tiny f}},\beta_{\textrm{\tiny f}},\Pi):={\Gamma}(\dfrac{\ell_{r}}{\mu_{r}}\boldsymbol{K}^{\prime},\sqrt{\mu_{r}\rho_{r}}\varkappa_{\textrm{\tiny f}}^{\prime},\alpha_{\textrm{\tiny f}},\beta_{\textrm{\tiny f}},\Pi),
(𝒈s,gf)(𝒈s,gf):=(rμr𝒈s,gf),\displaystyle({\boldsymbol{g}}_{s}^{\prime},g_{f}^{\prime})~{}\rightarrow~{}({\boldsymbol{g}}_{s},g_{f}):=\big{(}\dfrac{\ell_{r}}{\mu_{r}}{\boldsymbol{g}}_{s}^{\prime},g_{f}^{\prime}\big{)},
(𝒖,p)(𝒖,p):=(𝒖r,pμr).\displaystyle(\boldsymbol{u}^{\prime},p^{\prime})~{}\rightarrow~{}(\boldsymbol{u},p):=\big{(}\dfrac{\boldsymbol{u}^{\prime}}{\ell_{r}},\dfrac{p^{\prime}}{\mu_{r}}\big{)}.

2.2 Field equations

The incident field (𝒖ι,pι)Hloc1(3𝒢¯)3×Hloc1(3𝒢¯)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\!\setminus\!\overline{\mathscr{G}})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\!\setminus\!\overline{\mathscr{G}}) generated by (𝒈s,gf)L2(𝒢)3×L2(𝒢)({\boldsymbol{g}}_{s},g_{f})\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) in the background is governed by

(𝑪:𝒖ι)(𝝃)(αρfγ)pι(𝝃)+ω2(ρρf2γ)𝒖ι(𝝃)=𝒢δ(𝒚𝝃)𝒈s(𝒚)d𝒚,\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}^{\tiny\iota})(\boldsymbol{\xi})\,-\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla p^{\tiny\iota}(\boldsymbol{\xi})\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\gamma})\boldsymbol{u}^{\tiny\iota}(\boldsymbol{\xi})\,=\,-\int_{\mathscr{G}}\delta(\boldsymbol{y}-\boldsymbol{\xi}){\boldsymbol{g}}_{s}(\boldsymbol{y})\hskip 1.13809pt\textrm{d}{\boldsymbol{y}},\,\, 𝝃3𝒢¯,\displaystyle\boldsymbol{\xi}\in{\mathbb{R}^{3}}\!\setminus\!\overline{\mathscr{G}}, (3)
1γω22pι(𝝃)+M1pι(𝝃)+(αρfγ)𝒖ι(𝝃)=𝒢δ(𝒚𝝃)gf(𝒚)d𝒚,\displaystyle\dfrac{1}{\gamma\omega^{2}}\nabla^{2}p^{\tiny\iota}(\boldsymbol{\xi})\,+\,{M}^{-1}p^{\tiny\iota}(\boldsymbol{\xi})\,+\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\tiny\iota}(\boldsymbol{\xi})\,=\,-\int_{\mathscr{G}}\delta(\boldsymbol{y}-\boldsymbol{\xi})g_{f}(\boldsymbol{y})\hskip 1.13809pt\textrm{d}{\boldsymbol{y}}, 𝝃3𝒢¯,\displaystyle\boldsymbol{\xi}\in{\mathbb{R}^{3}}\!\setminus\!\overline{\mathscr{G}},

where 𝑪=λ𝑰2𝑰2+2μ𝑰4\boldsymbol{C}=\lambda\hskip 1.13809pt\boldsymbol{I}_{2}\!\otimes\boldsymbol{I}_{2}+2\mu\hskip 1.13809pt\boldsymbol{I}_{4} is the fourth-order drained elasticity tensor with 𝑰m(m=2,4)\boldsymbol{I}_{m}\,(m\!=\!2,4) denoting the mmth-order symmetric identity tensor, and

γ=ρaϕ2+ρfϕ+iωκ.\gamma~{}=~{}\frac{\rho_{a}}{\phi^{2}}\,+\,\frac{\rho_{f}}{\phi}\,+\,\frac{\textrm{i}}{\omega\kappa}.\vspace{-1mm} (4)

The interaction of (𝒖ι,pι)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota}) with Γ\Gamma gives rise to the scattered field (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}\!\setminus\!\overline{\Gamma}})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}\!\setminus\!\overline{\Gamma}}) solving

(𝑪:𝒖)(αρfγ)p+ω2(ρρf2γ)𝒖= 0\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u})\,-\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla p\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\gamma})\boldsymbol{u}\,=\,\boldsymbol{0}\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}\!\setminus\!\overline{\Gamma}}, (5)
1γω22p+M1p+(αρfγ)𝒖= 0\displaystyle\dfrac{1}{\gamma\omega^{2}}\nabla^{2}p\,+\,{M}^{-1}p\,+\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}\,=\,0\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}\!\setminus\!\overline{\Gamma}},
𝒕+α~fp𝒏=𝑲𝒖𝒕ια~fpι𝒏\displaystyle{\boldsymbol{t}}\,+\,\tilde{\alpha}_{\textrm{\tiny f}}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}p\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\boldsymbol{n}\,=\,\boldsymbol{K}\llbracket\boldsymbol{u}\rrbracket\,-\,{\boldsymbol{t}}^{\tiny\iota}\,-\,\tilde{\alpha}_{\textrm{\tiny f}}\hskip 1.13809ptp^{\tiny\iota}\boldsymbol{n}\quad on Γ,\displaystyle\,\,\,{\Gamma},
p+βf𝒕𝒏=knβfΠαfqpιβf𝒕ι𝒏\displaystyle\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}p\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\,+\,{\beta}_{\textrm{\tiny f}}\hskip 1.13809pt{\boldsymbol{t}}\!\cdot\!\boldsymbol{n}\,=\,-\dfrac{\textrm{k}_{n}{\beta}_{\textrm{\tiny f}}}{\Pi\alpha_{\textrm{\tiny f}}}\llbracket q\rrbracket\,-\,p^{\tiny\iota}\,-\,{\beta}_{\textrm{\tiny f}}\hskip 1.13809pt{\boldsymbol{t}}^{\tiny\iota}\!\cdot\!\boldsymbol{n}\quad on Γ,\displaystyle\,\,\,{\Gamma},
q=ϰfiωΠpqι,𝒕=𝟎\displaystyle\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}q\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\,=\,\dfrac{\varkappa_{\textrm{\tiny f}}}{\textrm{i}\omega\Pi}\llbracket p\rrbracket\,-\,q^{\tiny\iota},\quad\llbracket{\boldsymbol{t}}\rrbracket=\boldsymbol{0}\quad on Γ,\displaystyle\,\,\,{\Gamma},

where (𝒕,q)H1/2(Γ)3×H1/2(Γ)({\boldsymbol{t}},q)\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma) and (𝒕ι,qι)H1/2(Γ)3×H1/2(Γ)({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota})\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma) are defined by

[𝒕,q](𝒖,p):=[𝒏𝑪:𝒖αp𝒏,1γω2(p𝒏ρfω2𝒖𝒏)]\displaystyle[{\boldsymbol{t}},q](\boldsymbol{u},p)\,:=\,\big{[}\boldsymbol{n}\cdot\boldsymbol{C}\colon\!\nabla\boldsymbol{u}-\alpha p\hskip 1.13809pt\boldsymbol{n},\,\dfrac{1}{\gamma\omega^{2}}(\nabla p\!\cdot\!\boldsymbol{n}-\rho_{f}\omega^{2}\boldsymbol{u}\!\cdot\!\boldsymbol{n})\big{]}\,\,\, onΓ,\displaystyle\text{on}\,\,\,{\Gamma}, (6)
[𝒕ι,qι](𝒖ι,pι):=[𝒏𝑪:𝒖ιαpι𝒏,1γω2(pι𝒏ρfω2𝒖ι𝒏)]\displaystyle[{\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota}](\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})\,:=\,\big{[}\boldsymbol{n}\cdot\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota}-\alpha p^{\tiny\iota}\boldsymbol{n},\,\dfrac{1}{\gamma\omega^{2}}(\nabla p^{\tiny\iota}\!\cdot\!\boldsymbol{n}-\rho_{f}\omega^{2}\boldsymbol{u}^{\tiny\iota}\!\cdot\!\boldsymbol{n})\big{]}\,\,\, onΓ;\displaystyle\text{on}\,\,\,{\Gamma};

the unit normal 𝒏=𝒏\boldsymbol{n}=\boldsymbol{n}^{-} on Γ\Gamma is explicitly identified in Section 2.3;

α~f:=αfΠ1αfβf(1Π);\tilde{\alpha}_{\textrm{\tiny f}}\,:=\,\dfrac{\alpha_{\textrm{\tiny f}}\Pi}{1-\alpha_{\textrm{\tiny f}}\beta_{\textrm{\tiny f}}(1-\Pi)};

(𝒖,p,q)=(𝒖+𝒖,p+p,q+q)\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket p\rrbracket,\llbracket q\rrbracket\big{)}=\big{(}\boldsymbol{u}^{+}\!-\boldsymbol{u}^{-},p^{+}\!-p^{-},q^{+}\!-q^{-}\big{)} signifies the jump in (𝒖,p,q)(\boldsymbol{u},p,q) across Γ\Gamma, while

(𝒖,p,q)=12(𝒖++𝒖,p++p,q++q)\big{(}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\boldsymbol{u}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{p}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{q}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\big{)}=\dfrac{1}{2}\big{(}\boldsymbol{u}^{+}\!+\boldsymbol{u}^{-},p^{+}\!+p^{-},q^{+}\!+q^{-}\big{)}

is the respective mean fields on Γ\Gamma\,; 𝑲=𝑲(𝝃)\boldsymbol{K}=\boldsymbol{K}(\boldsymbol{\xi}), according to [35], is a symmetric and possibly complex-valued matrix of specific stiffnesses which in the fracture’s local coordinates (𝒆1,𝒆2,𝒏)(𝝃)(\boldsymbol{e}_{1},\boldsymbol{e}_{2},\boldsymbol{n})(\boldsymbol{\xi}) may be described as the following

𝑲:=kt𝒆1𝒆1+kt𝒆2𝒆2+α~fknαfΠ𝒏𝒏onΓ,\boldsymbol{K}\,:=\,\textrm{k}_{t}\hskip 1.13809pt\boldsymbol{e}_{1}\!\otimes\boldsymbol{e}_{1}\,+\,\textrm{k}_{t}\hskip 1.13809pt\boldsymbol{e}_{2}\!\otimes\boldsymbol{e}_{2}\,+\,\dfrac{\tilde{\alpha}_{\textrm{\tiny f}}\textrm{k}_{n}}{{\alpha}_{\textrm{\tiny f}}\Pi}\hskip 1.13809pt\boldsymbol{n}\!\otimes\boldsymbol{n}\quad\text{on}\,\,\,{\Gamma}, (7)

wherein kt(𝝃)\textrm{k}_{t}(\boldsymbol{\xi}) and kn(𝝃)\textrm{k}_{n}(\boldsymbol{\xi}) are known respectively as the tangential and normal drained elastic stiffnesses. It should be mentioned that the contact condition across Γ\Gamma makes use of the generalized Schoenberg’s model for a poroelatic interface of finite permeability [35].

Remark 2.1.

The contact law at high-permeability interfaces [35] is reduced to the following form

𝒕+αfp𝒏=𝑲u𝒕ιαfpι𝒏\displaystyle{\boldsymbol{t}}\,+\,{\alpha}_{\textrm{\tiny\emph{f}}}\hskip 1.13809ptp\boldsymbol{n}\,=\,\boldsymbol{K}_{\infty}\llbracket u\rrbracket\,-\,{\boldsymbol{t}}^{\tiny\iota}\,-\,{\alpha}_{\textrm{\tiny\emph{f}}}\hskip 1.13809ptp^{\tiny\iota}\boldsymbol{n}\quad on Γ,\displaystyle\,\,\,{\Gamma},
p+βf𝒕𝒏=knβfαfqpιβf𝒕ι𝒏\displaystyle p\,+\,{\beta}_{\textrm{\tiny\emph{f}}}\hskip 1.13809pt{\boldsymbol{t}}\!\cdot\!\boldsymbol{n}\,=\,-\textrm{\emph{k}}_{n}\dfrac{{\beta}_{\textrm{\tiny\emph{f}}}}{\alpha_{\textrm{\tiny\emph{f}}}}\llbracket q\rrbracket\,-\,p^{\tiny\iota}\,-\,{\beta}_{\textrm{\tiny\emph{f}}}\hskip 1.13809pt{\boldsymbol{t}}^{\tiny\iota}\!\cdot\!\boldsymbol{n}\quad on Γ,\displaystyle\,\,\,{\Gamma},
p= 0,𝒕=𝟎\displaystyle\llbracket p\rrbracket\,=\,0,\,\,\,\,\llbracket{\boldsymbol{t}}\rrbracket=\boldsymbol{0}\quad on Γ.\displaystyle\,\,\,{\Gamma}.

where 𝐊:=kt𝐞1𝐞1+kt𝐞2𝐞2+kn𝐧𝐧\boldsymbol{K}_{\infty}:=\textrm{\emph{k}}_{t}\hskip 1.13809pt\boldsymbol{e}_{1}\!\otimes\boldsymbol{e}_{1}\hskip 1.13809pt+\hskip 1.13809pt\textrm{\emph{k}}_{t}\hskip 1.13809pt\boldsymbol{e}_{2}\!\otimes\boldsymbol{e}_{2}\hskip 1.13809pt+\hskip 1.13809pt\textrm{\emph{k}}_{n}\hskip 1.13809pt\boldsymbol{n}\!\otimes\boldsymbol{n} on Γ{\Gamma}.

The formulation of the direct scattering problems may now be completed by requiring that (𝒖,p)(\boldsymbol{u},p) and (𝒖ι,pι)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota}) satisfy the radiation condition as |𝝃||\boldsymbol{\xi}|\rightarrow\infty [36]. On uniquely decomposing the incident and scattered fields into two irrotational parts (p1{\textrm{p}_{1}}, p2{\textrm{p}_{2}}) and a solenoidal part (s) as

(𝒖ι,pι)=(𝒖p1ι,pp1ι)+(𝒖p2ι,pp2ι)+(𝒖sι,psι),(𝒖,p)=(𝒖p1,pp1)+(𝒖p2,pp2)+(𝒖s,ps),(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=(\boldsymbol{u}^{\tiny\iota}_{\textrm{p}_{1}\!},p^{\tiny\iota}_{\textrm{p}_{1}})+(\boldsymbol{u}^{\tiny\iota}_{\textrm{p}_{2}\!},p^{\tiny\iota}_{\textrm{p}_{2}})+(\boldsymbol{u}^{\tiny\iota}_{\textrm{s}},p^{\tiny\iota}_{\textrm{s}}),\quad(\boldsymbol{u},p)=(\boldsymbol{u}_{\textrm{p}_{1}\!},p_{\textrm{p}_{1}})+(\boldsymbol{u}_{\textrm{p}_{2}\!},p_{\textrm{p}_{2}})+(\boldsymbol{u}_{\textrm{s}},p_{\textrm{s}}), (8)

the radiation condition can be stated as

uςrikςuς=o(r1eaςr)asr:=|𝝃|,aς:=(kς),ς=p1,p2,s,u=𝒖ι,𝒖,\displaystyle\frac{\partial{\textrm{\bf u}}_{\varsigma}}{\partial r}-\text{i}k_{\varsigma}{\textrm{\bf u}}_{\varsigma}=o\big{(}r^{-1}e^{-a_{\varsigma}r}\big{)}\quad\text{as}~{}~{}r:=|\boldsymbol{\xi}|\to\infty,\quad a_{\varsigma}:=\mathfrak{I}(k_{\varsigma}),\,\,\,\varsigma={\textrm{p}_{1}},{\textrm{p}_{2}},{\textrm{s}},\,\,\,\textrm{\bf u}=\boldsymbol{u}^{\tiny\iota},\boldsymbol{u}, (9)
pςrikςpς=o(r1eaςr)asr,ς=p1,p2,p=pι,p,\displaystyle\frac{\partial{\textrm{p}}_{\varsigma}}{\partial r}-\text{i}k_{\varsigma}{\textrm{p}}_{\varsigma}=o\big{(}r^{-1}e^{-a_{\varsigma}r}\big{)}\quad\text{as}~{}~{}r\to\infty,\qquad\varsigma={\textrm{p}_{1}},{\textrm{p}_{2}},\,\,\,\textrm{p}=p^{\tiny\iota},p,

uniformly with respect to 𝝃^:=𝝃/r\hat{\boldsymbol{\xi}}:=\boldsymbol{\xi}/r where kςk_{\varsigma}, ς=p1,p2,s\varsigma={\textrm{p}_{1}},{\textrm{p}_{2}},{\textrm{s}}, is the complex-valued wavenumber associated with the slow and fast p-waves and the transverse s-wave [37].

2.3 Function spaces

For clarity of discussion, it should be mentioned that the support of Γ\Gamma may be decomposed into NN smooth open subsets ΓnΓ\Gamma_{n}\!\subset\Gamma, n=1,Nn=1,\ldots N, such that Γ=n=1NΓn\Gamma\!=\!{\textstyle\bigcup_{n=1}^{N}}\Gamma_{n}. The support of Γn\Gamma_{n} may be arbitrarily extended to a closed Lipschitz surface Dn\partial\text{\sf D}_{n} of a bounded simply connected domain Dn\text{\sf D}_{n}. In this setting, the unit normal vector 𝒏\boldsymbol{n} to Γn\Gamma_{n} coincides with the outward normal vector to Dn\partial\text{\sf D}_{n}. Let D=n=1NDn\text{\sf D}={\textstyle\bigcup_{n=1}^{N}}\text{\sf D}_{n} be a multiply connected Lipschitz domain of bounded support such that ΓD\Gamma\subset\partial\text{\sf D}, then Γ\Gamma is assumed to be an open set relative to D\partial\text{\sf D} with a positive surface measure, and the closure of Γ\Gamma is denoted by Γ¯:=ΓΓ\overline{\Gamma}\colon\!\!\!=\Gamma\cup\partial\Gamma. Following [38], we define

H±12(Γ):={f|Γ:fH±12(D)},\displaystyle H^{\pm\frac{1}{2}}(\Gamma)~{}:=~{}\big{\{}f\big{|}_{\Gamma}\!\colon\,\,\,f\in H^{\pm\frac{1}{2}}(\partial\text{\sf D})\big{\}}, (10)
H~±12(Γ):={fH±12(D):supp(f)Γ¯},\displaystyle\tilde{H}^{\pm\frac{1}{2}}(\Gamma)~{}:=~{}\big{\{}f\in H^{\pm\frac{1}{2}}(\partial\text{\sf D})\colon\,\,\,\text{supp}(f)\subset\overline{\Gamma}\hskip 1.13809pt\big{\}},

and recall that H1/2(Γ)H^{-1/2}(\Gamma) and H~1/2(Γ)\tilde{H}^{-1/2}(\Gamma) are respectively the dual spaces of H~1/2(Γ)\tilde{H}^{1/2}(\Gamma) and H1/2(Γ)H^{1/2}(\Gamma). Accordingly, the following embeddings hold

H~12(Γ)H12(Γ)L2(Γ)H~12(Γ)H12(Γ).\tilde{H}^{\frac{1}{2}}(\Gamma)\,\subset\,H^{\frac{1}{2}}(\Gamma)\,\subset\,L^{2}(\Gamma)\,\subset\,\tilde{H}^{-\frac{1}{2}}(\Gamma)\,\subset\,H^{-\frac{1}{2}}(\Gamma). (11)

Note that since (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}), then by trace theorems (𝒖,p)H~1/2(Γ)3×H~1/2(Γ)(\llbracket\boldsymbol{u}\rrbracket,\llbracket p\rrbracket)\in\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma), (𝒕,q)H1/2(Γ)3×H1/2(Γ)({\boldsymbol{t}},q)\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma), and qH~1/2(Γ)\llbracket q\rrbracket\in\tilde{H}^{-1/2}(\Gamma).

Remark 2.2 (Wellposedness of the forward scattering problem).

In light of (5), let us define the interface operator 𝔓:H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\mathfrak{P}:\tilde{H}^{{1}/{2}}(\Gamma)^{3}\times\tilde{H}^{{1}/{2}}(\Gamma)\times\tilde{H}^{-{1}/{2}}(\Gamma)\rightarrow{H}^{-{1}/{2}}(\Gamma)^{3}\times{H}^{-{1}/{2}}(\Gamma)\times{H}^{{1}/{2}}(\Gamma) such that

𝔓(𝒖,p,q):=(𝒕+𝒕ι,q+qι,p+pι)onΓ.\mathfrak{P}(\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 0.56905ptp\hskip 0.56905pt\rrbracket,-\llbracket q\hskip 0.56905pt\rrbracket)\,:=\,({\boldsymbol{t}}+{\boldsymbol{t}}^{\tiny\iota},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{q}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+q^{\tiny\iota},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+p^{\tiny\iota})\qquad\text{on}\quad{\Gamma}. (12)

Then, (5)-(9) is wellposed provided that 𝔓\mathfrak{P} is bounded and

𝔓𝝋,𝝋Γ 0,𝝋H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ):𝝋𝟎,\Im\left<\mathfrak{P}{\boldsymbol{\varphi}},{\boldsymbol{\varphi}}\right>_{\Gamma}\,\leqslant\,0,\,\qquad\forall\hskip 1.13809pt{\boldsymbol{\varphi}}\in\tilde{H}^{{1}/{2}}(\Gamma)^{3}\times\tilde{H}^{{1}/{2}}(\Gamma)\times\tilde{H}^{-{1}/{2}}(\Gamma):~{}{\boldsymbol{\varphi}}\neq\boldsymbol{0}, (13)

given the duality product

,Γ=H12(Γ)3×H12(Γ)×H12(Γ),H~12(Γ)3×H~12(Γ)×H~12(Γ).\langle\cdot,\cdot\rangle_{\Gamma}~{}=~{}\big{\langle}{H}^{-\frac{1}{2}}(\Gamma)^{3}\times{H}^{-\frac{1}{2}}(\Gamma)\times{H}^{\frac{1}{2}}(\Gamma),\,\tilde{H}^{\frac{1}{2}}(\Gamma)^{3}\times\tilde{H}^{\frac{1}{2}}(\Gamma)\times\tilde{H}^{-\frac{1}{2}}(\Gamma)\big{\rangle}.

Detailed analysis is included in B.

2.4 Poroelastic scattering

In this section, the poroelastic scattering operator is introduced and its first and second factorizations are formulated. In what follows, the Einstein’s summation convention applies over the repeated indexes.

2.4.1 Incident wave function ​​

Observe that for a given density L2(𝒢)3×L2(𝒢)𝒈:=(𝒈s,gf)L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G})\ni{\boldsymbol{g}}:=({\boldsymbol{g}}_{s},g_{f}), the poroelastic incident field (𝒖ι,pι)Hloc1(3𝒢¯)3×Hloc1(3𝒢¯)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\mathscr{G}})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\mathscr{G}}) solving (3) may be recast in the integral form

[uiιpι](𝝃)=𝒢[U𝔰iju𝔣ip𝔰jp𝔣](𝒚,𝝃)[gsjgf](𝒚)d𝒚,𝝃3𝒢¯,i,j=1,2,3,\displaystyle\left[\!\!\begin{array}[]{l}\\[-15.50673pt] u^{\tiny\iota}_{i}\\[-0.28453pt] \hskip 1.13809pt\hskip 1.13809ptp^{\tiny\iota}\end{array}\!\!\right]\!(\boldsymbol{\xi})~{}=~{}\int_{{\mathscr{G}}}\left[\!\!\begin{array}[]{ll}{{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!ij}\!\!&\!\!{{\textrm{u}}^{\mathfrak{f}}}_{\!i}\\[0.28453pt] \hskip 1.13809pt{{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j}\!&\!\!{{\textrm{p}}^{\mathfrak{f}}}\\ \end{array}\!\!\right]\!(\boldsymbol{y},\boldsymbol{\xi})\,\!\cdot\!\left[\!\!\!\begin{array}[]{l}{\hskip 1.13809ptg_{s}^{\hskip 0.56905ptj}}\\[-1.42262pt] {\hskip 1.13809ptg_{f}}\end{array}\!\!\!\right]\!\!(\boldsymbol{y})\hskip 1.13809pt\text{d}{\boldsymbol{y}},\qquad\boldsymbol{\xi}\in\mathbb{R}^{3}\setminus\overline{\mathscr{G}},\qquad i,j=1,2,3, (14)

whose kernel is the poroelastodynamics fundamental solution tensor provided in A. Here, U𝔰ij(𝒚,𝝃){{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!ij}(\boldsymbol{y},\boldsymbol{\xi}) (i,j=1,2,3)(i,j=1,2,3) signifies the i thi^{\text{\hskip 0.56905ptth}} component of the solid displacement at 𝝃\boldsymbol{\xi} due to a unit total body force applied at 𝒚\boldsymbol{y} along the coordinate direction jj, while p𝔰j{{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j} is the associated pore pressure. Likewise, u𝔣i(𝒚,𝝃){{\textrm{u}}^{\mathfrak{f}}}_{\!i}(\boldsymbol{y},\boldsymbol{\xi}) (i=1,2,3)(i=1,2,3) stands for the solid displacement at 𝝃\boldsymbol{\xi} in the i thi^{\text{\hskip 0.56905ptth}} direction due to a unit fluid source i.e., volumetric injection at 𝒚\boldsymbol{y}, and p𝔣{{\textrm{p}}^{\mathfrak{f}}} is the induced pore pressure.

2.4.2 Poroelastic scattered field ​​

By way of the reciprocal theorem of poroelastodynamics [39, 40], one may show that the following integral representation holds for the scattered field (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) satisfying (5)-(9),

[uip](𝝃)=Γ[T𝔰ijq𝔰ip𝔰it𝔣jq𝔣p𝔣](𝒚,𝝃)[ujp-q](𝒚)dS𝒚,𝝃3Γ¯,i,j=1,2,3,\displaystyle\left[\!\!\begin{array}[]{l}\\[-16.64487pt] u_{i}\\[-2.84526pt] \hskip 1.13809pt\hskip 1.13809ptp\end{array}\!\!\right]\!(\boldsymbol{\xi})~{}=~{}\!\int_{\Gamma}\left[\!\!\begin{array}[]{lll}\\[-16.64487pt] {{\textrm{T}}^{\mathfrak{s}}}_{\!\!\!ij}\!\!&\!\!{{\textrm{q}}^{\mathfrak{s}}}_{\!\!\!i}\!\!&\!{{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!i}\\[0.0pt] {{\textrm{t}}^{\mathfrak{f}}}_{\!j}\!\!&\!\!{{\textrm{q}}^{\mathfrak{f}}}\!\!&\!{{\textrm{p}}^{\mathfrak{f}}}\end{array}\!\!\right]\!(\boldsymbol{y},\boldsymbol{\xi})\,\!\cdot\!\left[\!\!\!\begin{array}[]{l}\llbracket u_{j}\rrbracket\\[-1.42262pt] \hskip 1.13809pt\hskip 1.13809pt\llbracket\hskip 1.13809ptp\hskip 0.56905pt\rrbracket\\[-1.42262pt] \textrm{\small-}\llbracket\hskip 0.56905ptq\hskip 0.42677pt\rrbracket\end{array}\!\!\!\right]\!\!(\boldsymbol{y})\hskip 1.13809pt\text{d}S_{\boldsymbol{y}},\qquad\boldsymbol{\xi}\in\mathbb{R}^{3}\setminus\overline{\Gamma},\qquad i,j=1,2,3, (15)

whose kernel is derived from the fundamental solution tensor and provided in A. More specifically, T𝔰{{\textrm{\bf T}}^{\mathfrak{s}}} and t𝔣{{\textrm{\bf t}}^{\mathfrak{f}}} indicate the fundamental tractions on Γ\Gamma; q𝔰{{\textrm{\bf q}}^{\mathfrak{s}}} and q𝔣{{\textrm{q}}^{\mathfrak{f}}} signify the associated relative fluid-solid displacements across Γ\Gamma; p𝔰{{\textrm{\bf p}}^{\mathfrak{s}}} and p𝔣{{\textrm{p}}^{\mathfrak{f}}} are the fundamental pore pressure solutions.

2.4.3 Poroelastic scattering operator​​

The linearity of (5) implies that the scattered wave function may be expressed as a linear integral operator. To observe this, let 4𝒅:=(𝒅s,dp)\mathbb{R}^{4}\ni\boldsymbol{d}:=(\boldsymbol{d}_{s},d_{p}) be a constant vector characterizing an incident point source wherein 𝒅s\boldsymbol{d}_{s} indicates the amplitude and direction of the total body force, while dpd_{p} represents the magnitude of fluid volumetric source both applied at 𝒚𝒢\boldsymbol{y}\in\mathscr{G}. The resulting scattered field is observed in terms of seismic displacements 𝒖𝒅\boldsymbol{u}_{\boldsymbol{d}} and pore pressure p𝒅p_{\boldsymbol{d}} at 𝝃𝒢\boldsymbol{\xi}\in\mathscr{G}. Now, let us define the scattering kernel 𝕾(𝒚,𝝃)4×4\boldsymbol{\mathfrak{S}}(\boldsymbol{y},\boldsymbol{\xi})\in\mathbb{C}^{4\times 4} so that

𝕾(𝒚,𝝃)𝒅:=(𝒖𝒅,p𝒅)(𝝃),𝒚,𝝃𝒢.\boldsymbol{\mathfrak{S}}(\boldsymbol{y},\boldsymbol{\xi})\!\cdot\!\boldsymbol{d}~{}:=~{}(\boldsymbol{u}_{\boldsymbol{d}},\hskip 1.13809ptp_{\boldsymbol{d}})(\boldsymbol{\xi}),\qquad\boldsymbol{y},\boldsymbol{\xi}\in\mathscr{G}. (16)

Then, given 𝒈L2(𝒢)3×L2(𝒢){\boldsymbol{g}}\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}), one may easily verify that

(𝒖,p)(𝝃)=𝒢𝕾(𝒚,𝝃)𝒈(𝒚)dS𝒚,𝒚,𝝃𝒢,(\boldsymbol{u},p)(\boldsymbol{\xi})~{}=\,\int_{{\mathscr{G}}}\boldsymbol{\mathfrak{S}}(\boldsymbol{y},\boldsymbol{\xi})\!\cdot\!{\boldsymbol{g}}(\boldsymbol{y})\,\,\text{d}S_{\boldsymbol{y}},\qquad\boldsymbol{y},\boldsymbol{\xi}\in\mathscr{G}, (17)

where (𝒖,p)L2(𝒢)3×L2(𝒢)(\boldsymbol{u},p)\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) satisfies (5)-(9).

Accordingly, one may define the poroelastic scattering operator Λ:L2(𝒢)3×L2(𝒢)L2(𝒢)3×L2(𝒢)\Lambda:L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G})\to L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) by

Λ(𝒈)=(𝒖,p)|𝒢.\Lambda({\boldsymbol{g}})~{}=~{}(\boldsymbol{u},p)|_{\mathscr{G}}. (18)

2.4.4 Factorization of Λ\Lambda\!\!\!

With reference to (3) and (6), let us define the operator 𝒮:L2(𝒢)3×L2(𝒢)H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\mathscr{S}\colon L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G})\rightarrow H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma) given by

𝒮(𝒈):=(𝒕ι,qι,pι)onΓ,\mathscr{S}({\boldsymbol{g}})~{}:=~{}({\boldsymbol{t}}^{\iota},q^{\iota},p^{\iota})\quad~{}~{}\text{on}\quad\Gamma, (19)

where 𝒕ι{\boldsymbol{t}}^{\iota} specifies the incident field traction, qιq^{\iota} is the specific relative fluid-solid displacement across Γ\Gamma, and pιp^{\iota} is the incident pore pressure on Γ\Gamma. Next, define 𝒫:H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)L2(𝒢)3×L2(𝒢)\mathscr{P}\colon H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) as the map taking the incident traction, relative flow and pressure (𝒕ι,qι,pι)({\boldsymbol{t}}^{\iota},q^{\iota},p^{\iota}) over Γ\Gamma to the multiphase scattered data (𝒖,p)L2(𝒢)3×L2(𝒢)(\boldsymbol{u},p)\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) satisfying (5)-(9). Then, the scattering operator (18) may be factorized as

Λ=𝒫𝒮.\Lambda~{}=~{}\mathscr{P}\mathscr{S}. (20)
Lemma 2.1.

The adjoint operator 𝒮:H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)L2(𝒢)3×L2(𝒢)\mathscr{S}^{*}\colon\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) takes the form

𝒮(𝒂):=(𝒖𝒂,p𝒂)(𝝃),𝝃𝒢,\mathscr{S}^{*}(\boldsymbol{a})~{}:=~{}(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}})(\boldsymbol{\xi}),\qquad\boldsymbol{\xi}\in\mathscr{G}, (21)

where (𝐮𝐚,p𝐚)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) solves

(𝑪:𝒖𝒂)(αρfγ¯)p𝒂+ω2(ρρf2γ¯)𝒖𝒂= 0\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}_{\boldsymbol{a}})\,-\,(\alpha-\dfrac{\rho_{f}}{\bar{\gamma}})\nabla p_{\boldsymbol{a}}\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\bar{\gamma}})\boldsymbol{u}_{\boldsymbol{a}}\,=\,\boldsymbol{0}\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}\setminus\overline{\Gamma}}, (22)
1γ¯ω22p𝒂+M1p𝒂+(αρfγ¯)𝒖𝒂= 0\displaystyle\dfrac{1}{\bar{\gamma}\omega^{2}}\nabla^{2}p_{\boldsymbol{a}}\,+\,{M}^{-1}p_{\boldsymbol{a}}\,+\,(\alpha-\dfrac{\rho_{f}}{\bar{\gamma}})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}_{\boldsymbol{a}}\,=\,0\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}}\setminus\overline{\Gamma},
(𝒖𝒂,p𝒂,q𝒂)=𝒂,𝒕𝒂=𝟎\displaystyle(\llbracket\boldsymbol{u}_{\boldsymbol{a}}\rrbracket,\llbracket p_{\boldsymbol{a}}\rrbracket,-\llbracket q_{\boldsymbol{a}}\rrbracket)\,=\,\boldsymbol{a},\quad\llbracket{\boldsymbol{t}}_{\boldsymbol{a}}\rrbracket=\boldsymbol{0}\quad on Γ.\displaystyle\,\,\,{\Gamma}.

Here, the ‘bar’ indicates complex conjugate, and

[𝒕𝒂,q𝒂](𝒖𝒂,p𝒂):=[𝒏𝑪:𝒖𝒂αp𝒂𝒏,1γ¯ω2(p𝒂𝒏ρfω2𝒖𝒂𝒏)]onΓ.[{\boldsymbol{t}}_{\boldsymbol{a}},q_{\boldsymbol{a}}](\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}})\,:=\,\big{[}\boldsymbol{n}\cdot\boldsymbol{C}\colon\!\nabla\boldsymbol{u}_{\boldsymbol{a}}-\alpha\hskip 1.13809ptp_{\boldsymbol{a}}\boldsymbol{n},\,\dfrac{1}{\bar{\gamma}\omega^{2}}(\nabla p_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{n}-\rho_{f}\omega^{2}\boldsymbol{u}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{n})\big{]}\quad\text{on}\,\,\,{\Gamma}.

Note that (𝐮𝐚,p𝐚)(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}}) satisfy the radiation condition as |𝛏||\boldsymbol{\xi}|\to\infty, similar to the complex conjugate of (9).

Proof.

see C. ∎

On the basis of (15) and (21), the operator 𝒫\mathscr{P} can be further decomposed as 𝒫=𝒮¯T\mathscr{P}=\bar{\mathscr{S}}^{*}T where the middle operator T:H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)3T\colon H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)\rightarrow\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)^{3} is given by

T(𝒕ι,qι,pι):=(𝒖,p,q)onΓ,T({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota})~{}:=~{}\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 1.13809ptp\hskip 1.13809pt\rrbracket,-\llbracket\hskip 1.13809ptq\hskip 1.13809pt\rrbracket\big{)}\qquad\text{on}\,\,\,{\Gamma}, (23)

such that (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) satisfies (15) and qq is computed from (6). Thanks to this new decomposition of 𝒫\mathscr{P}, a second factorization of Λ:L2(𝒢)3×L2(𝒢)L2(𝒢)3×L2(𝒢)\Lambda\colon L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}) is obtained as

Λ=𝒮¯T𝒮.\Lambda~{}=~{}\bar{\mathscr{S}}^{*}\hskip 1.13809ptT\hskip 1.13809pt\mathscr{S}. (24)

It is worth noting that the asymmetry of the second factorization (24) stems from the non-selfadjoint nature of the Biot system which is evident from (22).

3 Key properties of the poroelastodynamic operators

Lemma 3.1.

Operator 𝒮:H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)L2(𝒢)3×L2(𝒢)\mathscr{S}^{*}\colon\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) in Lemma 2.1 is compact, injective, and has a dense range.

Proof.

See D. ∎

Lemma 3.2.

Operator T:H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)3T\colon H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)\rightarrow\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)^{3} in (23) is bounded and satisfies

𝝍,T𝝍Γ<0,𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ):𝝍𝟎.\Im\left<{\boldsymbol{\psi}},T{\boldsymbol{\psi}}\right>_{\Gamma}<0,\,\quad\forall\hskip 1.13809pt{\boldsymbol{\psi}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma):~{}{\boldsymbol{\psi}}\neq\boldsymbol{0}. (25)
Proof.

See E. ∎

Lemma 3.3.

Operator T:H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)3T\colon H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)\rightarrow\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)^{3}: (i) has a bounded (and thus continuous) inverse, and (ii) is coercive, i.e., there exists constant c>0c\!>\!0 independent of 𝛙{\boldsymbol{\psi}} such that

|𝝍,T(𝝍)|c𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)2,𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ).|\langle{\boldsymbol{\psi}},\,T({\boldsymbol{\psi}})\rangle|\,\,\geqslant\,\,\textrm{c}\parallel\!{\boldsymbol{\psi}}\!\parallel_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}^{2},\quad\forall{\boldsymbol{\psi}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma). (26)
Proof.

See F. ∎

Lemma 3.4.

Operator 𝒫=𝒮¯T:H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)L2(𝒢)3×L2(𝒢)\mathscr{P}=\bar{\mathscr{S}}^{*}T\,\colon H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) is compact over H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma).

Proof.

The claim follows immediately from Lemmas 3.1 and 3.2 establishing, respectively, the compactness of 𝒮\mathscr{S}^{*} and the boundedness of TT. ∎

Lemma 3.5.

The poroelastic scattering operator Λ:L2(𝒢)3×L2(𝒢)L2(𝒢)3×L2(𝒢)\Lambda:L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G})\to L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}) is injective, compact and has a dense range.

Proof.

See G. ∎

4 Inverse poroelastic scattering

This section presents an adaptation of the key results from sampling approaches to inverse scattering for the problem of poroelastic-wave imaging of finitely permeable interfaces. The fundamental idea stems from the nature of solution 𝒈=(𝒈s,gf)L2(𝒢)3×L2(𝒢){\boldsymbol{g}}=({\boldsymbol{g}}_{s},g_{f})\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) to the poroelastic scattering equation

Λ𝒈=𝚽L,Λ=𝒫𝒮=𝒮¯T𝒮,\Lambda{\boldsymbol{g}}~{}=~{}\boldsymbol{\Phi}_{\text{\tiny L}},\qquad\Lambda~{}=~{}\mathscr{P}\mathscr{S}~{}=~{}\bar{\mathscr{S}}^{*}\hskip 1.13809ptT\hskip 1.13809pt\mathscr{S}, (27)

where 𝚽L\boldsymbol{\Phi}_{\text{\tiny L}} is the near-field pattern of a trial poroelastodynamic field specified by Definition 1.

Definition 1.

With reference to (15), for every admissible density 𝐚H~1/2(L)3×H~1/2(L)×H~1/2(L)\boldsymbol{a}\!\in\!\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L) specified over a smooth, non-intersecting trial interface L3L\!\subset\!\mathscr{B}\!\subset\!\mathbb{R}^{3}, the induced near-field pattern 𝚽L:H~1/2(L)3×H~1/2(L)×H~1/2(L)L2(𝒢)3×L2(𝒢)\boldsymbol{\Phi}_{\text{\tiny L}}\colon\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L)\rightarrow L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) is given by

𝚽L(𝒂)(𝝃):=(𝒖L,pL)(𝝃)=L𝚺(𝒚,𝝃)𝒂(𝒚)dS𝒚,𝚺=[T𝔰q𝔰p𝔰t𝔣q𝔣p𝔣],𝝃𝒢.\displaystyle\boldsymbol{\Phi}_{\text{\tiny\emph{L}}}(\boldsymbol{a})(\boldsymbol{\xi})\,:=\,(\boldsymbol{u}_{\hskip 0.56905pt\text{\tiny\emph{L}}},p_{\hskip 0.56905pt\text{\tiny\emph{L}}})(\boldsymbol{\xi})\,=\int_{L}\,\boldsymbol{\Sigma}(\boldsymbol{y},\boldsymbol{\xi})\!\cdot\!\boldsymbol{a}(\boldsymbol{y})\,\,\text{d}S_{\boldsymbol{y}},\quad\boldsymbol{\Sigma}\,=\,\text{\emph{$\left[\!\!\begin{array}[]{lll}\\[-16.64487pt] {{\textrm{\bf T}}^{\mathfrak{s}}}\!\!&\!\!{{\textrm{\bf q}}^{\mathfrak{s}}}\!\!&\!{{\textrm{\bf p}}^{\mathfrak{s}}}\\[0.0pt] {{\textrm{\bf t}}^{\mathfrak{f}}}\!\!&\!\!{{\textrm{q}}^{\mathfrak{f}}}\!\!&\!{{\textrm{p}}^{\mathfrak{f}}}\end{array}\!\!\right]$}},\quad\boldsymbol{\xi}\in\mathscr{G}. (28)
Remark 4.1.

In light of (27), one may interpret the philosophy of sampling-based waveform tomography as the following. Let 𝖫{\sf L}\!\subset\!\mathscr{B} (containing the origin) denote a poroelastic discontinuity whose characteristic size is small relative to the length scales describing the forward scattering problem, and let L=𝐳+𝗥𝖫L=\boldsymbol{z}\!+\boldsymbol{\sf R}{\sf L} where 𝐳\boldsymbol{z}\!\in\mathscr{B} and 𝗥U(3)\boldsymbol{\sf R}\!\in\!U(3) is a unitary rotation matrix. Given a density 𝐚H~1/2(L)3×H~1/2(L)×H~1/2(L)\boldsymbol{a}\!\in\!\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L), solving the scattering equation (27) over a grid of trial pairs (𝐳,𝗥)(\boldsymbol{z},\boldsymbol{\sf R}) sampling ×U(3)\mathscr{B}\times U(3) is simply an effort to probe the range of operator Λ\Lambda (or that of 𝒮¯\bar{\mathscr{S}}^{*}), through synthetic reshaping of the illuminating wavefront, for fingerprints in terms of 𝚽L\boldsymbol{\Phi}_{\text{\tiny\emph{L}}}. As shown by Theorems 4.2, such fingerprint is found in the data if and only if LΓL\subset\Gamma. Otherwise, the norm of any approximate solution to (27) can be made arbitrarily large, which provides a criterion for reconstructing Γ\Gamma.

Theorem 4.1.

Under the assumptions of Remark 2.2 for the wellposedness of the forward scattering problem, for every smooth and non-intersecting trial dislocation L3L\!\subset\!\mathscr{B}\!\subset\!\mathbb{R}^{3} and density profile 𝐚(𝛏)H~1/2(L)3×H~1/2(L)×H~1/2(L)\,\boldsymbol{a}(\boldsymbol{\xi})\!\in\!\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L), one has

𝚽LRange(𝒮¯)LΓ.\boldsymbol{\Phi}_{\text{\tiny L}}\in Range(\bar{\mathscr{S}}^{*})~{}~{}\iff~{}~{}L\subset\Gamma.
Proof.

The argument directly follows that of [41, Theorem 6.1]. ∎

On the basis of Theorem 4.1, one arrives at the following statement which inspires the sampling-based poroelastic imaging indicators.

Theorem 4.2.

Under the assumptions of Theorem 4.1,

  • 1.

    If LΓL\subset\Gamma, there exists a density vector 𝒈ϵLL2(𝒢)3×L2(𝒢){\boldsymbol{g}}_{\epsilon}^{L}\!\in L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}) such that Λ𝒈ϵL𝚽LL2(𝒢)3×L2(𝒢)ϵ\|\Lambda\hskip 1.13809pt{\boldsymbol{g}}_{\epsilon}^{L}-\boldsymbol{\Phi}_{\text{\tiny L}}\|_{L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})}\leqslant\epsilon and
    lim supϵ0𝒮𝒈ϵLH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)<\limsup\limits_{\epsilon\rightarrow 0}\|\mathscr{S}{\boldsymbol{g}}_{\epsilon}^{L}\|_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}<\infty.

  • 2.

    If LΓL\not\subset\Gamma, then 𝒈ϵLL2(𝒢)3×L2(𝒢)\forall{\boldsymbol{g}}_{\epsilon}^{L}\!\in L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}) such that Λ𝒈ϵL𝚽LL2(𝒢)3×L2(𝒢)ϵ\parallel\!\Lambda\hskip 1.13809pt{\boldsymbol{g}}_{\epsilon}^{L}-\boldsymbol{\Phi}_{\text{\tiny L}}\!\parallel_{L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})}\,\leqslant\epsilon, one has
    limϵ0𝒮𝒈ϵLH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)=\lim\limits_{\epsilon\rightarrow 0}\parallel\!\mathscr{S}{\boldsymbol{g}}_{\epsilon}^{L}\!\parallel_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}\,\,=\infty.

Proof.

The argument directly follows that of [41, Theorem 6.2]. ∎

4.1 The multiphysics 𝔏\mathfrak{L} indicator ​​​

Theorem 4.2 poses two challenges in that: (i) the featured indicator 𝒮𝒈ϵLH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\|\mathscr{S}{\boldsymbol{g}}_{\epsilon}^{L}\|_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)} inherently depends on the unknown support Γ\Gamma of hidden scatterers, and (ii) construction of the wavefront density 𝒈ϵLL2(𝒢)3×L2(𝒢){\boldsymbol{g}}_{\epsilon}^{L}\in L^{2}({\mathscr{G}})^{3}\!\times\!L^{2}({\mathscr{G}}) is implicit in the theorem [20, 26]. These are conventionally addressed by replacing 𝒮𝒈ϵLH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\|\mathscr{S}{\boldsymbol{g}}_{\epsilon}^{L}\|_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)} with 𝒈ϵLL2(𝒢)3×L2(𝒢)\parallel\!\!{\boldsymbol{g}}_{\epsilon}^{L}\!\!\!\parallel_{L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})} which in turn is computed by way of Tikhonov regularization [42] as follows.

First, note that when the measurements are contaminated by noise (e.g., sensing errors and fluctuations in the medium properties), one has to deal with the noisy operator Λδ\Lambda^{\updelta}\! satisfying

ΛδΛδ,\parallel\!\Lambda^{\updelta}-\hskip 1.13809pt\Lambda\!\parallel\,\,\,\leqslant\,\hskip 1.13809pt\updelta,\vspace{-1 mm} (29)

where δ>0\updelta\!>\!0 is a measure of perturbation in data independent of Λ\Lambda. Assuming that Λδ\Lambda^{\updelta} is compact, the Tikhonov-regularized solution 𝒈η{\boldsymbol{g}}_{\upeta} to (27) is obtained by non-iteratively minimizing the cost functional Jη(𝚽L;):L2(𝒢)3×L2(𝒢)J_{\upeta}(\boldsymbol{\Phi}_{\text{\tiny L}};\,\cdot)\colon\hskip 1.13809ptL^{2}({\mathscr{G}})^{3}\hskip 1.13809pt\times\!L^{2}({\mathscr{G}})\rightarrow\mathbb{R} defined by

Jη(𝚽L;𝒈):=Λδ𝒈𝚽LL22+η𝒈L22,𝒈L2(𝒢)3×L2(𝒢),J_{\upeta}(\boldsymbol{\Phi}_{\text{\tiny L}};\,{\boldsymbol{g}})~{}\colon\!\!\!=~{}\!\!\parallel\!{\Lambda}^{\updelta}\hskip 1.13809pt{\boldsymbol{g}}\,-\,\boldsymbol{\Phi}_{\text{\tiny L}}\!\parallel^{2}_{L^{2}}\,+\,\,\,\upeta\!\parallel\!{\boldsymbol{g}}\!\parallel^{2}_{L^{2}},\qquad{\boldsymbol{g}}\in L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}),\vspace{-1 mm} (30)

where the regularization parameter η=η(δ,L)\upeta=\upeta(\updelta,L) is obtained by way of the Morozov discrepancy principle [42]. On the basis of (30), the well-known linear sampling indicator 𝔏\mathfrak{L} is constructed as

𝔏:=1𝒈ηL2,𝒈η:=min𝒈L2(𝒢)3×L2(𝒢)Jη(𝚽L;𝒈).\mathfrak{L}~{}\colon\!\!\!=~{}\!\!\frac{1}{\parallel\!{\boldsymbol{g}}_{\upeta}\!\parallel_{L^{2}}},\qquad{\color[rgb]{0,0,0}{\boldsymbol{g}}_{\upeta}~{}\colon\!\!\!=~{}\!\!\min_{{\boldsymbol{g}}\hskip 1.13809pt\in\hskip 1.13809ptL^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})}J_{\upeta}(\boldsymbol{\Phi}_{\text{\tiny L}};\,{\boldsymbol{g}}).}\vspace{-1 mm} (31)

By design, 𝔏\mathfrak{L} achieves its highest values at the loci of hidden scatterers Γ\Gamma. More specifically, the behavior of 𝔏\mathfrak{L} within \mathscr{B} may be characterized as the following,

ifLΓlim infη0𝔏(𝒈η)> 0,\displaystyle\text{if}\,\,\,L\subset\Gamma\quad\iff\quad\liminf\limits_{\eta\rightarrow 0}\hskip 1.13809pt\mathfrak{L}({\boldsymbol{g}}_{\upeta})\,>\,0, (32)
ifLΓ¯limη0𝔏(𝒈η)= 0.\displaystyle\text{if}\,\,\,L\subset\mathscr{B}\setminus\overline{\Gamma}\quad\iff\quad\lim\limits_{\eta\rightarrow 0}\mathfrak{L}({\boldsymbol{g}}_{\upeta})\,=\,0.

4.2 The modified 𝔊\mathfrak{G} indicator ​​​

In cases where the illumination frequency and/or the background’s permeability is sufficiently large so that γ0\Im{\gamma}\to 0 according to (4), one may deduce from (22)-(24) that the second factorization of the poroelastic scattering operator is symmetrized as follows,

Λ=𝒮T𝒮.\Lambda~{}=~{}{\mathscr{S}}^{*}\hskip 1.13809ptT\hskip 1.13809pt\mathscr{S}.\vspace{-1 mm} (33)

In this setting, one may deploy the more rigorous 𝔊\mathfrak{G} indicator [41] to dispense with the approximations underlying the 𝔏\mathfrak{L} imaging functional. This results in a more robust reconstruction especially with noisy data [26]. The 𝔊\mathfrak{G} indicator takes advantage of the positive and self-adjoint operator Λ:L2(𝒢)3×L2(𝒢)L2(𝒢)3×L2(𝒢)\Lambda_{\sharp}:\,L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})\rightarrow L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}) defined on the basis of the scattering operator Λ\Lambda by

Λ:=12|Λ+Λ|+|12i(ΛΛ)|,\Lambda_{\sharp}\,\colon\!\!\!=\,\frac{1}{2}\big{|}\hskip 1.13809pt\Lambda+\Lambda^{*}\big{|}\>+\>\big{|}\frac{1}{2\textrm{\emph{i}}}(\Lambda-\Lambda^{*})\big{|},\vspace{-1mm} (34)

with the affiliated factorization [43]

Λ=𝒮T𝒮,\Lambda_{\sharp}~{}=~{}\mathscr{S}^{*}\hskip 1.13809ptT_{\sharp}\hskip 1.13809pt\mathscr{S},\vspace{-0.5mm} (35)

where in light of Lemma 3.3, the middle operator TT_{\sharp} is coercive with reference to [41, Lemma 5.7] i.e., there exists a constant c>0c>0 independent of 𝛙=𝒮𝒈\boldsymbol{\uppsi}=\mathscr{S}{\boldsymbol{g}} such that 𝛙H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\forall\hskip 0.56905pt\boldsymbol{\uppsi}\in{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)},

(𝒈,Λ𝒈)L2(𝒢)3×L2(𝒢)=𝛙,T𝛙Γc𝛙H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)2.(\hskip 1.13809pt{\boldsymbol{g}},\hskip 1.13809pt\Lambda_{\sharp}\hskip 1.13809pt{\boldsymbol{g}})_{L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})}~{}=~{}\big{\langle}\boldsymbol{\uppsi},\,T_{\sharp}\hskip 0.56905pt\boldsymbol{\uppsi}\big{\rangle}_{\Gamma}~{}\geqslant~{}c\parallel\!\!\boldsymbol{\uppsi}\!\parallel^{2}_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}.\vspace{-0.5mm} (36)

Thanks to (36), the term 𝒮𝒈ϵH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)2\parallel\!\!\mathscr{S}{\boldsymbol{g}}_{\epsilon}\!\!\parallel^{2}_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)} in Theorem 4.2 may be safely replaced by (𝒈ϵ,Λ𝒈ϵ)L2(𝒢)3×L2(𝒢)(\hskip 1.13809pt{\boldsymbol{g}}_{\epsilon},\hskip 1.13809pt\Lambda_{\sharp}\hskip 1.13809pt{\boldsymbol{g}}_{\epsilon})_{L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}})} which is computable without prior knowledge of Γ\Gamma.

In presence of noise, the perturbed operator Λδ\Lambda_{\sharp}^{\updelta} is deployed satisfying

ΛδΛδ,\parallel\!\Lambda^{\updelta}_{\sharp}-\Lambda_{\sharp}\!\parallel\,\,\,\leqslant\,\,\updelta,\vspace{-1mm} (37)

which is assumed to be compact similar to Λδ\Lambda^{\updelta} in (29). Then, the regularized cost functional Jα,δ(𝚽L;):L2(𝒢)3×L2(𝒢)J_{\upalpha,\updelta}(\boldsymbol{\Phi}_{\text{\tiny L}};\,\cdot)\colon\hskip 1.13809ptL^{2}({\mathscr{G}})^{3}\hskip 1.13809pt\times L^{2}({\mathscr{G}})\rightarrow\mathbb{R} is constructed according to [23, Theorems 4.3],

Jα,δ(𝚽L;𝒈):=Λδ𝒈𝚽LL22+α(𝒈,Λδ𝒈)L2+αδ𝒈L22,𝒈L2(𝒢)3×L2(𝒢).J_{\upalpha,\updelta}(\boldsymbol{\Phi}_{\text{\tiny L}};\,{\boldsymbol{g}})~{}\colon\!\!\!=~{}\!\parallel\!\Lambda^{\updelta}{\boldsymbol{g}}\,-\,\boldsymbol{\Phi}_{\text{\tiny L}}\!\parallel^{2}_{L^{2}}+\,\,\upalpha(\hskip 1.13809pt{\boldsymbol{g}},\hskip 1.13809pt\Lambda^{\updelta}_{\sharp}\hskip 1.13809pt{\boldsymbol{g}})_{L^{2}}+\,\upalpha\hskip 1.13809pt\updelta\!\parallel\!{\boldsymbol{g}}\!\parallel^{2}_{L^{2}},\qquad{\boldsymbol{g}}\in L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}).\vspace{-1mm} (38)

Here, α>0\upalpha>0 represents the regularization parameter specified in terms of η\upeta from (30) as

α(δ,L):=η(δ,L)ΛL2+δ.\upalpha(\updelta,L)\,\,\colon\!\!\!=\,\,\frac{\upeta(\updelta,L)}{\parallel\!\Lambda\!\parallel_{L^{2}}+\,\,\updelta}.\vspace{-1mm} (39)

In addition, δ>0\updelta>0 signifies both a measure of perturbation in data and a regularization parameter that, along with α\upalpha, is designed to create a robust imaging indicator as per Theorems 4.3. It should be mentioned that Jα,δJ_{\upalpha,\updelta} is convex and that its minimizer 𝒈α,δL2(𝒢)3×L2(𝒢){\boldsymbol{g}}_{\upalpha,\updelta}\in L^{2}({\mathscr{G}})^{3}\times L^{2}({\mathscr{G}}) solves the linear system

Λδ(Λδ𝒈α,δ𝚽L)+α((Λδ)12(Λδ)12𝒈α,δ+δ𝒈α,δ)=𝟎,\Lambda^{\updelta*}(\Lambda^{\updelta}{\boldsymbol{g}}_{\upalpha,\updelta}-\,\boldsymbol{\Phi}_{\text{\tiny L}})~{}+~{}\upalpha\hskip 1.13809pt\big{(}\hskip 1.13809pt(\Lambda_{\sharp}^{\updelta})^{\frac{1}{2}*}(\Lambda_{\sharp}^{\updelta})^{\frac{1}{2}}\,{\boldsymbol{g}}_{\upalpha,\updelta}+\,\updelta\,{\boldsymbol{g}}_{\upalpha,\updelta}\hskip 1.13809pt\big{)}~{}=~{}\boldsymbol{0},\vspace{-0.5mm} (40)

which can be computed without iterations [23]. Within this framework, the following theorem rigorously establishes the relation between the range of operator 𝒮¯\bar{\mathscr{S}}^{*} and the norm of penalty term in (38).

Theorem 4.3.

Under the assumptions of Theorem 4.1 and additional hypothesis that Λδ\Lambda^{\updelta} and Λδ\Lambda_{\sharp}^{\updelta} are compact,

𝚽LRange(𝒮¯){lim supα0lim supδ0(|(𝒈α,δ,Λδ𝒈α,δ)|+δ𝒈α,δ2)<lim infα0lim infδ0(|(𝒈α,δ,Λδ𝒈α,δ)|+δ𝒈α,δ2)<},\boldsymbol{\Phi}_{\text{\tiny L}}\,\in\,Range(\bar{\mathscr{S}}^{*})~{}~{}\iff~{}~{}\Big{\{}\limsup\limits_{\upalpha\rightarrow 0}\limsup\limits_{\updelta\rightarrow 0}\big{(}\hskip 1.13809pt|({\boldsymbol{g}}_{\upalpha,\updelta},\Lambda^{\updelta}_{\sharp}\hskip 1.13809pt{\boldsymbol{g}}_{\upalpha,\updelta})|\,+\,\updelta\!\parallel\!{\boldsymbol{g}}_{\upalpha,\updelta}\!\parallel^{2}\big{)}\,<\,\infty\\ \iff~{}\liminf\limits_{\upalpha\rightarrow 0}\liminf\limits_{\updelta\rightarrow 0}\big{(}\hskip 1.13809pt|({\boldsymbol{g}}_{\upalpha,\updelta},\Lambda^{\updelta}_{\sharp}\hskip 1.13809pt{\boldsymbol{g}}_{\upalpha,\updelta})|\,+\,\updelta\!\parallel\!{\boldsymbol{g}}_{\upalpha,\updelta}\!\parallel^{2}\big{)}\,<\,\infty\Big{\}},\vspace{-1mm} (41)

where 𝐠α,δ{\boldsymbol{g}}_{\upalpha,\updelta} is a minimizer of the perturbed cost functional (38) in the sense of [23, Lemma 6.8].

In this setting, the imaging indicator 𝔊\mathfrak{G} takes the form

𝔊=1(Λδ)12𝒈α,δ2+δ𝒈α,δ2,\mathfrak{G}\,\,=\,\,\dfrac{1}{\sqrt{\parallel\!\!(\Lambda^{\updelta}_{\sharp})^{\frac{1}{2}}\hskip 1.13809pt{\boldsymbol{g}}_{\upalpha,\updelta}\!\parallel^{2}\hskip 1.13809pt+\,\,\updelta\parallel\!{\boldsymbol{g}}_{\upalpha,\updelta}\!\parallel^{2}}},\vspace{-1mm} (42)

with similar characteristics to 𝔏\mathfrak{L} as in (32) yet more robustness against noise [26].

Remark 4.2.

It is worth noting that the sampling-based characterization of Γ\Gamma from near-field data is deeply rooted in geometrical considerations, so that the fracture indicator functionals (31) and (42) may exhibit only a minor dependence on the complex contact condition – described according to (5) by the heterogeneous distribution of: the stiffness matrix 𝐊(𝛏)\boldsymbol{K}(\boldsymbol{\xi}), permeability coefficient ϰf(𝛏)\varkappa_{\textrm{\tiny f}}(\boldsymbol{\xi}), effective stress coefficient αf(𝛏)\alpha_{\textrm{\tiny f}}(\boldsymbol{\xi}), Skempton coefficient βf(𝛏)\beta_{\textrm{\tiny f}}(\boldsymbol{\xi}), and fluid pressure dissipation factor Π(𝛏)\Pi(\boldsymbol{\xi}) on 𝛏Γ\boldsymbol{\xi}\in\Gamma. This attribute may be traced back to Remark 4.1 where the opening displacement profile 𝐚H~1/2(L)3×H~1/2(L)×H~1/2(L)\boldsymbol{a}\!\in\!\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L) – which is intimately related to the interface law, is deemed arbitrary (within the constraints of admissibility). This quality makes the proposed imaging paradigm particularly attractive in situations where the interfacial parameters are unknown a priori, which opens possibilities for the sequential geometrical reconstruction and interfacial characterization of such anomalies e.g., see [44, 45].

Remark 4.3.

It should be mentioned that there are recent efforts to systematically adapt the 𝔊\mathfrak{G} indicator for application to asymmetric scattering operators [21]. These developments, however, do not lend themselves to poroelastic inverse scattering due to the non-selfadjoint nature of Λ\Lambda. A fundamental treatment of the poroelastic 𝔊\mathfrak{G} indicator in the general case where γ\gamma\in\mathbb{C} could be an interesting subject for a future study.

5 Computational treatment and results

This section examines the performance of multiphasic indicators (31) and (42) through a set of numerical experiments. In what follows, the synthetic data are generated within the FreeFem++ computational platform [46].

5.1 Testing configuration

Two test setups shown in Figs. 2 and 3 are considered where the background is a poroelastic slab P{P} of dimensions 22.322.3 ×\!\times\! 22.322.3 endowed with (evolving and stationary) hydraulic fracture networks. Following [47, 48], the properties of Pecos sandstone are used to characterize P{P}. On setting the reference scales μr=5.85\mu_{r}=5.85 GPA, ρr=103\rho_{r}=10^{3} kg/m3, and r=0.14\ell_{r}=0.14m for stress, mass density, and length, respectively, the non-dimensionalized quantities of Table 1 are obtained and used for simulations. Accordingly, the complex-valued wave speeds [16] affiliated with the modal decomposition in (8) read cs=0.66+8.8×106i\textrm{c}_{\textrm{s}}=0.66+8.8\!\times\!10^{-6}\textrm{i}, cp1=1.26+3×107i\textrm{c}_{\textrm{p}_{1}\!}=1.26+3\!\times\!10^{-7}\textrm{i}, and cp2=5.8×103+5.8×103i\textrm{c}_{\textrm{p}_{2}\!}=5.8\!\times\!10^{-3}+5.8\!\times\!10^{-3}\textrm{i}. The boundary condition on P\partial P is such that the total traction and pore pressure vanish for both incident and scattered fields i.e., (𝒏𝑪:𝒖ι,pι)=(𝒏𝑪:𝒖,p)=𝟎(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u},p)=\boldsymbol{0}. In Setup I, a hydraulic fracture network Γ1Γ9\Gamma_{1}-\Gamma_{9} is induced in four steps as shown in Fig. 2. A detailed description of scatterers including the center (xc,yc)(x_{c},y_{c}), length \ell, and orientation ϕ\phi of each crack Γκ\Gamma_{\kappa}, κ={1,2,,9}\kappa=\{1,2,...,9\} is provided in Table 2. All fractures in this configuration are highly permeable as per Remark 2.1 and characterized by the interfacial stiffness 𝑲(𝝃)=𝟎\boldsymbol{K}(\boldsymbol{\xi})=\boldsymbol{0}, effective stress coefficient αf(𝝃)=0.85{\alpha}_{\textrm{\tiny\emph{f}}}(\boldsymbol{\xi})=0.85, and Skempton coefficient βf(𝝃)=0.3{\beta}_{\textrm{\tiny\emph{f}}}(\boldsymbol{\xi})=0.3 on 𝝃κ=19Γκ\boldsymbol{\xi}\in\!\!{\textstyle\bigcup\limits_{\kappa=1}^{9}\!\!\Gamma_{\!\kappa}}. The latter quantities are taken from [35]. Note that in Setup I, the excitation and sensing grid 𝒢\mathscr{G} straddles two (vertical) monitoring wells and the horizontal section of the injection well. Depicted in Figs. 3 and 4, Setup II features hydraulic fractures Γ10Γ15\Gamma_{10}-\Gamma_{15} of distinct length scales as described in Table 3. The discontinuity interfaces in this configuration are modeled as thin poroelastic inclusions characterized by λf=0.1\lambda_{\text{\sf f}}=0.1, μf=0.2\mu_{\text{\sf f}}=0.2, Mf=0.33M_{\text{\sf f}}=0.33, κf=5×107\kappa_{\text{\sf f}}=5\times 10^{-7}, and ϕf=0.35\phi_{\text{\sf f}}=0.35, while the remaining material parameters are similar to their counterparts in the background as reported in Table 1. In Setup II, the excitation and measurements are solely conducted in the treatment well shown in Fig. 3.

Table 1: Poroelastic properties of the background.
property value dimensionless value
first Lamé parameter (drained) λ=\lambda^{\prime}= 2.74 GPA λ=\lambda= 0.47
drained shear modulus μ=\mu^{\prime}= 5.85 GPA μ=\mu= 1
Biot modulus M=M^{\prime}= 9.71 GPA M=M= 1.66
total density ρ=\rho^{\prime}= 2270 kg/m3 ρ=\rho= 2.27
fluid density ρf=\rho_{f}^{\prime}= 1000 kg/m3 ρf\rho_{f} = 1
apparent mass density ρa=\rho_{a}^{\prime}= 117   kg/m3 ρa=\rho_{a}= 0.117
permeability coefficient κ=\kappa^{\prime}= 0.8 mm4{}^{\text{4}}/N κ=\kappa= 24.5 ×\!\times\! 107{}^{-\text{7}}
porosity ϕ=\phi= 0.195 ϕ=\phi= 0.195
Biot effective stress coefficient α=\alpha= 0.83 α=\alpha= 0.83
frequency ω=\omega^{\prime}= 12 kHz ω=\omega= 3.91
Table 2: Process zone configuration illustrated in Fig. 2: center (xc,yc)(x_{c},y_{c}), length \ell, and orientation ϕ\phi (with respect to xx axis) of cracks Γκ\Gamma_{\kappa}, κ={1,2,,9}\kappa=\{1,2,...,9\}.
​​κ\kappa​​ 1 2 3 4 5 6 7 8 9
​​xc(Γκ)x_{\text{c}}(\Gamma_{\kappa})​​ ​​5.5-5.5​​ ​​0.25-0.25​​ ​​4.34.3​​ ​​3.3-3.3​​ ​​1.61.6​​ ​​4.3-4.3​​ ​​3.4-3.4​​ ​​2-2​​ ​​2.92.9​​
​​yc(Γκ)y_{\text{c}}(\Gamma_{\kappa})​​ ​​0​​ ​​0​​ ​​1-1​​ ​​0​​ ​​0.50.5​​ ​​0.50.5​​ ​​1.07-1.07​​ ​​0​​ ​​0​​
​​(Γκ)\ell\hskip 1.13809pt(\Gamma_{\kappa})​​ ​​33​​ ​​2.22.2​​ ​​33​​ ​​11​​ ​​2.752.75​​ ​​1.81.8​​ ​​1.251.25​​ ​​2.42.4​​ ​​2.22.2​​
​​ϕ(Γκ)\phi\hskip 1.13809pt(\Gamma_{\kappa})​​ ​​0.47π0.47\pi​​ ​​0.6π0.6\pi​​ ​​0.56π0.56\pi​​ ​​0.56π0.56\pi​​ ​​0.42π0.42\pi​​ ​​0.5π0.5\pi​​ ​​0.37π0.37\pi​​ ​​0.5π0.5\pi​​ ​​0.56π0.56\pi​​
Table 3: Process zone configuration illustrated in Figs. 34: center (xc,yc)(x_{c},y_{c}), length \ell, and orientation ϕ\phi (with respect to xx axis) of cracks Γκ\Gamma_{\kappa}, κ={10,11,,15}\kappa=\{10,11,...,15\}.
​​κ\kappa​​ 10 11 12 13 14 15
​​xc(Γκ)x_{\text{c}}(\Gamma_{\kappa})​​ ​​4.76-4.76​​ ​​1.13-1.13​​ ​​3.463.46​​ ​​3.06-3.06​​ ​​1.13-1.13​​ ​​1.161.16​​
​​yc(Γκ)y_{\text{c}}(\Gamma_{\kappa})​​ ​​0​​ ​​0​​ ​​0​​ ​​0​​ ​​0​​ ​​0​​
​​(Γκ)\ell\hskip 1.13809pt(\Gamma_{\kappa})​​ ​​2.352.35​​ ​​0.740.74​​ ​​7.037.03​​ ​​2.352.35​​ ​​0.740.74​​ ​​7.037.03​​
​​ϕ(Γκ)\phi\hskip 1.13809pt(\Gamma_{\kappa})​​ ​​0.3π0.3\pi​​ ​​0.5π0.5\pi​​ ​​0.46π0.46\pi​​ ​​0.3π0.3\pi​​ ​​0.5π0.5\pi​​ ​​0.46π0.46\pi​​

5.2 Forward scattering simulations

Every sensing step entails in-plane harmonic excitation via total body forces 𝒈s{\boldsymbol{g}}_{s} and fluid volumetric sources gfg_{f} at a set of points on 𝒢\mathscr{G}. The excitation frequency ω=3.91\omega=3.91 is set such that the induced shear wavelength λs\lambda_{s} in the specimen is approximately one, serving as a reference length scale. The incident waves interact with the hydraulic fracture network in each setup giving rise to the scattered field (𝒖,p)(\boldsymbol{u},p), governed by (5), whose pattern (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}) over the observation grid 𝒢\mathscr{G} is then computed. For both illumination and sensing purposes, 𝒢\mathscr{G} is sampled by a uniform grid of NN excitation and observation points. In Setup I, the H-shaped incident/observation grid is comprised of N=330N=330 source/receiver points, while in Setup II, the L-shaped support of injection well is uniformly sampled at N=130N=130 points for excitation and sensing.

5.3 Data Inversion

With the preceding data, one may generate the multiphasic indicator maps affiliated with (31) and (42) in three steps, namely by: (1) constructing the discrete scattering operators 𝚲\boldsymbol{\Lambda} and 𝚲δ\boldsymbol{\Lambda}^{\updelta} from synthetic data (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}), (2) computing the trial signature patterns 𝚽L\boldsymbol{\Phi}_{\text{\tiny L}} pertinent to a poroelastic host domain, and (3) evaluating the imaging indicators (𝔏\mathfrak{L} or 𝔊\mathfrak{G}) in the sampling area through careful minimization of the discretized cost functionals (30) and (38) as elucidated in the sequel.

Step 1: construction of the multiphase scattering operator

Given the in-plane nature of wave motion, i.e., that the polarization amplitude of excitation 𝒈s{\boldsymbol{g}}_{s} and the nontrivial components of associated scattered fields 𝒖obs\boldsymbol{u}^{\text{obs}} lay in the xyx-y plane of orthonormal bases (𝒆1,𝒆2)(\boldsymbol{e}_{1},\boldsymbol{e}_{2}), the discretized scattering operator 𝚲\boldsymbol{\Lambda} may be represented by a 3N×3N3N\!\times 3N matrix with components

𝚲(4i+1:4i+4, 4j+1:4j+4)=[u11u12u1u21u22u2\hdashlinep1p2p](𝝃i,𝒚j),i,j=0,N1,\boldsymbol{\Lambda}(4i+1\!:\!4i+4,\,4j+1\!:\!4j+4)~{}=\,\left[\!\begin{array}[]{cc:c}u_{11}\!&\!u_{12}\!&\textsf{u}_{1}\\[2.13394pt] u_{21}\!&\!u_{22}\!&\textsf{u}_{2}\\[3.55658pt] \hdashline\rule{0.0pt}{5.38193pt}p_{1}&p_{2}&\!\textsf{p}\end{array}\!\right](\boldsymbol{\xi}_{i},\boldsymbol{y}_{j}),\qquad i,j=0,\ldots N-1,\vspace{0 mm} (43)

where Λrs(𝝃i,𝒚j){\Lambda}_{rs}(\boldsymbol{\xi}_{i},\boldsymbol{y}_{j}) (r,s=1,2)(r,s\!=\!1,2) is the rthr^{\textrm{th}} component of the solid displacement measured at 𝝃i\boldsymbol{\xi}_{i} due to a unit force applied at 𝒚j\boldsymbol{y}_{j} along the coordinate direction ss and psp_{s} is the associated pore pressure measurement; also, u signifies the 2×12\!\times\!1 solid displacement at 𝝃i\boldsymbol{\xi}_{i} due to a unit injection at 𝒚j\boldsymbol{y}_{j} and p is the affiliated pore pressure. Note that here the general case is presented where excitations and measurements are conducted along all dimensions. 𝚲\boldsymbol{\Lambda} may be downscaled as appropriate according to the testing setup.

Noisy operator. To account for the presence of noise in measurements, we consider the perturbed operator

𝚲δ:=(𝑰+𝑵ϵ)𝚲,\boldsymbol{\Lambda}^{\updelta}\,\,\colon\!\!\!=\,(\boldsymbol{I}+\boldsymbol{N}_{\!\epsilon})\hskip 1.13809pt\boldsymbol{\Lambda},\vspace{1 mm} (44)

where 𝑰\boldsymbol{I} is the 3N×3N3N\times 3N identity matrix, and 𝑵ϵ\boldsymbol{N}_{\!\epsilon} is the noise matrix of commensurate dimension whose components are uniformly-distributed (complex) random variables in [ϵ,ϵ]2[-\epsilon,\,\epsilon]^{2}. In what follows, the measure of noise in data with reference to definition (29) is δ=𝑵ϵ𝚲=0.05\updelta=\,\parallel\!\!\boldsymbol{N}_{\!\epsilon}\hskip 1.13809pt\boldsymbol{\Lambda}\!\!\parallel\,=0.05.

Step 2: generation of a multiphysics library of scattering patterns

This step aims to construct a suitable right hand side for the discretized scattering equation in unbounded and bounded domains pertinent to the analytical developments of Section 4 and numerical experiments of this section, respectively.

Unbounded domain in 3\mathbb{R}^{3}. In this case, the poroelastic trial pattern 𝚽LL2(𝒢)3×L2(𝒢)\boldsymbol{\Phi}_{\text{\tiny L}}\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) is given by Definition 1 indicating that (a) the right hand side is not only a function of the dislocation geometry LL but also a function of the trial density 𝒂H~1/2(L)3×H~1/2(L)×H~1/2(L)\boldsymbol{a}\!\in\!\tilde{H}^{1/2}(L)^{3}\times\tilde{H}^{1/2}(L)\times\tilde{H}^{-1/2}(L), and (b) computing 𝚽L\boldsymbol{\Phi}_{\text{\tiny L}} generally requires an integration process at every sampling point 𝒙\boldsymbol{x}_{\small\circ}. Conventionally, one may dispense with the integration process by considering a sufficiently localized (trial) density function e.g., see [41, 23].

Bounded domain. This case corresponds to the numerical experiments of this section where the background is a poroelastic plate P{P} of finite dimensions. In this setting, following [49], it is straightforward to show that the trial patterns 𝚽L=(𝒖L,pL)\boldsymbol{\Phi}_{\text{\tiny L}}=(\boldsymbol{u}_{\hskip 0.56905pt\text{\tiny L}},p_{\hskip 0.56905pt\text{\tiny L}}) for a finite domain is governed by

(𝑪:𝒖L)(αρfγ)pL+ω2(ρρf2γ)𝒖L= 0\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}_{\text{\tiny L}})\,-\,(\alpha-\dfrac{\rho_{f}}{{\gamma}})\nabla p_{\hskip 0.56905pt\text{\tiny L}}\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{{\gamma}})\boldsymbol{u}_{\text{\tiny L}}\,=\,\boldsymbol{0}\quad in P\L¯,\displaystyle\,\,\,{P}\,\backslash\hskip 1.13809pt\overline{L}, (45)
1γω22pL+M1pL+(αρfγ)𝒖L= 0\displaystyle\dfrac{1}{{\gamma}\omega^{2}}\nabla^{2}p_{\hskip 0.56905pt\text{\tiny L}}\,+\,{M}^{-1}p_{\hskip 0.56905pt\text{\tiny L}}\,+\,(\alpha-\dfrac{\rho_{f}}{{\gamma}})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}_{\text{\tiny L}}\,=\,0\quad in P\L¯\displaystyle\,\,\,{P}\,\backslash\hskip 1.13809pt\overline{L}
(𝒏𝑪:𝒖L,pL)= 0\displaystyle(\boldsymbol{n}\cdot\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}_{\text{\tiny L}},p_{\hskip 0.56905pt\text{\tiny L}})\,=\,\boldsymbol{0}\quad on P,\displaystyle\,\,\partial P,
(𝒖L,pL,qL)=𝒂,𝒕L= 0\displaystyle(\llbracket\boldsymbol{u}_{\text{\tiny L}}\rrbracket,\llbracket\hskip 1.13809ptp_{\hskip 0.56905pt\text{\tiny L}}\rrbracket,-\llbracket\hskip 0.56905ptq_{\hskip 0.56905pt\text{\tiny L}}\rrbracket)\,=\,\boldsymbol{a},\quad\llbracket\hskip 0.56905pt{\boldsymbol{t}}_{\text{\tiny L}}\rrbracket\,=\,\boldsymbol{0}\quad on L.\displaystyle\,\,\,L.

In what follows, the trial signatures 𝚽𝒙n,ι=(u𝒙n,ι,p𝒙n,ι)(𝝃i)\boldsymbol{\Phi}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}=(\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota},\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota})(\boldsymbol{\xi}_{i}) over the observation grid 𝝃i𝒢\boldsymbol{\xi}_{i}\in\mathscr{G} are computed separately for every sampling point 𝒙LP\boldsymbol{x}_{\small\circ}\in L\subset P by solving

(𝑪:u𝒙n,ι)(αρfγ)p𝒙n,ι+ω2(ρρf2γ)u𝒙n,ι= 0\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota})\,-\,(\alpha-\dfrac{\rho_{f}}{{\gamma}})\nabla\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{{\gamma}})\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\,=\,\boldsymbol{0}\quad in P\L¯,\displaystyle\,\,\,{P}\,\backslash\hskip 1.13809pt\overline{L}, (46)
1γω22p𝒙n,ι+M1p𝒙n,ι+(αρfγ)u𝒙n,ι=(1ι)δ(𝒙𝝃)\displaystyle\dfrac{1}{{\gamma}\omega^{2}}\nabla^{2}\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\,+\,{M}^{-1}\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\,+\,(\alpha-\dfrac{\rho_{f}}{{\gamma}})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\,=\,-(1-\upiota)\hskip 1.13809pt\delta(\boldsymbol{x}_{\small\circ}\!-\boldsymbol{\xi})\quad in P\L¯\displaystyle\,\,\,{P}\,\backslash\hskip 1.13809pt\overline{L}
(𝒏𝑪:u𝒙n,ι,p𝒙n,ι)= 0\displaystyle(\boldsymbol{n}\cdot\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota},\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota})\,=\,\boldsymbol{0}\quad on P,\displaystyle\,\,\partial P,
ι(n𝑪:u𝒙n,ι|𝖫|1δ(𝒙𝝃)n,p𝒙n,ι)= 0\displaystyle\upiota\hskip 1.13809pt(\textrm{\bf{n}}\cdot\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}-|{\sf L}|^{-1}\delta(\boldsymbol{x}_{\small\circ}\!-\boldsymbol{\xi})\hskip 0.56905pt\textrm{\bf{n}},\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota})\,=\,\boldsymbol{0}\quad on 𝒙+𝗥𝖫,\displaystyle\,\,\,\boldsymbol{x}_{\small\circ}\!+\boldsymbol{\sf R}{\sf L},
p𝒙n,ι= 0,n𝑪:u𝒙n,ι= 0\displaystyle\llbracket\hskip 1.13809ptp_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\rrbracket\,=\,0,\,\,\,\,\llbracket\hskip 0.56905pt\textrm{\bf{n}}\cdot\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\hskip 0.56905pt\rrbracket\,=\,\boldsymbol{0}\quad on 𝒙+𝗥𝖫.\displaystyle\,\,\,\boldsymbol{x}_{\small\circ}\!+\boldsymbol{\sf R}{\sf L}.

where ι=0,1\upiota=0,1, and the trial dislocation LL is described by an infinitesimal crack L=𝒙+𝗥𝖫L=\boldsymbol{x}_{\small\circ}\!+\boldsymbol{\sf R}{\sf L} wherein 𝗥\boldsymbol{\sf R} is a unitary rotation matrix, and L represents a vanishing penny-shaped crack of unit normal 𝒏:={1,0,0}\boldsymbol{n}_{\small\circ}:=\{1,0,0\}, so that n=𝗥𝒏\textrm{\bf{n}}=\boldsymbol{\sf R}\boldsymbol{n}_{\small\circ}. On recalling (43), the three non-trivial components of (u𝒙n,ι,p𝒙n,ι)(𝝃i)(\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota},\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota})(\boldsymbol{\xi}_{i}) in the xyx-y plane, with orthonormal bases (𝒆1,𝒆2)(\boldsymbol{e}_{1},\boldsymbol{e}_{2}), are arranged into a 3N×13N\!\times\!1 vector for 𝝃i𝒢\boldsymbol{\xi}_{i}\in\mathscr{G} as the following

𝚽𝒙n,ι(3i+1:3i+3)=[u𝒙n,ι𝒆1u𝒙n,ι𝒆2p𝒙n,ι](𝝃i),i=0,N1.\boldsymbol{\Phi}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}(3i+1\!:\!3i+3)~{}=~{}\!\left[\!\begin{array}[]{c}\vspace{0.5 mm}\!\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\!\cdot\!\hskip 1.13809pt\boldsymbol{e}_{1}\!\!\\[0.7113pt] \!\textrm{\bf{u}}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\!\cdot\!\hskip 1.13809pt\boldsymbol{e}_{2}\!\!\\[0.7113pt] \!\textrm{{p}}_{\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota}\!\!\vspace{0.5 mm}\end{array}\!\right]\!\!(\boldsymbol{\xi}_{i}),\qquad i=0,\ldots N-1.\vspace{-0.0 mm} (47)

Sampling. With reference to Figs. 2 and 3, the search area i.e., the sampling region is a square [5,5]2P[-5,5]^{2}\subset P probed by a uniform 100×100100\times 100 grid of sampling points 𝒙\boldsymbol{x}_{\small\circ}\! where the featured indicator functionals are evaluated, while the unit circle spanning possible orientations for the trial dislocation LL is sampled by a grid of 1616 trial normal directions n=𝗥𝒏\textrm{\bf{n}}=\boldsymbol{\sf R}\boldsymbol{n}_{\circ}, and the excitation form – as a fluid source or an elastic force, is selected via ι\upiota. Accordingly, the multiphasic indicator maps 𝔏\mathfrak{L} and 𝔊\mathfrak{G} are constructed through minimizing the cost functionals (30) and (38) for a total of M=10000×8×2M=10000\times 8\times 2 trial triplets (𝒙,n,ι)(\boldsymbol{x}_{\small\circ},\textrm{\bf{n}},\iota).

Remark 5.1.

It is worth mentioning that in the discretized equation

𝚲δ𝒈𝒙n,ι=𝚽𝒙n,ι,\boldsymbol{\!\Lambda}^{\updelta}\,{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ}}~{}=~{}\boldsymbol{\Phi}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota},\vspace{-1 mm} (48)

the scattering operators 𝚲δ\boldsymbol{\Lambda}^{\updelta} is independent of any particular choice of L(𝐱,n)L(\boldsymbol{x}_{\small\circ},\textrm{\bf{n}}) or ι\upiota.Therefore, one may replace 𝚽𝐱n,ι\boldsymbol{\Phi}_{\hskip 0.56905pt\boldsymbol{x}_{\small\circ}}^{\textrm{\bf{n}},\upiota} in (48) with an assembled 3N×M3N\!\times\!M matrix, encompassing all variations of L(𝐱,n)L(\boldsymbol{x}_{\small\circ},\textrm{\bf{n}}) and ι\upiota, to solve only one equation for the reconstruction of hidden fractures which is computationally more efficient.

Step 3: computation of the multiphysics imaging functionals

𝔏\mathfrak{L} map. With reference to Theorem 4.2, (48) is solved by minimizing the regularized cost functional (30). It is well-known, however, that the Tikhonov functional JηJ_{\upeta} is convex and its minimizer can be obtained without iteration [42]. Thus-obtained solution 𝒈𝒙,𝔏n,ι{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{L}} to (48) is a 3N×13N\!\times\!1 vector (or 3N×M3N\!\times\!M matrix for all right-hand sides) identifying the structure of multiphasic wavefront densities on 𝒢\mathscr{G}. On the basis of (31) and (32), the multiphysics 𝔏\mathfrak{L} indicator is then computed as

𝔏(𝒙):=1𝒈𝒙𝔏L2,𝒈𝒙𝔏:=arg​min𝒈𝒙,𝔏n,ι𝒈𝒙,𝔏n,ιL22.\mathfrak{L}(\boldsymbol{x}_{\small\circ})\,\,:=\,\,\frac{1}{\parallel\!{\boldsymbol{g}}^{\mathfrak{L}}_{\boldsymbol{x}_{\small\circ}}\!\!\parallel_{L^{2}}},\qquad{\color[rgb]{0,0,0}{\boldsymbol{g}}^{\mathfrak{L}}_{\boldsymbol{x}_{\small\circ}}\!~{}\colon\!\!\!=~{}\!\!\text{arg\!}\min_{\!\!\!{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{L}}}\parallel\!{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{L}}\!\parallel^{2}_{L^{2}}.}\vspace{-1 mm} (49)

𝔊\mathfrak{G} map. In the case where γ0\Im{\gamma}\to 0, one may construct a more robust approximate solution to (48) by minimizing Jα,δJ_{\upalpha,\updelta} in (38). As indicated in Section 4, Jα,δJ_{\upalpha,\updelta} is also convex and its minimizer 𝒈𝒙,𝔊n,ι{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{G}} can be computed by solving the discretized linear system (40). Then, with reference to (42), the multiphase indicator 𝔊\mathfrak{G} is obtained as

𝔊(𝒙):=1(𝚲δ)12𝒈𝒙𝔊2+δ𝒈𝒙𝔊2,𝒈𝒙𝔊:=arg​min𝒈𝒙,𝔊n,ι𝒈𝒙,𝔊n,ιL22.\mathfrak{G}(\boldsymbol{x}_{\small\circ})\,\,:=\,\,\dfrac{1}{\sqrt{\parallel\!\!(\boldsymbol{\Lambda}^{\updelta}_{\sharp})^{\frac{1}{2}}\hskip 1.13809pt{\boldsymbol{g}}^{\mathfrak{G}}_{\boldsymbol{x}_{\small\circ}}\!\!\parallel^{2}\hskip 1.13809pt+\,\,\delta\!\parallel\!{\boldsymbol{g}}^{\mathfrak{G}}_{\boldsymbol{x}_{\small\circ}}\!\!\parallel^{2}}},\qquad{\boldsymbol{g}}^{\mathfrak{G}}_{\boldsymbol{x}_{\small\circ}}\!~{}\colon\!\!\!=~{}\!\!\text{arg\!}\min_{\!\!\!{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{G}}}\parallel\!{\boldsymbol{g}}^{\textrm{\bf{n}},\upiota}_{\boldsymbol{x}_{\small\circ},\mathfrak{G}}\!\parallel^{2}_{L^{2}}. (50)

5.4 Simulation results

Near-field tomography of an evolving network. With reference to Fig. 2, a hydraulic fracture network κ=19Γκ{\textstyle\bigcup\limits_{\kappa=1}^{9}\!\!\Gamma_{\!\kappa}} is induced in PP in four steps. Following each treatment, the numerical experiments are conducted as described earlier where the multiphasic scattering patterns (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}) are obtained over an H-shaped sensing grid 𝒢\mathscr{G} resulting in a 3N×3N=990×9903N\!\times\!3N=990\!\times\!990 scattering matrix. The latter is then used to compute the distribution of 𝔏\mathfrak{L} indicator in the sampling region, at every sensing step, to recover the sequential evolution of the network as shown in the figure.

Refer to caption
Figure 2: Reconstruction of an evolving hydraulic fracture network via the multiphysics 𝔏\mathfrak{L} indicator: (top row) sensing and network configurations at each treatment stage, and (bottom row) 𝔏\mathfrak{L} maps computed via (49) on the basis of solid displacements and pore pressures (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}) measured on an H-shaped grid 𝝃i𝒢,i={1,2,,330}\boldsymbol{\xi}_{i}\in\mathscr{G},i=\{1,2,...,330\} in the injection and two monitoring wells.
Refer to caption
Figure 3: Limited-aperture reconstruction via the multiphysics 𝔊\mathfrak{G} indicator: (panels 1 and 3) sensing and network configurations, and (panels 2 and 4) the affiliated 𝔊\mathfrak{G} maps constructed via (50) from the complete dataset i.e., solid displacements and pore pressures (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}) observed on an L-shaped grid 𝝃i𝒢,i={1,2,,130}\boldsymbol{\xi}_{i}\in\mathscr{G},i=\{1,2,...,130\} in the injection well.
Refer to caption
Figure 4: Near-field imaging via the pore fluid: (panels 1 and 3) sensing and network configurations similar to Fig. 3, and (panels 2 and 4) the associated 𝔊\mathfrak{G} maps computed via (50) when the excitation at every point 𝝃i𝒢,i={1,2,,130}\boldsymbol{\xi}_{i}\in\mathscr{G},i=\{1,2,...,130\} is a fluid volumetric source gfg_{f} and the measurements are in the form of pore pressure data pobsp^{\text{obs}}.

Limited-aperture imaging. Fig. 3 shows the reconstructed 𝔊\mathfrak{G} maps via (50) from the complete dataset, including both solid displacements and pore pressure measurements (𝒖obs,pobs)(\boldsymbol{u}^{\text{obs}},p^{\text{obs}}), on an L-shaped grid 𝒢\mathscr{G} associated with the injection (or treatment) well, which results in a 3N×3N=390×3903N\!\times\!3N=390\!\times\!390 poroelastic scattering matrix. Here, both networks κ=1012Γκ{\textstyle\bigcup\limits_{\kappa=10}^{12}\!\!\Gamma_{\!\kappa}} and κ=1315Γκ{\textstyle\bigcup\limits_{\kappa=13}^{15}\!\!\Gamma_{\!\kappa}} involve hydraulic fractures of different length scales e.g., Γ11\Gamma_{11} is of O(1)O(1) while Γ12\Gamma_{12} is of O(10)O(10). One may also observe the interaction between scatterers in the 𝔊\mathfrak{G} maps when the pair-wise distances between fractures is less than a few wavelengths. It should be mentioned that similar maps are obtained by deploying the 𝔏\mathfrak{L} imaging functional. However, it is interesting to note that the 𝔊\mathfrak{G} indicator is successful in reconstructing the fracture network even though γ\gamma\in\mathbb{C}. This may be due to the particular sensor arrangement germane to the hydraulic fracturing application where a subset of the sensing grid (within the injection well) is in fact in close proximity of scatterers Γκ\Gamma_{\kappa}, κ=1,2,,15\kappa=1,2,...,15.

Acoustic imaging via the pore fluid. In hydrofracking, it may be convenient to generate distributed fluid volumetric excitation within the treatment well and simultaneously measure thus-induced pore pressures on the same grid. Data inversion on the basis of acoustic excitation and measurements may expedite reconstruction of the induced fractures. In this spirit, Fig. 4 illustrates the reconstruction results for the network and sensing configurations similar to Fig. 3 where the excitation takes only the form of a fluid volumetric source, and the measurements are pore pressure data pobsp^{\text{obs}}. The 𝔏\mathfrak{L} indicator generates very similar maps.

6 Conclusions

A multiphysics data inversion framework is developed for near-field tomography of hydraulic fracture networks in poroelastic backgrounds. This imaging solution is capable of spatiotemporal reconstruction of discontinuous interfaces of arbitrary (and possibly fragmented) support whose poro-mechanical properties are unknown. The data is processed within an appropriate dimensional framework to allows for simultaneous inversion of elastic and acoustic (i.e., pore pressure) measurements, which are distinct in physics and scale. The proposed imaging indicator is (a) inherently non-iterative enabling fast inversion of dense data in support of real-time sensing, (b) flexible with respect to the sensing configuration and illumination frequency, and (c) robust against noise in data. Limited-aperture and partial data inversion – e.g., deep-well tomography as well as imaging based solely on pore pressure excitation and measurements – are explored. Both 𝔏\mathfrak{L} and 𝔊\mathfrak{G} indicators showed success in the numerical experiments germane to the limiting case of high-frequency illumination. It should be mentioned that this imaging modality can be naturally extended to more complex e.g., highly heterogeneous backgrounds. Given the multiphysics nature of data, a remarkable potential would be to deploy this approach within a hybrid data analytic platform to enable not only geometric reconstruction, but also interfacial characterization of discontinuity surfaces, e.g., involving the retrieval of permeability profile of hydraulic fractures from poroelastic data.

7 Acknowledgements

The corresponding author kindly acknowledges the support provided by the National Science Foundation (Grant #1944812). We would like to mention Rezgar Shakeri’s assistance with the preliminary studies on forward scattering by interfaces of infinite permeability. This work utilized resources from the University of Colorado Boulder Research Computing Group, which is supported by the National Science Foundation (awards ACI-1532235 and ACI-1532236), the University of Colorado Boulder, and Colorado State University.

Appendix A Poroelastodynamics fundamental solution

The poroelastodynamics fundamental solution [40, 50] takes the form

G(𝒚,𝝃):=[U𝔰iju𝔣ip𝔰jp𝔣](𝒚,𝝃),𝒚,𝝃3,𝝃𝒚,i,j=1,2,3,{\text{\bf G}}(\boldsymbol{y},\boldsymbol{\xi})\,:=\,\left[\!\!\begin{array}[]{ll}{{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!ij}\!\!&\!\!{{\textrm{u}}^{\mathfrak{f}}}_{\!i}\\[0.28453pt] \hskip 1.13809pt{{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j}\!&\!\!{{\textrm{p}}^{\mathfrak{f}}}\\ \end{array}\!\!\right](\boldsymbol{y},\boldsymbol{\xi}),\qquad\boldsymbol{y},\boldsymbol{\xi}\in\mathbb{R}^{3},\quad\boldsymbol{\xi}\neq\boldsymbol{y},\quad i,j=1,2,3, (51)

where (U𝔰,p𝔰)Hloc1(3{𝒚})3×3×Hloc1(3{𝒚})3({{\textrm{\bf U}}^{\mathfrak{s}}},{{\textrm{\bf p}}^{\mathfrak{s}}})\in H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\boldsymbol{y}\})^{3\times 3}\times H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\boldsymbol{y}\})^{3} satisfies

(𝑪:U𝔰)(𝒚,𝝃)(αρfγ)p𝔰(𝒚,𝝃)+ω2(ρρf2γ)U𝔰(𝒚,𝝃)=δ(𝒚𝝃)𝑰3×3,\displaystyle\nabla\!\cdot\!(\boldsymbol{C}\colon\!\nabla{{\textrm{\bf U}}^{\mathfrak{s}}})(\boldsymbol{y},\boldsymbol{\xi})-(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla{{\textrm{\bf p}}^{\mathfrak{s}}}(\boldsymbol{y},\boldsymbol{\xi})+\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\gamma}){{\textrm{\bf U}}^{\mathfrak{s}}}(\boldsymbol{y},\boldsymbol{\xi})\,=\,-\hskip 1.13809pt\delta(\boldsymbol{y}-\boldsymbol{\xi})\boldsymbol{I}_{3\times 3}, (52)
1γω22p𝔰(𝒚,𝝃)+M1p𝔰(𝒚,𝝃)+(αρfγ)U𝔰(𝒚,𝝃)= 0,𝝃3{𝒚},𝝃𝒚3,\displaystyle\dfrac{1}{\gamma\omega^{2}}\nabla^{2}{{\textrm{\bf p}}^{\mathfrak{s}}}(\boldsymbol{y},\boldsymbol{\xi})+{M}^{-1}{{\textrm{\bf p}}^{\mathfrak{s}}}(\boldsymbol{y},\boldsymbol{\xi})+(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla\!\cdot\!{{\textrm{\bf U}}^{\mathfrak{s}}}(\boldsymbol{y},\boldsymbol{\xi})\,=\,\boldsymbol{0},\quad\boldsymbol{\xi}\in\mathbb{R}^{3}\setminus\{\boldsymbol{y}\},\,\boldsymbol{\xi}\neq\boldsymbol{y}\in\mathbb{R}^{3},

while (u𝔣,p𝔣)Hloc1(3{𝒚})3×Hloc1(3{𝒚})({{\textrm{\bf u}}^{\mathfrak{f}}},{{\textrm{p}}^{\mathfrak{f}}})\in H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\boldsymbol{y}\})^{3}\times H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\boldsymbol{y}\}) solves

(𝑪:u𝔣)(𝒚,𝝃)(αρfγ)p𝔣(𝒚,𝝃)+ω2(ρρf2γ)u𝔣(𝒚,𝝃)= 0,\displaystyle\nabla\!\cdot\!(\boldsymbol{C}\colon\!\nabla{{\textrm{\bf u}}^{\mathfrak{f}}})(\boldsymbol{y},\boldsymbol{\xi})-(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla{{\textrm{p}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})+\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\gamma}){{\textrm{\bf u}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})\,=\,\boldsymbol{0}, (53)
1γω22p𝔣(𝒚,𝝃)+M1p𝔣(𝒚,𝝃)+(αρfγ)u𝔣(𝒚,𝝃)=δ(𝒚𝝃),𝝃3{𝒚},𝝃𝒚3,\displaystyle\dfrac{1}{\gamma\omega^{2}}\nabla^{2}{{\textrm{p}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})+{M}^{-1}{{\textrm{p}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})+(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla\!\cdot\!{{\textrm{\bf u}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})\,=\,-\hskip 1.13809pt\delta(\boldsymbol{y}-\boldsymbol{\xi}),\quad\boldsymbol{\xi}\in\mathbb{R}^{3}\setminus\{\boldsymbol{y}\},\,\boldsymbol{\xi}\neq\boldsymbol{y}\in\mathbb{R}^{3},

both subject to the relevant radiation conditions similar to (9). On setting

r:=|𝒚𝝃|,G(kς,r):=eikςr4πr,ς=s,p1,p2,[f],i:=fξi,i= 1,2,3,\displaystyle r\,:=\,|\boldsymbol{y}-\boldsymbol{\xi}|,\quad G(k_{\varsigma},r)\,:=\,\frac{e^{\text{i}k_{\varsigma}r}}{4\pi r},\,\,\varsigma\,=\,\textrm{s},\textrm{p}_{1},\textrm{p}_{2},\quad[f]_{,i}\,:=\,\frac{\partial f}{\partial\xi_{i}},\,\,i\,=\,1,2,3,
A1:=kp22(kp12Mγω2)γω2(kp12kp22),A2:=kp12(kp22Mγω2)γω2(kp22kp12),\displaystyle\qquad\quad\textrm{A}_{1}\,:=\,\frac{k_{\textrm{p}_{2}\!}^{2}(k_{\textrm{p}_{1}\!}^{2}M-\gamma\omega^{2})}{\gamma\omega^{2}(k_{\textrm{p}_{1}\!}^{2}-k_{\textrm{p}_{2}\!}^{2}\hskip 1.13809pt)},\quad\textrm{A}_{2}\,:=\,\frac{k_{\textrm{p}_{1}\!}^{2}(k_{\textrm{p}_{2}\!}^{2}M-\gamma\omega^{2})}{\gamma\omega^{2}(k_{\textrm{p}_{2}\!}^{2}-k_{\textrm{p}_{1}\!}^{2}\hskip 1.13809pt)},

(U𝔰,p𝔰)({{\textrm{\bf U}}^{\mathfrak{s}}},{{\textrm{\bf p}}^{\mathfrak{s}}}) permits the explicit form

U𝔰ij=γω2(ργρf2)[(G(ks,r)A1G(kp1,r)A2G(kp2,r)),ij+δijks2G(ks,r)],\displaystyle{{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!ij}\,=\,\dfrac{\gamma}{\omega^{2}(\rho\gamma-\rho_{f}^{2})}\Big{[}\Big{(}G(k_{\textrm{s}},r)-\textrm{A}_{1}G(k_{\textrm{p}_{1}\!},r)-\textrm{A}_{2}G(k_{\textrm{p}_{2}\!},r)\Big{)}_{\!,ij}\,+\,\delta_{ij}k_{\textrm{s}}^{2}G(k_{\textrm{s}},r)\Big{]}, (54)
p𝔰j=ω2(αγρf)(λ+2μ)(kp12kp22)(G(kp1,r)G(kp2,r)),j,i,j=1,2,3,\displaystyle\hskip 1.13809pt{{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j}\,=\,\frac{\omega^{2}(\alpha\gamma-\rho_{f})}{(\lambda+2\mu)(k_{\textrm{p}_{1}\!}^{2}-k_{\textrm{p}_{2}\!}^{2}\hskip 1.13809pt)}\Big{(}G(k_{\textrm{p}_{1}\!},r)-G(k_{\textrm{p}_{2}\!},r)\Big{)}_{\!,j},\quad i,j=1,2,3,

and (u𝔣,p𝔣)({{\textrm{\bf u}}^{\mathfrak{f}}},{{\textrm{p}}^{\mathfrak{f}}}) reads

u𝔣=p𝔰,p𝔣=γω2kp22kp12[(kp12ω2(ργρf2)γ(λ+2μ))G(kp1,r)(kp22ω2(ργρf2)γ(λ+2μ))G(kp2,r)],\displaystyle{{\textrm{\bf u}}^{\mathfrak{f}}}\,=\,-{{\textrm{\bf p}}^{\mathfrak{s}}},\qquad{{\textrm{p}}^{\mathfrak{f}}}\,=\,-\frac{\gamma\omega^{2}}{k_{\textrm{p}_{2}\!}^{2}-k_{\textrm{p}_{1}\!}^{2}}\Bigg{[}\Big{(}k_{\textrm{p}_{1}\!}^{2}-\frac{\omega^{2}(\rho\gamma-\rho_{f}^{2})}{\gamma(\lambda+2\mu)}\Big{)}G(k_{\textrm{p}_{1}\!},r)-\Big{(}k_{\textrm{p}_{2}\!}^{2}-\frac{\omega^{2}(\rho\gamma-\rho_{f}^{2})}{\gamma(\lambda+2\mu)}\Big{)}G(k_{\textrm{p}_{2}\!},r)\Bigg{]}, (55)

In this setting, the associated fundamental tractions T𝔰{{\textrm{\bf T}}^{\mathfrak{s}}} and t𝔣{{\textrm{\bf t}}^{\mathfrak{f}}} on a surface Γ3{𝒚}\Gamma\subset\mathbb{R}^{3}\setminus\{\boldsymbol{y}\} characterized by the unit normal vector 𝒏\boldsymbol{n} may be specified as

T𝔰ij(𝒚,𝝃)=nk(𝝃)CikςU𝔰j,ς(𝒚,𝝃)αni(𝝃)p𝔰j(𝒚,𝝃),\displaystyle{{\textrm{T}}^{\mathfrak{s}}}_{\!\!\!ij}(\boldsymbol{y},\boldsymbol{\xi})\,=\,n_{k}(\boldsymbol{\xi})C_{ik\ell\varsigma}{{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!\ell j,\varsigma}(\boldsymbol{y},\boldsymbol{\xi})-\alpha n_{i}(\boldsymbol{\xi}){{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j}(\boldsymbol{y},\boldsymbol{\xi}), (56)
t𝔣i(𝒚,𝝃)=nk(𝝃)Cikςu𝔣,ς(𝒚,𝝃)αni(𝝃)p𝔣(𝒚,𝝃),𝝃Γ,𝒚3Γ¯,\displaystyle\hskip 1.13809pt{{\textrm{t}}^{\mathfrak{f}}}_{\!i}(\boldsymbol{y},\boldsymbol{\xi})\,=\,n_{k}(\boldsymbol{\xi})C_{ik\ell\varsigma}{{\textrm{u}}^{\mathfrak{f}}}_{\!\ell,\varsigma}(\boldsymbol{y},\boldsymbol{\xi})-\alpha n_{i}(\boldsymbol{\xi}){{\textrm{p}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi}),\quad\,\,\boldsymbol{\xi}\in\Gamma,\,\boldsymbol{y}\in\mathbb{R}^{3}\setminus\overline{\Gamma},

where Einstein’s summation convention applies over the repeated indexes. In addition, the normal relative fluid-solid displacements q𝔰{{\textrm{\bf q}}^{\mathfrak{s}}} and q𝔣{{\textrm{q}}^{\mathfrak{f}}} across Γ\Gamma may be expressed as

q𝔰j(𝒚,𝝃)=1γω2(ni(𝝃)p𝔰j,i(𝒚,𝝃)ρfω2ni(𝝃)U𝔰ij(𝒚,𝝃)),\displaystyle{{\textrm{q}}^{\mathfrak{s}}}_{\!\!\!j}(\boldsymbol{y},\boldsymbol{\xi})\,=\,\dfrac{1}{\gamma\omega^{2}}\Big{(}n_{i}(\boldsymbol{\xi}){{\textrm{p}}^{\mathfrak{s}}}_{\!\!\!j,i}(\boldsymbol{y},\boldsymbol{\xi})\,-\,\rho_{f}\omega^{2}n_{i}(\boldsymbol{\xi}){{\textrm{U}}^{\mathfrak{s}}}_{\!\!\!ij}(\boldsymbol{y},\boldsymbol{\xi})\Big{)}, (57)
q𝔣(𝒚,𝝃)=1γω2(ni(𝝃)p𝔣,i(𝒚,𝝃)ρfω2ni(𝝃)u𝔣i(𝒚,𝝃)),𝝃Γ,𝒚3Γ¯.\displaystyle\hskip 1.13809pt{{\textrm{q}}^{\mathfrak{f}}}(\boldsymbol{y},\boldsymbol{\xi})\,=\,\dfrac{1}{\gamma\omega^{2}}\Big{(}n_{i}(\boldsymbol{\xi}){{\textrm{p}}^{\mathfrak{f}}}_{\!\!\!,i}(\boldsymbol{y},\boldsymbol{\xi})\,-\,\rho_{f}\omega^{2}n_{i}(\boldsymbol{\xi}){{\textrm{u}}^{\mathfrak{f}}}_{\!i}(\boldsymbol{y},\boldsymbol{\xi})\Big{)},\quad\,\,\boldsymbol{\xi}\in\Gamma,\,\boldsymbol{y}\in\mathbb{R}^{3}\setminus\overline{\Gamma}.

Appendix B Wellposedness of the poroelastic scattering problem

Given (𝒕ι,qι,pι)H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota})\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma) and the interface operator 𝔓:H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\mathfrak{P}:\tilde{H}^{{1}/{2}}(\Gamma)^{3}\times\tilde{H}^{{1}/{2}}(\Gamma)\times\tilde{H}^{-{1}/{2}}(\Gamma)\rightarrow{H}^{-{1}/{2}}(\Gamma)^{3}\times{H}^{-{1}/{2}}(\Gamma)\times{H}^{{1}/{2}}(\Gamma) satisfying (13), then the direct scattering problem (5)-(9) has a unique solution that continuously depends on (𝒕ι,qι,pι)({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota}). Note that qιq^{\tiny\iota} is continuously dependent on (𝒖ι𝒏,pι𝒏)(\boldsymbol{u}^{\tiny\iota}\!\cdot\!\boldsymbol{n},\nabla p^{\tiny\iota}\!\cdot\!\boldsymbol{n}) according to (6). First, observe that (5)-(9) may be expressed on Γ¯\mathscr{B}\setminus\overline{\Gamma} in the variational form in terms of (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) so that (u,𝚙)Hloc1(3Γ¯)3×Hloc1(3Γ¯)\forall(\texttt{\bf u}^{\prime},\mathtt{p}^{\prime})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) satisfying the radiation condition (9),

(𝒕ι+α~fpι𝒏,qι,pι),(u,p,ϑfp)Γ=\displaystyle\Big{\langle}({\boldsymbol{t}}^{\tiny\iota}+\tilde{\alpha}_{\textrm{\tiny f}}\hskip 1.13809ptp^{\tiny\iota}\hskip 1.13809pt\boldsymbol{n},\,q^{\tiny\iota},\,p^{\tiny\iota}),(\llbracket{{\texttt{\bf u}}}^{\prime}\rrbracket,\llbracket{{\texttt{p}}}^{\prime}\rrbracket,\vartheta_{\textrm{\tiny f}}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt{{\texttt{p}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}})\Big{\rangle}_{\Gamma}\,=\, (58)
Γ{u¯:𝑪:𝒖+1γω2p¯p(αρfγ)(u¯p+p¯𝒖)}dV𝝃\displaystyle\int_{\mathscr{B}\setminus\Gamma}\big{\{}\nabla\bar{{\texttt{\bf u}}}^{\prime}\hskip 1.13809pt\colon\!\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}\,+\,\frac{1}{\gamma\omega^{2}}\nabla\bar{{\texttt{p}}}^{\prime}\!\cdot\!\nabla p\,-\,(\alpha-\frac{\rho_{f}}{\gamma})\hskip 1.13809pt(\nabla\!\cdot\!\bar{{\texttt{\bf u}}}^{\prime}\hskip 0.56905ptp\hskip 1.13809pt+\hskip 1.13809pt\bar{{\texttt{p}}}^{\prime}\nabla\!\cdot\!\boldsymbol{u})\big{\}}\,\text{d}V_{\boldsymbol{\xi}}~{}-
Γ{ω2(ρρf2γ)u¯𝒖+M1p¯p}dV𝝃+ρfγΓ(pu¯+pu¯+p¯𝒖+p¯𝒖)𝒏dS𝝃\displaystyle\int_{\mathscr{B}\setminus\Gamma}\Big{\{}\omega^{2}(\rho-\frac{\rho_{f}^{2}}{\gamma})\hskip 1.13809pt\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}+M^{-1}\bar{{\texttt{p}}}^{\prime}p\Big{\}}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\frac{\rho_{f}}{\gamma}\!\int_{\Gamma}\Big{(}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\llbracket\hskip 0.56905pt\bar{{\texttt{\bf u}}}^{\prime}\rrbracket+\llbracket\hskip 0.56905ptp\hskip 0.56905pt\rrbracket\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\bar{{\texttt{\bf u}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\bar{{\texttt{p}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\llbracket\boldsymbol{u}\rrbracket+\llbracket\hskip 0.56905pt\bar{{\texttt{p}}}^{\prime}\rrbracket\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\boldsymbol{u}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\Big{)}\!\cdot\!\boldsymbol{n}\,\text{d}S_{\boldsymbol{\xi}}\,\,-
(𝒖𝑲+α~fp𝒏,ϰfiωΠp,p+βf1α~fβf𝒏𝑲𝒖),(u,p,ϑfp)Γ\displaystyle\Big{\langle}\big{(}-\llbracket\boldsymbol{u}\rrbracket\!\cdot\!\boldsymbol{K}+\tilde{\alpha}_{\textrm{\tiny f}}\hskip 1.13809pt\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\boldsymbol{n},-\dfrac{\varkappa_{\textrm{\tiny f}}}{\textrm{i}\omega\Pi}\llbracket\hskip 0.56905ptp\rrbracket,\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+\frac{{\beta}_{\textrm{\tiny f}}}{1-\tilde{\alpha}_{\textrm{\tiny f}}{\beta}_{\textrm{\tiny f}}}\boldsymbol{n}\!\cdot\!\boldsymbol{K}\llbracket\boldsymbol{u}\rrbracket\big{)},\big{(}\llbracket{{\texttt{\bf u}}}^{\prime}\rrbracket,\llbracket{{\texttt{p}}}^{\prime}\rrbracket,\vartheta_{\textrm{\tiny f}}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt{{\texttt{p}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\big{)}\Big{\rangle}_{\Gamma}\,\,-
{u¯𝒕+1γω2p¯p𝒏+ρfγu¯𝒏p}dS𝝃,\displaystyle\int_{\partial\mathscr{B}}\Big{\{}\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\hskip 1.13809pt{\boldsymbol{t}}\,+\,\frac{1}{\gamma\omega^{2}}\bar{{\texttt{p}}}^{\prime}\hskip 1.13809pt\nabla p\!\cdot\!\boldsymbol{n}\,+\,\frac{\rho_{f}}{\gamma}\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\boldsymbol{n}\hskip 1.13809ptp\Big{\}}\,\text{d}S_{\boldsymbol{\xi}},

where ϑfL(Γ)\vartheta_{\textrm{\tiny f}}\in L^{\infty}(\Gamma) is defined by

ϑf:=Παfknβf(1α~fβf)¯onΓ,{\vartheta}_{\textrm{\tiny f}}~{}:=~{}\overline{\frac{\Pi\alpha_{\textrm{\tiny f}}}{\textrm{k}_{n}{\beta}_{\textrm{\tiny f}}}({1-\tilde{\alpha}_{\textrm{\tiny f}}{\beta}_{\textrm{\tiny f}}})}\qquad\text{on}\,\,\,\,\Gamma,

with the ‘bar’ indicating complex conjugate. Note that in light of (9), one has

limR{u¯𝒕+1γω2p¯p𝒏+ρfγu¯𝒏p}dS𝝃=0.\lim\limits_{R\rightarrow\infty}\int_{\partial\mathscr{B}}\hskip 1.13809pt\Big{\{}\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\hskip 1.13809pt{\boldsymbol{t}}\,+\,\frac{1}{\gamma\omega^{2}}\bar{{\texttt{p}}}^{\prime}\hskip 1.13809pt\nabla p\!\cdot\!\boldsymbol{n}\,+\,\frac{\rho_{f}}{\gamma}\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\boldsymbol{n}\hskip 1.13809ptp\Big{\}}\,\text{d}S_{\boldsymbol{\xi}}~{}=~{}0.

It is worth mentioning that (58) is obtained by premultiplying the first of (5) by u¯\bar{\texttt{\bf u}}^{\prime} and its second by 𝚙¯\bar{\mathtt{p}}^{\prime}; the results are then added and integrated by parts over Γ¯\mathscr{B}\setminus\overline{\Gamma}. The duality pairing ,Γ\left<\cdot,\cdot\right>_{\Gamma} on the left hand side of (58) is continuous with respect to 𝒕ι{\boldsymbol{t}}^{\tiny\iota}, qιq^{\tiny\iota}, and pιp^{\tiny\iota} owing to the continuity of the trace mapping (u,p)(u,p,p)({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime})\to(\llbracket{{\texttt{\bf u}}}^{\prime}\rrbracket,\llbracket\hskip 0.56905pt{{\texttt{p}}}^{\prime}\rrbracket,\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}\hskip 0.56905pt{\texttt{p}}^{\prime}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}) from Hloc1(3Γ¯)3×Hloc1(3Γ¯)H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) into H~1/2(Γ)3×H~1/2(Γ)×H1/2(Γ)\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times{H}^{1/2}(\Gamma). Now let us define (u,𝚙)Hloc1(3Γ¯)3×Hloc1(3Γ¯)\forall(\texttt{\bf u}^{\prime},\mathtt{p}^{\prime})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) subject to (9), the coercive operator

A((𝒖,p),(u,p)):=Γ{u¯:𝑪:𝒖+1γω2p¯p\displaystyle\textrm{A}\big{(}(\boldsymbol{u},p),({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime})\big{)}~{}:=\,\int_{\mathscr{B}\setminus\Gamma}\big{\{}\nabla\bar{{\texttt{\bf u}}}^{\prime}\hskip 1.13809pt\colon\!\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}\,+\,\frac{1}{\gamma\omega^{2}}\nabla\bar{{\texttt{p}}}^{\prime}\!\cdot\!\nabla p\,\,- (59)
(αρfγ)(u¯p+p¯𝒖)+p¯p+ρf2γω2u¯𝒖}dV𝝃.\displaystyle\hskip 119.50148pt(\alpha-\frac{\rho_{f}}{\gamma})\hskip 1.13809pt(\nabla\!\cdot\!\bar{{\texttt{\bf u}}}^{\prime}\hskip 0.56905ptp\hskip 1.13809pt+\hskip 1.13809pt\bar{{\texttt{p}}}^{\prime}\nabla\!\cdot\!\boldsymbol{u})\,+\,\bar{{\texttt{p}}}^{\prime}p\,+\,\frac{\rho_{f}^{2}}{\gamma}\hskip 1.13809pt\omega^{2}\hskip 1.13809pt\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}\big{\}}\,\text{d}V_{\boldsymbol{\xi}}.

On invoking μr=λ\mu_{r}=\lambda^{\prime} i.e., λ=1\lambda=1 from Section 2.1 and recalling from (4) that |αρf/γ|<1|\alpha-{\rho_{f}}/{\gamma}|<1, one may observe

(A)((𝒖,p),(𝒖,p))Γ{(λ|αρfγ|)𝒖2+μ𝒖2+γγ2ω2p2+\displaystyle\mathfrak{R}(\textrm{A})((\boldsymbol{u},p),(\boldsymbol{u},p))\,\geqslant\,\,\int_{\mathscr{B}\setminus\Gamma}\big{\{}\big{(}\lambda-|\alpha-\frac{\rho_{f}}{\gamma}|\big{)}\!\parallel\!\nabla\!\cdot\!\boldsymbol{u}\!\parallel^{2}+\,\,\mu\parallel\!\nabla\boldsymbol{u}\!\parallel^{2}+\,\,\frac{\mathfrak{R}\gamma}{\gamma^{2}\omega^{2}}\parallel\!\nabla p\!\parallel^{2}\,+\, (60)
(1|αρfγ|)p2+ρf2γ2ω2γ𝒖2}dV𝝃c(𝒖H1(Γ)32+pH1(Γ)2),\displaystyle\hskip 73.97733pt\big{(}1-|\alpha-\frac{\rho_{f}}{\gamma}|\big{)}\parallel\!p\!\parallel^{2}+\,\,\frac{\rho_{f}^{2}}{\gamma^{2}}\hskip 1.13809pt\omega^{2}\mathfrak{R}\gamma\parallel\!\boldsymbol{u}\!\parallel^{2}\!\big{\}}\,\text{d}V_{\boldsymbol{\xi}}\,\geqslant\,\,\textrm{c}\hskip 1.13809pt\big{(}\!\parallel\!\boldsymbol{u}\!\parallel^{2}_{H^{1}(\mathscr{B}\setminus\Gamma)^{3}}+\parallel\!p\!\parallel^{2}_{H^{1}(\mathscr{B}\setminus\Gamma)}\!\big{)},

for a constant c>0\textrm{c}>0 independent of (𝒖,p)(\boldsymbol{u},p). Next, define (u,𝚙)Hloc1(3Γ¯)3×Hloc1(3Γ¯)\forall(\texttt{\bf u}^{\prime},\mathtt{p}^{\prime})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) subject to (9), the compact operator

B((𝒖,p),(u,p))=(𝒖𝑲+α~fp𝒏,ϰfiωΠp,p+βf1α~fβf𝒏𝑲𝒖),(u,p,ϑfp)Γ\displaystyle\textrm{B}\big{(}(\boldsymbol{u},p),({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime})\big{)}~{}=\,-\big{\langle}\big{(}-\llbracket\boldsymbol{u}\rrbracket\!\cdot\!\boldsymbol{K}+\tilde{\alpha}_{\textrm{\tiny f}}\hskip 1.13809pt\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\boldsymbol{n},-\dfrac{\varkappa_{\textrm{\tiny f}}}{\textrm{i}\omega\Pi}\llbracket\hskip 0.56905ptp\rrbracket,\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+\frac{{\beta}_{\textrm{\tiny f}}}{1-\tilde{\alpha}_{\textrm{\tiny f}}{\beta}_{\textrm{\tiny f}}}\boldsymbol{n}\!\cdot\!\boldsymbol{K}\llbracket\boldsymbol{u}\rrbracket\big{)},\big{(}\llbracket{{\texttt{\bf u}}}^{\prime}\rrbracket,\llbracket{{\texttt{p}}}^{\prime}\rrbracket,\vartheta_{\textrm{\tiny f}}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt{{\texttt{p}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\big{)}\big{\rangle}_{\Gamma}\,\,- (61)
Γ{ρω2u¯𝒖+(M1+1)p¯p}dV𝝃+ρfγΓ(pu¯+pu¯+p¯𝒖+p¯𝒖)𝒏dS𝝃,\displaystyle\qquad\quad\int_{\mathscr{B}\setminus\Gamma}\big{\{}\rho\hskip 1.13809pt\omega^{2}\bar{{\texttt{\bf u}}}^{\prime}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}\,+\,(M^{-1}+1)\hskip 0.56905pt\bar{{\texttt{p}}}^{\prime}p\big{\}}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\frac{\rho_{f}}{\gamma}\!\int_{\Gamma}\big{(}\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\llbracket\hskip 0.56905pt\bar{{\texttt{\bf u}}}^{\prime}\rrbracket+\llbracket\hskip 0.56905ptp\hskip 0.56905pt\rrbracket\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\bar{{\texttt{\bf u}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}+\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\bar{{\texttt{p}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\llbracket\boldsymbol{u}\rrbracket+\llbracket\hskip 0.56905pt\bar{{\texttt{p}}}^{\prime}\rrbracket\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\boldsymbol{u}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\big{)}\!\cdot\!\boldsymbol{n}\,\text{d}S_{\boldsymbol{\xi}},

where

|B((𝒖,p),(u,p))|c{𝒖L2(Γ)3uL2(Γ)3+pL2(Γ)pL2(Γ)+\displaystyle\big{|}\textrm{B}\big{(}(\boldsymbol{u},p),({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime})\big{)}\big{|}\leqslant\,\text{c}\hskip 1.13809pt\Big{\{}\parallel\!\!\boldsymbol{u}\!\!\parallel_{L^{2}(\mathscr{B}\setminus\Gamma)^{3}}\,\parallel\!{\texttt{\bf u}}^{\prime}\!\!\parallel_{L^{2}(\mathscr{B}\setminus\Gamma)^{3}}+\parallel\!p\!\parallel_{L^{2}(\mathscr{B}\setminus\Gamma)}\,\parallel\!{\texttt{p}}^{\prime}\!\!\parallel_{L^{2}(\mathscr{B}\setminus\Gamma)}+
𝒖L2(Γ)3uL2(Γ)3+pL2(Γ)pL2(Γ)+pL2(Γ)pL2(Γ)+\displaystyle\quad\,\parallel\!\!\llbracket\boldsymbol{u}\rrbracket\!\!\parallel_{L^{2}(\Gamma)^{3}}\,\parallel\!\!\llbracket{\texttt{\bf u}}^{\prime}\rrbracket\!\!\parallel_{L^{2}(\Gamma)^{3}}+\parallel\!\!\llbracket\hskip 1.13809ptp\hskip 1.13809pt\rrbracket\!\!\parallel_{L^{2}(\Gamma)}\,\parallel\!\!\llbracket\hskip 0.56905pt{\texttt{p}}^{\prime}\rrbracket\!\!\parallel_{L^{2}(\Gamma)}+\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 1.13809ptp\hskip 0.56905pt}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)}\,\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt{\texttt{p}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)}+
pL2(Γ)uL2(Γ)3+pL2(Γ)u¯L2(Γ)3+𝒖L2(Γ)3pL2(Γ)+𝒖L2(Γ)3pL2(Γ)},\displaystyle\quad\,\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)}\hskip 1.13809pt\parallel\!\!\llbracket{\texttt{\bf u}}^{\prime}\rrbracket\!\!\parallel_{L^{2}(\Gamma)^{3}}+\parallel\!\!\llbracket\hskip 0.56905ptp\rrbracket\!\!\parallel_{L^{2}(\Gamma)}\hskip 1.13809pt\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\bar{{\texttt{\bf u}}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)^{3}}+\parallel\!\!\llbracket\boldsymbol{u}\rrbracket\!\!\parallel_{L^{2}(\Gamma)^{3}}\hskip 1.13809pt\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt{\texttt{p}}^{\prime}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)}+\parallel\!\!\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905pt\boldsymbol{u}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}\!\!\parallel_{L^{2}(\Gamma)^{3}}\hskip 1.13809pt\parallel\!\!\llbracket\hskip 0.56905pt{\texttt{p}}^{\prime}\rrbracket\!\!\parallel_{L^{2}(\Gamma)}\!\!\Big{\}},

for a constant c independent of (𝒖,p)(\boldsymbol{u},p) and (u,p)({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime}). Then, the compact embedding of H1(Γ¯)H^{1}(\mathscr{B}\setminus\overline{\Gamma}) into L2(Γ¯)L^{2}(\mathscr{B}\setminus\overline{\Gamma}) and the compactness of the trace operator indicates that B defines a compact perturbation of A((𝒖,p),(u,p))\textrm{A}\big{(}(\boldsymbol{u},p),({{\texttt{\bf u}}}^{\prime},{{\texttt{p}}}^{\prime})\big{)}. Therefore, on letting RR\to\infty, one may observe that the problem (58) is of Fredholm type, and thus, is well-posed as soon as the uniqueness of a solution is guaranteed. Assume that (𝒕ι,qι,pι)=𝟎({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0}. Then, on invoking (4) and (12), one may show that when RR\to\infty,

ωκ𝒒L2(3Γ)32+𝔓(𝒖,p,q),(𝒖,p,q)Γ= 0,-\frac{\omega}{\kappa}\parallel\!{\boldsymbol{q}}\!\parallel^{2}_{L^{2}(\mathbb{R}^{3}\setminus\Gamma)^{3}}+~{}\Im\hskip 1.13809pt\big{\langle}\mathfrak{P}\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 1.13809pt{p}\hskip 0.56905pt\rrbracket,-\llbracket q\rrbracket\big{)},\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 1.13809pt{p}\hskip 0.56905pt\rrbracket,-\llbracket q\rrbracket\big{)}\big{\rangle}_{\Gamma}\,=\,0, (62)

where

𝒒:=1γω2(pρfω2𝒖)in3Γ¯.{\boldsymbol{q}}\,\,:=\,\,\dfrac{1}{\gamma\omega^{2}}\big{(}\nabla p\,-\rho_{f}\omega^{2}\boldsymbol{u}\hskip 0.56905pt\big{)}\qquad\text{in}\,\,\,{\mathbb{R}^{3}}\setminus\overline{\Gamma}. (63)

It should be mentioned that (62) is obtained by (a) premultiplying the first of (5) by 𝒖¯Hloc1(3Γ¯)3\bar{\boldsymbol{u}}\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}, (b) conjugating the second of (5) and multiplying the results by pHloc1(3Γ¯)-{p}\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}), and (c) integrating by parts over Γ¯{\mathscr{B}}\setminus\overline{\Gamma} and adding the results. On letting RR\to\infty, one finds (62) indicating that 𝒒=𝟎{\boldsymbol{q}}=\boldsymbol{0} in 3Γ¯\mathbb{R}^{3}\setminus\overline{\Gamma} by the premise of Remark 2.2. As a result, the first equation in (5) will be reduced to the Navier equation that is subject to the (exponentially decaying) poroelastodynamics radiation conditions (9) which implies that 𝒖=𝟎\boldsymbol{u}=\boldsymbol{0} in 3Γ¯\mathbb{R}^{3}\setminus\overline{\Gamma}. The second equation in (5) then reads p=0p=0 in 3Γ¯\mathbb{R}^{3}\setminus\overline{\Gamma} which completes the proof for the uniqueness of a scattering solution, and thus, substantiates the wellposedness of the forward problem.

Appendix C Proof of Lemma 2.1

Let us define

𝝈ι:=𝑪:𝒖ιαpι𝑰3×3,ζι:=M1pι+α𝒖ι\displaystyle\boldsymbol{\sigma}^{\iota}~{}:=~{}\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}^{\iota}\,-\,\alpha\hskip 1.13809ptp^{\iota}\boldsymbol{I}_{3\times 3},\quad\,\,\,\,\,\zeta^{\iota}~{}:=~{}M^{-1}p^{\iota}\,+\,\alpha\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\iota}\,\,\, in3Γ¯,\displaystyle\text{in}\,\,\,{\mathbb{R}^{3}}\setminus\overline{\Gamma},
𝝈𝒂:=𝑪:𝒖𝒂αp𝒂𝑰3×3,ζ𝒂:=M1p𝒂+α𝒖𝒂\displaystyle\boldsymbol{\sigma}_{\boldsymbol{a}}~{}:=~{}\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}_{\boldsymbol{a}}\,-\,\alpha\hskip 1.13809ptp_{\boldsymbol{a}}\boldsymbol{I}_{3\times 3},\quad\zeta_{\hskip 1.13809pt\boldsymbol{a}}~{}:=~{}M^{-1}p_{\boldsymbol{a}}\,+\,\alpha\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}_{\boldsymbol{a}}\,\,\, in3Γ¯,\displaystyle\text{in}\,\,\,{\mathbb{R}^{3}}\setminus\overline{\Gamma},

where (𝒖ι,pι)Hloc1(3𝒢¯)3×Hloc1(3𝒢¯)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\mathscr{G}})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\mathscr{G}}) and (𝒖𝒂,p𝒂)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}})\in H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}({\mathbb{R}^{3}}\setminus\overline{\Gamma}) solve (3) and (22) respectively. Then, the poroelastic reciprocal theorem [40] reads

𝒖¯𝒂:𝝈ι+ζ¯𝒂pι=𝝈¯𝒂:𝒖ι+p¯𝒂ζι.\nabla\bar{\boldsymbol{u}}_{\boldsymbol{a}}\colon\boldsymbol{\sigma}^{\iota}\,+\,\bar{\zeta}_{\hskip 1.13809pt\boldsymbol{a}}\hskip 1.13809ptp^{\tiny\iota}~{}=~{}\bar{\boldsymbol{\sigma}}_{\boldsymbol{a}}\colon\!\nabla\boldsymbol{u}^{\iota}\,+\,\bar{p}_{\boldsymbol{a}}\hskip 1.13809pt\zeta^{\iota}. (64)

Now, on setting

𝒒ι:=1γω2(pιρfω2𝒖ι),𝒒𝒂:=1γ¯ω2(p𝒂ρfω2𝒖𝒂)in3Γ¯,{\boldsymbol{q}}^{\tiny\iota}\,\,:=\,\,\dfrac{1}{\gamma\omega^{2}}\big{(}\nabla p^{\tiny\iota}\,-\,\rho_{f}\omega^{2}\boldsymbol{u}^{\tiny\iota}\big{)},\quad{\boldsymbol{q}}_{\boldsymbol{a}}\,\,:=\,\,\dfrac{1}{\bar{\gamma}\omega^{2}}\big{(}\nabla p_{\boldsymbol{a}}\,-\,\rho_{f}\omega^{2}\boldsymbol{u}_{\boldsymbol{a}}\big{)}\quad\text{in}\,\,\,{\mathbb{R}^{3}}\setminus\overline{\Gamma}, (65)

observe that

3Γ{𝒖¯𝒂:𝝈ι+ζ¯𝒂pι}dV𝝃:=Γ{𝒖¯𝒂𝒕ιq¯𝒂pι}dS𝝃+𝒢𝒖¯𝒂𝒈sdS𝝃+\displaystyle\int_{\mathbb{R}^{3}\setminus\Gamma}\big{\{}\nabla\bar{\boldsymbol{u}}_{\boldsymbol{a}}\colon\boldsymbol{\sigma}^{\iota}\,+\,\bar{\zeta}_{\hskip 1.13809pt\boldsymbol{a}}\hskip 1.13809ptp^{\tiny\iota}\big{\}}\,\text{d}V_{\boldsymbol{\xi}}~{}:=~{}\!-\int_{\Gamma}\big{\{}\llbracket\bar{\boldsymbol{u}}_{\boldsymbol{a}}\rrbracket\!\cdot\!\hskip 1.13809pt{\boldsymbol{t}}^{\tiny\iota}\,-\,\llbracket\hskip 1.13809pt\bar{q}_{\boldsymbol{a}}\rrbracket\hskip 1.13809ptp^{\tiny\iota}\big{\}}\,\text{d}S_{\boldsymbol{\xi}}\,+\int_{\mathscr{G}}\bar{\boldsymbol{u}}_{\boldsymbol{a}}\!\cdot\!{\boldsymbol{g}}_{s}\,\text{d}S_{\boldsymbol{\xi}}\,\,+
ρω23Γ𝒖¯𝒂𝒖ιdV𝝃+ρfω23Γ{𝒖¯𝒂𝒒ι+𝒒¯𝒂𝒖ι}dV𝝃+γω23Γ𝒒¯𝒂𝒒ιdV𝝃,\displaystyle\quad\qquad\qquad\rho\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\bar{\boldsymbol{u}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\tiny\iota}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\rho_{f}\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\big{\{}\bar{\boldsymbol{u}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt{\boldsymbol{q}}^{\tiny\iota}\,+\,\bar{{\boldsymbol{q}}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\tiny\iota}\big{\}}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\gamma\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\bar{{\boldsymbol{q}}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt{\boldsymbol{q}}^{\tiny\iota}\,\text{d}V_{\boldsymbol{\xi}},
3Γ{𝝈¯𝒂:𝒖ι+p¯𝒂ζι}dV𝝃:=Γp¯𝒂qιdS𝝃𝒢p¯𝒂𝒈fdS𝝃+\displaystyle\int_{\mathbb{R}^{3}\setminus\Gamma}\big{\{}\bar{\boldsymbol{\sigma}}_{\boldsymbol{a}}\colon\!\nabla\boldsymbol{u}^{\iota}\,+\,\bar{p}_{\boldsymbol{a}}\hskip 1.13809pt\zeta^{\iota}\big{\}}\,\text{d}V_{\boldsymbol{\xi}}~{}:=~{}\!\int_{\Gamma}\llbracket\hskip 1.13809pt\bar{p}_{\boldsymbol{a}}\rrbracket\hskip 1.13809ptq^{\tiny\iota}\,\text{d}S_{\boldsymbol{\xi}}\,-\int_{\mathscr{G}}\bar{p}_{\boldsymbol{a}}\!\cdot\!{\boldsymbol{g}}_{f}\,\text{d}S_{\boldsymbol{\xi}}\,\,+
ρω23Γ𝒖¯𝒂𝒖ιdV𝝃+ρfω23Γ{𝒖¯𝒂𝒒ι+𝒒¯𝒂𝒖ι}dV𝝃+γω23Γ𝒒¯𝒂𝒒ιdV𝝃.\displaystyle\quad\qquad\qquad\rho\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\bar{\boldsymbol{u}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\tiny\iota}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\rho_{f}\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\big{\{}\bar{\boldsymbol{u}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt{\boldsymbol{q}}^{\tiny\iota}\,+\,\bar{{\boldsymbol{q}}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\tiny\iota}\big{\}}\,\text{d}V_{\boldsymbol{\xi}}\,+\,\gamma\omega^{2}\!\!\int_{\mathbb{R}^{3}\setminus\Gamma}\bar{{\boldsymbol{q}}}_{\boldsymbol{a}}\!\cdot\!\hskip 1.13809pt{\boldsymbol{q}}^{\tiny\iota}\,\text{d}V_{\boldsymbol{\xi}}.

Then, by invoking (19), (21), and (22), it is evident that

𝒮(𝒈),𝒂Γ=(𝒈,𝒮(𝒂))L2(𝒢),\displaystyle\big{\langle}\mathscr{S}({\boldsymbol{g}}),\boldsymbol{a}\big{\rangle}_{\Gamma}~{}=~{}\big{(}{\boldsymbol{g}},\mathscr{S}^{*}(\boldsymbol{a})\big{)}_{L^{2}(\mathscr{G})},

which completes the proof.

Appendix D Proof of Lemma 3.1

Observe that 𝒮\mathscr{S}^{*} of Lemma 2.1 possesses the integral form

𝒮(𝒂)=Γ𝚺¯(𝝃,𝒚)𝒂(𝒚)dS𝒚,𝚺=[T𝔰q𝔰p𝔰t𝔣q𝔣p𝔣],𝝃𝒢.\displaystyle\mathscr{S}^{*}(\boldsymbol{a})~{}=~{}\int_{\Gamma}\bar{\boldsymbol{\Sigma}}(\boldsymbol{\xi},\boldsymbol{y})\cdot\boldsymbol{a}(\boldsymbol{y})\,\text{d}S_{\boldsymbol{y}},\qquad\boldsymbol{\Sigma}\,=\,\left[\!\!\begin{array}[]{lll}\\[-16.64487pt] {{\textrm{\bf T}}^{\mathfrak{s}}}\!\!&\!\!{{\textrm{\bf q}}^{\mathfrak{s}}}\!\!&\!{{\textrm{\bf p}}^{\mathfrak{s}}}\\[0.0pt] {{\textrm{\bf t}}^{\mathfrak{f}}}\!\!&\!\!{{\textrm{q}}^{\mathfrak{f}}}\!\!&\!{{\textrm{p}}^{\mathfrak{f}}}\end{array}\!\!\right],\qquad\boldsymbol{\xi}\in\mathscr{G}. (66)

Note that the kernel 𝚺(𝝃,𝒚){\boldsymbol{\Sigma}}(\boldsymbol{\xi},\boldsymbol{y}) is derived from the poroelastodynamics fundamental solution of A which also appeared in the integral representation (15). In this setting, the smoothness of 𝚺¯\bar{\boldsymbol{\Sigma}} substantiates the compactness of 𝒮\mathscr{S}^{*} as an application from H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma) into L2(𝒢)3×L2(𝒢)L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}). Next, suppose that there exists 𝒂H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)\boldsymbol{a}\in\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma) such that 𝒮(𝒂)=𝟎\mathscr{S^{*}}(\boldsymbol{a})=\boldsymbol{0}. Then, the vanishing trace of (𝒖𝒂,p𝒂)(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}}) on 𝒢\mathscr{G} implies, by the unique continuation principle, that (𝒖𝒂,p𝒂)=𝟎(\boldsymbol{u}_{\boldsymbol{a}},p_{\boldsymbol{a}})=\boldsymbol{0} in 3\Γ¯\mathbb{R}^{3}\backslash\overline{\Gamma}, and whereby (𝒖𝒂,p𝒂,q𝒂)=𝒂=𝟎(\llbracket\boldsymbol{u}_{\boldsymbol{a}}\rrbracket,\llbracket p_{\boldsymbol{a}}\rrbracket,-\llbracket q_{\boldsymbol{a}}\rrbracket)=\boldsymbol{a}=\boldsymbol{0} which guarantees the injectivity of 𝒮\mathscr{S}^{*}. The densness of the range of 𝒮\mathscr{S}^{*} is established via the injectivity of 𝒮\mathscr{S}. To observe the latter, let (𝒈s,gf)L2(𝒢)3×L2(𝒢)({\boldsymbol{g}}_{s},g_{f})\in L^{2}(\mathscr{G})^{3}\times L^{2}(\mathscr{G}) such that 𝒮(𝒈s,gf)=𝟎\mathscr{S}({\boldsymbol{g}}_{s},g_{f})=\boldsymbol{0}. The definition (19) then reads (𝒏𝑪:𝒖ι,pι)=𝟎(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0} on Γ\Gamma with (𝒖ι,pι)(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota}) satisfying (3). Now, assume that Γ\hskip 1.13809pt\Gamma contains M1M\!\geqslant\!1 (possibly disjoint) analytic surfaces ΓmΓ\Gamma_{m}\!\subset\,\Gamma, m=1,Mm=1,\ldots M, and consider the unique analytic continuation Dm\partial D_{m} of Γm\>\Gamma_{m} identifying an “interior” domain DmD_{m}\!\subset\mathscr{B}. In this setting, (𝒏𝑪:𝒖ι,pι)=𝟎(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0} also on Γm\Gamma_{m} which thanks to the analyticity of 𝒏𝑪:𝒖ι\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota} with respect to the surface coordinates on Dm\partial D_{m} leads to (𝒏𝑪:𝒖ι,pι)=𝟎(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0} on Dm\partial D_{m}. Keeping in mind that ω\omega\in\mathbb{R}, observe that ω\omega may not be an eigenfrequency associated with

(𝑪:𝒖m)(αρfγ)pm+ω2(ρρf2γ)𝒖m= 0\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}^{m})\,-\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla p^{m}\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\gamma})\boldsymbol{u}^{m}\,=\,\boldsymbol{0}\,\,    in Dm,\displaystyle D_{m}, (67)
1γω22pm+M1pm+(αρfγ)𝒖m= 0\displaystyle\dfrac{1}{\gamma\omega^{2}}\nabla^{2}p^{m}\,+\,{M}^{-1}p^{m}\,+\,(\alpha-\dfrac{\rho_{f}}{\gamma})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{m}\,=\,\boldsymbol{0}\,\,    in Dm,\displaystyle D_{m},
(𝒏𝑪:𝒖m,pm)= 0\displaystyle(\boldsymbol{n}\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{m},p^{m})\,=\,\boldsymbol{0}    on Dm,\displaystyle\partial D_{m},

which are inherently complex due to the complexity of γ\gamma, then (𝒖ι,pι)=𝟎(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0} in DmD_{m}. The unique continuation principle then implies that (𝒖ι,pι)=𝟎(\boldsymbol{u}^{\tiny\iota},p^{\tiny\iota})=\boldsymbol{0} in 3\mathbb{R}^{3} and thus (𝒈s,gf)=𝟎({\boldsymbol{g}}_{s},g_{f})=\boldsymbol{0} according to (3) which completes the proof.

Appendix E Proof of Lemma 3.2

The boundedness of TT stems from the well-posedness of the forward problem (5) and trace theorems. To observe (25), let 𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ){\boldsymbol{\psi}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma) and consider (𝒖,p)(\boldsymbol{u},p) satisfying (5) with (𝒕ι,qι,pι)=𝝍({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota})={\boldsymbol{\psi}}. Following similar steps used to obtain (62), one finds 𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ)\forall\hskip 1.13809pt{\boldsymbol{\psi}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma),

𝝍,T𝝍Γ=ωκ𝒒L2(3Γ)2+𝔓(𝒖,p,q),(𝒖,p,q)Γ.\Im\left<{\boldsymbol{\psi}},T{\boldsymbol{\psi}}\right>_{\Gamma}~{}=~{}-\frac{\omega}{\kappa}\parallel\!{\boldsymbol{q}}\!\parallel^{2}_{L^{2}(\mathbb{R}^{3}\setminus\Gamma)}+~{}\Im\hskip 1.13809pt\big{\langle}\mathfrak{P}\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 1.13809pt{p}\hskip 0.56905pt\rrbracket,-\llbracket q\rrbracket\big{)},\big{(}\llbracket\boldsymbol{u}\rrbracket,\llbracket\hskip 1.13809pt{p}\hskip 0.56905pt\rrbracket,-\llbracket q\rrbracket\big{)}\big{\rangle}_{\Gamma}. (68)

Then, the Lemma’s claim follows immediately from (13).

Appendix F Proof of Lemma 3.3

  1. 1.

    First, observe that TT given by (23) is Fredholm with index zero. The argument directly follows the proof of [51, Lemma 2.3] and makes use of the integral representation theorems and asymptotic behavior of the poroelastodynamics fundamental solution on Γ\Gamma [40]. One then further demonstrate the injectivity of TT by assuming that there exists 𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ){\boldsymbol{\psi}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma) so that T(𝝍)=𝟎T({\boldsymbol{\psi}})=\boldsymbol{0}. Consequently, (15) reads that (𝒖,p)=𝟎(\boldsymbol{u},p)\!=\!\boldsymbol{0} in 3\Γ¯\mathbb{R}^{3}\backslash\overline{\Gamma}, which implies that (𝒕,q)=𝟎({\boldsymbol{t}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}q\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}})\!=\!\boldsymbol{0} on Γ\Gamma according to (6). Then, the third to fifth of (5) indicate that 𝝍=𝟎{\boldsymbol{\psi}}=\boldsymbol{0}, and thus the Lemma’s claim follows.

  2. 2.

    We proceed using a contradiction argument. Assume that estimation (26) does not hold true, then one can find a sequence (𝝍n)nH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)({{\boldsymbol{\psi}}}_{n})_{n\in\mathbb{N}^{*}}\subset H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma) such that for all nn\in\mathbb{N}^{*},

    |𝝍n,T(𝝍n)|<1n,𝝍n=1.|\langle{{\boldsymbol{\psi}}}_{n},T({{\boldsymbol{\psi}}}_{n})\rangle|\,\,<\,\,\frac{1}{n},\qquad\quad\|{{\boldsymbol{\psi}}}_{n}\|=1. (69)

    Since (𝝍n)({{\boldsymbol{\psi}}}_{n}) is bounded, one may assume up to extracting a subsequence that (𝝍n)({{\boldsymbol{\psi}}}_{n}) converges weakly to some 𝝍H1/2(Γ)3×H1/2(Γ)×H1/2(Γ){{\boldsymbol{\psi}}}\in H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma). Now, let (𝒖n,pn)Hloc1(3Γ¯)3×Hloc1(3Γ¯)({\boldsymbol{u}}_{n},p_{n})\in H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma})^{3}\times H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma}) solve the poroelastic scattering problem (5) for (𝒕ι,qι,pι)=𝝍n({\boldsymbol{t}}^{\tiny\iota},q^{\tiny\iota},p^{\tiny\iota})={{\boldsymbol{\psi}}}_{n}. In this setting, (68) reads that for 𝝍=𝝍n{{\boldsymbol{\psi}}}={{\boldsymbol{\psi}}}_{n},

    |𝝍n,T(𝝍n)|ωκ𝒒nL2(3Γ¯)32𝔓(𝒖n,pn,qn),(𝒖n,pn,qn)Γ,|\langle{{\boldsymbol{\psi}}}_{n},T({{\boldsymbol{\psi}}}_{n})\rangle|~{}\geqslant~{}\frac{\omega}{\kappa}\parallel\!{\boldsymbol{q}}_{n}\!\parallel^{2}_{L^{2}(\mathbb{R}^{3}\setminus\overline{\Gamma})^{3}}-~{}\Im\hskip 1.13809pt\big{\langle}\mathfrak{P}\big{(}\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 1.13809pt{p_{n}}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\rrbracket\big{)},\big{(}\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 1.13809pt{p_{n}}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\rrbracket\big{)}\big{\rangle}_{\Gamma}, (70)

    wherein

    𝒒n:=1γω2(pnρfω2𝒖n)\displaystyle{\boldsymbol{q}}_{n}~{}:=~{}\frac{1}{\gamma\omega^{2}}\big{(}\nabla p_{n}\,-\,\rho_{f}\omega^{2}\hskip 1.13809pt{\boldsymbol{u}}_{n}\big{)} in3Γ¯,\displaystyle\text{in}\,\,\,\,\mathbb{R}^{3}\setminus\overline{\Gamma}, (71)
    qn:=1γω2(pn𝒏ρfω2𝒖n𝒏)\displaystyle q_{n}~{}:=~{}\frac{1}{\gamma\omega^{2}}\big{(}\nabla p_{n}\!\cdot\!\hskip 0.56905pt\boldsymbol{n}\,-\,\rho_{f}\omega^{2}\hskip 1.13809pt{\boldsymbol{u}}_{n}\!\cdot\!\hskip 0.56905pt\boldsymbol{n}\big{)} onΓ.\displaystyle\text{on}\,\,\,\,\Gamma.

    In light of (13), (69) and (70), observe that 𝒒n{\boldsymbol{q}}_{n} converges strongly to zero in L2(3Γ¯)3L^{2}(\mathbb{R}^{3}\setminus\overline{\Gamma})^{3}. Next, from the wellposedness of the scattering problem and the continuity of the trace operator, one finds that

    (𝒖n,pn,qn)H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)\displaystyle\big{\|}{\big{(}\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 1.13809pt{p_{n}}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\rrbracket\big{)}}\big{\|}_{\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma)}{} c(𝒖n,pn)Hloc1(3Γ¯)3×Hloc1(3Γ¯)\displaystyle\leq\,\text{c}\,\|({\boldsymbol{u}}_{n},p_{n})\|_{H^{1}_{\text{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma})^{3}\times H^{1}_{\text{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma})} (72)
    c𝝍nH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)=c,\displaystyle\leq\,\text{c}^{\prime}\,\|{{\boldsymbol{\psi}}}_{n}\|_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}~{}=~{}\text{c}^{\prime},

    for a constant c>0\text{c}>0 (resp. c>0\text{c}^{\prime}>0) independent of (𝒖n,pn)({\boldsymbol{u}}_{n},p_{n}) (resp. 𝝍n{{\boldsymbol{\psi}}}_{n}). The boundedness of sequences 𝒖n\llbracket\boldsymbol{u}_{n}\rrbracket, pn\llbracket\hskip 1.13809pt{p_{n}}\hskip 0.56905pt\rrbracket and qn\llbracket q_{n}\rrbracket indicates, up to extracting a subsequence, that they respectively converge weakly to some 𝒖\llbracket\boldsymbol{u}\rrbracket, p\llbracket\hskip 1.13809pt{p}\hskip 0.56905pt\rrbracket and q\llbracket q\rrbracket. Then, the compactness of 𝒮¯\bar{\mathscr{S}}^{*}, mapping (𝒖n,pn,qn)(\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 1.13809pt{p_{n}}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\rrbracket) to (𝒖n,pn)({\boldsymbol{u}}_{n},p_{n}) according to (15), implies that (𝒖n,pn)({\boldsymbol{u}}_{n},p_{n}) converges strongly to some (𝒖,p)Hloc1(3Γ¯)3×Hloc1(3Γ¯)(\boldsymbol{u},p)\in H^{1}_{\text{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma})^{3}\times H^{1}_{\text{loc}}(\mathbb{R}^{3}\setminus\overline{\Gamma}). From (5) and (9), one may show that 𝒖n{\boldsymbol{u}}_{n} satisfies

    (𝑪:𝒖n)+(ραρf)ω2𝒖n=(αγρf)ω2𝒒n\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}_{n})\,+\,(\rho-\alpha\rho_{f})\hskip 1.13809pt\omega^{2}\boldsymbol{u}_{n}~{}=~{}(\alpha\gamma-\rho_{f})\hskip 1.13809pt\omega^{2}{\boldsymbol{q}}_{n}\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}\setminus\overline{\Gamma}}, (73)
    uςrikςuς=o(r1e(kς)r),ς=p1,p2,s,u=𝒖n\displaystyle\frac{\partial{\textrm{\bf u}}_{\varsigma}}{\partial r}\,-\,\text{i}k_{\varsigma}{\textrm{\bf u}}_{\varsigma}~{}=~{}o\big{(}r^{-1}e^{-\mathfrak{I}(k_{\varsigma})r}\big{)},\quad\,\,\,\varsigma={\textrm{p}_{1}},{\textrm{p}_{2}},{\textrm{s}},\,\,\,\textrm{\bf u}=\boldsymbol{u}_{n}\quad as r:=|𝝃|.\displaystyle\,\,\,r:=|\boldsymbol{\xi}|\to\infty.

    As nn\to\infty, one obtains that 𝒖\boldsymbol{u} satisfies

    (𝑪:𝒖)+(ραρf)ω2𝒖=𝟎\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u})\,+\,(\rho-\alpha\rho_{f})\hskip 1.13809pt\omega^{2}\boldsymbol{u}~{}=~{}\boldsymbol{0}\quad in 3Γ¯,\displaystyle\,\,\,{\mathbb{R}^{3}\setminus\overline{\Gamma}}, (74)
    𝒖ςrikς𝒖ς=o(r1e(kς)r),ς=p1,p2,s,\displaystyle\frac{\partial{\boldsymbol{u}}_{\varsigma}}{\partial r}\,-\,\text{i}k_{\varsigma}{\boldsymbol{u}}_{\varsigma}~{}=~{}o\big{(}r^{-1}e^{-\mathfrak{I}(k_{\varsigma})r}\big{)},\quad\,\,\,\varsigma={\textrm{p}_{1}},{\textrm{p}_{2}},{\textrm{s}}, as r:=|𝝃|,\displaystyle\,\,\,r:=|\boldsymbol{\xi}|\to\infty,

    implying that 𝒖=𝟎\boldsymbol{u}=\boldsymbol{0}. Note that the conservative Navier system in first of (74) does not generate the exponentially decaying radiation pattern in the second of (74) (which is associated with the lossy Biot system). In this setting, (71) indicates that pp is constant in 3Γ¯\mathbb{R}^{3}\setminus\overline{\Gamma} and since pp is subject to a radiation condition similar to the second of (74), one deduces that p=0p=0. Now, from the wellposedness of the forward scattering problem, one finds that 𝝍=0{{\boldsymbol{\psi}}}=0. On the other hand, observe from (12) that

    𝝍nH1/2(Γ)3×H1/2(Γ)×H1/2(Γ)2=𝝍n,𝔓(𝒖n,pn,qn)Γ𝝍n,(𝒕n,qn,pn)Γ.\|{{\boldsymbol{\psi}}}_{n}\|^{2}_{H^{-1/2}(\Gamma)^{3}\times H^{-1/2}(\Gamma)\times H^{1/2}(\Gamma)}~{}=~{}\big{\langle}{{\boldsymbol{\psi}}}_{n},\mathfrak{P}(\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 0.56905ptp_{n}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\hskip 0.56905pt\rrbracket)\big{\rangle}_{\Gamma}\,-\,\big{\langle}{{\boldsymbol{\psi}}}_{n},({\boldsymbol{t}}_{n},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{q_{n}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp_{n}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}})\big{\rangle}_{\Gamma}. (75)

    Then, the regularity of the trace operator implies that (𝒖n,pn,qn)(\llbracket\boldsymbol{u}_{n}\rrbracket,\llbracket\hskip 0.56905ptp_{n}\hskip 0.56905pt\rrbracket,-\llbracket q_{n}\hskip 0.56905pt\rrbracket) and (𝒕n,qn,pn)({\boldsymbol{t}}_{n},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{q_{n}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}{\hskip 0.56905ptp_{n}}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}}) converges strongly to zero. Consequently, (75) reads that 𝝍n2\|{{\boldsymbol{\psi}}}_{n}\|^{2} goes to zero as nn goes to infinity. This result contradicts (69) and establishes the lemma’s statement.

Appendix G Proof of Lemma 3.5

The injectivity of Λ\Lambda is implied by the injectivity of operators 𝒮{\mathscr{S}}^{*}, TT, and 𝒮\mathscr{S} in the factorization Λ=𝒮¯T𝒮\Lambda=\bar{\mathscr{S}}^{*}\hskip 1.13809ptT\hskip 1.13809pt\mathscr{S} (see Lemmas 3.1 and 3.3). The compactness of Λ\Lambda follows immediately from the compactness of 𝒮\mathscr{S}^{*} – and thus that of 𝒮\mathscr{S} (Lemma 3.1), and the boundedness of TT (Lemma 3.2). The last claim may be verified by establishing the injectivity of Λ=𝒮𝒫\Lambda^{*}=\mathscr{S}^{*}\mathscr{P}^{*} which in light of Lemma 3.1, results from the injectivity of 𝒫\mathscr{P}^{*}. To demonstrate the latter, one first finds from the definition of 𝒫\mathscr{P} that the adjoint operator 𝒫:L2(𝒢)3×L2(𝒢)H~1/2(Γ)3×H~1/2(Γ)×H~1/2(Γ)\mathscr{P}^{*}\colon L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G})\rightarrow\tilde{H}^{1/2}(\Gamma)^{3}\times\tilde{H}^{1/2}(\Gamma)\times\tilde{H}^{-1/2}(\Gamma) is given by 𝒫(𝒈)=(𝒖,p,q)\mathscr{P}^{*}({\boldsymbol{g}})=(\llbracket\hskip 0.56905pt\boldsymbol{u}^{\star}\rrbracket,\llbracket\hskip 1.13809pt{p}^{\star}\hskip 0.56905pt\rrbracket,-\llbracket\hskip 0.56905ptq^{\star}\rrbracket) on Γ\Gamma where (𝒖,p)Hloc1(3{Γ¯𝒢¯})3×Hloc1(3{Γ¯𝒢¯})(\boldsymbol{u}^{\star},p^{\star})\in H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\overline{\Gamma}\cup\overline{\mathscr{G}}\})^{3}\times H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\{\overline{\Gamma}\cup\overline{\mathscr{G}}\}) solves

(𝑪:𝒖)(𝝃)(αρfγ¯)p(𝝃)+ω2(ρρf2γ¯)𝒖(𝝃)=𝒢δ(𝒚𝝃)𝒈s(𝒚)dS𝒚\displaystyle\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt(\boldsymbol{C}\hskip 1.13809pt\colon\!\nabla\boldsymbol{u}^{\star})(\boldsymbol{\xi})\,-\,(\alpha-\dfrac{\rho_{f}}{\bar{\gamma}})\nabla p^{\star}(\boldsymbol{\xi})\,+\,\omega^{2}(\rho-\dfrac{\rho_{f}^{2}}{\bar{\gamma}})\boldsymbol{u}^{\star}(\boldsymbol{\xi})\,=\,-\int_{\mathscr{G}}\delta(\boldsymbol{y}-\boldsymbol{\xi}){\boldsymbol{g}}_{s}(\boldsymbol{y})\hskip 1.13809pt\textrm{d}S_{\boldsymbol{y}}\quad in 3{Γ¯𝒢¯},\displaystyle\,\,\,{\mathbb{R}^{3}\setminus\{\overline{\Gamma}\cup\overline{\mathscr{G}}\}}, (76)
1γ¯ω22p(𝝃)+M1p(𝝃)+(αρfγ¯)𝒖(𝝃)=𝒢δ(𝒚𝝃)gf(𝒚)dS𝒚\displaystyle\dfrac{1}{\bar{\gamma}\omega^{2}}\nabla^{2}p^{\star}(\boldsymbol{\xi})\,+\,{M}^{-1}p^{\star}(\boldsymbol{\xi})\,+\,(\alpha-\dfrac{\rho_{f}}{\bar{\gamma}})\nabla\hskip 1.13809pt\!\cdot\!\hskip 1.13809pt\boldsymbol{u}^{\star}(\boldsymbol{\xi})\,=\,-\int_{\mathscr{G}}\delta(\boldsymbol{y}-\boldsymbol{\xi})g_{f}(\boldsymbol{y})\hskip 1.13809pt\textrm{d}S_{\boldsymbol{y}}\quad in 3{Γ¯𝒢¯},\displaystyle\,\,\,{\mathbb{R}^{3}\setminus\{\overline{\Gamma}\cup\overline{\mathscr{G}}\}},
(𝒕,q,p)=𝔓¯T(𝒖,p,q),𝒕=𝟎\displaystyle({\boldsymbol{t}}^{\star},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}\hskip 1.13809ptq^{\star}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}},\mathopen{\hbox{${\langle}$}\kern-2.52776pt\leavevmode\hbox{${\langle}$}}\hskip 1.13809ptp^{\star}\mathclose{\hbox{${\rangle}$}\kern-2.52776pt\leavevmode\hbox{${\rangle}$}})~{}=~{}\bar{\mathfrak{P}}^{\textrm{T}}(\llbracket\hskip 0.56905pt\boldsymbol{u}^{\star}\rrbracket,\llbracket\hskip 1.13809pt{p}^{\star}\hskip 0.56905pt\rrbracket,-\llbracket\hskip 0.56905ptq^{\star}\rrbracket),\quad\llbracket\hskip 0.56905pt{\boldsymbol{t}}^{\star}\rrbracket~{}=~{}\boldsymbol{0}\quad on Γ,\displaystyle\,\,\,{\Gamma},

wherein ‘T’ indicates transpose, and

[𝒕,q](𝒖,p):=[𝒏𝑪:𝒖αp𝒏,1γ¯ω2(p𝒏ρfω2𝒖𝒏)]onΓ.[{\boldsymbol{t}}^{\star},q^{\star}](\boldsymbol{u}^{\star},p^{\star})\,:=\,\big{[}\boldsymbol{n}\cdot\boldsymbol{C}\colon\!\nabla\boldsymbol{u}^{\star}-\alpha\hskip 1.13809ptp^{\star}\boldsymbol{n},\,\dfrac{1}{\bar{\gamma}\omega^{2}}(\nabla p^{\star}\!\cdot\!\hskip 1.13809pt\boldsymbol{n}-\rho_{f}\omega^{2}\boldsymbol{u}^{\star}\!\cdot\!\hskip 1.13809pt\boldsymbol{n})\big{]}\qquad\text{on}\,\,\,{\Gamma}.

Note that (𝒖,p)(\boldsymbol{u}^{\star},p^{\star}) satisfy the radiation condition as |𝝃||\boldsymbol{\xi}|\rightarrow\infty, similar to the complex conjugate of (9). It should be mentioned that 𝒫\mathscr{P}^{*} is obtained by applying the poroelastodynamics reciprocal theorem between (𝒖,p)(\boldsymbol{u}^{\star},p^{\star}) of (76) and (𝒖,p)(\boldsymbol{u},p) satisfying (5) similar to C. Now, let 𝒈L2(𝒢)3×L2(𝒢){\boldsymbol{g}}\in L^{2}({\mathscr{G}})^{3}\times L^{2}(\mathscr{G}), and assume that 𝒫(𝒈)=𝟎\mathscr{P}^{*}({\boldsymbol{g}})=\boldsymbol{0}, i.e., (𝒖,p,q)=𝟎(\llbracket\hskip 0.56905pt\boldsymbol{u}^{\star}\rrbracket,\llbracket\hskip 1.13809pt{p}^{\star}\hskip 0.56905pt\rrbracket,-\llbracket\hskip 0.56905ptq^{\star}\rrbracket)=\boldsymbol{0}. In this case, (𝒖,p)(\boldsymbol{u}^{\star},p^{\star}) and qq^{\star} are continuous across Γ\Gamma independent of 𝔓¯T\bar{\mathfrak{P}}^{\textrm{T}}. Therefore, (𝒖,p)Hloc1(3𝒢¯)3×Hloc1(3𝒢¯)(\boldsymbol{u}^{\star},p^{\star})\in H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\overline{\mathscr{G}})^{3}\times H^{1}_{\tiny\textrm{loc}}(\mathbb{R}^{3}\setminus\overline{\mathscr{G}}) and (𝒕,q,p)=𝟎({\boldsymbol{t}}^{\star},q^{\star},p^{\star})=\boldsymbol{0}. Thanks to this result, one can show as in the proof of Lemma 3.1 that (𝒖,p)=𝟎(\boldsymbol{u}^{\star},p^{\star})=\boldsymbol{0}, and consequently 𝒈=𝟎{\boldsymbol{g}}=\boldsymbol{0} which concludes the proof.

References

  • [1] H. Hofmann, G. Zimmermann, M. Farkas, E. Huenges, A. Zang, M. Leonhardt, G. Kwiatek, P. Martinez-Garzon, M. Bohnhoff, K.-B. Min, P. Fokker, R. Westaway, F. Bethmann, P. Meier, K. S. Yoon, J. W. Choi, T. J. Lee, K. Y. Kim, First field application of cyclic soft stimulation at the pohang enhanced geothermal system site in korea, Geophysical Journal International 217 (2) (2019) 926–949.
  • [2] R. A. Caulk, I. Tomac, Reuse of abandoned oil and gas wells for geothermal energy production, Renewable Energy 112 (2017) 388–397.
  • [3] Y. Masson, B. Romanowicz, Box tomography: Localized imaging of remote targets buried in an unknown medium, a step forward for understanding key structures in the deep earth, Geophysical Journal International (2017) ggx141.
  • [4] S. Maxwell, Microseismic imaging of hydraulic fracturing: Improved engineering of unconventional shale reservoirs, Society of Exploration Geophysicists, 2014.
  • [5] L. Hou, D. Elsworth, Mechanisms of tripartite permeability evolution for supercritical co2 in propped shale fractures, Fuel 292 (2021) 120188.
  • [6] J.-E. L. Snee, M. D. Zoback, Multiscale variations of the crustal stress field throughout north america, Nature communications 11 (1) (2020) 1–9.
  • [7] M. D. Zoback, A. H. Kohli, Unconventional reservoir geomechanics, Cambridge University Press, 2019.
  • [8] N. Watanabe, G. Blocher, M. Cacace, S. Held, T. Kohl, Geoenergy modeling III enhanced geothermal systems, Springer, 2017.
  • [9] K. Wapenaar, R. Snieder, S. de Ridder, E. Slob, Green’s function representations for marchenko imaging without up/down decomposition, arXiv preprint arXiv:2103.07734 (2021).
  • [10] H. Sun, L. Demanet, Extrapolated full-waveform inversion with deep learning, Geophysics 85 (3) (2020) R275–R288.
  • [11] L. Métivier, A. Allain, R. Brossier, Q. Mérigot, E. Oudet, J. Virieux, Optimal transport for mitigating cycle skipping in full-waveform inversion: A graph-space transform approach, Geophysics 83 (5) (2018) R515–R540.
  • [12] J. P. Verdon, S. A. Horne, A. Clarke, A. L. Stork, A. F. Baird, J.-M. Kendall, Microseismic monitoring using a fiber-optic distributed acoustic sensor array, Geophysics 85 (3) (2020) KS89–KS99.
  • [13] A. Reshetnikov, S. Buske, S. A. Shapiro, Seismic imaging using microseismic events: results from the san andreas fault system at safod, Journal of Geophysical Research: Solid Earth 115 (B12) (2010) B12324.
  • [14] M. Calò, C. Dorbath, F. H. Cornet, N. Cuenot, Large-scale aseismic motion identified through 4-dp-wave tomography, Geophysical Journal International 186 (3) (2011) 1295–1314.
  • [15] W. Gajek, J. Verdon, M. Malinowski, J. Trojanowski, Imaging seismic anisotropy in a shale gas reservoir by combining microseismic and 3d surface reflection seismic data, in: 79th EAGE Conference & Exhibition, Paris, France, 2017.
  • [16] S. A. Shapiro, Fluid-induced seismicity, Cambridge University Press, 2015.
  • [17] A. Baig, T. Urbancic, Microseismic moment tensors: A path to understanding frac growth, The Leading Edge 29 (3) (2010) 320–324.
  • [18] D. Teodor, C. Cesare, F. Khosro Anjom, R. Brossier, V. S. Laura, J. Virieux, Challenges in shallow targets reconstruction by 3d elastic full-waveform inversion: Which initial model?, Geophysics 86 (4) (2021) 1–91.
  • [19] P.-T. Trinh, R. Brossier, L. Métivier, L. Tavard, J. Virieux, Efficient time-domain 3d elastic and viscoelastic full-waveform inversion using a spectral-element method on flexible cartesian-based mesh, Geophysics 84 (1) (2019) R75–R97.
  • [20] L. Audibert, H. Haddar, A generalized formulation of the linear sampling method with exact characterization of targets in terms of farfield measurements, Inverse Problems 30 (2014) 035011.
  • [21] L. Audibert, H. Haddar, The generalized linear sampling method for limited aperture measurements, SIAM Journal on Imaging Sciences 10 (2) (2017) 845–870.
  • [22] M. Bonnet, F. Cakoni, Analysis of topological derivative as a tool for qualitative identification, Inverse Problems 35 (10) (2019) 104007.
  • [23] F. Pourahmadian, H. Haddar, Differential tomography of micromechanical evolution in elastic materials of unknown micro/macrostructure, SIAM Journal on Imaging Sciences 13 (3) (2020) 1302–1330.
  • [24] F. Cakoni, D. Colton, H. Haddar, Inverse Scattering Theory and Transmission Eigenvalues, SIAM, 2016.
  • [25] F. Pourahmadian, Experimental validation of differential evolution indicators for ultrasonic waveform tomography, arxiv preprint arXiv:2010.01813 (2020).
  • [26] F. Pourahmadian, H. Yue, Laboratory application of sampling approaches to inverse scattering, Inverse Problems 37 (5) (2021) 055012.
  • [27] J. G. Rubino, T. M. Müller, L. Guarracino, M. Milani, K. Holliger, Seismoacoustic signatures of fracture connectivity, Journal of Geophysical Research: Solid Earth 119 (3) (2014) 2252–2271.
  • [28] J. G. Rubino, L. Guarracino, T. M. Müller, K. Holliger, Do seismic waves sense fracture connectivity?, Geophysical Research Letters 40 (4) (2013) 692–696.
  • [29] J. Allard, N. Atalla, Propagation of sound in porous media: modeling sound absorbing materials, John Wiley & Sons, 2009.
  • [30] X. Qiao, Z. Shao, W. Bao, Q. Rong, Fiber bragg grating sensors for the oil industry, Sensors 17 (3) (2017) 429.
  • [31] M. A. Biot, Theory of propagation of elastic waves in a fluid-saturated porous solid. ii. higher frequency range, The Journal of the acoustical Society of America 28 (2) (1956) 179–191.
  • [32] M. A. Biot, Mechanics of deformation and acoustic propagation in porous media, Journal of applied physics 33 (4) (1962) 1482–1498.
  • [33] J. D. Hyman, J. Jiménez-Martínez, H. S. Viswanathan, J. W. Carey, M. L. Porter, E. Rougier, S. Karra, Q. Kang, L. Frash, L. Chen, Z. Lei, D. O’Malley, N. Makedonska, Understanding hydraulic fracturing: a multi-scale problem, Philos Trans A Math Phys Eng Sci 374 (2078) (2016) 20150426.
  • [34] G. I. Barenblatt, Scaling (Cambridge texts in applied mathematics), Cambridge University Press, Cambridge, UK, 2003.
  • [35] S. Nakagawa, M. A. Schoenberg, Poroelastic modeling of seismic boundary conditions across a fracture, The Journal of the Acoustical Society of America 122 (2) (2007) 831–847.
  • [36] A. N. Norris, Radiation from a point source and scattering theory in a fluid-saturated porous solid, The Journal of the Acoustical Society of America 77 (6) (1985) 2012–2023.
  • [37] T. Bourbié, O. Coussy, B. Zinszner, M. C. Junger, Acoustics of porous media, Acoustical Society of America, 1992.
  • [38] W. McLean, Strongly Elliptic Systems and Boundary Integral Equations, Cambridge University Press, Cambridge, 2000.
  • [39] A. H.-D. Cheng, T. Badmus, D. E. Beskos, Integral equation for dynamic poroelasticity in frequency domain with bem solution, Journal of engineering mechanics 117 (5) (1991) 1136–1157.
  • [40] M. Schanz, Wave propagation in viscoelastic and poroelastic continua: a boundary element approach, Vol. 2, Springer Science & Business Media, 2012.
  • [41] F. Pourahmadian, B. B. Guzina, H. Haddar, Generalized linear sampling method for elastic-wave sensing of heterogeneous fractures, Inverse Problems 33 (5) (2017) 055007.
  • [42] R. Kress, Linear integral equation, Springer, Berlin, 1999.
  • [43] A. Kirsch, N. Grinberg, The factorization methods for inverse problems, Oxford University Press, Oxford, 2008.
  • [44] F. Pourahmadian, B. B. Guzina, H. Haddar, A synoptic approach to the seismic sensing of heterogeneous fractures: from geometric reconstruction to interfacial characterization, Computer Methods in Applied Mechanics and Engineering 324 (2017) 395–412.
  • [45] F. Pourahmadian, B. B. Guzina, On the elastic anatomy of heterogeneous fractures in rock, International Journal of Rock Mechanics and Mining Sciences 106 (2018) 259–268.
  • [46] F. Hecht, New development in freefem++, Journal of Numerical Mathematics 20 (3-4) (2012) 251–265.
    URL https://freefem.org/
  • [47] B. Ding, A. H.-D. Cheng, Z. Chen, Fundamental solutions of poroelastodynamics in frequency domain based on wave decomposition, Journal of Applied Mechanics 80 (6) (2013).
  • [48] C. H. Yew, P. N. Jogi, Study of wave motions in fluid-saturated porous rocks, The Journal of the Acoustical Society of America 60 (1) (1976) 2–8.
  • [49] T.-P. Nguyen, B. B. Guzina, Generalized linear sampling method for the inverse elastic scattering of fractures in finite bodies, Inverse Problems 35 (10) (2019) 104002.
  • [50] A. H.-D. Cheng, Poroelasticity, Vol. 27, Springer, 2016.
  • [51] F. Cakoni, D. Colton, The linear sampling method for cracks, Inverse Problems. 19 (2003) 279–295.