SUMMARY

Many seismic imaging methods use wavefield extrapolation operators to redatum sources and receivers from the surface into the subsurface. We discuss wavefield extrapolation operators that account for internal multiple reflections, in particular propagator matrices, transfer matrices and Marchenko focusing functions. A propagator matrix is a square matrix that ‘propagates’ a wavefield vector from one depth level to another. It accounts for primaries and multiples and holds for propagating and evanescent waves. A Marchenko focusing function is a wavefield that focuses at a designated point in space at zero time. Marchenko focusing functions are useful for retrieving the wavefield inside a heterogeneous medium from the reflection response at its surface. By expressing these focusing functions in terms of the propagator matrix, the usual approximations (such as ignoring evanescent waves) are avoided. While a propagator matrix acts on the full wavefield vector, a transfer matrix (according to the definition used in this paper) ‘transfers’ a decomposed wavefield vector (containing downgoing and upgoing waves) from one depth level to another. It can be expressed in terms of decomposed Marchenko focusing functions. We present propagator matrices, transfer matrices and Marchenko focusing functions in a consistent way and discuss their mutual relations. In the main text we consider the acoustic situation and in the appendices we discuss other wave phenomena. Understanding these mutual connections may lead to new developments of Marchenko theory and its applications in wavefield focusing, Green’s function retrieval and imaging.

1 INTRODUCTION

In many seismic imaging methods, wavefield extrapolation is used to redatum sources and receivers from the surface to a depth level in the subsurface. In most cases the redatuming process is based on one-way wavefield extrapolation operators, which only account for primaries. To account for internal multiple reflections in redatuming, more advanced wavefield extrapolation operators are required. This paper is not about the redatuming process itself, but about wavefield extrapolation operators that account for internal multiples. In particular, we discuss propagator and transfer matrices, Marchenko focusing functions and their mutual relations.

In elastodynamic wave theory, a propagator matrix is a square matrix that ‘propagates’ a wavefield vector from one depth level to another. It was originally introduced in geophysics for horizontally layered media (Thomson 1950; Haskell 1953; Gilbert & Backus 1966) and later extended for laterally varying media (Kennett 1972). It has been used for modelling surface waves (Woodhouse 1974) and reflection and transmission responses of heterogeneous media (Haines 1988; Kennett et al. 1990; Koketsu et al. 1991; Takenaka et al. 1993). It has also been proposed as an operator for accurate seismic imaging schemes, accounting for high propagation angles (Kosloff & Baysal 1983) and internal multiple reflections (Wapenaar & Berkhout 1986). The wavefield vector that the propagator matrix acts on contains components of the full wavefield (e.g. particle velocity and stress). Here ‘full’ means that the wavefield implicitly consists of downgoing and upgoing, propagating and evanescent waves.

A Marchenko focusing function is a wavefield that focuses at a designated point in space at zero time, accounting for primaries and multiples. Marchenko focusing functions were originally introduced to retrieve the wavefield inside a horizontally layered medium from the reflection response at the boundary of that medium (Rose 2001, 2002; Broggini & Snieder 2012; Slob et al. 2014). This has been extended for laterally varying media (Wapenaar et al. 2013), under the assumption that the wavefield inside the medium can be decomposed into downgoing and upgoing components and that evanescent waves can be neglected. It has recently been shown that the propagator matrix can be expressed in terms of Marchenko focusing functions and vice versa (Wapenaar & de Ridder 2022). Via this relation, the usual assumptions underlying the focusing functions (such as ignoring evanescent waves) are circumvented.

In this paper, we define a transfer matrix as a square matrix that ‘transfers’ decomposed wavefield vectors (explicitly containing downgoing and upgoing waves) from one depth level to another (Born & Wolf 1965; Katsidis & Siapkas 2002). It is different from the propagator matrix, which acts on full wavefield vectors (but please note that in the literature there is not a clear distinction between the use of the terminologies ‘propagator matrix’ and ‘transfer matrix’). It has recently been shown that the transfer matrix can be expressed in terms of decomposed Marchenko focusing functions (Dukalski et al. 2022a, b), an insight that is expected to be useful in further analysis of the minimum-phase property of elastodynamic focusing functions (Reinicke et al. 2023) and beyond.

The aim of this paper is to present propagator matrices, transfer matrices and Marchenko focusing functions in a consistent way and to discuss their mutual relations. We aim to set up the theory as general as possible, accounting for lateral and vertical variations of the medium parameters, accounting for evanescent waves and taking dissipation into account. Whereas in the main text we consider acoustic waves, in the appendices we generalize the theory for other wave phenomena. The numerical examples, which are meant as illustrations of the different quantities and their relations, are restricted to oblique acoustic plane waves in a lossless horizontally layered medium.

We hope that this consistent treatment will contribute to the understanding of the mutual connections and provide insight in the assumptions and approximations that underlie Marchenko-type wavefield retrieval schemes and how to cope with them (Slob 2016; Dukalski et al. 2019, 2022b; Reinicke et al. 2020, 2023; Elison et al. 2020; Diekmann & Vasconcelos 2021; Wapenaar et al. 2021; Kiraz et al. 2023). Moreover, we hope to stimulate new research directions.

The setup of this paper is as follows. In Section 2, we discuss the 2 × 2 propagator matrix for acoustic wavefields and its relation with acoustic Marchenko focusing functions. The advantage of concentrating on the acoustic situation is that all expressions are relatively simple and yet contain all essential aspects. In Section 3, we discuss the 2 × 2 transfer matrix for acoustic wavefields and its relation with decomposed acoustic Marchenko focusing functions. In Section 4, we present some conclusions.

Appendices  A and  B are generalisations of Sections 2 and 3 for other wave phenomena. Here, the propagator and transfer matrices are N × N matrices, with N ranging from 2 for acoustic waves to 12 for seismoelectric waves; the Marchenko focusing functions are N2×N2 matrices. We derive their mutual relations by exploiting general symmetry properties, which are derived in Appendix  C. The appendices not only cover classical waves, but also quantum mechanical waves obeying the Schrödinger equation (N = 2) and the Dirac equation (N = 4).

2 ACOUSTIC PROPAGATOR MATRIX AND FOCUSING FUNCTIONS

2.1 Acoustic matrix–vector wave equation

Our starting point is the following matrix–vector wave equation in the space–frequency domain

(1)

(Woodhouse 1974; Corones 1975; Ursin 1983; Kosloff & Baysal 1983; Fishman & McCoy 1984; Wapenaar & Berkhout 1986; de Hoop 1996). In Appendix  A, we discuss this equation for a range of wave phenomena. Here we consider acoustic waves. For this situation, q is a vector containing the wavefield components p (acoustic pressure) and v3 (vertical component of the particle velocity), both as a function of the space coordinate vector x = (x1, x2, x3) (with positive x3 denoting depth) and the angular frequency ω, hence,

(2)

Operator ∂3 stands for the partial differential operator ∂/∂x3. The space- and frequency-dependent operator matrix A is defined as

(3)

where κ(x, ω) is the compressibility, ρ(x, ω) the mass density and i the imaginary unit. Operator ∂α stands for the partial differential operator ∂/∂xα. Greek subscripts take on the values 1 and 2 and Einstein’s summation convention applies to repeated subscripts, unless otherwise noted. In general the medium may be dissipative, meaning that κ and ρ may be frequency-dependent and complex-valued, with (for positive ω) ℑ(κ) ≥ 0 and ℑ(ρ) ≥ 0, where ℑ denotes the imaginary part. For later convenience we rewrite the operator matrix as follows:

(4)

Here H2(x,ω) is the Helmholtz operator, defined as

(5)

with wavenumber k(x, ω) defined via

(6)

(Brekhovskikh 1960; Wapenaar et al. 2001). Finally, vector d in eq. (1) contains source terms, according to

(7)

Here f^α(x,ω) and f^3(x,ω) are the horizontal and vertical components, respectively, of the external force density (the hats are used to distinguish external force components from focusing functions), and q^(x,ω) is the volume injection-rate density (where q^ is to be distinguished from the wavefield vector q). From here onward we simplify the notation by not explicitly mentioning the frequency-dependency in the argument lists.

2.2 Acoustic propagator matrix

We define a boundary DF at depth level x3 = x3, F. We define a coordinate vector xF at this boundary as xF = (x1, F, x2, F, x3, F) (with fixed x3, F). We introduce the propagator matrix W(x, xF) as a solution of wave eq. (1) for the source-free situation, according to

(8)

with boundary condition

(9)

where I is the identity matrix and xH and xH, F denote the horizontal coordinates of x and xF, respectively, hence xH = (x1, x2) and xH, F = (x1, F, x2, F). Since eqs (1) and (8) are both linear, Huygens’ superposition principle can be applied to get a representation for q(x) in terms of W(x, xF). For a given depth level x3, assuming there are no sources for q(x) between x3, F and x3, we obtain

(10)

(Gilbert & Backus 1966; Kennett 1972; Woodhouse 1974). Note that eq. (10) expresses the ‘propagation’ of q from depth level x3, F to depth level x3, which is why W(x, xF) is called the propagator matrix. It is partitioned as follows:

(11)

where Wp,p, Wp,v, Wv,p and Wv,v are the scalar components of the propagator matrix. For each of these components, the second superscript refers to the wavefield component (p or v3) it acts on at xF, whereas the first superscript refers to the wavefield component it contributes to at x. Eq. (10) is illustrated in the upper-left frame of Fig. 1. The solid line at x3, F denotes the boundary DF (not necessarily a physical boundary). The medium below DF may be inhomogeneous and dissipative. The dashed line at x3 indicates an arbitrary depth level inside the inhomogeneous medium.

Relations between the propagator matrix W(x, xF), the transfer matrix ${{{\rm{\boldsymbol {\it T}}}}}({\bf x},{\bf x}_F)$ and the Marchenko focusing functions $F^p$(x, xF) and $F^v$(x, xF) [right-hand column of Y(x, xF)]. The green and yellow double-sided arrows indicate full wavefields (implicitly consisting of downgoing and upgoing components), whereas the red and blue single-sided arrows indicate decomposed downgoing and upgoing wavefields, respectively.
Figure 1.

Relations between the propagator matrix W(x, xF), the transfer matrix T(x,xF) and the Marchenko focusing functions Fp(x, xF) and Fv(x, xF) [right-hand column of Y(x, xF)]. The green and yellow double-sided arrows indicate full wavefields (implicitly consisting of downgoing and upgoing components), whereas the red and blue single-sided arrows indicate decomposed downgoing and upgoing wavefields, respectively.

By applying eq. (10) recursively, it follows that W obeys the following recursive expression

(12)

where D is a horizontal boundary at a constant depth level x3. By taking x3=x3,F, we obtain from eqs (9) and (12)

(13)

from which it follows that W(xF, x) is the inverse of W(x, xF).

The propagator matrix W(x, xF) accounts for primaries and multiples between x3, F and x3 and it holds for propagating and evanescent waves (for example, Woodhouse 1974, uses the elastodynamic version of the propagator matrix to analyse surface waves). Evanescent field components may lead to instability and should be handled with care (Kennett & Kerry 1979). Since the underlying wave equation is based on the explicit Helmholtz operator H2 (rather than on its square-root, appearing in one-way wave equations), Kosloff & Baysal (1983) argue that the numerical evaluation of eq. (10) converges much faster and for higher propagation angles than schemes based on one-way wave equations. They exploit this property in wide-angle imaging of seismic reflection responses. They use filters to eliminate evanescent and downgoing waves, so they do not exploit the fact that the propagator matrix can handle multiply reflected and evanescent waves. Wapenaar & Berkhout (1986) propose a seismic imaging scheme based on the propagator matrix that handles internal multiple reflections. Since their scheme is very sensitive to the chosen background model it has not found broad applications. In Section 2.3, we show that the propagator matrix can be expressed in terms of Marchenko focusing functions. For a lossless medium, these focusing functions can be derived from seismic reflection data and a smooth background model (Section 2.4). Hence, this leads to a propagator matrix that can be used for seismic imaging, which properly handles internal multiple reflections without being highly sensitive to the background model.

We conclude this section with a numerical illustration of the propagator matrix for the horizontally layered lossless medium of Fig. 2(a). In each layer the propagation velocity c=1/κρ is shown (in m s−1). We define the spatial Fourier transformation of a function u(x, ω) along the horizontal coordinate xH for constant x3 as

(14)

where s = (s1, s2) is the horizontal slowness vector and R is the set of real numbers. For a horizontally layered medium, this transformation decomposes u(x, ω) into independent plane waves, with propagation angle θ (with respect to the vertical axis) obeying sin θ = c|s|. We apply this transformation to the propagator matrix W(x, xF), choosing xF = (0, 0, x3, F). This yields the transformed propagator matrix W~(s,x3,x3,F), with boundary condition W~(s,x3,F,x3,F)=I. Analogous to eq. (12) it obeys the recursive expression

(15)

Next, we define the inverse temporal Fourier transformation for constant s and x3 as

(16)

where ℜ denotes the real part and τ is the intercept time (Stoffa 1989). Applying this transformation to W~(s,x3,x3,F) we obtain W(s, x3, x3, F, τ), with boundary condition W(s, x3, F, x3, F, τ) = Iδ(τ) and with W(s, x3, x3, F, τ) obeying the recursive expression

(17)

where the inline asterisk denotes temporal convolution. Although the numerical modelling is most efficiently done in the slowness-frequency domain (using eq. 15), the results are more conveniently interpreted when displayed in the slowness intercept-time domain. Setting s2 = 0, the components Wp,p(s1, x3, x3, F, τ) and Wp,v(s1, x3, x3, F, τ), with boundary conditions Wp,p(s1, x3, F, x3, F, τ) = δ(τ) and Wp,v(s1, x3, F, x3, F, τ) = 0, are shown in Figs 2(b) and (c) for fixed s1 = 1/3500 s m−1, as a function of intercept time τ and depth x3. To get a smooth display, at each depth the components are convolved with a Ricker wavelet with a central frequency of 50 Hz. The upper traces at x3 = x3, F = 0 m represent the aforementioned boundary conditions. Note that Wp,p and Wp,v are, for each depth x3, even and odd functions, respectively, of intercept time τ. The recursive character, described by eq. (17), is manifest in Figs 2(b) and (c). The propagation velocity in the layer between x3 = 760 m and x3 = 800 m equals 3600 m s−1, which implies that for the chosen horizontal slowness s1 = 1/3500 s m−1 we have sin θ = cs1 > 1 (i.e. θ is complex-valued). Hence, waves become ‘evanescent’ in this layer. The wavefield tunnels through this layer and the amplitudes below this layer are higher than above it. In general, evanescent field components of the propagator matrix should be handled with care, because next to exponentially decaying terms they contain exponentially growing terms that may cause numerical inaccuracies (Kennett & Kerry 1979). In practice this means that beyond a certain horizontal slowness the wavefield should be tapered to zero.

(a) Horizontally layered medium. (b) Propagator matrix component $W^{p,p}$(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (c) Propagator matrix component $W^{p,v}$(s1, x3, x3, F, τ).
Figure 2.

(a) Horizontally layered medium. (b) Propagator matrix component Wp,p(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (c) Propagator matrix component Wp,v(s1, x3, x3, F, τ).

2.3 Relation between acoustic propagator matrix and Marchenko focusing functions

From here onward we assume that the medium at and above DF is homogeneous and may be dissipative, with mass density ρ0 and propagation velocity c0. The medium below DF may be inhomogeneous and dissipative, and it is source free.

In preparation for defining the focusing functions, we decompose operator matrix A (in the space–frequency domain) as follows:

(18)

with

(19)
(20)

(Corones 1975; Fishman & McCoy 1984; Wapenaar & Berkhout 1986; de Hoop 1996). The square-root operator H21/2 is symmetric in the following sense:

(21)

(Wapenaar et al. 2001), where g(xH) and h(xH) are test functions in the horizontal plane with ‘sufficient decay at infinity’. Operator H1, as defined in eq. (19) is not symmetric, but operator 1ρH1 and its inverse, both appearing in eq. (20), are symmetric. We use the operator matrix L to express the wavefield vector q(x) in terms of downgoing and upgoing waves p+(x) and p(x) via

(22)

with

(23)

Note that these equations imply

(24)

hence, the downgoing and upgoing waves p+ and p are pressure-normalized. An advantage of pressure-normalized (or, more generally, field-normalized) decomposition is that the decomposed quantities simply add up to a field quantity [acoustic pressure in the case of eq. (24)]. This property does not apply to flux-normalized decomposed wavefields (Frasier 1970; Kennett et al. 1978; Ursin 1983). On the other hand, an advantage of flux-normalized decomposition is that the underlying equations obey more simple symmetry properties. For a comprehensive discussion on field-normalized versus flux-normalized decomposition in inhomogeneous media, see de Hoop (1996) and Wapenaar (2020). In this paper, we use field-normalized decomposition. In the remainder of Section 2, we apply decomposition only at and above DF, where the medium is assumed to be homogeneous. In Section 3, we will apply decomposition also inside the inhomogeneous medium.

We use eq. (22) at DF to derive focusing functions and express them in the components of the propagator matrix and vice versa. Substituting eq. (22), with x replaced by xF, into the right-hand side of eq. (10) gives

(25)

for x3x3, F, with

(26)

or

(27)

The operators ±1ωρ0H1(xF) in eq. (27) act, via eq. (25), on p±(xF). However, since these operators are symmetric [in the sense of eq. (21)], we may replace the actions of these operators on p±(xF) by actions on the elements Wp,v(x, xF) and Wv,v(x, xF). To be more specific, if we partition Y(x, xF) as follows:

(28)

we obtain from eq. (27) for the elements of this matrix

(29)
(30)

We analyse these expressions one by one. First, we consider the element Yp,. From eqs (25) and (28) it can be seen that the superscript p refers to the acoustic pressure p(x) contained in q(x) and superscript − refers to the upgoing wavefield component p(xF) in p(xF). Using eqs (9), (11) and (29) we obtain

(31)

which is a focusing condition. Hence, we define

(32)

with Fp(x, xF) denoting a focusing function for the acoustic pressure p, which focuses as an upgoing field at x = xF and continues as an upgoing field in the homogeneous upper half-space, see the lower frame of Fig. 1. Next, we consider the element Yv,. Superscript v refers to the vertical particle velocity v3(x) contained in q(x) and superscript − refers again to the upgoing wavefield component p(xF) in p(xF). Using eqs (9), (11) and (30) we obtain

(33)

which is also a focusing condition, but somewhat more complicated than eq. (31) because of the mix of the involved wavefield components v3(x) and p(xF). Hence, we define

(34)

with Fv(x, xF) denoting the particle velocity counterpart of the focusing function Fp(x, xF) (note that the definition of Fv(x, xF) is different from that in Wapenaar (2022), to facilitate the derivations below). The focusing functions Fp(x, xF) and Fv(x, xF), which together form the right-hand column of matrix Y(x, xF), are illustrated in the lower frame of Fig. 1. They resemble the focusing function f2 introduced in previous work (Wapenaar et al. 2013; Slob et al. 2014), which also focuses at the upper boundary (as opposed to the focusing function f1, which focuses inside the medium). However, there are also some notable differences. First, f2(x, xF) is defined in a truncated version of the actual medium and is obtained from a superposition of downgoing and upgoing components, f2+(x,xF) and f2(x,xF) respectively, at x inside the medium (at the lower boundary of the truncated medium). Moreover, representations involving f2+ and f2 ignore evanescent waves at x3, F and x3. In contrast, Fp(x, xF) and Fv(x, xF) are defined in the actual (i.e. untruncated) medium and represent the full pressure and vertical particle velocity at x of a field that focuses at xF at the upper boundary. Since they are derived from the propagator matrix, these focusing functions account for evanescent waves (this will be demonstrated below with a numerical example). The only decomposition takes place at the boundary DF, where the medium is homogeneous. This decomposition, formulated by eqs (32) and (34), accounts for evanescent waves. Last but not least, Fp and Fv hold for dissipative media and they are normalized differently from f2.

Before we analyse the elements in the left-hand column of matrix Y(x, xF), we introduce an adjoint medium, with parameters κ¯(x)=κ(x) and ρ¯(x)=ρ(x). The bar denotes the adjoint medium and the superscript asterisk denotes complex conjugation. When the original medium is dissipative, the adjoint medium is effectual, with (for positive ω) (κ¯)0 and (ρ¯)0. Waves propagating through an effectual medium gain energy (Bojarski 1983; de Hoop 1988). Adjoint media are usually associated to a computational state. The operator matrix A¯ and the Helmholtz operator H¯2 of the adjoint medium are defined similarly as A and H2 in eqs (4) and (5), respectively, but with κ(x) and ρ(x) replaced by κ¯(x) and ρ¯(x), respectively. Hence, H¯2=H2. Analogous to eqs (8) and (9), we define the propagator matrix W¯(x,xF) of the adjoint medium as the solution of 3W¯(x,xF)=A¯(x)W¯(x,xF), with boundary condition W¯(x,xF)|x3=x3,F=Iδ(xHxH,F). In Appendix  C we derive

(35)

(eq. C4). For the square-root operator we have, similar as for the Helmholtz operator,

(36)

(Wapenaar et al. 2001). Using eqs (35) and (36) in eqs (29) and (30), we find Y¯p,+(x,xF)=Yp,(x,xF) and Y¯v,+(x,xF)=Yv,(x,xF). Hence, using eqs (32) and (34), we find for the elements in the left-hand column of matrix Y(x, xF)

(37)
(38)

For matrix Y(x, xF) we thus obtain

(39)

Note that Fp, Fv, F¯p and F¯v are expressed in terms of the components of the propagator matrix W(x, xF) via equations (32), (34), (37) and (38). Conversely, we can express the components of the propagator matrix W(x, xF) in terms of the focusing functions Fp, Fv, F¯p and F¯v. Inverting eq. (26) yields

(40)

with L1 defined in eq. (20). Since operator 1ρH1 is symmetric, its inverse H11ρ is symmetric as well. Hence, in eq. (40) these operators can be taken to act on the elements of matrix Y(x, xF). This yields

(41)
(42)
(43)
(44)

Note that up to this point the medium may be dissipative (and its adjoint effectual), and evanescent wave modes are accounted for, inside the medium as well as at the boundary DF. Hence, the expressions in this section are more general than their counterparts in Wapenaar (2022), which were derived for a lossless medium, under the assumption that evanescent waves can be ignored at DF. If we make the same assumptions here, we can omit the bars on Fp and Fv. For this situation eqs (41)–(44) simplify to

(45)
(46)
(47)
(48)

We illustrate the focusing function and its relation with the propagator matrix with a numerical example. Applying the transformations of eqs (14) and (16) to eq. (32) (assuming a laterally invariant medium), taking xF = (0, 0, x3, F) and s2 = 0, we obtain

(49)

with vertical slowness s3,0=1/c02s12 being the spatial Fourier transform of 1ωH1 at x3, F for the laterally invariant medium (here we assumed s12<1/c02). Eq. (49) shows how a weighted superposition of the even component Wp,p of Fig. 2(b) and the odd component Wp,v of Fig. 2(c) yields the focusing function Fp(s1, x3, x3, F, τ). This focusing function is shown in Fig. 3(a) for s1 = 1/3500 m s−1. The upper trace at x3 = x3, F = 0 m represents the focusing condition Fp(s1, x3, F, x3, F, τ) = δ(τ). At and above x3, F the focusing function is an upgoing field. Note that, similar as in Fig. 2, the wavefield tunnels through the high-velocity layer between x3 = 760 m and x3 = 800 m, which confirms that this focusing function accounts for evanescent waves inside the medium. The time-reversed focusing function Fp(s1, x3, x3, F, −τ) is shown in Fig. 3(b). The focusing function of Fig. 3(a) and its time-reversed version of Fig. 3(b) can be combined to give components of the propagator matrix. To this end, eqs (45) and (46) are transformed to (assuming s12<1/c02)

(50)
(51)

For the acoustic case all components are symmetric in s1, that is Fp( − s1, x3, x3, F, −τ) = Fp(s1, x3, x3, F, −τ), etc. Hence, eqs (50) and (51) show how the even and odd components Wp,p(s1, x3, x3, F, τ) and Wp,v(s1, x3, x3, F, τ) of Figs 2(b) and (c) are obtained from the focusing function Fp(s1, x3, x3, F, τ) and its time-reversal Fp(s1, x3, x3, F, −τ) of Fig. 3.

(a) Focusing function $F^p$(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (b) Time-reversed focusing function $F^p$(s1, x3, x3, F, −τ).
Figure 3.

(a) Focusing function Fp(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (b) Time-reversed focusing function Fp(s1, x3, x3, F, −τ).

2.4 Representations with acoustic Marchenko focusing functions

Substituting the expressions for q(x), p(xF) and Y(x, xF) (eqs 2, 23 and 39) into eq. (25), gives the following representations for the acoustic pressure p(x) and the vertical particle velocity v3(x) inside the inhomogeneous medium

(52)
(53)

for x3x3, F. These expressions are exact and hold for dissipative media. Eq. (52) is a generalisation of eq. (17) of Wapenaar & de Ridder (2022) for dissipative media.

We use eqs (52) and (53) to derive representations for Green’s functions between the boundary DF and any position x inside the medium. To this end, we define a unit point source of vertical force at xS just above DF. For the downgoing field at DF (i.e. just below the source), we then have p+(xF)=12δ(xH,FxH,S), where xH, S denotes the horizontal coordinates of xS. The upgoing field at DF is the reflection response to this downgoing source field, hence p(xF)=12R(xF,xS). The field at x inside the medium is the Green’s response to the source at xS, hence p(x) = Gp,f(x, xS) and v3(x) = Gv,f(x, xS). Here the second superscript (f) refers to the vertical force source at xS, whereas the first superscripts (p and v) refer to the observed quantities (pressure and vertical particle velocity) at x. Substitution of these expressions for p±(xF), p(x) and v3(x) into eqs (52) and (53) gives

(54)
(55)

for x3x3, F. Slob (2016) derived similar representations for decomposed wavefields in dissipative media. In the present derivation we only used decomposition at the boundary DF [similar as Diekmann & Vasconcelos (2021, 2023) and Wapenaar et al. (2021)]. This implies that inside the medium the wavefield does not need to be decomposed into downgoing and upgoing waves and that evanescent waves can be present.

When the medium is lossless and evanescent waves are neglected at DF, the bars on Fp and Fv in representations (54) and (55) can be omitted. Using the Marchenko method, these focusing functions can then be retrieved from the reflection response R(xF, xS) and a smooth background model (Wapenaar et al. 2013; Elison et al. 2020). Since representations (54) and (55) account for evanescent waves inside the medium, the retrieved focusing functions potentially also account for evanescent waves inside the medium (this is subject of current research [Brackenhoff & Wapenaar 2023)]. Once the focusing functions are found, they can be used to retrieve the Green’s functions Gp,f(x, xS) and Gv,f(x, xS) [from eqs (54) and (55)] and all components of the propagator matrix W(x, xF) [from eqs (45) to (48)].

3 ACOUSTIC TRANSFER MATRIX AND DECOMPOSED FOCUSING FUNCTIONS

3.1 Acoustic transfer matrix

We introduce the transfer matrix as follows. Given the downgoing and upgoing fields p+(xF) and p(xF) at the boundary DF, we ‘transfer’ these fields to downgoing and upgoing fields p+(x) and p(x) at any depth level x3 inside the medium using the following expression:

(56)

for x3x3, F. Vectors p(xF) and p(x) contain the downgoing and upgoing fields at depth levels x3, F and x3 (eq. 23). We call T(x,xF) the transfer matrix, which we partition as follows:

(57)

For each component of this matrix, the superscripts refer to the propagation direction at x and at xF, respectively. Eq. (56) is illustrated in the upper-right frame of Fig. 1.

For horizontally layered media, the transfer matrix is usually built up recursively from interface to interface (Born & Wolf 1965; Katsidis & Siapkas 2002; Elison 2020; Dukalski et al. 2022a, b). Here we follow a different approach to derive an expression for T(x,xF) for laterally varying media. Substituting eq. (22) into eq. (10) we obtain eq. (56), with

(58)

with L(x) and its inverse defined in eq. (20). Eq. (58), which relates the transfer matrix T(x,xF) to the propagator matrix W(x, xF), is illustrated in the upper half of Fig. 1. In the next section we show that the transfer matrix can be expressed in terms of decomposed Marchenko focusing functions.

3.2 Relation between acoustic transfer matrix and decomposed Marchenko focusing functions

From eqs (26) and (58) we find

(59)

According to eq. (39), the right-hand column of Y(x, xF) contains Fp(x, xF) and Fv(x, xF), that is, the pressure and vertical particle velocity components at x of the focusing function. Hence, analogous to p(x)=L1(x)q(x), we obtain for the right-hand column of T(x,xF)

(60)

with F+(x, xF) and F(x, xF) being the downgoing and upgoing parts at x of the focusing function Fp(x, xF).

According to eq. (39), the left-hand column of Y(x, xF) contains F¯p(x,xF) and F¯v(x,xF). Hence, for the left-hand column of T(x,xF) we obtain

(61)

or, using H1=H¯1 (eq. 36) and ρ=ρ¯,

(62)

Comparing this with eq. (60) we find that this gives a vector with F¯(x,xF) and F¯+(x,xF). This is the left-hand column of T(x,xF). Hence, we have obtained

(63)

see the upper-right frame of Fig. 1. Hence, the transfer matrix for an inhomogeneous dissipative acoustic medium is expressed in terms of decomposed focusing functions of the medium and its adjoint.

We consider the special case of a horizontally layered medium. Applying the transformations of eqs (14) and (16) to eq. (63), taking xF = (0, 0, x3, F), we obtain

(64)

Dukalski et al. (2022a, b) used a recursive approach and obtained an expression similar to eq. (64). In their derivation they used a path-reversal operator P, which is equivalent with (i) taking the adjoint medium, (ii) taking the complex conjugate (or in the time domain taking the time-reversal) and (iii) changing the sign of the horizontal slowness. Hence, P{F±(s,x3,x3,F,τ)} is equivalent with F¯±(s,x3,x3,F,τ).

For the lossless medium of Fig. 2(a), the decomposed focusing functions F(s1, x3, x3, F, τ) and F+(s1, x3, x3, F, τ) for s1 = 1/3500 m s−1 and s2 = 0 are shown in Figs 4(a) and (b), respectively. For each x3, the function F(s1, x3, x3, F, τ) can be seen as the intricate field that needs to be emitted upward from x3 to arrive as a single upward propagating pulse at the focal depth x3, F at τ = 0. For the same x3, the function F+(s1, x3, x3, F, τ) is the downward reflected response to F(s1, x3, x3, F, τ). Figs 4(b) and (a) together form the right-hand column of the transformed transfer matrix T(s1,x3,x3,F,τ). Their superposition gives the focusing function Fp(s1, x3, x3, F, τ), shown in Fig. 3(a).

(a) Decomposed focusing function F−(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (b) Decomposed focusing function F+(s1, x3, x3, F, τ).
Figure 4.

(a) Decomposed focusing function F(s1, x3, x3, F, τ) (for fixed s1 = 1/3500 m s−1). (b) Decomposed focusing function F+(s1, x3, x3, F, τ).

3.3 Representations with decomposed acoustic Marchenko focusing functions

Substituting the expressions for p(x) and T(x,xF) (eqs 23 and 63) into eq. (56), gives the following representations for the downgoing and upgoing components of the acoustic pressure, p+(x) and p(x), respectively, inside the inhomogeneous medium

(65)
(66)

for x3x3, F. These expressions are exact and hold for dissipative media. Making similar substitutions as in Section 2.4 we obtain

(67)
(68)

for x3x3, F. Here G±, f(x, xS) stands for the downgoing (+) and upgoing (−) part of the Green’s function Gp,f(x, xS).

When the medium is lossless and when evanescent waves are neglected at DF and at depth level x3 inside the medium, the bars on F+ and F in representations (67) and (68) can be omitted. Using the Marchenko method, these decomposed focusing functions can then be retrieved from the reflection response R(xF, xS) and a smooth background model (Wapenaar et al. 2013; Slob et al. 2014). Once the focusing functions are found, they can be used to retrieve the decomposed Green’s functions G+, f(x, xS) and G−, f(x, xS) [from eqs (67) and (68)] and all components of the transfer matrix T(x,xF) [from eq. (63)].

4 CONCLUSIONS

We have derived relations between acoustic propagator matrices, transfer matrices and Marchenko focusing functions. In the appendices we generalize the expressions for other wave phenomena. All relations hold for a heterogeneous dissipative medium below a homogeneous upper half-space and account for propagating and evanescent waves. Only for the transfer matrix beyond the acoustic approach (Appendix  B) we assume that there are no lateral variations at the depth level inside the medium where decomposition takes place.

The derived relations provide insight in the connections between the propagator matrices, transfer matrices and Marchenko focusing functions and may lead to new modelling algorithms for these quantities. Moreover, several of the derived relations may be useful to develop improved Marchenko-type wavefield retrieval and imaging schemes for different wave phenomena, possibly accounting for evanescent waves inside the medium.

ACKNOWLEDGEMENTS

We thank Ivan Vasconcelos and an anonymous reviewer for their useful comments, which helped us to improve this paper. The research of KW has received funding from the European Union’s Horizon 2020 research and innovation programme: European Research Council (grant agreement 742703).

DATA AVAILABILITY

No data have been used for this study.

References

Bojarski
N.N.
,
1983
.
Generalized reaction principles and reciprocity theorems for the wave equations, and the relationship between the time-advanced and time-retarded fields
,
J. acoust Soc. Am.
,
74
,
281
285
.

Born
M.
,
Wolf
E.
,
1965
.
Principles of Optics
,
Pergamon Press
.

Brackenhoff
J.
,
Wapenaar
K.
,
2023
.
On evanescent wave field retrieval with the Marchenko method in 2D settings
,
preprint
(arXiv:2304.03238).

Brekhovskikh
L.M.
,
1960
.
Waves in Layered Media
,
Academic Press

Broggini
F.
,
Snieder
R.
,
2012
.
Connection of scattering principles: a visual and mathematical tour
,
Eur. J. Phys.
,
33
,
593
613
.

Corones
J.P.
,
1975
.
Bremmer series that correct parabolic approximations
,
J. Math. Anal. Appl.
,
50
,
361
372
.

da Costa Filho
C.A.
,
Ravasi
M.
,
Curtis
A.
,
Meles
G.A.
,
2014
.
Elastodynamic Green’s function retrieval through single-sided Marchenko inverse scattering
,
Phys. Rev. E
,
90
,
doi:10.1103/PhysRevE.90.063201
.

de Hoop
A.T.
,
1988
.
Time-domain reciprocity theorems for acoustic wave fields in fluids with relaxation
,
J. acoust Soc. Am.
,
84
,
1877
1882
.

de Hoop
M.V.
,
1996
.
Generalization of the Bremmer coupling series
,
J. Math. Phys.
,
37
,
3246
3282
.

Diekmann
L.
,
Vasconcelos
I.
,
2021
.
Focusing and Green’s function retrieval in three-dimensional inverse scattering revisited: a single-sided Marchenko integral for the full wave field
,
Phys. Rev. Res.
,
3
,
doi:10.1103/PhysRevResearch.3.013206
.

Diekmann
L.
,
Vasconcelos
I.
,
2023
.
Direct, wave-equation-based modeling of Marchenko-type focusing functions
,
JASA Express Lett.
,
3
,
doi:10.1121/10.0017096
.

Dukalski
M.
,
Mariani
E.
,
de Vos
K.
,
2019
.
Handling short-period scattering using augmented Marchenko autofocusing
,
Geophys. J. Int.
,
216
,
2129
2133
.

Dukalski
M.
,
Reinicke
C.
,
Wapenaar
K.
,
2022a
.
Implications of evanescent waves for the Marchenko method through the lens of the transfer-scattering matrix relation
, in
Proceedings of the 83rd EAGE Annual Conference & Exhibition
, Vol.
2022
, pp.
1
4
.,
European Association of Geoscientists & Engineers
.

Dukalski
M.
,
Reinicke
C.
,
Wapenaar
K.
,
2022b
.
Towards understanding the impact of the evanescent elastodynamic mode coupling in Marchenko equation-based demultiple methods
, in
Proceedings of the Second International Meeting for Applied Geoscience and Energy
,
Houston, TX, USA
,
Paper Number: SEG-2022-3745607
.

Elison
P.
,
2020
.
Data-driven focusing and two-way wave modeling with applications to seismic processing and imaging
,
PhD thesis
,
ETH Zürich
,
doi:10.3929/ethz-b-000408969
.

Elison
P.
,
Dukalski
M.S.
,
de Vos
K.
,
van Manen
D.J.
,
Robertsson
J. O.A.
,
2020
.
Data-driven control over short-period internal multiples in media with a horizontally layered overburden
,
Geophys. J. Int.
,
221
,
769
787
.

Fishman
L.
,
McCoy
J.J.
,
1984
.
Derivation and application of extended parabolic wave theories. I. The factorized Helmholtz equation
,
J. Math. Phys.
,
25
(
2
),
285
296
.

Frasier
C.W.
,
1970
.
Discrete time solution of plane P-SV waves in a plane layered medium
,
Geophysics
,
35
,
197
219
.

Gelinsky
S.
,
Shapiro
S.A.
,
1997
.
Poroelastic Backus averaging for anisotropic layered fluid- and gas-saturated sediments
,
Geophysics
,
62
(
6
),
1867
1878
.

Gilbert
F.
,
Backus
G.E.
,
1966
.
Propagator matrices in elastic wave and vibration problems
,
Geophysics
,
31
(
2
),
326
332
.

Haartsen
M.W.
,
Pride
S.R.
,
1997
.
Electroseismic waves from point sources in layered media
,
J. geophys. Res.
,
102
,
24 745
24 769
.

Haines
A.J.
,
1988
.
Multi-source, multi-receiver synthetic seismograms for laterally heterogeneous media using F-K domain propagators
,
Geophys. J. Int.
,
95
,
237
260
.

Haskell
N.A.
,
1953
.
The dispersion of surface waves on multilayered media
,
Bull. seism. Soc. Am.
,
43
,
17
34
.

Katsidis
C.C.
,
Siapkas
D.I.
,
2002
.
General transfer-matrix method for optical multilayer systems with coherent, partially coherent, and incoherent interference
,
Appl. Opt.
,
41
(
19
),
3978
3987
.

Kennett
B.L.N.
,
1972
.
Seismic waves in laterally inhomogeneous media
,
Geophys. J. R. astr. Soc.
,
27
,
301
325
.

Kennett
B. L.N.
,
Kerry
N.J.
,
1979
.
Seismic waves in a stratified half-space
,
Geophys. J. R. astr. Soc.
,
57
,
557
584
.

Kennett
B. L.N.
,
Kerry
N.J.
,
Woodhouse
J.H.
,
1978
.
Symmetries in the reflection and transmission of elastic waves
,
Geophys. J. R. astr. Soc.
,
52
,
215
229
.

Kennett
B. L.N.
,
Koketsu
K.
,
Haines
A.J.
,
1990
.
Propagation invariants, reflection and transmission in anisotropic, laterally heterogeneous media
,
Geophys. J. Int.
,
103
,
95
101
.

Kiraz
M.S.R.
,
Snieder
R.
,
Wapenaar
K.
,
2023
.
The role of the background velocity model for the Marchenko focusing of reflected and refracted waves
,
preprint (arXiv:2304.09780)
.

Koketsu
K.
,
Kennett
B. L.N.
,
Takenaka
H.
,
1991
.
2-D reflectivity method and synthetic seismograms for irregularly layered structures - II. Invariant embedding approach
,
Geophys. J. Int.
,
105
,
119
130
.

Kosloff
D.D.
,
Baysal
E.
,
1983
.
Migration with the full acoustic wave equation
,
Geophysics
,
48
(
6
),
677
687
.

Løseth
L.O.
,
Ursin
B.
,
2007
.
Electromagnetic fields in planarly layered anisotropic media
,
Geophys. J. Int.
,
170
,
44
80
. https://doi.org/10.1111/j.1365-246X.2007.03390.x

Reinicke
C.
,
Wapenaar
K.
,
2019
.
Elastodynamic single-sided homogeneous Green’s function representation: theory and numerical examples
,
Wave Motion
,
89
,
245
264
.

Reinicke
C.
,
Dukalski
M.
,
Wapenaar
K.
,
2020
.
Comparison of monotonicity challenges encountered by the inverse scattering series and the Marchenko demultiple method for elastic waves
,
Geophysics
,
85
(
5
),
Q11
Q26
.

Reinicke
C.
,
Dukalski
M.
,
Wapenaar
K.
,
2023
.
Minimum-phase property and reconstruction of elastodynamic dereverberation matrix operators
,
Geophys. J. Int.
,
235
,
1
11
.

Rose
J.H.
,
2001
.
“Single-sided” focusing of the time-dependent Schrödinger equation
,
Phys. Rev. A
,
65
,
doi:10.1103/PhysRevA.65.012707
.

Rose
J.H.
,
2002
.
‘Single-sided’ autofocusing of sound in layered materials
,
Inverse Probl.
,
18
,
1923
1934
.

Sakurai
J.J.
,
1967
.
Advanced Quantum Mechanics
,
Pearson Education
.

Slob
E.
,
2016
.
Green’s function retrieval and Marchenko imaging in a dissipative acoustic medium
,
Phys. Rev. Lett.
,
116
,
doi:10.1103/PhysRevLett.116.164301
.

Slob
E.
,
Wapenaar
K.
,
Broggini
F.
,
Snieder
R.
,
2014
.
Seismic reflector imaging using internal multiples with Marchenko-type equations
,
Geophysics
,
79
(
2
),
S63
S76
.

Stoffa
P.L.
,
1989
.
Tau-p - A Plane Wave Approach to the Analysis of Seismic Data
,
Kluwer Academic Publishers

Takenaka
H.
,
Kennett
B. L.N.
,
Koketsu
K.
,
1993
.
The integral operator representation of propagation invariants for elastic waves in irregularly layered media
,
Wave Motion
,
17
,
299
317
.

Thomson
W.T.
,
1950
.
Transmission of elastic waves through a stratified solid medium
,
J. appl. Phys.
,
21
,
89
93
.

Ursin
B.
,
1983
.
Review of elastic and electromagnetic wave propagation in horizontally layered media
,
Geophysics
,
48
,
1063
1081
.

Van Stralen
M.J.N.
,
1997
.
Directional decomposition of electromagnetic and acoustic wave-fields—applications in integrated optics, exploration seismics and underwater acoustics
,
PhD thesis
,
Delft University of Technology
,
Delft
(repository.tudelft.nl)
.

Wapenaar
C.P.A.
,
Berkhout
A.J.
,
1986
.
Wave-field extrapolation techniques for inhomogeneous media which include critical angle events. Part II: methods using the two-way wave equation
,
Geophys. Prospect.
,
34
(
2
),
147
179
.

Wapenaar
C.P.A.
,
Dillen
M.W.P.
,
Fokkema
J.T.
,
2001
.
Reciprocity theorems for electromagnetic or acoustic one-way wave fields in dissipative inhomogeneous media
,
Radio Sci.
,
36
,
851
863
.

Wapenaar
K.
,
2019
.
Unified matrix-vector wave equation, reciprocity and representations
,
Geophys. J. Int.
,
216
,
560
583
.

Wapenaar
K.
,
2020
.
Reciprocity and representation theorems for flux- and field-normalised decomposed wave fields
,
Adv. Math. Phys.
,
2020
,
doi:10.1155/2020/9540135
.

Wapenaar
K.
,
2022
.
Wave-field representations with Green’s functions, propagator matrices, and Marchenko-type focusing functions
,
J. acoust. Soc. Am.
,
151
(
1
),
587
608
.

Wapenaar
K.
,
de Ridder
S.
,
2022
.
On the relation between the propagator matrix and the Marchenko focusing function
,
Geophysics
,
87
(
2
),
A7
A11
.

Wapenaar
K.
,
Slob
E.
,
2014
.
On the Marchenko equation for multicomponent single-sided reflection data
,
Geophys. J. Int.
,
199
,
1367
1371
.

Wapenaar
K.
,
Broggini
F.
,
Slob
E.
,
Snieder
R.
,
2013
.
Three-dimensional single-sided Marchenko inverse scattering, data-driven focusing, Green’s function retrieval, and their mutual relations
,
Phys. Rev. Lett.
,
110
,
doi:10.1103/PhysRevLett.110.084301
.

Wapenaar
K.
,
Snieder
R.
,
de Ridder
S.
,
Slob
E.
,
2021
.
Green’s function representations for Marchenko imaging without up/down decomposition
,
Geophys. J. Int.
,
227
,
184
203
.

White
B.S.
,
Zhou
M.
,
2006
.
Electroseismic prospecting in layered media
,
SIAM J. Appl. Math.
,
67
,
69
98
.

Woodhouse
J.H.
,
1974
.
Surface waves in a laterally varying layered structure
,
Geophys. J. R. astr. Soc.
,
37
,
461
490
.

APPENDIX A: UNIFIED PROPAGATOR MATRIX AND FOCUSING FUNCTIONS

In this appendix, we extend the theory of Section 2 to unified wave fields.

A1 Unified matrix-vector wave equation

Consider matrix-vector wave eq. (1). We partition the N × 1 wave field vector q, the N × 1 source vector d and the N × N operator matrix A as follows

(A1)

This includes the acoustic situation (for N = 2) discussed in Section 2. The vectors and operator matrix for other wave phenomena can be found in various references (Woodhouse 1974; Ursin 1983; Van Stralen 1997; Gelinsky & Shapiro 1997; Haartsen & Pride 1997; White & Zhou 2006; Løseth & Ursin 2007). A comprehensive overview is given by Wapenaar (2019) for acoustic waves (N = 2), quantum mechanical waves obeying the Schrödinger equation (N = 2), electromagnetic waves (N = 4), elastodynamic waves (N = 6), poroelastodynamic waves (N = 8), piezoelectric waves (N = 10) and seismoelectric waves (N = 12). For all these wave phenomena, the operator matrix A obeys the following symmetry properties

(A2)
(A3)
(A4)

with

(A5)

where O and I are zero and identity matrices of appropriate size. Superscript t denotes transposition of the matrix and the operators contained in it, with αt=α. Superscript † denotes transposition and complex conjugation. The bar on an operator denotes that it is defined in the adjoint medium. For further details we refer to the aforementioned references.

We show that, with some modifications, eqs (A1) – (A4) also hold for the matrix-vector form of the Dirac equation for the 4 × 1 spinor ψ. The Dirac equation is given by (Sakurai 1967)

(A6)

(summation over μ from 1 to 4), with the Dirac spinor partitioned as

(A7)

and with

(A8)
(A9)
(A10)

with V(x) the space-dependent potential, h Planck’s constant, c the speed of light, e the electron charge and σ1, σ2 and σ3 the Pauli matrices, defined as

(A11)

Assuming a time-dependence exp ( − iEt/ℏ), we replace ∂4 by −(EeV)/ℏc. Eq. (A6) can be rewritten as follows

(A12)

or

(A13)
(A14)

Pre-multiplying all terms by iσ3 and bringing the ∂3-operators to the left-hand side, using σ3σ1=iσ2, σ3σ2=iσ1 and σ3σ3=I, yields a set of equations which can be recast in the form of eq. (1), with 4 × 1 vectors q and d and 4 × 4 operator matrix A partitioned as in eq. (A1), with

(A15)
(A16)
(A17)
(A18)

With these definitions of the operator submatrices, matrix A obeys symmetry relations (A2) – (A4), with N, K and J defined as

(A19)

Although there are no direct applications for geophysics, the Schrödinger and Dirac equations are included in all derivations below, since this comes almost for free. When we speak of the ‘medium’, for the Schrödinger and Dirac equations it should be understood as the ‘potential’.

A2 Unified propagator matrix

We define the unified N × N propagator matrix W(x, xF) as the solution of wave eq. (8), with boundary condition (9) and with the operator matrix A being the unified operator matrix discussed in Appendix A1. Using eq. (10), the unified wave field vector q(x) can be propagated from x3, F to any depth level x3, assuming there are no sources between these depth levels. We partition W(x, xF) as

(A20)

where W11, W12, W21 and W22 are N2×N2 submatrices of W. For each of these submatrices, the second subscript refers to the wave field component (q1 or q2) it acts on at xF, whereas the first subscript refers to the wave field component it contributes to at x. W(x, xF) obeys the recursive relation (12), and from eq. (13) it follows that W(xF, x) is the inverse of W(x, xF).

A3 Relation between unified propagator matrix and Marchenko focusing functions

We assume again that the medium at and above DF is homogeneous and may be dissipative. The medium below DF may be inhomogeneous and dissipative, and it is source-free. In preparation for defining the focusing functions, in the upper half-space (i.e., at and above DF) we apply eigenvalue decomposition to matrix A~ (the spatial Fourier transform of operator matrix A), as follows

(A21)

with

(A22)

with S3+ and S3 being diagonal matrices containing vertical slownesses for downgoing and upgoing waves, respectively. We express the Fourier transformed wave field vector q~(s,x3) in terms of downgoing and upgoing waves p~+(s,x3) and p~(s,x3) via

(A23)

with

(A24)

Note that these equations imply q~1=L~1+p~++L~1p~. Similar as in Section 2.3, eq. (24), we continue with downgoing and upgoing waves q~1+ and q~1 which are field-normalized such that q~1=q~1++q~1. To this end, we define q~1±=L~1±p~± and we replace eq. (A23) by

(A25)

where

(A26)
(A27)

with

(A28)

Whereas there is ambiguity in the normalization of the matrices L~1± and L~2±, the matrix D~1± is uniquely defined (for each wave phenomenon). Note that for the acoustic situation we have L~1±=1, hence D~1±=L~2±=±s3,0/ρ0. Consequently, D~=L~ and b~=p~. Some other examples of matrix D~1± (for electromagnetic and elastodynamic waves) are given by Wapenaar (2022). In Appendix  C we derive for any wave phenomenon

(A29)

with J11 and J22 being the N2×N2 submatrices of N × N matrix J. From eq. (A5) we have for all wave phenomena except for the Dirac equation J11 = −J22 = I, and from eq. (A19) we have for the Dirac equation J11=J22=σ2.

We use eq. (A25) at DF and the properties of matrix D¯~1±(s,x3) to derive unified focusing functions and express them in the components of the unified propagator matrix and vice-versa. First we aim to substitute eq. (A25) for x3 = x3, F into a transformed version of eq. (10). This equation contains the propagator matrix W(x, xF). For a function of two space variables, u(x, xF) (with xF at DF), we define the spatial Fourier transformation along the horizontal components of the second space variable as

(A30)

and its inverse as

(A31)

Note that the sign in the exponential of eq. (A30) is opposite to that in eq. (14). Using these definitions and Parseval’s theorem, we rewrite eq. (10) as

(A32)

with W~(x,s,x3,F) obeying the boundary condition

(A33)

Substitution of eq. (A25) into eq. (A32) gives

(A34)

for x3x3, F, with

(A35)

We partition matrix Y~(x,s,x3,F) as follows

(A36)

Using equation (A26) and the spatial Fourier transform of eq. (A20), we obtain

(A37)
(A38)

We analyse these expressions one by one. First consider Y~1(x,s,x3,F). Via eq. (A34) it can be seen that subscript 1 refers to wavefield component q1 at x and superscript − refers to the upgoing wavefield component q~1 at x3, F. Moreover, for x3 = x3, F we obtain, using equation (A33), Y~1(x,s,x3,F)|x3=x3,F=Iexp{iωsxH}, or, applying the inverse spatial Fourier transformation defined in eq. (A31), Y1(x,xF)|x3=x3,F=Iδ(xHxH,F), which is a focusing condition. Hence, we define

(A39)

with F~1(x,s,x3,F) denoting the spatial Fourier transform of the focusing function F1(x, xF) for wavefield component q1, which focuses as an upgoing field at x = xF and continues as an upgoing field in the homogeneous upper half-space. Note that the focusing function is a N2×N2 matrix. Next, we consider Y~2(x,s,x3,F). Subscript 2 refers to wavefield component q2 at x and superscript − refers again to the upgoing wavefield component q~1 at x3, F. For x3 = x3, F we obtain, using eq. (A33), Y~2(x,s,x3,F)|x3=x3,F=D~1(s,x3,F)exp{iωsxH}, which is a focusing condition, but somewhat more complicated than for Y~1(x,s,x3,F) because of the mix of wavefield components q2 and q~1. Hence, we define

(A40)

with F~2(x,s,x3,F) denoting the spatial Fourier transform of the focusing function F2(x, xF) for wavefield component q2, which focuses as an upgoing field at x = xF and continues as an upgoing field in the homogeneous upper half-space (note that the definition of F~2 is different from that in Wapenaar (2022), to facilitate the derivations below). The focusing functions F~1(x,s,x3,F) and F~2(x,s,x3,F) together form the right-hand column of matrix Y~(x,s,x3,F).

For the analysis of the submatrices in the left-hand column of Y~(x,s,x3,F), we use symmetry relation (A29) and we need a similar relation for the submatrices of W~(x,s,x3,F). In Appendix  C we derive

(A41)

From the spatial Fourier transform of this equation we obtain for the submatrices of W~(x,s,x3,F)

(A42)

(no summation for repeated subscripts). Substituting eqs (A29) and (A42) into eqs (A37) and (A38) yields

(A43)
(A44)

Hence, using eqs (A39) and (A40), we find for the submatrices in the left-hand column of Y~(x,s,x3,F)

(A45)
(A46)

Hence, matrix Y~(x,s,x3,F) becomes

(A47)

or, using the inverse Fourier transformation defined in eq. (A31),

(A48)

This is a generalisation of eq. (39). Note that F~1, F~2, F¯~1 and F¯~2 are expressed in terms of the submatrices of the propagator matrix W~(x,s,x3,F) via eqs (A37) – (A40), (A45) and (A46). Conversely, we can express the submatrices of the propagator matrix W~(x,s,x3,F) in terms of the focusing functions F~1, F~2, F¯~1 and F¯~2. To this end, we start with inverting eq. (A35), according to

(A49)

with

(A50)
(A51)

Using eqs (A47) and (A50), we obtain

(A52)
(A53)

(no summation for repeated subscripts). These expressions are a generalisation of eqs (41) – (44). Those equations follow as a special case from eqs (A52) and (A53) by substituting J11 = −J22 = 1, D~1±(s,x3,F)=±s3,0/ρ0, {Δ~1(s,x3,F)}1=ρ0/2s3,0, and applying an inverse spatial Fourier transformation, which involves replacing s3, 0 by operator 1ωH1(xF).

A4 Representations with unified Marchenko focusing functions

Applying Parseval’s theorem to eq. (A34) and substituting the expressions for q(x), b(xF) and Y(x, xF) (eqs (A1), (A27) and (A48)), gives the following representation for the quantities q1(x) and q2(x) inside the inhomogeneous medium

(A54)

(no summation for repeated subscripts) for x3x3, F. This is a generalisation of eqs (52) and (53).

We use eq. (A54) to derive representations for Green’s functions between the boundary DF and any position x inside the medium. We define a unit d2-type source (see eq. (A1)) at xS just above DF. The N2×N2 Green’s matrix G12(x, xS) stands for the q1-type field observed at x, in response to this source. The spatial Fourier transform of the downgoing component at DF (i.e., just below the source) is proportional to the upper-right submatrix of the decomposition operator of eq. (A50), according to

(A55)

(Wapenaar 2022). To compensate for the effects of the inverse matrix {Δ~1(s,x3,F)}1, we define a modified Green’s matrix as

(A56)

such that its downgoing component at DF is given by

(A57)

or, after applying an inverse spatial Fourier transformation

(A58)

The upgoing response at DF to this downgoing source field is by definition the reflection response, hence

(A59)

The field at x inside the medium consists of Γ12(x, xS) and Γ22(x, xS), where Γ~22(x,s,x3,S)=G~22(x,s,x3,S)Δ~1(s,x3,F), with G~22(x,s,x3,S) being the Green’s function for the q2-type field observed at x. Substituting qα(x) = Γα2(x, xS) and q1±(xF)=Γ12±(xF,xS) into eq. (A54), using eqs (A58) and (A59), we obtain

(A60)

(no summation for repeated subscripts) for x3x3, F. This is a generalisation of eqs (54) and (55) and a starting point for developing a unified Marchenko method for full wave fields, accounting for evanescent waves inside the medium. Once the focusing functions are found, they can be used to retrieve the Green’s matrices Γα2(x, xS) for α = 1, 2 (from eq. (A60)) and all components of the propagator matrix W(x, xF) (from eqs (A52) and (A53)).

APPENDIX B: UNIFIED TRANSFER MATRIX AND DECOMPOSED FOCUSING FUNCTIONS

In this appendix, we extend the theory of Section 3 to unified wave fields.

B1 Unified transfer matrix

We introduce the unified transfer matrix as follows. Given the downgoing and upgoing fields q+(xF) and q(xF) contained in vector b(xF) at the boundary DF, we transfer these fields to depth level x3 via

(B1)

for x3x3, F. The unified transfer matrix T(x,xF) is partitioned as follows

(B2)

with T±,± being N2×N2 submatrices. Analogous to eq. (58), matrix T(x,xF) is related to the unified propagator matrix W(x, xF) of eq. (A20) via

(B3)

with D(xF) and D1(x) being the inverse spatial Fourier transforms of D~(s,x3,F) and {D~(s,x3)}1, defined in eqs (A26) and (A50), respectively. Unlike in the acoustic situation, where L(xF) and L1(x) in eq. (58) account for lateral variations of the medium parameters, the unified matrices D~(s,x3,F) and {D~(s,x3)}1 are defined for laterally invariant medium parameters at depths x3, F and x3. For D~(s,x3,F) this is not a restriction, since x3, F is the depth of the boundary DF between the inhomogeneous medium and the homogeneous upper half-space. However, for {D~(s,x3)}1 it implies that this operator can only be applied at depths where no lateral variations occur.

B2 Relation between unified transfer matrix and decomposed Marchenko focusing functions

Assuming there are no lateral variations at a specific depth level x3, we use the spatial Fourier transformation defined in eq. (14) along the horizontal components of the first space variable to express the transfer matrix (analogous to eq. (59)) as

(B4)

with {D~(s,x3)}1 defined in eq. (A50) and Y~(s,x3,xF) being the Fourier transform of Y(x, xF) defined in eq. (A48). Analogous to eq. (60), we obtain for the right-hand column of T~(s,x3,xF)

(B5)

with F~+(s,x3,xF) and F~(s,x3,xF) being the downgoing and upgoing parts at x3 of F~1(s,x3,xF). For the left-hand column of T~(s,x3,xF) we analyse the following expression

(B6)

Using eqs (A29) and (C16) in eq. (B6) gives

(B7)

By comparing this with eq. (B5) we find that the expression in eq. (B7) is equal to

(B8)

Combining the right-hand column (eq. B5) and the left-hand column (eq. B8), we obtain the following expression for the unified transfer matrix

(B9)

or, in the space domain,

(B10)

This is the generalisation of eq. (63).

B3 Representations with decomposed unified Marchenko focusing functions

Substituting the expressions for b(x) and T(x,xF) into eq. (B1) gives the following representations for the downgoing and upgoing fields, q1+(x) and q1(x) respectively, inside the inhomogeneous medium

(B11)
(B12)

for x3x3, F. These expressions are exact and hold for dissipative media. Making similar substitutions as in Section A4 we obtain

(B13)
(B14)

for x3x3, F. Here Γ12+(x,xS) and Γ12(x,xS) stand for the downgoing and upgoing part of the Green’s function Γ12(x, xS). These equations are generalisations of eqs (67) and (68) and form a starting point for developing a unified Marchenko method for decomposed wave fields. Once the focusing functions are found, they can be used to retrieve the decomposed Green’s functions Γ12+(x,xS) and Γ12(x,xS) (from eqs (B13) and (B14)) and all components of the transfer matrix T(x,xF) (from eq. (B10)). Versions of the Marchenko method based on expressions similar to eqs (B13) and (B14) have already been implemented for the retrieval of decomposed elastodynamic Green’s functions in lossless media, ignoring evanescent waves (Wapenaar & Slob 2014; da Costa Filho et al. 2014; Reinicke & Wapenaar 2019; Reinicke et al. 2020).

APPENDIX C: SYMMETRY PROPERTIES OF THE PROPAGATOR AND DECOMPOSITION MATRICES

Let N × N propagator matrices W(x, xA) and W(x, xB) be two independent solutions of eq. (8) (with unified operator matrix A(x) defined in eq. (A1)), with boundary condition (9), modified for coordinate vectors xA and xB. We show that the quantity R2Wt(xH,x3,xA)NW(xH,x3,xB)d2xH is a ‘propagation invariant’, meaning that it is independent of x3 (Haines 1988; Kennett et al. 1990; Koketsu et al. 1991; Takenaka et al. 1993). To this end we take the derivative in the x3-direction, apply the product rule for differentiation, use eq. (8) and symmetry relation (A2), according to

(C1)

This confirms that the integral is a propagation invariant. In a similar way, using symmetry relation (A3), it can be shown that R2W¯(xH,x3,xA)KW(xH,x3,xB)d2xH is also a propagation invariant. Using boundary condition (9), modified for x3 = x3, A and x3 = x3, B, we find from the first propagation invariant

(C2)

and from the second propagation invariant

(C3)

From eqs (C2) and (C3), using KN−1 = J, we find

(C4)

This equation is used in Section 2.3 and Appendix A3 in the derivation of the relation between the propagator matrix and the Marchenko focusing functions.

To derive a symmetry property for D~1±=L~2±(L~1±)1 (eq. (A28)), we start by Fourier transforming symmetry relations (A2) – (A4), assuming the medium is laterally invariant at the depth level where the transformation is applied. This gives

(C5)
(C6)
(C7)

The eigenvalue decomposition of matrix A~ is defined in eq. (A21), with the partitioning of Λ~ and L~ defined in eq. (A22). For all wave phenomena mentioned in Appendix A1, the eigenvalue matrix Λ~ obeys the following symmetry relations

(C8)
(C9)

Given eqs (A21) and (C5) – (C9), matrix L~ can be normalized such that

(C10)
(C11)

From the latter two equations, we obtain

(C12)

or, using the partitioning of L~ as defined in eq. (A22),

(C13)
(C14)

with J11 and J22 being the upper-left and lower-right submatrices of matrix J, and K12 being the upper-right (= lower-left) submatrix of matrix K. From eqs (C13) and (C14) it follows that D~1± as defined in eq. (A28) obeys the following symmetry relation

(C15)

Finally, using Δ~1=D~1+D~1 (equation A51), we obtain

(C16)
This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://dbpia.nl.go.kr/journals/pages/open_access/funder_policies/chorus/standard_publication_model)