Paper The following article is Open access

Global uniqueness in a passive inverse problem of helioseismology

, and

Published 24 April 2020 © 2020 IOP Publishing Ltd
, , Citation A D Agaltsov et al 2020 Inverse Problems 36 055004 DOI 10.1088/1361-6420/ab77d9

0266-5611/36/5/055004

Abstract

We consider the inverse problem of recovering the spherically symmetric sound speed, density and attenuation in the Sun from the observations of the acoustic field randomly excited by turbulent convection. We show that observations at two heights above the photosphere and at two frequencies above the acoustic cutoff frequency uniquely determine the solar parameters. We also present numerical simulations which confirm this theoretical result.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

1.1. Acoustic field in the Sun and its measurement

Solar oscillations are excited near the solar surface by turbulent convection. We consider the approximate equations describing these oscillations in three-dimensions at fixed angular frequency ω > 0 proposed in [1]:

Equation (1)

where ξω is the spatial displacement vector, c is the sound speed, ρ is the density, γ is the attenuation, fω is the random source field due to turbulent convection5, ψω describes the pressure oscillations, and ∇ is the gradient with respect to position vector $x\in {\mathbb{R}}^{3}$. In this work we consider this model under an additional spherical symmetry assumption: c = c(|x|), ρ = ρ(|x|), γ = γ(|x|). We also assume that

Equation (2)

where ${\mathbb{R}}_{+}=\left(0,\infty \right)$ and ${W}^{2,\infty }\left({\mathbb{R}}_{+}\right)$ denotes the Sobolev space of functions defined on ${\mathbb{R}}_{+}$which belong to ${L}^{\infty }\left({\mathbb{R}}_{+}\right)$ together with their first two derivatives. We suppose that in the upper atmosphere |x| ⩾ Ra the sound speed is constant, the density is exponentially decreasing (which corresponds to the adiabatic approximation, see [2, section 5.4]), and there is no attenuation6:

Equation (3)

where Ra = R + ha, R = 6.957 × 105 km is the solar radius (the radius of the photosphere), ha is the altitude at which the (conventional) interface between the lower and upper parts of the atmosphere is located, and H is called density scale height. The first two assumptions of formula (3) follow the model of [3], which extends a standard solar model of [4] to the upper atmosphere. In this article we do not fix exact values of the above parameters, but recall that in [3] they are given by the following table:

 ValueMeaning
ha500 kmAltitude (above the photosphere) of the interface
c06855  ms−1Sound speed at the interface
ρ02.886 × 10−6 kg m−3Density at the interface
H125 kmDensity scale height in the upper atmosphere

In reality, the Sun is surrounded by a hot corona whose base is located at about hc = 2000 km above the surface and which is highly inhomogeneous, for more details see [5]. Here we neglect this complication, and assume that the sound speed remains constant and that the density decays exponentially above the interface.

The exponential decay of density in the atmosphere leads to the trapping of acoustic waves with frequencies less than the cutoff frequency ωctf = c0/(2H), which is about ωctf/2π ≈ 5.2 mHz for the Sun, and to the quantization of their admissible frequencies. These mode frequencies measured at the solar surface constitute the basic input data for helioseismological studies, see, e.g., [2].

On the other hand, acoustic waves with frequencies above the cutoff, that is, such that

Equation (4)

propagate away into the upper atmosphere. Several motivations for using measurements above the acoustic cutoff frequency for helioseismic inversions have been given in the literature. In particular, it has been shown that this high-frequency data can be used to study the sources of wave excitation [6], as well as the properties of the medium above the solar surface [7], see also [3,8].

The motions near the solar surface due to the solar acoustic waves can be measured through the Doppler shifts of an absorption line in the solar spectrum. Different instruments use different absorption lines. For example, the Michelson Doppler imager (MDI) instrument onboard the solar and heliospheric observatory (SOHO) satellite measures the line-of-sight Doppler velocity 200 km above the phostosphere, while the helioseismic and magnetic imager (HMI) instrument onboard the solar dynamics observatory (SDO) satellite observes solar oscillations at a height of 125 km.

In the present work we assume that the cross-covariance measurements of the solar acoustic field can be performed simultaneously at two different heights above the surface. This was the case during one year in 2010, when SOHO/MDI and SDO/HMI were both in operation. Also, the ground-based network GONG can be used in combination with the satellites. As recently shown in [9], it is also possible to perform multi-height measurements by combining six raw HMI filtergrams in different ways. In the present work we do not discuss details of measurements of the solar acoustic field, for more information on this subject see [1,10,11].

The main theoretical results of this work are presented in section 2. Related proofs are given in section 3. Numerical reconstructions confirming our theoretical conclusions are given in section 4.

2. Main results

2.1. Extracting the imaginary part of the Green's function

Under the assumptions (2), (3), (4) equation (1) at fixed ω can be rewritten as the Schrödinger equation

Equation (5)

where the indices indicating dependence on ω are suppressed,

Equation (6)

$v\in {L}^{\infty }\left({\mathbb{R}}^{3}\right)$, and v(x) = 1/(H|x|) for |x| ⩾ Ra. Note that the term k2 is included in the definition of v to guarantee that v decays at infinity.

In this article we consider equation (5) with a general complex-valued potential v such that

Equation (7)

If the potential v satisfies (7), then the resolvent ${\left({L}_{v}-{k}^{2}\right)}^{-1}$ is a meromorphic operator-valued function of $k\in {\mathbb{C}}_{+}=\left\{z\in \mathbb{C}\,:\Im z{ >}0\right\}$ with the distributional kernel Gv(x, x') = Gv(x, x'; k) admitting a unique meromorphic continuation across the positive real axis. The restriction to $k\in {\mathbb{R}}_{+}$ of the distributional kernel Gv(x, x') is called the radiation Green's function for equation (5). In addition, Gv(x, x') is a distributional solution to equation (Lvk2)Gv(⋅, x') = δx', where δx' denotes the Dirac delta function centered at x'. We also suppose that $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$, where

Equation (8)

Basic properties of Gv can be found in [12, 13]. In particular, at fixed k the function Gv is jointly continuous outside the diagonal ${\Delta}=\left\{\left(x,{x}^{\prime }\right)\in {\mathbb{R}}^{3}{\times}{\mathbb{R}}^{3}\,:\,x={x}^{\prime }\right\}$. Besides, ${\left({L}_{v}-{\left(k+\text{i}0\right)}^{2}\right)}^{-1}\in \mathcal{L}\left({L}_{1+\varepsilon }^{2}\left({\mathbb{R}}^{3}\right),{L}_{-1-\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)\right)$, $\varepsilon \in \left(0,\frac{1}{2}\right]$, where ${L}_{\delta }^{2}\left({\mathbb{R}}^{3}\right)$ denotes the Hilbert space of measurable functions u in ${\mathbb{R}}^{3}$ with the finite norm

Accordingly, for any $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ equation (5) with $f\in {L}_{1+\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)$, $\varepsilon \in \left(0,\frac{1}{2}\right)$, admits a unique radiation (limiting absorption), solution7 given by

Equation (9)

Spherical symmetry of the potential v allows to separate variables in equation (5), reducing it to an equivalent multi-channel Schrödinger equation on the half-line ${\mathbb{R}}_{+}$ with non-coupled channels. More precisely, consider the orthogonal expansions in normalized spherical harmonics ${Y}_{\ell }^{m}$:

Equation (10)

where r > 0, $\vartheta \in {S}_{1}^{2}$ and ${S}_{R}^{2}=\left\{x\in {\mathbb{R}}^{3}\,:\,\vert x\vert =R\right\}$. Plugging these expansions into formula (5), we get the radial equations

Equation (11)

where ⩾ 0, |m| ⩽ . Besides, it follows from formulas (9), (10) that if ψv is a unique radiation solution of equation (5), then ${\varphi }_{\ell }^{m}$ can be expressed as:

Equation (12)

where Gv,(r, r') is the coefficient in the spherical harmonics expansion of Gv:

Equation (13)

One can show that Gv, is indeed a radiation Green's function for equation (11), see [12] and section 3.2.

In this article we consider equation (5) with a random source function f. In this case the radiation solution ψv is also a random function, as well as functions ${\varphi }_{v,\ell }^{m}$ and ${f}_{\ell }^{m}$ in the spherical harmonics expansions of formula (10). Following [1], we assume that the power spectrum (power spectral density) of ψv, defined as ${\mathcal{P}}_{v,\ell }^{m}\left(r\right)=\mathbb{E}\vert {\varphi }_{v,\ell }^{m}\left(r\right){\vert }^{2}$, can be measured experimentally. However, in contrast to [1], where the power spectral density is assumed to be known at the solar surface r = R, we assume that it can be measured at two different observation radii ${R}_{o}^{{\dagger}}{ >}{R}_{o}{\geqslant}{R}_{\odot }$. These measurements can be roughly achieved by using concurrent MDI and HMI Dopplergrams [15], or multi-height measurements from raw HMI filtergrams [9].

Our first result relates cross correlations $\mathbb{E}\left(\overline{{\varphi }_{v,\ell }^{m}\left({r}_{1}\right)}{\varphi }_{v,\ell }^{m}\left({r}_{2}\right)\right)$ to the Green's function Gv,(r1, r2). We prove the following proposition:

Proposition 1. Let v be a complex-valued potential satisfying (7) and let $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ be fixed. Assume that the random functions ${f}_{\ell }^{m}$ satisfy the condition

Equation (14)

for some Π > 0, RRa. Then the following formula is valid at fixed r1, r2 > 0:

Equation (15)

Proposition 1 is proved in section 3.3. This proposition is a variation of a well-known result, see, e.g., [1, 16] and references therein. The main difference is that we consider long range potentials and the radiation Green's function, whereas in the literature the Green's function with an artificial boundary condition imposed at r = R and approximating the Sommerfeld radiation condition is used. The approximate radiation boundary condition allows to get rid of the error term $O\left(\frac{1}{R}\right)$ in the formula (15) but complicates the further analysis. In addition, note that in general the Sommerfeld radiation condition does not apply for long range potentials.

Assumption (14) requires that the random sources be uncorrelated in space, excited throughout the volume with a power proportional to $-\Im \tilde {v}$, and excited at the surface r = R with a power proportional to k.

Remark 1. Recall that equation (5) arises, in particular, by rewriting equation (1) under the assumptions (2)–(4). In this case k and v are given by formulas (6), and proposition 1 has a physical interpretation. Taking into account that $-\Im \tilde {v}=2\omega \gamma $, condition (14) implies proportionality of the power spectral density of random excitations to the local attenuation (energy dissipation) rate. It has long been known in physics that this condition is related to the possibility to extract the imaginary part of the point-source response function (Green's function) from the power spectral density of the randomly excited field, which is expressed by relation (15) in our setting. In physical literature similar relations are established in fluctuation-dissipation theorems, see, e.g., [17].

2.2. Uniqueness results

Proposition 1 allows to retrieve Im Gv,(r, r) approximately from the power spectral density of noise ${\mathcal{P}}_{v,\ell }^{m}\left(r\right)=\mathbb{E}\vert {\varphi }_{v,\ell }^{m}\left(r\right){\vert }^{2}$ at fixed r. Next, we prove that Im Gv,(r, r) known exactly for all ⩾ 0 and at two different r uniquely determines v. Equivalently, taking into account the orthogonal expansion (13), v is uniquely determined by Im Gv known on ${S}_{r}^{2}{\times}{S}_{r}^{2}$ at two different r.

Theorem 1. Let v1, v2 be two complex-valued potentials satisfying (7) and let $k\in {\mathbb{R}}_{+}{\backslash}\left({{\Sigma}}_{{v}_{1}}^{P}\cup {{\Sigma}}_{{v}_{2}}^{P}\right)$ be fixed. Assume that that one of the following conditions holds true:

  • (a)  
    ${G}_{{v}_{1}}={G}_{{v}_{2}}$ on ${M}_{{R}_{o}}^{4}=\left({S}_{{R}_{o}}^{2}{\times}{S}_{{R}_{o}}^{2}\right){\backslash}{\Delta}$ for some Ro > Ra;
  • (b)  
    $\Im {G}_{{v}_{1}}=\Im {G}_{{v}_{2}}$ on ${M}_{{R}_{o}}^{4}\cup {M}_{{R}_{o}^{{\dagger}}}^{4}$ for some ${R}_{o}^{{\dagger}}{ >}{R}_{o}{\geqslant}{R}_{a}$ such that ${R}_{o}^{{\dagger}}\notin {{\Sigma}}_{\alpha ,k,{R}_{o}}^{S}$, where ${{\Sigma}}_{\alpha ,k,{R}_{o}}^{S}\subset \left[{R}_{o},\infty \right)$ is a discrete set without finite accumulation points defined by (34b), (35).

Then v1 = v2 in ${L}^{\infty }\left({\mathbb{R}}^{3}\right)$.

Remark 2. If v is some potential satisfying (7) then the restriction of Gv(x, x') to ${M}_{R}^{4}$ depends on |xx'| only because v is spherically symmetric. In particular, theorem 1 remains valid if the four-dimensional manifolds ${M}_{R}^{4}$ are replaced by the one-dimensional manifolds

for some fixed x1, ${x}_{2}\in {S}_{R}^{2}$ with x1x2 = 0.

Theorem 1 is proved in section 3 and the proof consists of the following steps presented in sections 3.1, 3.2, 3.4 and 3.5. In section 3.1 we separate variables in the equation (Lvk2)ψ = 0 and establish auxiliary results for the regular solutions of the arising radial Schrödinger equations. In section 3.2 we derive an appropriate relation between the Green's function Gv and the Green's functions Gv, of the radial equations. This relation will allow to extract the diagonal values Gv,(R, R) from Gv on ${M}_{R}^{4}$. Using this relation, in section 3.4 we show that the scattering matrix elements sv, can be extracted from Gv on ${M}_{{R}_{o}}^{4}$ or from the imaginary part Im Gv only on ${M}_{{R}_{o}}^{4}\cup {M}_{{R}_{o}^{{\dagger}}}^{4}$, under the assumption that ${R}_{o}^{{\dagger}}\notin {{\Sigma}}_{\alpha ,k,{R}_{o}}^{S}$. In section 3.5 we prove that the scattering matrix elements sv, determine the Dirichlet-to-Neumann map Λv,R for potential v in some ball ${B}_{R}^{3}$, RRa, where

In section 3.6 we combine these results together with the uniqueness theorem for the Dirichlet-to-Neumann map from [18] to uniquely determine v. This will prove theorem 1.

Corollary 1. Let c, ρ, γ, and c', ρ', γ' be two sets of parameters satisfying (2), (3) and define the corresponding potentials v = vω, v' =v'ω and the wavenumber k = kω according to formula (6) at fixed ω. Let ω1ω2 be two positive frequencies satisfying (4) and such that ${k}_{\omega }\in {\mathbb{R}}_{+}{\backslash}\left({{\Sigma}}_{{v}_{\omega }}^{P}\cup {{\Sigma}}_{{v}_{\omega }^{\prime }}^{P}\right)$ for ω = ω1, ω2. Let ${G}_{{v}_{\omega }}$, ${G}_{{v}_{\omega }^{\prime }}$ be the radiation Green's functions at fixed ω for the potentials v, v' respectively.

Suppose that $\Im {G}_{{v}_{\omega }}=\Im {G}_{{v}_{\omega }^{\prime }}$ on ${M}_{{R}_{0}}^{4}\cup {M}_{{R}_{o}^{{\dagger}}}^{4}$ for some ${R}_{o}^{{\dagger}}{ >}{R}_{o}{\geqslant}{R}_{a}$ such that ${R}_{o}^{{\dagger}}\notin {{\Sigma}}_{1/H,{k}_{\omega },{R}_{o}}^{S}$, where ω = ω1, ω2. Then c = c', ρ = ρ', γ = γ' in ${L}^{\infty }\left({\mathbb{R}}_{+}\right)$.

Proof. Under the assumptions of corollary 1 it follows from theorem 1 that vω = v'ω for ω = ω1, ω2. Using that $\mathfrak{R}{v}_{\omega }=\mathfrak{R}{v}_{\omega }^{\prime }$ for ω = ω1, ω2 one can show that c = c', ρ = ρ', see, e.g., the proof of [19, theorem 2.9]. Then, recalling that Im vω = −2ωγ/c2, Im v'ω = −2ωγ'/(c')2, it follows from the equality Im vω = Im v'ω for ω = ω1, ω2 together with the equality c = c' that γ = γ', concluding the proof of corollary 1. □

Remark 3. In corollary 1 it is assumed for simplicity of proofs and clarity of presentation that the measurements are performed in the homogeneous part of the solar atmosphere. This assumption can be relaxed (the proofs must be modified in a straightforward way) by requiring that the sound speed, density, and attenuation are known functions in the inhomogeneous part of the solar atmosphere, where the measurements are to be performed.

In section 4 we shall present numerical simulations which confirm the uniqueness results of theorem 1 and corollary 1.

3. Proof of the main results

3.1. Properties of regular radial solutions

In this subsection we shall establish some auxiliary results regarding regular solutions of the radial Schrödinger equation which arises by separation of variables in the homogeneous equation (Lvk2)ψ = 0.

As the potential v is spherically symmetric, this equation separates in spherical coordinates. We seek a solution ${\psi }_{v,\ell }^{m}\in {H}_{\text{loc}}^{2}\left({\mathbb{R}}^{3}\right)$ of the form

Equation (16)

which leads to the equation

Equation (17)

together with the condition that φv, vanishes at the origin. One can show that this determines ${\varphi }_{v,\ell }\in {H}_{\text{loc}}^{2}\left({\mathbb{R}}_{+}\right)$ uniquely up to a multiplicative factor, see, e.g., [12]. We impose the boundary condition

Equation (18)

which fixes φv, uniquely.

Note that φv,(r) does not depend on the values of $\tilde {v}$ in the region [r, +), see [20, formula (12.4)]. This analysis implies the following lemma.

Lemma 1. Let k > 0 and let v be a complex-valued potential satisfying (7). Then the Dirichlet problem

where ${Y}_{\ell }^{m}={Y}_{\ell }^{m}\left(x/\vert x\vert \right)$, $x\in {S}_{R}^{2}$, has a unique solution $\psi \in {H}^{2}\left({B}_{R}^{3}\right)$ if and only if φv,λ(R) ≠ 0 for all integer λ ⩾ 0. In addition, this solution is given by the formula

Equation (19)

Next we shall derive an expression for the regular solution φv, in the domain rR in terms of the Coulomb wave functions and of the so-called scattering matrix element sv,. First we recall the definition and some basic properties of the Coulomb wave functions from [21,  22].

The Coulomb wave functions ${H}_{\ell }^{{\pm}}\left(\eta ,kr\right)$, η = α/(2k), are the unique solutions of equation (17) with $\tilde {v}\left(r\right)=\alpha /r$ specified by the following asymptotics as r → +:

Equation (20a)

Equation (20b)

where σ(η) = arg   Γ( + 1 + iη) is the Coulomb phase shift and Γ denotes the usual gamma function. Functions ${H}_{\ell }^{+}\left(\eta ,kr\right)$ and ${H}_{\ell }^{-}\left(\eta ,kr\right)$ are complex conjugates of each other and are linearly independent. Using (20a) together with the relation

given in [23], one can show that

Equation (21)

Note that this derivation of formula (21) uses the identity

which is a corollary of the recurrence relation for the gamma function:

Using these properties, we shall prove the following result.

Lemma 2. Let v be a complex-valued potential satisfying (7), let $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ be fixed and let η = α/(2k). Then the function φv, defined by (17), (18) admits in the region rRa the representation

Equation (22)

with unique ${b}_{\ell }={b}_{\ell }\left(k\right)\in \mathbb{C}{\backslash}\left\{0\right\}$ and ${s}_{v,\ell }={s}_{v,\ell }\left(k\right)\in \mathbb{C}{\backslash}\left\{0\right\}$.

Proof. As the Coulomb wave functions ${H}_{\ell }^{{\pm}}\left(\eta ,kr\right)$ are linearly independent, any solution to equation (17) in the region rRa is given by their linear combination with unique coefficients. In particular, φv, can be expressed in the region rR in the form

for some av,, ${b}_{v,\ell }\in \mathbb{C}$. We shall show that av, ≠ 0, bv, ≠ 0.

Recall from [13] that for any $k\in {\mathbb{R}}_{+}$ outside the singular set ${{\Sigma}}_{v}^{P}$ and for any $f\in {L}_{1+\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)$ the Schrödinger equation (5) admits the unique solution $\psi \in {H}_{\text{loc}}^{2}\left({\mathbb{R}}^{3}\right)\cap {L}_{-1-\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)$, $\varepsilon \in \left(0,\frac{1}{2}\right]$, satisfying the radiation condition

Equation (23a)

Equation (23b)

Now assume that bv, = 0. Then it follows from (20a), (21) that the function defined by (19) is a non-zero solution of class ${H}_{\text{loc}}^{2}\left({\mathbb{R}}^{3}\right)\cap {L}_{-1-\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)$, $\varepsilon \in \left(0,\frac{1}{2}\right]$, to (Lvk2)ψ = 0 in ${\mathbb{R}}^{3}$ satisfying the radiation condition (23a). This contradicts the assumption $k\notin {{\Sigma}}_{v}^{P}$.

Now assume that av, = 0 and let ψ be defined by (19). Then it follows from (20a), (21) that $\overline{\psi }$ is of class ${H}_{\text{loc}}^{2}\left({\mathbb{R}}^{3}\right)\cap {L}_{-1-\varepsilon }^{2}\left({\mathbb{R}}^{3}\right)$, $\varepsilon \in \left(0,\frac{1}{2}\right]$, and satisfies $\left({L}_{\overline{v}}-{k}^{2}\right)\overline{\psi }=0$ in ${\mathbb{R}}^{3}$ together with the radiation condition (23a). Taking into account definition (8), this also contradicts the assumption $k\notin {{\Sigma}}_{v}^{P}$ and concludes the proof of lemma 2. □

Remark 4. The coefficient sv, = sv,(k) in lemma 2 is called the th scattering matrix element of the potential v.

3.2. Green's functions

In this subsection we shall express the radiation Green's function Gv, for equation (17) in the region rRa in terms of the Coulomb wave functions. Then we shall give a formula for extracting the diagonal values Gv,(R, R) from the radiation Green's function Gv for equation (5) known at ${M}_{R}^{4}$.

In addition to the regular solution φv, of equation (5) specified by the boundary condition (18), we consider the outgoing solution ${\varphi }_{v,\ell }^{+}$ which is specified by the asymptotics

Equation (24)

The outgoing Green's function Gv,(r, r') for equation (17) is defined as a distributional solution to the equation (Lv,k2)Gv,(⋅, r') = δr' specified by the following boundary conditions at fixed r' > 0:

for some non-zero constant c = c(r', η, k). If the regular solution φv,(r) and the outgoing solution ${\varphi }_{v,\ell }^{+}\left(r\right)$ are linearly independent, this Green's function exists, is unique and is given by the explicit formula

Equation (25)

where r< = min(r, r'), r> = max(r, r') and $\left[{\varphi }_{v,\ell },{\varphi }_{v,\ell }^{+}\right]$ is the Wronskian:

Equation (26)

By a standard theory of ordinary differential equations, this Wronskian is independent of r since the homogeneous second-order equation satisfied by φv, and ${\varphi }_{v,\ell }^{+}$ does not contain first-order terms, see [24, p 83, lemma 3.11]. Also note that formula (25) is a standard result from the theory of Sturm–Liouville problems, see, e.g., [24, p 158, formula (5.65)].

Lemma 3. If $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$, then Gv, is well-defined and is given in the region rRa, r' ⩾ Ra by the formula

Equation (27)

Proof. It follows from lemma 2 that under the assumption $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ the functions φv,(r) and ${H}_{\ell }^{+}\left(\eta ,kr\right)$ are linearly independent in the region rRa. Using the relation $\left[{H}_{\ell }^{-}\left(\eta ,kr\right),{H}_{\ell }^{+}\left(\eta ,kr\right)\right]=2\text{i}k$, given in [21], and formula (22) we get $\left[{\varphi }_{v,\ell },{\varphi }_{v,\ell }^{+}\right]=2\text{i}k{b}_{v,\ell }$. Together with (25), this implies (27). □

Next we shall show how the diagonal values Gv,(R, R) can be extracted from the Green's function Gv restricted to ${M}_{R}^{4}$.

First recall that the Legendre polynomials P can be defined using the formal generating identity [25]

Equation (28)

Using the Laplace formula [25, theorem 8.21.2] and the Dirichlet convergence test one can show that at fixed t = 1 this series converges pointwise for all s ∈ (−1, 1). Besides, the Legendre polynomials form a complete orthogonal system in L2(−1, 1) such that

Equation (29)

where δℓm is the Kronecker delta.

Now we recall [12] that at fixed $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ the Green's function Gv(x, x') is continuous outside of the diagonal x = x' and Gv(x, x') = O(|xx'|−1) as xx'. Besides, the series expansion

Equation (30)

converges for x ≠ ±x' and Gv,(R, R) has the asymptotics

Equation (31)

Note that series expansion (30) follows from the spherical harmonics expansion (13), taking into account the well known addition theorem (see, e.g., [12]):

Lemma 4. Let $k\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v}^{P}$ and fix x', ${x}^{{\prime\prime}}\in {S}_{R}^{2}$ such that x' ⋅ x'' = 0. Then for each ⩾ 0

Proof. Using (28) and (30) we get a pointwise convergent series expansion

which, in view of (29) and (31), also converges in L2(−1, 1). Recalling that Legendre polynomials form a complete orthogonal system in L2(−1, 1), we get lemma 4. □

3.3. Extracting the Green's function from cross correlations

In this subsection we prove proposition 1. First, recall that the Green's function Gv, satisfies the reciprocity relation

Equation (32)

To prove it, consider the equations

Multiplying the first equation by Gv,(⋅, r2), subtracting the second equation multiplied by Gv,(⋅, r1), and integrating over (0, R), R > r1, r2, we obtain

where [−, −] denotes the Wronskian defined according to (26) and the notation ${x\vert }_{a}^{b}=x\left(b\right)-x\left(a\right)$ is used. The next step is to show that the term on the left-hand side vanishes. Using formulas (18), (25) and the estimate $\frac{\partial }{\partial r}{\varphi }_{v,\ell }\left(r\right)=O\left({r}^{\ell }\right)$, r → +0, given in [12, theorem 3.3], we get [Gv,(⋅, r1), Gv,(⋅, r2)](+0) = 0. Using formulas (20a), (20b), (21), (25), we also get $\left[{G}_{v,\ell }\left(\cdot ,{r}_{1}\right),{G}_{v,\ell }\left(\cdot ,{r}_{2}\right)\right]\left(R\right)=O\left(\frac{1}{R}\right)$, R → +. As R tends to +, we get formula (32).

To prove (15), we follow a similar scheme. We start from the equations

Multiplying the first equation by Gv,(⋅, r2), subtracting the second equation multiplied by $\overline{{G}_{v,\ell }\left(\cdot ,{r}_{1}\right)}$, integrating over (0, R), R > r1, r2, we get

Equation (33)

In a similar way with the proof of formula (32), one can show that the Wronskian vanishes at zero and that

Combining this with formulas (12), (14), (33), (32) to compute $\mathbb{E}\left(\overline{{\psi }_{\ell }^{m}\left({r}_{1}\right)}{\psi }_{\ell }^{m}\left({r}_{2}\right)\right)$, we get formula (15), which concludes the proof of proposition 1.

3.4. Recovering the scattering matrix elements

In this subsection we shall show that the scattering matrix elements sv, for the potential v can be extracted from the Green's function Gv on ${M}_{{R}_{o}}^{4}$ or from its imaginary part Im Gv only on ${M}_{{R}_{o}}^{4}\cup {M}_{{R}_{o}^{{\dagger}}}^{4}$, where ${R}_{a}{\leqslant}{R}_{o}{< }{R}_{o}^{{\dagger}}$.

Recall that the Coulomb function ${H}_{\ell }^{+}\left(\eta ,kr\right)$ does not vanish for r > 0, since ${H}_{\ell }^{+}\left(\eta ,kr\right)$ and its complex conjugate ${H}_{\ell }^{-}\left(\eta ,kr\right)$ form a basis of solutions of equation (17) with $\tilde {v}\left(r\right)=\alpha /r$. Together with lemmas 3 and 4, this leads to the following result.

Lemma 5. Let v1, v2 be two complex-valued potentials satisfying (7), let $k\in {\mathbb{R}}_{+}{\backslash}\left({{\Sigma}}_{{v}_{1}}^{P}\cup {{\Sigma}}_{{v}_{2}}^{P}\right)$ be fixed and let RoRa. Suppose that ${G}_{{v}_{1}}={G}_{{v}_{2}}$ on ${M}_{{R}_{o}}^{4}$. Then ${G}_{{v}_{1},\ell }\left({R}_{o},{R}_{o}\right)={G}_{{v}_{2},\ell }\left({R}_{o},{R}_{o}\right)$ and ${s}_{{v}_{1},\ell }={s}_{{v}_{2},\ell }$ for all ⩾ 0.

Next we shall show that the scattering matrix elements sv, can be extracted from Im Gv only, i.e., without knowing $\mathfrak{R}{G}_{v}$. However, the values of Im Gv must be given not only on ${M}_{{R}_{o}}^{4}$ but also on ${M}_{{R}_{o}^{{\dagger}}}^{4}$ for some ${R}_{o}^{{\dagger}}{ >}{R}_{o}$.

Note that formula (27) implies that

Equation (34a)

Equation (34b)

where rRa, η = α/(2k). Using (34a) with r = Ro, ${R}_{o}^{{\dagger}}$ one can see that sv,l is uniquely determined from Im Gv,(Ro, Ro) and $\Im {G}_{v,\ell }\left({R}_{o}^{{\dagger}},{R}_{o}^{{\dagger}}\right)$ if and only if $\mathrm{sin}\left({\vartheta }_{\ell }\left(\eta ,k{R}_{o}^{{\dagger}}\right)-{\vartheta }_{\ell }\left(\eta ,k{R}_{o}\right)\right)\ne 0$. This justifies the definition of the singular set

Equation (35)

Lemma 6. Let v1, v2 be two complex-valued potentials satisfying (7), let $k\in {\mathbb{R}}_{+}{\backslash}\left({{\Sigma}}_{{v}_{1}}^{P}\cup {{\Sigma}}_{{v}_{2}}^{P}\right)$, and let ${R}_{o}^{{\dagger}}{ >}{R}_{o}{\geqslant}{R}_{a}$ be such that ${R}_{o}^{{\dagger}}\notin {{\Sigma}}_{\alpha ,k,{R}_{o}}^{S}$. Suppose that $\Im {G}_{{v}_{1}}=\Im {G}_{{v}_{2}}$ on ${M}_{{R}_{o}}^{4}\cup {M}_{{R}_{o}^{{\dagger}}}^{4}$. Then ${s}_{{v}_{1},\ell }={s}_{{v}_{2},\ell }$ for all ⩾ 0. Besides, the set ${{\Sigma}}_{\alpha ,k,{R}_{o}}^{S}$ is discrete and does not have finite accumulation points.

Proof. Under the assumptions of lemma 6 it follows from lemma 4 that for all ⩾ 0 the equality $\Im {G}_{{v}_{1},\ell }\left(r,r\right)=\Im {G}_{{v}_{2},\ell }\left(r,r\right)$ holds true for r = Ro, ${R}_{o}^{{\dagger}}$. Together with the discussion before lemma 6 it implies that ${s}_{{v}_{1},\ell }={s}_{{v}_{2},\ell }$ for all ⩾ 0. This concludes the proof of the first assertion of lemma 6.

Next we shall prove the second assertion. The following formulas are valid as → + at fixed η and ρ ≠ 0:

see formulas (33.2.11), (33.5.8), (33.5.9) of [21]. It follows that

Equation (36)

locally uniformly in $r\in {\mathbb{R}}_{+}$. Besides, it follows from formula (33.2.11) of [21] and from discussion below it that at fixed ⩾ 0 the set of solutions $r\in {\mathbb{R}}_{+}$ of the equation sin(ϑ(η, kr) − ϑ(η, kRo)) = 0 is discrete and does not have finite accumulation points, as the zero-set of a non-zero analytic function. Together with (36), this concludes the proof of lemma 6. □

3.5. Recovering the Dirichlet-to-Neumann map

In this subsection we shall show that the scattering matrix elements sv, uniquely determine the Dirichlet-to-Neumann map for v is some ball ${B}_{R}^{3}$, RRa.

Note that lemma 1 justifies the definition of the singular set

Lemma 7. Let v be a complex-valued potential satisfying (7) and let $k\in {\mathbb{R}}_{+}$ be fixed. Then $R\in {\mathbb{R}}_{+}{\backslash}{{\Sigma}}_{v,k}^{D}$ if and only if for any $g\in {H}^{3/2}\left({S}_{R}^{2}\right)$ the Dirichlet problem

Equation (37)

is uniquely solvable for $\psi \in {H}^{2}\left({B}_{R}^{3}\right)$. Besides, the set ${{\Sigma}}_{v,k}^{D}$ is discrete and does not have finite accumulation points.

Proof. It follows from lemma 1 that if the Dirichlet problem (37) is uniquely solvable for any $g\in {H}^{3/2}\left({S}_{R}^{2}\right)$ then $R\notin {{\Sigma}}_{v,k}^{D}$.

Now suppose that $R\notin {{\Sigma}}_{v,k}^{D}$. First we shall show that (37) does not admit a non-zero solution for g = 0. Assuming that $\psi \in {H}_{0}^{2}\left({B}_{R}^{3}\right)$ is a solution to (37) with g = 0 one can see (taking into account the spherical symmetry of v) that its partial wave components

belong to ${H}_{0}^{1}\left({B}_{R}^{3}\right)$ and satisfy (37) weakly. Because of the boundary elliptic regularity [26] they also belong to ${H}^{2}\left({B}_{R}^{3}\right)$. Then it follows from lemma 1 that all ${\psi }_{\ell }^{m}$ vanish and ψ = 0.

Next we recall that the operator

Equation (38)

is Fredholm of index zero from ${H}^{2}\left({B}_{R}^{3}\right)$ to ${L}^{2}\left({B}_{R}^{3}\right){\times}{H}^{3/2}\left({S}_{R}^{2}\right)$, see [27]. Together with already established uniqueness for the Dirichlet problem (37) for $R\notin {{\Sigma}}_{v,k}^{D}$, this proves the first assertion of lemma 7.

Next we shall show that ${{\Sigma}}_{v,k}^{D}$ is a discrete set without finite accumulation points. It can be shown [12] that the regular solution φv,(r) defined by (17), (18) satisfies the estimate

Equation (39)

uniformly in r ∈ (0, R] at fixed R > 0. Besides, zeros of each φv, are discrete and do not have finite accumulation points. This concludes the proof of the second assertion of lemma 7. □

Remark 5. Recalling from [27] that the operator (38) is Fredholm of index zero from ${H}^{2}\left({B}_{R}^{3}\right)$ to ${L}^{2}\left({B}_{R}^{3}\right){\times}{H}^{3/2}\left({S}_{R}^{3}\right)$ one can see that if the potential $v\in {L}^{\infty }\left({B}_{R}^{3}\right)$ is such that the Dirichlet problem (37) with $g\in {H}^{3/2}\left({S}_{R}^{2}\right)$ is uniquely solvable for $\psi \in {H}^{2}\left({B}_{R}^{3}\right)$, then

for some constant Cv,k,R > 0. In addition, the trace theorem [27] leads to the estimate

for some constant ${C}_{v,k,R}^{\prime }$C'v,k,R > 0, where ${C}_{v,k,R}^{\prime }$ is the derivative of ψ in the radial direction.

Under the assumption that the Dirichlet problem (37) is uniquely solvable for $\psi \in {H}^{2}\left({B}_{R}^{3}\right)$ for all $g\in {H}^{3/2}\left({S}_{R}^{2}\right)$, we define the Dirichlet-to-Neumann map ${{\Lambda}}_{v,R}\in \mathcal{L}\left({H}^{3/2}\left({S}_{R}^{2}\right),{H}^{1/2}\left({S}_{R}^{2}\right)\right)$ by ${{\Lambda}}_{v,R}\varphi ={\frac{\partial \psi }{\partial r}\vert }_{{S}_{R}^{2}}$. Next we shall show that the partial scattering matrix elements sv, known for all ⩾ 0 uniquely determine the Dirichlet-to-Neumann map Λv,R.

Lemma 8. Let v1, v2 be two complex-valued potentials satisfying (7) and let $k\in {\mathbb{R}}_{+}{\backslash}\left({{\Sigma}}_{{v}_{1}}^{P}\cup {{\Sigma}}_{{v}_{2}}^{P}\right)$ be fixed. Besides, let $R\in \left[{R}_{a},+\infty \right){\backslash}\left({{\Sigma}}_{{v}_{1},k}^{D}\cup {{\Sigma}}_{{v}_{2},k}^{D}\right)$. Suppose that ${s}_{{v}_{1},\ell }={s}_{{v}_{2},\ell }$ for all ⩾ 0. Then ${{\Lambda}}_{{v}_{1},R}={{\Lambda}}_{{v}_{2},R}$.

Proof. It follows from lemmas 1 and 2 and from continuity of φv,(r) at r = R that ${{\Lambda}}_{{v}_{1},R\vert {\mathcal{H}}_{\ell ,R}}={{\Lambda}}_{{v}_{2},R\vert {\mathcal{H}}_{\ell ,R}}$,where ${\mathcal{H}}_{\ell ,R}$ denotes the space of restrictions to ${S}_{R}^{2}$ of harmonic polynomials of degree , spanned by the spherical harmonics ${Y}_{\ell }^{m}={Y}_{\ell }^{m}\left(\frac{x}{\vert x\vert }\right)$, |m| ⩽ . More precisely, the following explicit formula for ${{\Lambda}}_{{v}_{j},R\vert {\mathcal{H}}_{\ell ,R}}$ is valid:

Equation (40)

where η = α/(2k) and the denominator is non-zero as $R\notin {{\Sigma}}_{{v}_{j},k}^{D}$.

Now let $g\in {H}^{3/2}\left({S}_{R}^{2}\right)$ and denote by ${\psi }_{{v}_{j}}\in {H}^{2}\left({B}_{R}^{3}\right)$ the unique solution of the Dirichlet problem (37) with v = vj. Besides, define gN by

so that gNg in ${H}^{3/2}\left({S}_{R}^{2}\right)$ and, according to remark 5, ${{\Lambda}}_{{v}_{j},R}{g}_{N}\to {{\Lambda}}_{{v}_{j},R}g$ in ${H}^{1/2}\left({B}_{R}^{3}\right)$ as N. Together with (40), which shows that ${{\Lambda}}_{{v}_{1},R}{g}_{N}={{\Lambda}}_{{v}_{2},R}{g}_{N}$, this implies that ${{\Lambda}}_{{v}_{1},R}g={{\Lambda}}_{{v}_{2},R}g$ and concludes the proof of lemma 8. □

3.6. Demonstration of the uniqueness theorem

Now we combine the preliminary results established in sections 3.1, 3.2, 3.4 and 3.5 to prove theorem 1.

Under the assumptions of theorem 1 it follows from lemma 5 in case (a) and from lemma 6in case (b) that ${s}_{{v}_{1},\ell }={s}_{{v}_{2},\ell }$ for all ⩾ 0. Using lemma 8 we conclude that ${{\Lambda}}_{{v}_{1},R}={{\Lambda}}_{{v}_{2},R}$ for any R in the non-empty set $\left[{R}_{a},\infty \right){\backslash}\left({{\Sigma}}_{{v}_{1},k}^{D}\cup {{\Sigma}}_{{v}_{2},k}^{D}\right)$. It follows from the uniqueness theorem of [18], where the proof does not use that the potential is real-valued, that v1 = v2 a.e. This proves theorem 1.

4. Reconstruction

4.1. Reconstruction scheme for exact simulated data

The possibility to use measurements of the solar acoustic field at two heigths above the surface to recover the sound speed, density and attenuation inside of the Sun is confirmed by our numerical simulations. In this subsection we shall briefly describe the reconstruction algorithm that we use.

We assume that the unknown solar parameters q = (c, ρ, γ) are perturbations of some known background quantities q0 = (c0, ρ0, γ0) such that

Equation (41)

where R = 6.957 × 105 km is the solar radius, and we assume that both parameter sets q and q0 satisfy (2), (3). Let ${\Omega}\subset {\mathbb{R}}_{+}$ be a finite set of admissible frequencies such that ${\Omega}\cap \left({{\Sigma}}_{q}^{P}\cup {{\Sigma}}_{{q}^{0}}^{P}\right)=\varnothing $, where

${{\Sigma}}_{{v}_{\omega }}^{P}$ is defined according to (8), and the potentials vω and ${v}_{\omega }^{0}$ are defined using formula (6) with parameters q and q0, respectively. We recall that c0/(2H) is the acoustic cutoff frequency, which separates the regime of oscillations at eigenfrequencies and the scattering regime.

Put ${G}_{q}\left(h,\ell ,\omega \right)={G}_{{v}_{\omega },\ell }\left({R}_{\odot }+h,{R}_{\odot }+h\right)$. In view of lemma 4, as initial data for inversions from exact data we use the imaginary part of the Green's function Im Gq(h, , ω) measured at two different non-negative altitudes h ∈ {h1, h2}, at angular degrees ∈ {min, ..., max}, and at all admissible frequencies ω ∈ Ω. From this data we recover the solar parameters q as follows.

The first step of the algorithm consists in recovering the scattering matrix elements ${s}_{q}\left(\ell ,\omega \right)={s}_{{v}_{\omega },\ell }$, ∈ {min, ..., max}, ω ∈ Ω, which are defined according to lemma 2. This reconstruction is done by considering equation (34a) with r = R + h, h ∈ {h1, h2} as a linear system for finding $\mathfrak{R}{s}_{q}\left(\ell ,\omega \right)$, Im sq(, ω) at each fixed , ω.

At the next step the scattering matrix elements sq(, ω) are used to recover the map ${u}_{q}\,:\,I\to {\mathbb{R}}^{3}$(where I is the interval defined in (41)) which is defined as follows:

Equation (42)

and such that

This reconstruction is done by applying the iteratively regularized Gauss–Newton method, going back to [28], to the forward map

Equation (43)

where |Ω| denotes the number of elements in the set Ω. For more details on the iteratively regularized Gauss–Newton method and for sufficient conditions of its convergence see, e.g., [29]. Note that we do not iterate directly over c, ρ, γ as it would require a more complicated forward map defined on a Sobolev-type space, see formula (42).

Remark 6. For constructing vω from the Dirichlet-to-Neumann map one can also use methods developed, for example, in [30]. However, in the present work we restrict ourselves to the iteratively regularized Gauss–Newton method.

The last step is to determine q = (c, ρ, γ) from uq. Note that definitions (42) lead to the following explicit formulas for determining c and γ:

Also note that definitions (42) lead to the following problem for determining ρ:

which is solved for the unknown function ${\rho }^{-\frac{1}{2}}$. This step concludes the reconstruction algorithm from exact data.

4.2. Numerical example without noise

We consider the background sound speed and density from the model of [3], which extends the standard solar model of [4] to the upper atmosphere. We suppose that the background attenuation is equal to γ0/2π = 102.5 μHz inside of the Sun and decays to zero smoothly in the region [R, R + ha], where R = 6.957 × 105 km is the solar radius and ha = 500 km is the height above the photosphere at which the (conventional) interface between the lower and upper atmosphere is located. Note that this approximate value for the background attenuation can be obtained by analysing the observed full width at half maximum (FWHM) of acoustic modes, see [1, section 7.3] for more details8. We assume that the unknown perturbations to the background values of solar parameters are supported in the interval [0.9R, 0.95R]. Besides, the background sound speed used in our reconstructions is by a factor of 1.1 larger in the region [0, 0.5R] than the sound speed used to generate data.

The initial data for reconstructions is the imaginary part of the radiation Green's function Im Gq(h, , ω) at heights9 h ∈ {105, 144} km, angular degrees ∈ {50, ..., 215}, and frequencies ω/2π ∈ {5.3, 5.4} mHz. Note that observations of the acoustic field at these heights approximately correspond to measuring Doppler velocities at line center (three-point approximation) and using the center of gravity (six-point approximation) of the HMI absorption line [9].

We consider perturbations δc, δρ, δγ of the sound speed, density and attenuation of maximal relative magnitudes of 5%, 9%, and 100%, respectively. The perturbations are represented by piecewise constant functions with 200 equidistant nodes. Figure 1 shows exact profiles of perturbations δc, δρ, δγ, reconstructed profiles of perturbations δc, δρ, δγ, and relative L2 reconstruction errors e(δc, δc), e(δρ, δρ), e(δγ, δγ). Here,

This reconstruction example confirms the uniqueness results of section 2.2.

Figure 1.

Figure 1. Perturbations δc, δρ, δγ of solar parameters, reconstructed approximations δc, δρ, δγ and relative L2 reconstruction errors e(c, c), e(ρ, ρ), e(γ, γ) shown in %.

Standard image High-resolution image

Recall that the reconstructions are obtained using the data for angular degrees ∈ {50, ..., 215}. Note that:

  • (a)  
    data for lower angular degrees are sensitive to perturbations in the region [0, 0.5R] where our background model is wrong.
  • (b)  
    data for larger angular degrees are insensitive to perturbations in the region of interest [0.9R, 0.95R].

The reason is that the penetration depth of acoustic modes decreases with , as indicated by formula (39) and confirmed by plotting the regular solutions φv, for different , see figure 2.

Figure 2.

Figure 2. Regular solutions φv, for the background model normalized by their value at the measurement height R + 105 km. The frequency is ω/2π = 5.27 mHz.

Standard image High-resolution image

4.3. Reconstruction scheme for noisy data

The power spectrum ${P}_{{v}_{\omega },\ell }^{m}\left(r\right)=\mathbb{E}\vert {\varphi }_{{v}_{\omega },\ell }^{m}{\left(r\right)\vert }^{2}$ introduced in section 2.1 can be approximated from the data. A standard approach is to parse the time series of acoustic oscillations into N segments of equal duration T, compute for each segment the sample power spectrum, and then take the arithmetic average ${\hat{P}}_{{v}_{\omega },\ell }^{m}\left(r\right)$.

The duration T is chosen to achieve a desired frequency resolution. We recall that for a time series segment of duration T the frequency resolution of the sample power spectrum is equal to 1/T. Besides, for a time resolution (cadence) equal to ΔT, the sample power spectrum can be computed for frequencies up to the Nyquist frequency 1/(2ΔT).

It is a standard assumption going back to [31] (see also [32]) that

Equation (44)

where χ2(2N) denotes the chi-squared distribution with 2N degrees of freedom and r, ω, , m are fixed. We use relation (44) to simulate noisy data. These simulated data are then used to extract the diagonal values of the imaginary part of the radial Green's function $\Im {G}_{q}\left(h,\ell ,\omega \right)=\Im {G}_{{v}_{\omega },\ell }\left({R}_{\odot }+h,{R}_{\odot }+h\right)$ using formula (15) with the $O\left(\frac{1}{R}\right)$ term dropped and assuming that Π = 1. Note that in reality Π is a function of frequency which is not directly accessible to measurements; for a possible model of this function see, e.g., [1, section 7.5].

Then the reconstruction proceeds as described in section 4.1.

4.4. Numerical examples with noisy data

We consider a model situation where the solar oscillations are observed for a total period of eight years with the time resolution of ΔT = 45 s. We recall that the HMI instrument observes the solar oscillations continuously with this cadence since April 30, 2010. We assume that the sample power spectra are computed from three-day intervals (that is, T = 3 × 24 × 3600s), so that the frequency resolution is equal to 1/T ≈ 3.86 μHz and the Nyquist frequency is 1/(2ΔT) ≈ 11.11 mHz.

We simulate ${\hat{P}}_{{v}_{\omega },\ell }^{m}\left({R}_{\odot }+h\right)$ according to (44) with N = 974 (which is the number of three-day intervals constituting eight years of observations) for angular degrees ∈ {50, ..., 250}, azimuthal degree m = 0, observation heights h ∈ {105, 144} km and six equidistant frequencies ω/2π from the range [5.27, 5.38] mHz. We use the same assumptions on the unknown parameters and the same background models for data generation and reconstructions and as in section 4.2.

In contrast to reconstructions with exact data, simultaneous recovery of all the parameters from noisy data fails when these realistic settings are used. The point is that the (numerically computed) singular values of the forward map of formula (43) decrease exponentially fast, which leads to a severe ill-posedness of the inverse problem.

However, if two out of three parameters are known a priori, the third parameter is recovered with reasonable precision. Figure 3 shows parameters δc, δρ, δγ, their reconstructions δc, δρ, δγ from noisy data and relative L2 reconstruction errors for one realization of data. The mean relative L2 reconstruction errors for parameters δc, δρ, δγ are equal to

Besides, standard deviations of relative L2 reconstruction errors are equal to 1.85%, 10.62%, 2.1%. The statistics are computed from 60 simulations for each parameter. We emphasize that in each of these examples two out of three parameters are known a priori and fixed, and we reconstruct the remaining parameter.

Figure 3.

Figure 3. Perturbations δc, δρ, δγ of solar parameters, reconstructed approximations δc, δρ, δγ and relative L2 reconstruction errors for noisy data e(δc, δc), e(δρ, δρ), e(δγ, δγ) shown in %. In each case two out of three parameters are known a priori and fixed.

Standard image High-resolution image

These simulations show that reconstructions from noisy simulated data and, as a corollary, from experimental data, require a separate and detailed treatment, which is beyond the scope of the present article.

Recall that we use the data for six equidistant frequencies ω/2π in the range [5.27, 5.38] mHz. We do not use a larger range of frequencies because it would require to take into account the frequency dependence of attenuation which is not assumed by our model10. Our simulations also show that for a power law frequency dependence of attenuation

Equation (45)

as in [1], higher frequency waves have smaller penetration and are less sensitive to the medium properties in the region [0.9R, 0.95R], see figure 4. Note that the actual frequency dependence of attenuation is a subject of active research in helioseismology.

Figure 4.

Figure 4. Regular solutions φv, for the background model normalized by their value at the measurement height R + 105 km. The angular degree is = 100. The attenuation satisfies (45).

Standard image High-resolution image

5. Conclusion

We considered the inverse problem of recovering the radially symmetric sound speed, density and attenuation in the Sun from the measurements of the solar acoustic field at two heights above the photosphere and for a finite number of frequencies above the acoustic cutoff frequency. We showed that this problem reduces to recovering a long range potential (with a Coulomb-type decay at infinity) in a Schrödinger equation from the measurements of the imaginary part of the radiation Green's function at two distances from zero. We demonstrated that generically this inverse problem for the Schrödinger equation admits a unique solution, and that the original inverse problem for the Sun admits a unique solution when measurements are performed at least for two different frequencies above the acoustic cutoff frequency. These uniqueness results are confirmed by numerical experiments with simulated data without noise. However, simulations also show that the inverse problem is severly ill-posed, and only a single solar physical quantity can be reconstructed with precision (for example density at fixed sound-speed and attenuation) using a standard iterative reconstruction method (IRGNM) for realistic noise levels.

Acknowledgments

The authors would like to thank Prof. L Gizon for his helpful remarks and reference suggestions.

Footnotes

  • In section 2 we will assume that fω has a prescribed form of the cross-covariance function as in [1].

  • In reality the attenuation in the atmosphere results from the interaction with the magnetic field. In the present article we consider a simplified model without the magnetic field.

  • See also a preprint [14] where the properties of the outgoing solutions arising in a similar helioseismic model are investigated.

  • Our simulations show that another reasonable choice for the attenuation constant in the Sun is γ0/2π = 39.8 μHz. For this attenuation coefficient the difference between the observed by HMI and the modeled power spectra (after a linear transformation) is minimal.

  • A more realistic approximation is that the Doppler velocity signal is formed at a range of heights described by a bell-type contribution function with maximum at the 'formation height', see [9] for more details. Therefore the measurement heights can't be very close to avoid the undesirable correlations in data.

  • 10 

    The frequency dependence of attenuation can be incorporated in the model in a straightforward way.

Please wait… references are loading.
10.1088/1361-6420/ab77d9