Paper The following article is Open access

A frame decomposition of the atmospheric tomography operator

and

Published 31 August 2020 © 2020 The Author(s). Published by IOP Publishing Ltd
, , Special Issue on Modern Challenges in Imaging Citation Simon Hubmer and Ronny Ramlau 2020 Inverse Problems 36 094001 DOI 10.1088/1361-6420/aba4fe

0266-5611/36/9/094001

Abstract

We consider the problem of atmospheric tomography, as it appears for example in adaptive optics systems for extremely large telescopes. We derive a frame decomposition, i.e., a decomposition in terms of a frame, of the underlying atmospheric tomography operator, extending the singular-value-type decomposition results of Neubauer and Ramlau (2017 SIAM J. Appl. Math. 77 838–853) by allowing a mixture of both natural and laser guide stars, as well as arbitrary aperture shapes. Based on both analytical considerations as well as numerical illustrations, we provide insight into the properties of the derived frame decomposition and its building blocks.

Export citation and abstract BibTeX RIS

Original 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

The imaging quality of earthbound astronomical telescopes like the extremely large telescope (ELT) [11] of the European Southern Observatory (ESO), currently under construction in the Atacama desert in Chile, suffers from aberrations due to turbulences in the atmosphere, which result in blurred images. This is commonly counteracted by the use of adaptive optics (AO) systems, which use the measurements of one or more wavefront sensors (WFS) to suitably adjust deformable mirrors (DMs) in such a way that the incoming wavefronts are corrected (flattened) after reflection on the mirrors; see figure 1 (left). Since the atmosphere is constantly changing, this has to be done in real time. For more details on AO we refer to [7, 29, 30].

Figure 1.

Figure 1. Correction of an incoming wavefront by a deformable mirror (left, image taken from [2]) and sketch of an SCAO system (right, image taken from [6]).

Standard image High-resolution image

There are a number of different types of AO systems, the simplest one being single conjugate adaptive optics (SCAO). Thereby, the light of a so-called natural guide star (NGS), a bright star in the vicinity of an object of interest, is used to adjust the DM to obtain a corrected wavefront and thus a sharp image of the NGS and the nearby object. See figure 1 (right) for a schematic drawing of an SCAO system.

In case that there is no bright enough NGS available in the vicinity of the object of interest, or if one wants to achieve a good correction over either a larger or multiple fields of view, one needs to resort to different, more complex AO systems. A good unidirectional correction in the absence of a NGS in the vicinity of an object of interest is for example achieved by laser tomography adaptive optics (LTAO), while a correction over a large or multiple fields of view is achieved by multiconjugate adaptive optics (MCAO) and multiobject adaptive optics (MOAO), respectively [1, 5, 19, 24, 28]. These are schematically depicted in figure 2. In common with all of those different AO systems is the use of so-called laser guide stars (LGS), artificial stars created by powerful laser beams in the sodium layer of the atmosphere, which are used to increase the number of guide stars available for correction and thus enhance the imaging quality.

Figure 2.

Figure 2. Different fundamental types of AO systems. Magenta spirals stand for astronomical objects of interest, and red and green stars for (the location of) NGS and LGS, respectively. The light blue areas correspond to those directions of view, for which the AO systems aim at achieving a correction. Image taken from [2].

Standard image High-resolution image

Since an SCAO system succeeds at enhancing the imaging quality in one direction of view only, measurements from a single WFS are enough to compute suitable correction shapes of the DM, because wavefront aberrations from two objects close to each other, in this case from the reference NGS and the considered object of interest, are approximately the same. However, this is not the case for LTAO, MCAO, and MOAO, where the NGS and LGS are far away from the object of interest, or a good correction has to be achieved over a large or multiple fields of view. Hence, one has to use multiple WFSs (typically one for each guide star) and DMs, whose correcting shapes have to be determined from the turbulence profile of the atmosphere, which in turn has to be calculated from the WFS measurements. This gives rise to the problem of atmospheric tomography.

Unfortunately, and in particular since the separation of the NGSs and LGSs is low (e.g., 1 arcmin for MCAO and 3.5 arcmin for MOAO), the problem of atmospheric tomography falls into the category of limited-angle tomography, which is known to be a severely ill-posed problem [4, 21]. In addition, the number of available guide stars is relatively small as well (e.g., 6 LGSs for the ELT), which in combination with the severe ill-posedness makes the reconstruction of the full atmospheric turbulence above the telescope a hopeless endeavour. Hence, one works with the commonly accepted assumption that the atmosphere contains only a limited number of turbulent layers, which are infinitely thin and located at predetermined heights. The problem of atmospheric tomography then becomes the task of reconstructing the turbulences on only a finite number of turbulence layers from the available WFS measurements. For a schematic depiction with three layers, NGSs, and WFSs see figure 3 (left).

Figure 3.

Figure 3. Schematic depiction of an atmospheric tomography setup with three turbulence layers, NGSs, and WFSs (left). The coloured areas are those parts of the turbulence layers which are 'seen' by each of the different WFSs. Illustration of the cone effect (right) for a single LGS. Images taken from [38].

Standard image High-resolution image

A number of numerical reconstruction approaches have been proposed and developed for the atmospheric tomography problem, among them a minimum mean square error method [13], a back-projection algorithm [14], conjugate gradient type iterative reconstruction methods with suitable preconditioning [9, 15, 18, 36, 37], the fractal iterative method [3335], the finite element wavelet hybrid algorithm [20, 3840], and Kaczmarz iteration [27, 31]; see also [8, 16, 17, 23, 25, 26, 32] and the references therein. All of these methods work comparatively well, each with its own peculiar advantages and drawbacks, and the resulting reconstructions have been successfully used to enhance the overall imaging quality of the corresponding AO systems. However, these numerical approaches themselves do not provide any deeper insight into the atmospheric tomography problem itself.

Hence, the authors of [22] set out to provide a mathematical analysis of the atmospheric tomography operator (defined below) underlying the problem, which is derived from the Radon transform using the layered and limited-angle structure of the problem [7, 13]. In particular, they derived a singular-value-type decomposition of the operator, which not only provides the basis for efficient numerical reconstruction methods but also allowed to gain insight into the ill-posedness of the problem itself. In contrast to the already known singular-value decomposition of the limited-angle tomography operator [4, 21], the singular values of this decomposition could be computed explicitly.

However, the singular-value-type decomposition derived in [22] is only valid under a very restrictive set of assumptions. In particular, leaving aside some technicalities, it is only valid for square aperture shapes and tomography settings with only NGSs and no LGSs. This is obviously problematic for two reasons: firstly, the aperture shapes of telescopes is usually not square, and secondly, as we have already seen above, most of the AO systems which rely on atmospheric tomography include LGSs as an integral part of their design. Hence, many of the interesting theoretical results derived in [22] no longer hold for those (practically) important cases. Furthermore, while numerical routines based on this singular-value-type decomposition can in principle be adapted via measurement extension to (partly) circumvent the restriction of square aperture shapes, an adaption to include also LGSs is not possible in any straightforward way. This is mainly due to the so-called cone effect: since a LGS is created by a powerful laser beam inside the sodium layer of the atmosphere, it is, in contrast to an 'infinitely far away' NGS, located at a finite height. Thus, light travelling from the LGS to the telescope pupil passes through larger areas in lower atmospheric layers than in higher ones; see figure 3 (right) for an illustration. Mathematically, this results in the addition of a layer and guide star dependent scaling parameter in the atmospheric tomography operator (see below), which causes a number of complications.

The aim of this paper is to overcome the restrictions of the singular-value-type decomposition in [22] of the atmospheric tomography operator noted above. In particular, we want to find a decomposition which allows both LGSs and arbitrary aperture shapes. This is done in two steps: first, we consider the case of a tomography setup with only LGSs and no NGSs. Setting aside for the moment considerations on the practicality of such a setup (e.g., the tip/tilt problem), this is at the same time both a completion and a natural extension of the results of [22], and an ideal starting point for deriving the main results of this paper. Then, we provide a decomposition of the atmospheric tomography operator for general problem settings including both a mixture of NGSs and LGSs, as well as arbitrary aperture shapes. This decomposition is done in terms of a set of functions which form a frame, with important implications for both theoretical as well as numerical aspects of the tomography problem.

The outline of this paper is as follows. In section 2 we describe the precise mathematical setting of atmospheric tomography considered in this paper, and in section 3 we review some necessary material on frames in Hilbert spaces. In section 4 we derive the decompositions for the atmospheric tomography operator mentioned above, first in the case of square domains and LGSs only, and then for a mixture of NGSs and LGSs as well as arbitrary aperture shapes. Section 5 presents some numerical results based on the obtained analytical derivations and is followed by a short conclusion in section 6.

2. Mathematical setting

In this section, we describe the precise mathematical setting of the atmospheric tomography problem considered in this paper. For the sake of consistency, we mainly use the same notations as in [22].

The atmospheric tomography problem is a limited-angle tomography problem with only finitely many directions of view αg , g = 1, ..., G, where G denotes the total number of NGSs and LGSs. The directions ${\alpha }_{g}\in {\mathbb{R}}^{2}$ are such that if $\left({\alpha }_{g}^{x},{\alpha }_{g}^{y},1\right)\in {\mathbb{R}}^{3}$ points from the center of the telescope to the guide star g, then ${\alpha }_{g}=\left({\alpha }_{g}^{x},{\alpha }_{g}^{y}\right)\in {\mathbb{R}}^{2}$. We denote the telescope aperture with ${{\Omega}}_{A}\subset {\mathbb{R}}^{2}$ and assume that the atmosphere contains L layers, where each layer is a plane at height hl , l = 1, ..., L parallel to ΩA .

Since we do not only consider NGSs but also LGSs in this paper, we need to, in particular, take the cone effect into account (see above). For this, we need to define the parameters cl,g . Assuming that GNGS and GLGS denote the number of NGSs and LGSs, respectively, such that G = GNGS + GLGS, we set

where hLGS denotes the height of the LGSs. Since hl < hLGS for all l = 1, ..., L, we have that cl,g ⩽ 1. Furthermore, since we assume that the hl are in ascending order and that hL < hLGS, we have that ${c}_{l,g}{\geqslant}1-\frac{{h}_{L}}{{h}_{\text{LGS}}}{ >}0$.

For every layer l, we can now define the domains

where

The domains Ωl are exactly those parts of the layers which are 'seen' by the WFSs (compare with figure 3) and are therefore those parts of the atmosphere on which one can expect to reconstruct the atmospheric turbulences.

Denoting by $\phi ={\left({\phi }_{l}\right)}_{l=1,\dots ,L}$ the turbulence layers and by $\varphi ={\left({\varphi }_{g}\right)}_{g=1,\dots ,G}$ the incoming wavefronts, the atmospheric tomography operator can now by defined as follows:

Equation (2.1)

where ${{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$ denotes the Cartesian product space ${\prod }_{g=1}^{G}{L}_{2}\left({{\Omega}}_{A}\right)$. On the definition and image spaces of A we define the canonic inner products

Equation (2.2)

Completely analogous to [22, theorem 3.1], we have that the operator A with respect to the above scalar products is not compact, and hence, a singular system does not necessarily need to exist for A.

However, the authors of [22] managed to derive a singular-value-type decomposition of what they called the periodic atmospheric tomography operator Ã, given by

Equation (2.3)

where ΩT := [−T,T]2 with T sufficiently large, such that

and under the assumption that functions in L2T ) are periodic. This specific extension of the operator A for the case of only NGSs allowed the use (of the special properties) of the functions

Equation (2.4)

which form an orthonormal basis of L2T ). However, the derived decomposition introduces artefacts due to the periodicity assumptions and in particular does not cover the case of LGSs and mixtures of NGSs and LGSs. Furthermore, in practice, wavefront measurements are not given on the extended domain ΩT but on ΩA only. Thus, for applying the decomposition derived in [22], these measurements have to be extrapolated to ΩT in some way, which is also not desirable.

Hence, in section 4, we extend the singular-value-type decomposition from [22] by deriving a frame decomposition of the original operator A, based on the ideas of [22] but avoiding some of their shortcomings. For this, we first review some necessary materials on frames in Hilbert spaces below.

3. Frames in Hilbert spaces

For the upcoming analysis, we need to recall some known results on frames in Hilbert spaces, which can for example be found in [3]. First, recall the definition of a frame:

Definition 3.1. A sequence ${\left\{{e}_{k}\right\}}_{k\in K}$ for some countable index set K in a Hilbert space H is a frame, if and only if there exist numbers B1, B2 > 0 such that for all fH we have

Equation (3.1)

The numbers B1, B2 are called frame bounds. The frame is called tight if B1 = B2.

For a given frame ${\left\{{e}_{k}\right\}}_{k\in K}$, one can consider the so-called frame (analysis) operator F and its adjoint F*, or synthesis-operator, defined by

Equation (3.2)

Due to (3.1) and the general fact that ∥F∥ = ∥F*∥, there holds that

Equation (3.3)

Furthermore, one can define the operator S := F*F, i.e.,

and it follows that S is a bounded linear operator with B1 ISB2 I, where I denotes the identity operator. Furthermore, S is invertible and ${B}_{2}^{-1}I{\leqslant}{S}^{-1}{\leqslant}{B}_{1}^{-1}I$. It follows that if one defines tilde ek := S−1 ek , then the set ${\left\{{\tilde {e}}_{k}\right\}}_{k\in K}$ also forms a frame, with bounds ${B}_{2}^{-1},{B}_{1}^{-1}$, which is called the dual frame of ${\left\{{e}_{k}\right\}}_{k\in K}$. It is also known that every fH can be written in the form

Equation (3.4)

It is not always possible to compute the dual frame tilde ek explicitly. However, since it holds that [3]

Equation (3.5)

where $R{:=}I-\frac{2}{A+B}S$, the elements of the dual frame can be approximated by only summing up to a finite index N, i.e.,

Equation (3.6)

The induced error of this approximation is controlled by the frame bounds A, B, i.e.,

Equation (3.7)

For a numerical implementation, (3.6) can also be written in the following recursive form, which allows for an efficient numerical implementation in practice

Equation (3.8)

Although for frames the decomposition of f in terms of the functions ek is not unique, the representation in (3.4) is the most economical, in the sense the following

Proposition 3.1. If $f=\sum _{k\in K}{a}_{k}{e}_{k}$ for some sequence ${\left\{{a}_{k}\right\}}_{k\in K}\in {\ell }_{2}\left(K\right)$ and if ${a}_{k}\ne \left\langle \enspace f,{\tilde {e}}_{k}\enspace \right\rangle $ for some kK, then

Proof. See for example proposition 3.2.4 of [3].□

Note that if ${\left\{{e}_{k}\right\}}_{k\in K}$ is a tight frame with bounds B1 = B2 = B, then we have that S−1 = B−1 I, tilde ek = ek /B and therefore

Equation (3.9)

Using these results, we now proceed to derive a frame decomposition of the atmospheric tomography operator A below.

4. Frame decomposition

In this section, we first derive a singular-value-type decomposition of the operator à in the case of only LGSs, following the ideas of [22]. Afterwards, we derive a frame decomposition of the operator A as defined in (2.1) where, in contrast to [22], we do not restrict ourselves to only NGSs, but allow a mixture of both NGSs and LGSs, as well as (almost) arbitrary aperture shapes ΩA .

4.1. The pure LGS case

In this section, we consider the periodic tomography operator (2.3) from [22], but now for the case that instead of only NGSs, we consider a setting using only LGSs. Since in this case cl,g = cl independent of the guide star direction g, the operator à can now be written in the form

Equation (4.1)

where, as before, ΩT = [−T, T], however now with T sufficiently large, such that

We now derive a singular-value-type decomposition of the adapted operator à using the ideas from [22]. First, due to the presence of the constants cl , it makes sense to use, for ever layer l, a different orthonormal basis of L2(cl ΩT ), namely the functions

Equation (4.2)

Any function ϕl L2(cl ΩT ) can be written in the form

Equation (4.3)

and for $\phi ={\left({\phi }_{l}\right)}_{l=1}^{L}\in {\prod }_{l=1}^{L}{L}_{2}\left({c}_{l}\enspace {{\Omega}}_{T}\right)$ we collect the expansion coefficients ϕjk,l in the vectors ${\phi }_{jk}{:=}{\left({\phi }_{jk,l}\right)}_{l=1}^{L}\in {\mathbb{C}}^{L}$. With this, we now get

Proposition 4.1. Let à be defined as in (4.1) and let the G × L matrices Ãjk be defined by

Then there holds

Equation (4.4)

Proof. Using the definition (4.1) of à and the expansion (4.3), we get

from which the assertion immediately follows after interchanging the series, which is allowed since the norm of Ãjk is bounded independently of j, k.□

As in [22] we consider the singular system for each of the matrices Ãjk , i.e., the vectors ${v}_{jk,n}\in {\mathbb{C}}^{L}$, ${u}_{jk,n}\in {\mathbb{C}}^{G}$ and numbers σjk,n , n = 1, ..., rjk ⩽ min {G, L} such that

Equation (4.5)

where the superscript H denotes the Hermitian adjoint (i.e. the complex transpose), rjk is the rank of the matrix Ãjk , and the ${\sigma }_{jk,n}^{2}$ are the positive eigenvalues of the matrices ${\tilde{A}}_{jk}^{H}{\tilde{A}}_{jk}$ and ${\tilde{A}}_{jk}{\tilde{A}}_{jk}^{H}$, respectively.

Hence, the decomposition of à in terms of the functions wjk is given by

Equation (4.6)

which is completely analogous to [22, equation (10)], however with a different singular systems (σjk,n , ujk,n , vjk,n ). Thus, for incoming wavefronts

the best-approximate solution of the equation

is given by

where à denotes the Moore–Penrose generalized inverse of à (see for example [10, 22]), which is well-defined if and only if the following Picard-condition holds

4.2. The general case

In this section, we derive a frame decomposition for the general atmospheric tomography operator A as defined in (2.1), i.e., for the case of arbitrary aperture domains ΩA and both NGSs and LGSs.

The main idea of [22], where some of the ideas of the upcoming analysis are taken from, was to use the special properties of the functions wjk (2.4), in particular that they form an orthogonal basis of L2A ). Unfortunately, for general domains ΩA , the functions wjk do not necessarily form a basis of L2A ). However, they do form a (tight) frame in the sense of definition 3.1, as we see in the following

Lemma 4.2. Let wjk be defined as in (2.4). If T is large enough such that ΩA ⊂ ΩT , then the system ${\left\{{w}_{jk}\right\}}_{j,k\in \mathbb{Z}}$ forms a tight frame over L2A ) with frame bound 1, i.e.,

Proof. We start by defining $\tilde {\psi }{:=}\psi {I}_{{{\Omega}}_{A}}$ where ${I}_{{{\Omega}}_{A}}$ denotes the indicator function of ΩA . Now since $\tilde {\psi }\in {L}_{2}\left({{\Omega}}_{T}\right)$ and $\tilde {\psi }=\psi $ on ΩA and $\tilde {\psi }=0$ on ΩT A , we get

where we have used that since the wjk form an orthonormal basis of L2T ), they are also a tight frame with frame bounds B1 = B2 = 1. This proves the assertion.□

Now, since ${\left\{{w}_{jk}\right\}}_{jk\in \mathbb{Z}}$ forms a tight frame with bound 1, it is also its own dual frame. Hence, (3.4) implies that any function ψL2A ) can be written in the form

Equation (4.7)

In particular, we have that the incoming wavefronts φ can be written in the form

Equation (4.8)

We also want to expand functions on L2l ) in terms of frames. It is not difficult to find frames for L2l ); for example, for large enough T, the sets ${\left\{{w}_{jk}\right\}}_{jk\in \mathbb{Z}}$ or ${\left\{{w}_{jk,l}\right\}}_{jk\in \mathbb{Z}}$ already form frames over L2l ). However, for the upcoming analysis, those frames do not satisfy the specific needs of the problem under consideration. Hence, we use a different, problem tailored frame, which we build from the functions

Equation (4.9)

with wjk defined as in (2.4), where we choose T large enough such that

Equation (4.10)

which we assume to hold from now on. That these functions can indeed form frames can be seen from the following

Proposition 4.3. Let wjk,lg be defined as in (4.9) and let T be large enough such that (4.10) holds. Then, for fixed l, the set ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},\enspace g=1,\dots ,G}$ forms a frame over L2l ) with frame bounds 1 ⩽ B1B2G.

Proof. Let ψL2l ) be arbitrary but fixed and start by looking at

Equation (4.11)

where we have used that cl,g ΩA + αg hl ⊂ Ωl cl,g ΩT . Now since for each (fixed) g the functions ${c}_{l,g}^{-1}{w}_{jk}\left(\left(x,y\right)/{c}_{l,g}\right)$ form an orthonormal basis of L2(cl,g ΩT ) (and thus a tight frame with bound 1), we get that

which, together with (4.11), yields

Equation (4.12)

Now since ${{\Omega}}_{l}={\bigcup }_{g=1}^{G}\left({c}_{l,g}{{\Omega}}_{A}+{\alpha }_{g}{h}_{l}\right)$, we get that

Combing this together with (4.12), we get that

which proves the assertion.□

Now, denoting for fixed l the dual frame of ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},\enspace g=1,\dots ,G}$ by ${\left\{{\tilde {w}}_{jk,lg}\right\}}_{jk\in \mathbb{Z},\enspace g=1,\dots ,G}$, the above proposition together with (3.4) implies that any function ϕl L2l ) can be written in the form

Equation (4.13)

Next, we want to find an expansion of in terms of the frame ${\left\{{w}_{jk}\right\}}_{jk\in \mathbb{Z}}$. Due to (4.7) and since L2A ), such an expansion is given by

Equation (4.14)

As we already know, even though this might not be the only possible expansion, it is the most economical one in the sense of proposition 3.1. Furthermore, due to (3.9), this expansion allows us to express ∥φ∥ in terms of the expansion coefficients of and φ, which is important for determining an (approximate) solution of = φ. We now derive an explicit expression for the coefficients ${\left\langle \enspace {\left(A\phi \right)}_{g},{w}_{jk}\enspace \right\rangle }_{{L}_{2}\left({{\Omega}}_{A}\right)}$ in terms of the coefficients ϕjk,lg in the following

Proposition 4.4. Let $\phi ={\left({\phi }_{l}\right)}_{l=1,\dots ,L}$ with ϕl L2l ) for all l = 1, ..., L, and let A, wjk , and ϕjk,lg be defined as in (2.1), (2.4), and (4.13), respectively. Then there holds

Equation (4.15)

Proof. Due to the definition of A, and setting r = (x, y) we have

Using substitution in the integral yields

Since cl,g ΩA + αg hl ⊂ Ωl , due to (4.9) and (4.13), we get

Combining the above results, we get

which directly yields the assertion.□

Hence, combining (4.14) and the above proposition, we get that

Equation (4.16)

Thus, if we define (with a slight abuse of notation), the vectors

Equation (4.17)

and the ${\mathbb{C}}^{G{\times}\left(L\cdot G\right)}$ matrices

Equation (4.18)

then we get the following expansion of the tomography operator

Equation (4.19)

The above expansion is obviously similar to (4.4) for the pure LGS case, however now with different (and slightly larger) matrices Ajk . Hence, we can again consider the singular value decomposition of each of the matrices Ajk , i.e., with another small abuse of notation, the vectors ${v}_{jk,n}\in {\mathbb{C}}^{L\cdot G}$, ${u}_{jk,n}\in {\mathbb{C}}^{G}$ and numbers σjk,n , n = 1, ..., rjk G such that

Equation (4.20)

to get, in complete analogy to (4.6), the following frame decomposition of the atmospheric tomography operator A defined in (2.1):

Equation (4.21)

Using this decomposition, we are now able to define the operator

Equation (4.22)

which we use to find a solution to the equation = φ in theorem 4.8 below. Before we proceed to that though, we observe that the structure of the matrices Ajk allows to compute their singular-value-decomposition explicitly, which leads to the following

Proposition 4.5. Let ${\left({\sigma }_{jk,n},{u}_{jk,n},{v}_{jk,n}\right)}_{n=1}^{{r}_{jk}}$, for $j,k\in \mathbb{Z}$ be the singular systems of the matrices Ajk as defined in (4.20). Then there holds

Equation (4.23)

where ${ \overrightarrow {e}}_{n}$ denotes the nth unit vector in ${\mathbb{C}}^{G}$.

Proof. Due to the structure of Ajk , we have that

from which the formula for σjk,n , ${u}_{jk,n}={ \overrightarrow {e}}_{n}$ and rjk = G immediately follow. Furthermore, since there holds

the expression for (vjk,n ) immediately follows, which concludes the proof.□

Using the explicit representation derived above, we get the following

Corollary 4.6. The operator $\mathcal{A}$ defined in (4.22) can be written in the form

Equation (4.24)

with the explicit form of the singular-values σjk,g as derived in proposition 4.5.

Proof. Note that by proposition 4.5, it follows that

from which, together with the expression for σjk,n from proposition 4.5, there follows

Equation (4.25)

Furthermore, again due to proposition 4.5, we get that

Equation (4.26)

which, together with (4.25) and the definition of $\mathcal{A}$ immediately yields the assertion.□

Concerning the well-definedness of $\mathcal{A}$, we can now derive the following

Lemma 4.7. Let $\varphi \in {{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$ and let $\mathcal{A}$ be defined as in (4.22). Then $\mathcal{A}\varphi $ is well-defined and there holds

Equation (4.27)

Proof. Let ${\tilde {F}}_{l}$ be the frame operator [compare with (3.2)] corresponding to the dual frame ${\left\{{\tilde {w}}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$ of ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$ and let ${\tilde {F}}_{l}^{{\ast}}$ be its adjoint. Since the dual frame is also a frame but with inverse frame bounds, it follows from proposition 4.3 together with (3.3) that

Hence, for any sequence ${a}_{l}={\left\{{a}_{jk,\mathrm{lg}}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}\in {\ell }_{2}$ there holds

Equation (4.28)

which we now use for the choice

Equation (4.29)

where ${\left\{{\varphi }_{jk}\right\}}_{jk\in \mathbb{Z}}$ are the expansion coefficients of φ defined in (4.8). Since due to (4.24) and the choice of al there holds

Equation (4.30)

we now obtain from (4.28) that

Now summing over l we get that

and therefore,

where the last equality follows from the definition (4.8) of the coefficients φjk together with lemma 4.2. Due to proposition 4.5 and since 0 < cl,g ⩽ 1 the singular values σjk,g are independent of j, k and bounded away from 0. Hence, since $\varphi \in {{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$, it follows that $\mathcal{A}\varphi $ is well-defined. Furthermore, the above estimate together with the explicit expression for σjk,g from proposition 4.5 implies that

which, together with 0 < cl,g ⩽ 1, immediately yields the assertion.□

With this, we are now able to prove the main theorem of this section.

Theorem 4.8. Let $\varphi \in {{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$ be given and let ${\left\{{\varphi }_{jk}\right\}}_{jk\in \mathbb{Z}}$ be its expansion coefficients defined as in (4.8). Furthermore, let ${\left({\sigma }_{jk,n},{u}_{jk,n},{v}_{jk,n}\right)}_{n=1}^{{r}_{jk}}$ for $j,k\in \mathbb{Z}$ be the singular systems of the matrices Ajk defined in (4.20). Moreover, for all $l\in \left\{1,\dots ,L\right\}$ let al be defined as in (4.29) and satisfy al R(Fl ), where Fl denotes the frame operator corresponding to ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$. Then the function $\mathcal{A}\varphi $ is a solution of the equation = φ. Furthermore, among all other solutions $\psi \in \mathcal{D}\left(A\right)$ of = φ there holds

Equation (4.31)

Moreover, among all other possible expansions of $\mathcal{A}\varphi $ in terms of the functions ${\tilde {w}}_{jk,lg}$, (4.22) is the most economical one in the sense of proposition 3.1.

Proof. Let $\varphi \in {{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$ and let ${\varphi }_{jk,g}={\left\langle \enspace {\varphi }_{g},{w}_{jk}\enspace \right\rangle }_{{L}_{2}\left({{\Omega}}_{A}\right)}$ be its canonical expansion coefficients, collected in the vectors φjk ; compare with (4.8). Furthermore, let $\phi \in \mathcal{D}\left(A\right)$ and let ${\phi }_{jk,lg}={\left\langle \enspace {\phi }_{l},{w}_{jk,lg}\enspace \right\rangle }_{{L}_{2}\left({{\Omega}}_{l}\right)}$ be its canonical expansion coefficients, collected in the vectors ϕjk ; compare with (4.13) and (4.17). Due to proposition 4.4 there holds,

Equation (4.32)

where the matrices Ajk are as in (4.18). Since due to lemma 4.2 the set ${\left\{{w}_{jk}\right\}}_{jk\in \mathbb{Z}}$ forms a tight frame over L2A ) with frame bound 1, it follows with (3.9) that

Hence, together with

we obtain

Equation (4.33)

Thus, ϕ is a solution of = φ if and only if its expansion coefficients ϕjk satisfy

Equation (4.34)

The solutions of these matrix-vector systems can be characterized via the singular systems ${\left({\sigma }_{jk,n},{u}_{jk,n},{v}_{jk,n}\right)}_{n=1}^{{r}_{jk}}$ of the matrices Ajk . Since in proposition 4.5 we showed that rjk = G and ${u}_{jk,n}={ \overrightarrow {e}}_{n}$, it follows that at least one solution of (4.34) exists. Hence, by the properties of the SVD it follows that the vectors

Equation (4.35)

are the unique solution of (4.34) with minimal 2 norm. Hence, if ϕ can be found such that ${\phi }_{jk}={A}_{jk}^{{\dagger}}{\varphi }_{jk}$, then for all other solutions ψ of = φ there holds

Equation (4.36)

We now show that the choice $\phi {:=}\mathcal{A}\varphi $ is exactly such that

Equation (4.37)

Since due to (4.17) there holds ${\phi }_{jk,lg}={\left({\phi }_{jk}\right)}_{l+\left(g-1\right)L}$, this is equivalent to showing

For this, note first that as in the proof of corollary 4.6 we have that

and thus together with the definition of al in (4.29) there holds

Equation (4.38)

Hence, we now have to show that

To do so, note that in (4.30) in the proof of lemma 4.7 we saw that

where ${\tilde {F}}_{l}$ is the frame operator of the dual frame ${\left\{{\tilde {w}}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$. Hence, we obtain

where we used the fact (see [3]) that ${F}_{l}{\tilde {F}}_{l}^{{\ast}}={P}_{l}$, with Pl denoting the orthogonal projector in 2 onto $R\left({\tilde {F}}_{l}\right)=R\left({F}_{l}\right)$, and that by assumption al R(Fl ). However, since this implies that

it now follows that (4.37) holds, and thus $\mathcal{A}\varphi $ is a solution of = φ. Together with (4.36), this also implies the coefficient inequality (4.31). Note that due to lemma 4.7, $\mathcal{A}\varphi $ is well-defined since $\varphi \in {{L}_{2}\left({{\Omega}}_{A}\right)}^{G}$. Finally, since ${\phi }_{jk,lg}={\left\langle \enspace {\phi }_{l},{w}_{jk,lg}\enspace \right\rangle }_{{L}_{2}\left({{\Omega}}_{l}\right)}={\left({a}_{l}\right)}_{jk,g}$, it follows that the expansion (4.22) is the most economical one in the sense of proposition 3.1, which concludes the proof.□

Remark. In the above theorem we saw that $\mathcal{A}\varphi $ is a solution of the atmospheric tomography problem given that al R(Fl ), which can be interpreted as a consistency condition. On the other hand, if instead we only assume that = φ is solvable, then it follows from (4.33) that for any solution ϕ there holds Ajk ϕjk = φjk , and thus that

Since the complete set of solutions is given by ϕ + N(A), where ϕN(A) denotes the minimum-norm solution of = φ, and since by (4.33) for any $\tilde {\phi }\in N\left(A\right)$ there holds ${\tilde {\phi }}_{jk}\in N\left({A}_{jk}\right)$, it follows that

Noting that this can be rewritten as

Equation (4.39)

and since due to (4.30) and (4.38) there holds

it now follows from (4.39) that

which implies that if al R(Fl ) one can still consider $\mathcal{A}\varphi $ as an approximate solution.

Remark. The explicit representation of $\mathcal{A}$ given in (4.24) can be used as the basis of an efficient numerical routine for (approximately) solving the atmospheric tomography problem = φ. Replacing the infinite sum over the indices j, k by a finite sum, all one needs for implementation are the vectors φjk and the evaluation of the functions ${\tilde {w}}_{jk,lg}\left(x,y\right)$ on a predetermined grid. For this, note that since the functions ${\left\{{\tilde {w}}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$ form the dual frame of the functions ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$, they can be efficiently numerically approximated via the approximation formula (3.8). Furthermore, the coefficients φjk can be efficiently computed via (4.8) using the Fourier transform. Thus, since apart from φjk all other quantities can be precomputed independently of the input data φ, the computation of $\mathcal{A}\varphi $ via formula (4.24) can be efficiently numerically implemented.

Remark. In lemma 4.7, we have derived that ${\Vert}\mathcal{A}{\Vert}{\leqslant}1/L$, i.e., that $\mathcal{A}$ is bounded. In particular, for any noisy measurement φδ of the incoming wavefronts φ this implies

which shows that the (approximate) solution of the atmospheric tomography problem via the application of the operator $\mathcal{A}$ is stable with respect to noise in the data.

At first, this result seems counter-intuitive, since the atmospheric tomography operator is derived from the Radon transform under the assumption of a layered atmosphere. Hence, since inverting the Radon transform is known to be an ill-posed problem, one expects $\mathcal{A}$ to become unbounded with an increasing number of layers L. However, note that on the one hand $\mathcal{A}\varphi $ is only an approximate solution for al R(Fl ), and on the other hand, for L going to infinity, the atmospheric tomography operator A as defined in (2.1) does not tend towards the (limited-angle) Radon transform.

To see this, note that the sum in the definition of A stems from the discretization of the line integrals of the Radon transform via the simple quadrature rule

where the weights (xk xk−1) were dropped. Assuming that (xk xk−1) = 1/L, which would correspond to equidistant atmospheric layers, the above quadrature rule becomes

which indicates that unless the atmospheric tomography operator is multiplied by 1/L in this case, it does not converge to the desired limit as the number of layers tends to infinity. Incidentally, if A is replaced by (1/L)A, then $\mathcal{A}$ has to be replaced by $L\mathcal{A}$, and it then follows from (4.27) that

Equation (4.40)

resulting in an upper bound for the solution operator $L\mathcal{A}$ which tends towards infinity for an increasing number of layers L, as one would expect.

Note that further adaptations to the tomography operator A would have to be made in order to ensure its convergence to the Radon transform in the limit for L. These correspond to the integration weights having to be adapted to different guide star directions. However, since this is practically not relevant as L is generally fixed, we do not go into further details here. In summary, we want to emphasize that the upper bound on $\mathcal{A}$ does not contradict the ill-posedness of the general tomography problem.

Remark. Equation (4.31) implies that $\mathcal{A}\varphi $ can be seen as the minimum-coefficient solution of = φ. Since the sets ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$ do not form tight frames over L2l ), $\mathcal{A}\varphi $ is not necessarily also the minimum-norm solution of the equation.

One way to obtain a minimum-norm solution is to take Ml piecewise disjoint domains Ωl,m ⊂ Ωl satisfying

and to use, instead of ${\left\{{w}_{jk,lg}\right\}}_{jk\in \mathbb{Z},g=1,\dots ,G}$, the set of functions

where

Similarly as in the proof of proposition 4.3, it can be shown that for fixed l, these functions form a tight frame with frame bound G for L2l ). Hence, in complete analogy to (4.13), it is possible to decompose ϕl as follows:

Equation (4.41)

Following the same steps as in the proof of proposition 4.4, one finds that

from which it follows that

Defining (again with a small abuse of notation) the vectors

Equation (4.42)

we get, in analogy to (4.19), that A can be written as

Equation (4.43)

but now with

where

and

Denoting again by ${\left({\sigma }_{jk,n},{u}_{jk,n},{v}_{jk,n}\right)}_{n=1}^{{r}_{jk}}$ the singular value decompositions of the matrices Ajk , we arrive at the same decomposition of A as in (4.21), but obviously with the new singular values and functions. All results derived above hold analogously for this new decomposition, with the difference that, due to the tightness of the frame, $\mathcal{A}\varphi $ is then not only the minimum-coefficient solution of = φ, but also the minimum-norm solution. Furthermore, one can again derive the singular values of Ajk explicitly, namely

In addition, as in (4.7) one can again show that $\mathcal{A}\varphi $ is well-defined and that

However, note that using this approach, the number of subdomains Ωl,m can increase strongly with the number of guide stars G. Furthermore, note that the number of frames per layer considered in the two approaches presented in this section are directly related to G and Ml , respectively. Hence, since the used frames have discontinuities, using the approach based on the frames {wjk,lgm } potentially introduces more discontinuities into the solution in a numerical implementation than the approach based on the frames {wjk,lg }, which is not necessarily desirable in practice.

Remark. In some situations, it is desirable to, instead of the standard L2 inner product (2.2), work with the weighted inner product

Equation (4.44)

where ${\left\{{\gamma }_{l}\right\}}_{l=1}^{L}$ denotes a normalized, non-zero sequence of weights. Although we do not focus on this issue in our current investigation, it should be noted that adapting the frame decomposition to include this weighted inner product should be straightforward.

5. Numerical illustrations

In this section, we present some numerical illustrations showcasing the building blocks of the frame decomposition of the atmospheric tomography operator derived above, namely the frame functions wjk,lg and their corresponding dual frame functions ${\tilde {w}}_{jk,lg}$. While the functions wjk,lg have been defined explicitly and thus can be easily visualized, this is not the case for the functions ${\tilde {w}}_{jk,lg}$, which are only implicitly defined. However, since they are the foundation of the frame decomposition, it is of interest and importance to get some idea about their visual representation, which we aim to do in this section.

It was already noted above that the dual frame functions ${\tilde {w}}_{jk,lg}$ can be numerically approximated via formula (3.8), and that this iterative process can be implemented efficiently using the Fourier transform. However, the functions ${\tilde {w}}_{jk,lg}$ are problem-adapted, i.e., they are dependent on the concrete parameter setting of the considered atmospheric tomography problem, which we thus first need to decide upon. Hence, for our purposes, we take the same setup as in [22], which in turn is adapted from the MAORY setup [5, 12] as planned for the ELT currently being built by ESO. More precisely, we consider a telescope with a diameter of 42 m (the originally planned ELT size), which leads to a circular aperture domain ΩA , as well as 6 NGS positioned uniformly on a circle with a diameter of 1 arcmin, with the corresponding directions αg given in the following table (ρ = 0.000 290 888, a = ρ ⋅ 0.5, b = ρ ⋅ 0.866 025):

g 123456
αg (ρ, 0)(a, b)(−a, b)(−ρ, 0)(−a, −b)(a, −b)

For defining the layer profile we choose the ESO standard atmosphere, which consists of 9 layers located at the heights hl given in the following table:

l 123456789
hl (m)0140281562112522504500900018000

For the numerical implementation, and in particular for the approximation of the dual frame functions ${\tilde {w}}_{jk,lg}$ via formula (3.8), which in our present case reads as

Equation (5.1)

one can only consider a finite number of indices j, k (and thus also of p, q). Hence, for our numerical illustrations, we restricted ourselves to $j,k,p,q\in \left\{-63,\dots ,63\right\}$, which is motivated by the fact that the WFSs of MAORY cannot reconstruct higher frequencies which might be present in the wavefronts anyway. For computing the dual frame functions ${\tilde {w}}_{jk,lg}$ depicted below, we implemented the iterative procedure (5.1) in Matlab (R2019a) on a desktop computer running on a Mac OS with an 8 core processor (Intel Xeon W-2140B CPU@3.20GHz) and 32GB RAM. The iteration itself was stopped after 500 iterations for each of the functions, which is more than sufficient as suggested by the error estimate (3.7) and the fact that already after much fewer iterations the magnitudes of the updates become almost negligible. The computation time for a single dual frame function was more than two hours, although this can be significantly reduced to a couple of seconds by more sophisticated implementations (currently under development). Note again that the computation of the dual frame functions has to be done only once for each AO setting and can be carried out in advance of the actual atmospheric tomography reconstruction.

The resulting frame and dual frame functions wjk,lg and ${\tilde {w}}_{jk,lg}$, for different values of j, k, l, g are depicted in figures 4 and 5, which show absolute value plots of the real part of those functions in linear scale and (for the dual frame functions) also in logarithmic scale. While the frame functions depicted in figure 4 are living on the same layer (l = 2) and differ only in the guide star direction (g = 2 and g = 5), the frame functions depicted in figure 5 vary also in the other parameters. In particular, they are living on atmospheric layers far apart from each other (l = 2 and l = 9). One can clearly (especially from the log-plots) see that even though by definition the frame functions wjk,lg have local support, namely exactly the domains ΩA (αg hl ), the resulting dual frame functions ${\tilde {w}}_{jk,lg}$ spread over the whole domain Ωl , which is due to the fact that all the frame functions wjk,lg , also for all other values of g, contribute to their definition. The influence of all the different guide star directions on each of the dual frame functions ${\tilde {w}}_{jk,lg}$ is also apparent in the appearance of ring-like structures corresponding to the the domains ΩA (αg hl ). Furthermore, the nicely visible periodicity of the frame functions wjk,lg due to their definition involving an exponential function manifests itself in a periodic pattern of the dual frame functions ${\tilde {w}}_{jk,lg}$.

Figure 4.

Figure 4. Frame functions wjk,lg (left) and dual frame functions ${\tilde {w}}_{jk,lg}$ (middle, right) for j = 3, k = 2, l = 8, with g = 2 (top) and g = 5 (bottom), respectively. Plotted is the absolute value of the real part in linear (left, middle) and logarithmic scale (right).

Standard image High-resolution image
Figure 5.

Figure 5. Frame functions wjk,lg (left) and dual frame functions ${\tilde {w}}_{jk,lg}$ (middle, right) for j = 3, k = 2, l = 2, g = 2 (top) and j = 10, k = 4, l = 9, g = 4 (bottom). Plotted is the absolute value of the real part in linear (left, middle) and logarithmic scale (right).

Standard image High-resolution image

6. Conclusion and outlook

In this paper, we considered the problem of atmospheric tomography by analysing the underlying tomography operator, and derived a frame decomposition for different problem settings. In particular, in contrast to [22], we considered settings with a mixture of both NGS and LGS, and did not place any restrictions on the shape of the aperture. The resulting decomposition yields information on the behaviour of the operator and provides a stable reconstruction algorithm for obtaining a minimum-coefficient least-squares solution of the atmospheric tomography problem. Furthermore, we presented numerical illustrations showcasing some of the building blocks of the frame decomposition. In a forthcoming publication currently under preparation, we plan to present numerical reconstruction results from both simulated and experimental AO data. For this, we are in particular developing efficient, accurate, and stable numerical methods for the computation of the dual frame functions.

Acknowledgments

S Hubmer and R Ramlau were (partly) funded by the Austrian Science Fund (FWF): F6805-N36. The authors would like to thank Dr. Stefan Kindermann for valuable discussions on some theoretical questions which arose during the writing of this manuscript.

Please wait… references are loading.
10.1088/1361-6420/aba4fe