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

Inversion and Symmetries of the Star Transform

G. Ambartsoumian and M. J. Latifi Jebelli
Department of Mathematics, University of Texas at Arlington
Department of Mathematics, University of Arizona
Abstract

The star transform is a generalized Radon transform mapping a function of two variables to its integrals along “star-shaped” trajectories, which consist of a finite number of rays emanating from a common vertex. Such operators appear in mathematical models of various imaging modalities based on scattering of elementary particles. The paper presents a comprehensive study of the inversion of the star transform. We describe the necessary and sufficient conditions for invertibility of the star transform, introduce a new inversion formula and discuss its stability properties. As an unexpected bonus of our approach, we prove a conjecture from algebraic geometry about the zero sets of elementary symmetric polynomials.

1 Introduction

The star transform is a generalized Radon transform mapping a function in 2\mathbb{R}^{2} to its weighted integrals along “star-shaped” trajectories, which consist of a finite number of rays emanating from a common vertex. Such operators appear in mathematical models of single-scattering optical tomography and single-scattering X-ray CT [9, 10, 11, 16, 26]. Particular attention has been drawn to the transforms with only two rays, usually called either broken ray transforms or V-line transforms [1, 3, 5, 6, 8, 11, 12, 15, 16, 20, 21, 25]. In addition to that, multiple authors have studied various generalizations of these transforms to higher dimensions (see the review article [2] and the references there).

All works cited above deal with cases, where the vertices of the stars or V-lines are inside the image domain. This paper also studies the star transform in that setup. A whole other world of applications and mathematical results exists in the case, when the vertices of the V-lines (or cones in higher dimensions) are restricted to the boundary of the image domain (e.g. see [13, 18, 19, 23, 24]). To find out more about the latter, we refer the interested reader to the review articles [2, 22] and the references there.

While there are multiple interesting formulas and procedures for inversion of the V-line transform, the star transform with three or more branches has been studied only in the pioneering paper [26]. The method presented in that article uses Fourier analysis techniques to reduce the inversion of the star transform to a problem of solving an infinite system of linear equations. The latter is solved using various truncations of the system and fast iterative computations of the Tikhonov-regularized pseudo-inverses of the resulting finite-size rectangular matrices. The paper also discusses the stability of the proposed inversion algorithms, formulating an algebraic condition that is necessary and sufficient for the stable reconstruction. Finally, several special geometric setups of the star transform are studied in relation to that condition.

This paper presents a comprehensive study of the inversion of the star transform using a totally different approach. We introduce a new exact closed-form inversion formula for the star transform, by establishing its connection with the (ordinary) Radon transform. It is much simpler than the inversion algorithm described in [26] and provides a very intuitive geometric insight to the problem. Our formula leads to a description of necessary and sufficient conditions for invertibilty of the star transform based on its geometric configuration, which was not known before. It also naturally yields the necessary and sufficient condition for the stability of inversion described in [26]. A detailed analysis of this condition allows us to classify the geometric setups of the transform according to their stability and provides a recipe for building star configurations with the “most stable” inversions. As an unexpected bonus of our analysis, we prove a long-standing conjecture from algebraic geometry (see [7]) about the zero sets of elementary symmetric polynomials.

The rest of the article is organized as follows. In Section 2 we define the main concepts used in the paper. In Section 3 we present a new inversion formula for the star transform. In Section 4 we study the injectivity of that transform. In Section 5 we discuss the so-called singular directions of the star transform that play an important role in the stability of numerical reconstructions. In Section 6 we state and prove the aforementioned conjecture from algebraic geometry. In Section 7 we describe mathematical models of some imaging modalities that use the star transform. In Section 8 we present various numerical simulations demonstrating the efficiency of our inversion method. We finish the paper with some additional remarks in Section 9 and acknowledgements in Section 10.

2 Definitions

Let f(x)Cc(2)f(x)\in C_{c}\left(\mathbb{R}^{2}\right) be a compactly supported continuous function, and let γ\gamma be a unit vector in the plane.

Definition 1

The divergent beam transform 𝒳γ\mathcal{X_{\gamma}} of ff at x2x\in\mathbb{R}^{2} is defined as:

𝒳γf(x)=0f(x+tγ)𝑑t.\mathcal{X}_{\gamma}f(x)=\int_{0}^{\infty}f(x+t\gamma)\,dt. (1)

One can use a linear combination of divergent beam transforms with a set of fixed distinct directions γi𝕊1\gamma_{i}\in\mathbb{S}^{1} and non-zero constants ci,i=1,,mc_{i}\in\mathbb{R},\,i=1,\ldots,m to define the star transform of ff. Namely,

Definition 2

The (weighted) star transform 𝒮\mathcal{S} of ff at x2x\in\mathbb{R}^{2} is defined as:

𝒮f(x)=i=1mci𝒳γif(x)=i=1mci0f(x+tγi)𝑑t.\mathcal{S}f(x)=\sum_{i=1}^{m}\;{c_{i}\;\mathcal{X}_{\gamma_{i}}f(x)}=\sum_{i=1}^{m}\;{c_{i}\int_{0}^{\infty}{f(x+t\gamma_{i})\,dt}}. (2)

In particular, if c1==cm=1c_{1}=\ldots=c_{m}=1, the star transform 𝒮\mathcal{S} puts into correspondence to a given function f(x)f(x) its integrals 𝒮f(x)\mathcal{S}f(x) along a two-parameter family of so-called “stars”. Each star consists of a union of rays emanating from a common vertex xx along fixed directions γi,i=1,,m\gamma_{i},\,i=1,\ldots,m (see Figure 1(a)).

Refer to caption
(a) A star centered at point xx
with rays emanating along fixed
unit vectors γ1,,γ4\gamma_{1},\ldots,\gamma_{4}.
Refer to caption
(b) If suppfD1\operatorname{supp}f\subseteq D_{1}, then supp𝒮fΩ\operatorname{supp}\mathcal{S}f\subseteq\Omega
and 𝒮f|D2Ω\mathcal{S}f|_{\partial D_{2}\cap\Omega} uniquely defines 𝒮f\mathcal{S}f in
2\D2\mathbb{R}^{2}\backslash D_{2}.
Figure 1: (The color version is available online.) A star and the support of the corresponding transform.
Definition 3

Let the ray directions of the star be parameterized as

γj=(aj,bj)2,j=1,,m.\gamma_{j}=(a_{j},b_{j})\in\mathbb{R}^{2},\;\;j=1,\ldots,m. (3)

We call

a=(a1,,am) and b=(b1,,bm)a=(a_{1},\ldots,a_{m})\mbox{ and }\;b=(b_{1},\ldots,b_{m}) (4)

the aperture vectors of that star. Since γj\gamma_{j}’s are unit vectors

aj2+bj2=1,j=1,,m.a_{j}^{2}+b_{j}^{2}=1,\;\;j=1,\ldots,m. (5)

Every star transform corresponds to an mm-gon (polygon with mm sides) with vertices c11γ1,,cm1γmc_{1}^{-1}\gamma_{1},\dots,c_{m}^{-1}\gamma_{m}. Many properties of the star transform can be characterised by geometric features of the corresponding polygon. The following special star plays a prominent role in some of our proofs.

Definition 4

We call a star transform symmetric, if m=2km=2k for some kk\in\mathbb{N} and (after possible re-indexing) γi=γk+i\gamma_{i}=-\gamma_{k+i} with ci=ck+ic_{i}=c_{k+i} for all i=1,,ki=1,\ldots,k. The corresponding star is also called symmetric.

Definition 5

We call a star transform regular, if c1==cm=1c_{1}=\ldots=c_{m}=1 and the ray directions γi\gamma_{i}, i=1,,mi=1,\ldots,m correspond to the radius vectors of the vertices of a regular mm-gon (see Figure 3(a)). The corresponding star is also called regular.

We will show that the symmetric star transforms are the only non-invertible configurations of 𝒮\mathcal{S}, while the regular star transforms with an odd number of vertices have the most stable inversions.

For regular stars with m=2k+1m=2k+1 rays, we parameterize the ray directions as follows (see Figure 3(a)):

γj=(aj,bj)=(cosαj,sinαj),\gamma_{j}=(a_{j},b_{j})=(\cos\alpha_{j},\sin\alpha_{j}), (6)

where

α1=0,α2j=2πj2k+1 and α2j+1=2πj2k+1,j=1,,k.\alpha_{1}=0,\;\;\alpha_{2j}=\frac{2\pi j}{2k+1}\mbox{ and }\alpha_{2j+1}=-\frac{2\pi j}{2k+1},\;j=1,\ldots,k. (7)

Since a2j=a2j+1a_{2j}=a_{2j+1} and b2j=b2j+1b_{2j}=-b_{2j+1} for j=1,,kj=1,\ldots,k, the aperture vectors of the regular star with 2k+12k+1 rays can be expressed as

a=(1,a2,a2,,a2k,a2k),b=(0,b2,b2,,b2k,b2k).a=(1,a_{2},a_{2},\ldots,a_{2k},a_{2k}),\;\;\;\;b=(0,b_{2},-b_{2},\ldots,b_{2k},-b_{2k}). (8)

Notice that if aperture vectors are fixed, the star is uniquely defined by its vertex xx. If one does not fix the directions of rays and allows arbitrary locations of its vertices, then the star transform will have m+2m+2 degrees of freedom. The problem of inverting such a transform will be overdetermined. We are interested in the inversion of the two-dimensional star transform 𝒮\mathcal{S} with mm rays in fixed directions γi\gamma_{i} and non-zero weights ci,i=1,,mc_{i},\,i=1,\ldots,m.

3 An Exact Inversion of the Star Transform

Let us assume that suppfD1\operatorname{supp}f\subseteq D_{1}, where D1D_{1} is an open disc of radius r1r_{1} centered at the origin. Then 𝒮f\mathcal{S}f is supported in an unbounded domain Ω\Omega (see Figure 1(b)). However, there exists a closed disc D2¯\overline{D_{2}} of some finite radius r2>r1r_{2}>r_{1} centered at the origin such that 𝒮f\mathcal{S}f is constant along the directions of rays γi\gamma_{i} inside the corresponding “strips” of 2\D2¯\mathbb{R}^{2}\backslash\overline{D_{2}}. In other words, the restriction of 𝒮f\mathcal{S}f to D2¯\overline{D_{2}} completely defines it everywhere else. Throughout the rest of the text we will assume that 𝒮f(x)\mathcal{S}f(x) is known for all xD2¯x\in\overline{D_{2}}.

In this section we formulate and prove an exact closed-form inversion formula for 𝒮\mathcal{S} with arbitrary geometry and arbitrary non-zero weights. The intuitive idea of this inversion is the following. The Radon transform \mathcal{R} of g=𝒮fg=\mathcal{S}f at (ψ,t)(\psi,{t}) is the integral of the star transform data gg along the line l(ψ,t)={x2|x,ψ=t}l(\psi,{t})=\{x\in\mathbb{R}^{2}|\,\langle x,\psi\rangle={t}\} with normal vector ψ\psi and distance t{t} from the origin. For a generic line, this gives a sum of weighted integrals of ff over half-planes in 2\mathbb{R}^{2}, with different weights on different sides of l(ψ,t)l(\psi,{t}). By differentiating that quantity with respect to variable t{t}, we recover the Radon data of ff along l(ψ,t)l(\psi,{t}). The inversion of 𝒮\mathcal{S} is then accomplished by inverting the Radon transform.

Before proving the theorem we discuss two auxiliary statements. They are geometric in nature and help with outlining the main ideas in the proof of Theorem 1.

3.1 Half Plane Transform and Its Derivative

For a unit vector ψ2\psi\in\mathbb{R}^{2} and t{t}\in\mathbb{R} define the half plane transform of ff as

Fψ(t)=x,ψtf𝑑x.F_{\psi}({t})=\int\limits_{\langle x,\psi\rangle\leq{t}}f\,dx. (9)

Geometrically, Fψ(t)F_{\psi}({t}) is the integral of ff over the half plane on one side of the line l(ψ,t)l(\psi,{t}) (see Fig. 2).

Refer to caption
Figure 2: A setup for the half-plane, divergent beam and Radon transforms.

Now, let’s look at the derivative of FF as a function of t{t}

dFψ(t)dt=limh0Fψ(t+h)Fψ(t)h.\frac{dF_{\psi}({t})}{d{t}}=\lim\limits_{h\rightarrow 0}\frac{F_{\psi}({t}+h)-F_{\psi}({t})}{h}.

Note that Fψ(t+h)Fψ(t)F_{\psi}({t}+h)-F_{\psi}({t}) is the integral of ff over the infinite strip between the lines l(ψ,t)l(\psi,{t}) and l(ψ,t+h)l(\psi,{t}+h), which plays a central role in our construction. Using Fubini’s theorem and the fundamental theorem of calculus, we get

dFψ(t)dt\displaystyle\frac{dF_{\psi}({t})}{d{t}} =ddtx,ψtf𝑑x=ddttf(ψ,s)𝑑s=f(ψ,t),\displaystyle=\frac{d}{d{t}}\int\limits_{\langle x,\psi\rangle\leq{t}}f\,dx=\frac{d}{d{t}}\int_{-\infty}^{t}\mathcal{R}f(\psi,{s})\,d{s}=\mathcal{R}f(\psi,{t}),

where \mathcal{R} is the standard Radon transform. Thus we have proved the following

Lemma 1

For function fCc(D1)f\in C_{c}{({D_{1}})} we have

dFψ(t)dt=f(ψ,t).\frac{dF_{\psi}({t})}{d{t}}=\mathcal{R}f(\psi,{t}). (10)

A key step in our proof is the possibility of expressing the half plane transform FψF_{\psi} of ff in terms of its star transform 𝒮f\mathcal{S}f. We show this by first finding a relation between FψF_{\psi} and 𝒳γ(f)\mathcal{X}_{\gamma}(f).

If ψ,γ0\langle\psi,\gamma\rangle\neq 0 we can integrate 𝒳γ(x)\mathcal{X}_{\gamma}(x) along the line l(ψ,t)l(\psi,{t}) to get an expression in terms of Fψ(t)F_{\psi}({t}). Namely

Lemma 2

Let fCc(D1)f\in C_{c}({D_{1}}) and ψ,γ0\langle\psi,\gamma\rangle\neq 0. Then

ddt(𝒳γf)(ψ,t)=1ψ,γdFψ(t)dt.\frac{d}{d{t}}\mathcal{R}(\mathcal{X}_{\gamma}f)(\psi,{t})=\frac{-1}{\langle\psi,\gamma\rangle}\frac{dF_{\psi}({t})}{d{t}}. (11)

Proof. Depending on whether ψ,γ\langle\psi,\gamma\rangle is positive or negative, (𝒳γf)(ψ,t)\mathcal{R}(\mathcal{X}_{\gamma}f)(\psi,{t}) is integrating ff with the weight ψ,γ1\langle\psi,\gamma\rangle^{-1} over the half plane on one or the other side of the line l(ψ,t)l(\psi,{t}) (see Figure 2). This is an implication of a modified version of Fubini’s theorem (e.g. see the Moving Sections Lemma in [3]). If ψ,γ>0\langle\psi,\gamma\rangle>0 we have

(𝒳γf)(ψ,t)=1ψ,γ[2f𝑑xFψ(t)]\mathcal{R}(\mathcal{X}_{\gamma}f)(\psi,{t})=\frac{1}{\langle\psi,\gamma\rangle}\left[\int_{\mathbb{R}^{2}}f\,dx-F_{\psi}({t})\right]

and if ψ,γ<0\langle\psi,\gamma\rangle<0

(𝒳γf)(ψ,t)=1ψ,γFψ(t).\mathcal{R}(\mathcal{X}_{\gamma}f)(\psi,{t})=\frac{-1}{\langle\psi,\gamma\rangle}F_{\psi}({t}).

In either case, taking the derivative of both sides yields the same expression and finishes the proof. \hfill\blacksquare

3.2 Main Result and Some Corollaries

Definition 6

Let 𝒮\mathcal{S} be the star transform defined in (2). We call

𝒵1=j=1m{ψ:ψ,γj=0}\mathcal{Z}_{1}=\bigcup\limits_{j=1}^{m}\{\psi:\langle\psi,\gamma_{j}\rangle=0\} (12)

the set of singular directions of Type 1 for 𝒮\mathcal{S}, and

𝒵2={ψ:j=1mcjijψ,γj=0}\mathcal{Z}_{2}=\left\{\psi:\sum_{j=1}^{m}c_{j}\prod_{i\neq j}\langle\psi,\gamma_{j}\rangle=0\right\} (13)

the set of singular directions of Type 2 for 𝒮\mathcal{S}. Now define

w(ψ)=i=1mciψ,γi,ψ𝕊1𝒵1,q(ψ)=1w(ψ),ψ𝕊1(𝒵1𝒵2).w(\psi)=\displaystyle\sum_{i=1}^{m}\frac{c_{i}}{\langle\psi,\gamma_{i}\rangle},\;\;\psi\in\mathbb{S}^{1}\setminus\mathcal{Z}_{1},\;\;\;\;\;q(\psi)=\frac{-1}{w(\psi)},\;\;\psi\in\mathbb{S}^{1}\setminus(\mathcal{Z}_{1}\cup\mathcal{Z}_{2}). (14)
Theorem 1

Let fCc(D1)f\in C_{c}(D_{1}), 𝒮\mathcal{S} be the star transform and \mathcal{R} be the Radon transform of a function in 2\mathbb{R}^{2}. Then for any ψ𝕊1(𝒵1𝒵2)\psi\in\mathbb{S}^{1}\setminus(\mathcal{Z}_{1}\cup\mathcal{Z}_{2}) we have:

f(ψ,t)=q(ψ)ddt(𝒮f)(ψ,t).\mathcal{R}f(\psi,{t})=q(\psi)\frac{d}{d{t}}\mathcal{R}(\mathcal{S}f)(\psi,{t}). (15)

Proof. Using Lemmas 1 and 2 we get

ddt(𝒳γf)(ψ,t)=1ψ,γf(ψ,t).\frac{d}{d{t}}\mathcal{R}(\mathcal{X}_{\gamma}f)(\psi,{t})=\frac{-1}{\langle\psi,\gamma\rangle}\mathcal{R}f(\psi,{t}).

Now we write 𝒮\mathcal{S} in terms of 𝒳γi\mathcal{X}_{\gamma_{i}}’s

ddt(𝒮f)(ψ,t)=\displaystyle\frac{d}{d{t}}\mathcal{R}(\mathcal{S}f)(\psi,{t})= ddt(i=1mci𝒳γif)(ψ,t)=i=1mciddt(𝒳γif)(ψ,t)\displaystyle\frac{d}{d{t}}\mathcal{R}\left(\sum_{i=1}^{m}{c_{i}\,\mathcal{X}_{\gamma_{i}}f}\right)(\psi,{t})=\sum_{i=1}^{m}{c_{i}\,\frac{d}{d{t}}\mathcal{R}(\mathcal{X}_{\gamma_{i}}f)(\psi,{t})}
=\displaystyle= i=1mciψ,γif(ψ,t)=w(ψ)f(ψ,t),\displaystyle\sum_{i=1}^{m}\frac{-c_{i}}{\langle\psi,\gamma_{i}\rangle}\;\mathcal{R}f(\psi,{t})\;=-w(\psi)\;\mathcal{R}f(\psi,{t}),

which ends the proof. \hfill\blacksquare

It is easy to notice that if 𝒵1𝒵2\mathcal{Z}_{1}\cup\mathcal{Z}_{2} is finite, then the singularities appearing in the right hand side of formula (15) are removable. Namely,

Remark 1

Let 𝒵1𝒵2={ζi}i=1M\mathcal{Z}_{1}\cup\mathcal{Z}_{2}=\left\{\zeta_{i}\right\}_{i=1}^{M}. Then by formula (15) and continuity of f\mathcal{R}f we have

limψζi[q(ψ)ddt(𝒮f)(ψ,t)]=f(ζi,t)<.\lim_{\psi\to\zeta_{i}}\left[q(\psi)\frac{d}{d{t}}\mathcal{R}(\mathcal{S}f)(\psi,{t})\right]=\mathcal{R}f(\zeta_{i},{t})<\infty. (16)

In other words, the full Radon data can be recovered from the star transform.

Corollary 1

If 𝒵1𝒵2\mathcal{Z}_{1}\cup\mathcal{Z}_{2} is finite, we can apply 1\mathcal{R}^{-1} to recover ff.

We finish this section with statements of two special cases of Theorem 1 and discussion of some pertinent relations. In the case of m=2m=2, Theorem 1 yields a new and simple inversion for the V-line transform.

Corollary 2

An inversion formula for the V-line transform 𝒱=c1𝒳γ1+c2𝒳γ2\mathcal{V}=c_{1}\mathcal{X}_{\gamma_{1}}+c_{2}\mathcal{X}_{\gamma_{2}} with fixed non-collinear ray directions γ1,γ2\gamma_{1},\,\gamma_{2} and nonzero weights c1,c2c_{1},\,c_{2} is given by

f=1[ψ,γ1ψ,γ2c2ψ,γ1+c1ψ,γ2ddt(𝒱f)(ψ,t)].f=\mathcal{R}^{-1}\left[\frac{-\langle\psi,\gamma_{1}\rangle\langle\psi,\gamma_{2}\rangle}{c_{2}\langle\psi,\gamma_{1}\rangle+c_{1}\langle\psi,\gamma_{2}\rangle}\frac{d}{d{t}}\mathcal{R}(\mathcal{V}f)(\psi,{t})\right]. (17)

A curious special case of Theorem 1 is the following.

Corollary 3

An inversion of the divergent beam transform 𝒳γ\mathcal{X_{\gamma}} is given by the formula

f=1[ψ,γddt(𝒳γf)(ψ,t)].f=\mathcal{R}^{-1}\left[-\langle\psi,\gamma\rangle\frac{d}{d{t}}\mathcal{R}(\mathcal{X_{\gamma}}f)(\psi,{t})\right]. (18)

The directional derivative DγD_{-\gamma}, given by Dγh(x)=γh(x)D_{\gamma}h\,(x)=\gamma\cdot\nabla h\,(x), is the natural inverse of 𝒳γ\mathcal{X_{\gamma}}. Hence, for any hUγ={𝒳γ(f):fCc(2)}h\in U_{\gamma}=\left\{\mathcal{X}_{\gamma}(f):\,f\in C_{c}(\mathbb{R}^{2})\right\} in the range of 𝒳γ\mathcal{X}_{\gamma}, one can write the following identity

Dγh(x)=1[ψ,γddt]h(x).D_{\gamma}h\,(x)=\mathcal{R}^{-1}\left[\langle\psi,\gamma\rangle\frac{d}{d{t}}\mathcal{R}\right]h\,(x). (19)

The identity (19) is not really new. It is equivalent to a known relation

P(De1,De2)=1[P(ψ1ddt,ψ2ddt)],P(D_{e_{1}},D_{e_{2}})=\mathcal{R}^{-1}\,\left[P\left(\psi_{1}\frac{d}{dt},\psi_{2}\frac{d}{dt}\right)\right]\,\mathcal{R}, (20)

where eie_{i}’s are standard orthonormal vectors and PP is a given polynomial of two variables. One can get (20) by repeatedly applying (19) and canceling 1\mathcal{R}^{-1}\mathcal{R} in the resulting telescoping identity. The other direction of equivalence is trivial.

4 Injectivity of the Star Transform

In certain configurations of the star transform, the function q(ψ)q(\psi) is not defined for any ψ\psi. For example, this happens when m=2m=2, c1=c2=1c_{1}=c_{2}=1, and γ1=γ2=(1,0)\gamma_{1}=-\gamma_{2}=(1,0). In this case ψ,γ1+ψ,γ20\langle\psi,\gamma_{1}\rangle+\langle\psi,\gamma_{2}\rangle\equiv 0 for any ψ𝕊1\psi\in\mathbb{S}^{1}.

Consider a similar configuration with more rays: m=4m=4, c1==c4=1c_{1}=\ldots=c_{4}=1, γ1=γ3\gamma_{1}=-\gamma_{3} and γ2=γ4\gamma_{2}=-\gamma_{4}. While it is not obvious that q(ψ)q(\psi) is not defined at any ψ\psi, it is easy to notice that in this case the star transform is not injective, since it provides less information than the ordinary Radon transform restricted to two directions. Interestingly enough, such configurations are the only ones, for which the star transform is not injective.

Theorem 2

The star transform 𝒮=i=1mci𝒳γi\mathcal{S}=\sum_{i=1}^{m}c_{i}\mathcal{X}_{\gamma_{i}} is invertible if and only if it is not symmetric.

An important object in our inversion is the function q(ψ)q(\psi) (or equivalently its reciprocal w(ψ)w(\psi)). It leads to the geometric condition that is necessary for the star transform to be non-invertible. To describe the connection between q(ψ)q(\psi) and the geometric condition we need to consider the elementary symmetric polynomial of degree m1m-1 in mm variables y=(y1,,ym)y=(y_{1},\ldots,y_{m}) (see [17])

em1(y1,,ym)=i=1mjiyj.e_{m-1}(y_{1},\dots,y_{m})=\sum_{i=1}^{m}\prod_{j\neq i}y_{j}. (21)

Using the above notation, one can re-write formula (14) as

q(ψ)=c11cm1em1(ψ,c11γ1,,ψ,cm1γm)ψ,γ1ψ,γm.q(\psi)=\frac{c_{1}^{-1}\ldots c_{m}^{-1}}{\displaystyle\frac{e_{m-1}\left(\langle\psi,c_{1}^{-1}\gamma_{1}\rangle,\dots,\langle\psi,c_{m}^{-1}\gamma_{m}\rangle\right)}{\langle\psi,\gamma_{1}\rangle\dots\langle\psi,\gamma_{m}\rangle}}. (22)

The proof of Theorem 2 is based on the description of zero sets of em1e_{m-1} and the fact that the star transform 𝒮\mathcal{S} is invertible if em1(ψ,c11γ1,,ψ,cm1γm)e_{m-1}(\langle\psi,c_{1}^{-1}\gamma_{1}\rangle,\dots,\langle\psi,c_{m}^{-1}\gamma_{m}\rangle) is not identically zero as a function of ψ\psi.

If ψ^=(r,s)2\hat{\psi}=(r,s)\in\mathbb{R}^{2} (not necessarily on 𝕊1\mathbb{S}^{1}) and cc1cmc\doteq c_{1}\ldots c_{m}, we define a polynomial in two variables (r,s)(r,s) as follows:

P2(ψ^)=cem1(ψ^,c11γ1,,ψ^,cm1γm).P_{2}(\hat{\psi})=c\,e_{m-1}(\langle\hat{\psi},c_{1}^{-1}\gamma_{1}\rangle,\dots,\langle\hat{\psi},c_{m}^{-1}\gamma_{m}\rangle). (23)
Lemma 3

The polynomial P2(r,s)P_{2}(r,s) either has finitely many zeros on 𝕊1\mathbb{S}^{1} or it is identically zero.

Proof. Let us start with noticing that em1(y)e_{m-1}(y) is a homogeneous polynomial of degree d1d-1, i.e. em1(λy)=λm1em1(y)e_{m-1}(\lambda y)=\lambda^{m-1}e_{m-1}(y). Hence, P2(r,s)P_{2}(r,s) is also a homogeneous polynomial of degree m1m-1, which either has finitely many (projective) roots or is identically zero. \hfill\blacksquare

Proof (of Theorem 2). Assume that for a fixed choice of γ1,,γm\gamma_{1},\dots,\gamma_{m} there is no inversion for the corresponding star transform. By Theorem 1, P2(ψ)P_{2}(\psi) has to be zero on an infinite subset of 𝕊1\mathbb{S}^{1}. Therefore, by Lemma 3, P2(ψ)0P_{2}(\psi)\equiv 0.

Using the notation introduced in (14)

w(ψ)=i=1mciψ,γi=P2(ψ)i=1mψ,γi.w(\psi)=\displaystyle\sum_{i=1}^{m}\frac{c_{i}}{\langle\psi,\gamma_{i}\rangle}=\frac{P_{2}(\psi)}{\prod\limits_{i=1}^{m}\langle\psi,\gamma_{i}\rangle}.

If w(ψ)=0w(\psi)=0 away from the poles, then

0=limψγjψ,γjw(ψ).0=\lim\limits_{\psi\to\gamma_{j}^{\perp}}\langle\psi,\gamma_{j}\rangle w(\psi).

This is cjc_{j} if there is no other γi\gamma_{i} parallel to γj\gamma_{j} and is equal to cjcjc_{j}-c_{j^{\prime}} if γj=γj\gamma_{j^{\prime}}=-\gamma_{j}. The first case can not hold since cj0c_{j}\neq 0 so the second must hold, which implies that 𝒮\mathcal{S} is symmetric. \hfill\blacksquare

Theorem 2 immediately implies the following.

Corollary 4

Any star transform with an odd number of rays is invertible.

5 Singular Directions of the Star Transform

The number and location of singular directions affect the quality of numerical reconstructions. The singular directions of Type 2 correspond to “division by zero” of the processed data ddt(𝒮f)\frac{d}{d{t}}\mathcal{R}(\mathcal{S}f), while those of Type 1 correspond to “multiplication by zero”. Hence, it is natural to expect that singular directions of Type 2 will create instability and adversely impact the reconstruction. Our numerical experiments confirm these expectations (see Section 8). It is also interesting, that the (totally different) algorithm for inversion of the star transform obtained in [26] produces a relation equivalent to the defining relation of 𝒵2\mathcal{Z}_{2} in (13) as a necessary and sufficient condition for the instability of that algorithm (see formula (51) on page 18 in [26]).

While the geometric meaning of singular directions of Type 1 is obvious for any mm, there is no easy interpretation of set 𝒵2\mathcal{Z}_{2} for m3m\geq 3. However, the singular directions of Type 2 are more crucial for the quality of reconstruction.

In Section 5.1 we discuss the existence of singular directions of Type 2 in star transforms with uniform weights. In Section 5.2 we show the absence of Type 2 singularities in regular star transforms with an odd number of rays. The case of star transforms with non-uniform weights is discussed in Section 5.3.

5.1 Star Transforms with Uniform Weights

Theorem 3

Consider the star transform 𝒮=i=1m𝒳γi\mathcal{S}=\sum_{i=1}^{m}\mathcal{X}_{\gamma_{i}} with uniform weights.

  1. 1.

    If mm is even, 𝒮\mathcal{S} must contain a singular direction of Type 2.

  2. 2.

    When mm is odd, there exist configurations of 𝒮\mathcal{S} that contain singular directions of Type 2, as well as configurations that do not contain them.

Refer to caption
(a) A star with 2k+12k+1 symmetric rays and without a singular direction of Type 2.
Refer to caption
(b) A star with 2k+12k+1 rays that has a singular direction of Type 2.
Figure 3: Stars of 2k+12k+1 rays with and without singular directions of Type 2.

Proof (of Part 1). Let m=2km=2k. Then polynomial P2P_{2} defined in (23) is an odd function on the unit circle, i.e.

P2(ψ)=P2(ψ),ψ𝕊1.P_{2}(-\psi)=-P_{2}(\psi),\;\;\psi\in\mathbb{S}^{1}.

If P20P_{2}\not\equiv 0, then there exists ψ𝕊1\psi^{*}\in\mathbb{S}^{1} such that P2(ψ)=P2(ψ)>0P_{2}(-\psi^{*})=-P_{2}(\psi^{*})>0. Hence, by continuity of P2P_{2} it has to be zero on each arc of 𝕊1\mathbb{S}^{1} between ψ\psi^{*} and ψ-\psi^{*}. \hfill\blacksquare

Proof (of Part 2a).

Now let us show that for an odd number of rays, there exist star transforms with singularities of Type 2. In other words, for any m=2k+1m=2k+1, kk\in\mathbb{N} there exist configurations of ray directions γi\gamma_{i}, i=1,,mi=1,\ldots,m, such that P2(ψ)=0P_{2}(\psi)=0 for some ψ𝕊1\psi\in\mathbb{S}^{1}, where P2P_{2} is the polynomial defined in (23).

A trivial example is when a pair of rays point to opposite directions, i.e. γi=γj\gamma_{i}=-\gamma_{j} for some indices ii and jj. Then it is easy to see that the direction normal to them is a singular direction of Type 2. A more interesting example is discussed below.

Consider an arbitrary open half-plane H={x:x,n<0}H=\left\{x:\langle x,n\rangle<0\right\}, where nn is the outward normal to its boundary passing through the origin. Choose a set of distinct ray directions {γi}i=1m\left\{\gamma_{i}\right\}_{i=1}^{m} so that γiH\gamma_{i}\in H for all i=1,,mi=1,...,m. Without loss of generality, let us assume that γi\gamma_{i}’s are indexed according to the growth of their polar angle (see Figure 3(b)).

Consider ψ=γ2\psi=\gamma_{2}^{\perp} oriented so that ψ,γ1>0\langle\psi,\gamma_{1}\rangle>0. Then ψ,γ2=0\langle\psi,\gamma_{2}\rangle=0 and ψ,γi<0\langle\psi,\gamma_{i}\rangle<0 for all i>2i>2. Estimating the terms of the sum in (21), it is easy to notice that j2ψ,γj<0\prod_{j\neq 2}\langle\psi,\gamma_{j}\rangle<0, while all of the other terms are 0. Thus, for ψ=γ2\psi=\gamma_{2}^{\perp} we get P2(ψ)<0P_{2}(\psi)<0.

On the other hand, if ψ=n\psi=n, then all terms of the sum in (21) are positive and P2(ψ)>0P_{2}(\psi)>0. By continuity of P2(ψ)P_{2}(\psi), there has to exist a value of ψ\psi, for which P2(ψ)=0P_{2}(\psi)=0. \hfill\blacksquare

Proof (of Part 2b). The regular star transform 𝒮\mathcal{S} with m=2k+1m=2k+1 rays does not have a singular direction of Type 2. See Section 5.2. \hfill\blacksquare

5.2 Regular Star Transforms with an Odd Number of Rays

In this section we show that any regular star transform with an odd number of rays (see Figure 3(a)) has a stable inversion. More specifically,

Theorem 4

The regular star transform 𝒮\mathcal{S} with m=2k+1m=2k+1 rays does not have a singular direction of Type 2.

Proof. We need to show that for the regular star with m=2k+1m=2k+1 rays, P2(ψ)0P_{2}(\psi)\neq 0 for any ψ𝕊1\psi\in\mathbb{S}^{1}, where P2P_{2} is the polynomial defined in (23).

Using the parameterization ψ(α)=(cosα,sinα)\psi(\alpha)=(\cos\alpha,\sin\alpha), we denote

F(α)=P2(ψ(α)).F(\alpha)=P_{2}(\psi(\alpha)). (24)

We will show that F(α)0F(\alpha)\neq 0 for any α[0,2π]\alpha\in[0,2\pi]. Our argument is based on the following statement, proved after the proof of Theorem 4.

Lemma 4

There exists a polynomial of one variable P1P_{1} of degree m1\leq m-1, such that

F(α)=P1(cosα),F(\alpha)=P_{1}(\cos\alpha), (25)

Due to the symmetry of P2(ψ)P_{2}(\psi) with respect to γi\gamma_{i}’s and the symmetric distributions of γi\gamma_{i}’s on 𝕊1\mathbb{S}^{1}, it follows that F(α)F(\alpha) is periodic with period 2π/m2\pi/m. Therefore all of its Fourier coefficients FkF_{k} are zero unless kk is divisible by mm. Since FF is a trigonometric polynomial, it is equal to its Fourier expansion, and since it is of degree at most (m1)(m-1), the only index in its Fourier expansion divisible by mm is k=0k=0, thus FF is constant. By Theorem 2 that constant can not be zero, which completes the proof of Theorem 4. \hfill\blacksquare

Proof of Lemma 4: Recall Definition 3 and Equation (8). Using our previous notations ψ(α)=(cosα,sinα)(r,s)\psi(\alpha)=(\cos\alpha,\sin\alpha)\doteq(r,s) we can write

F(α)\displaystyle F(\alpha) =P2(ψ(α))=em1(ψ,γ1,,ψ,γm)=em1(ra+sb)\displaystyle=P_{2}(\psi(\alpha))=e_{m-1}(\langle\psi,\gamma_{1}\rangle,\dots,\langle\psi,\gamma_{m}\rangle)=e_{m-1}(ra+sb)
=i=1mji(raj+sbj)=j1(raj+sbj)+[j2(raj+sbj)+j3(raj+sbj)]\displaystyle=\sum_{i=1}^{m}\prod_{j\neq i}(ra_{j}+sb_{j})=\prod_{j\neq 1}(ra_{j}+sb_{j})+\left[\prod_{j\neq 2}(ra_{j}+sb_{j})+\prod_{j\neq 3}(ra_{j}+sb_{j})\right]
++[j2k(raj+sbj)+j2k+1(raj+sbj)]\displaystyle+\ldots+\left[\prod_{j\neq 2k}(ra_{j}+sb_{j})+\prod_{j\neq 2k+1}(ra_{j}+sb_{j})\right]

It easy to notice, that after appropriate cancellations inside the brackets using (8), the last expression does not contain any odd degree of ss. Substituting s2=1r2s^{2}=1-r^{2} we obtain the desired result. \hfill\blacksquare

In the case of m=3m=3, there is a simpler proof of Theorem 4. That proof also shows that the transform 𝒮\mathcal{S} corresponding to the regular star has (in some sense) the most stable inversion. It is based on the following result from [7]:

Theorem The zero set of e2(y1,y2,y3)e_{2}(y_{1},y_{2},y_{3}) is the circular cone

e21(0)={y3:cos2(y,u)=13},e_{2}^{-1}(0)=\left\{y\in\mathbb{R}^{3}:\,\cos^{2}{(y,u)}=\frac{1}{3}\right\}, (26)

where u=(1,1,1)u=(1,1,1) and (y,u)(y,u) denotes the angle between vectors yy and uu.

An Alternative Proof of Theorem 4 when m=3m=3.

We need to show that e2(ra+sb)0e_{2}(ra+sb)\neq 0, where aa and bb are the aperture vectors of the star transform defined by (6), (7), (8) as

a=(1,12,12),b=(0,32,32),a=\left(1,-\frac{1}{2},-\frac{1}{2}\right),\;\;b=\left(0,\frac{\sqrt{3}}{2},-\frac{\sqrt{3}}{2}\right), (27)

and r2+s2=1r^{2}+s^{2}=1. In light of (26), it is equivalent to proving that the plane TT spanned by aperture vectors aa and bb intersects the circular cone defined in (26) only at the origin. Since

a,u=0 and b,u=0,\langle a,u\rangle=0\mbox{ and }\langle b,u\rangle=0,

it is clear that uTu\perp T. Hence, Te21(0)={0}T\bigcap e^{-1}_{2}(0)=\{0\}. Moreover, the aperture vectors aa and bb corresponding to the regular star, span the plane that is “as far as possible” from the zero set of e2e_{2}. \hfill\blacksquare

5.3 Star Transforms with Non-Uniform Weights

The regular stars with an odd number of rays are not the only configurations of the transform 𝒮\mathcal{S} that do not have a singular direction of Type 2. Due to continuous dependence of P2(ψ)P_{2}(\psi) on αi\alpha_{i}’s, small perturbations of γi\gamma_{i}’s from symmetry positions will not introduce a singular direction of Type 2. Similarly, small perturbations of uniform weights c1==cm=1c_{1}=\ldots=c_{m}=1 to non-uniform values will not introduce singular directions of Type 2. Moreover,

Remark 2

The polynomial P2(ψ)P_{2}(\psi) does not change if one simultaneously replaces γi\gamma_{i} by γi-\gamma_{i} and cic_{i} by ci-c_{i} for any ii. This provides a recipe for creating star transforms with weights of mixed algebraic signs that have stable inversions.

For example, modifying the regular star transform of 3 rays (defined by formula (27)) through replacement of γ2\gamma_{2} by γ2-\gamma_{2} and changing the sign of c2=1c_{2}=1, we get a stable configuration with weights c1=1c_{1}=1, c2=1c_{2}=-1, c3=1c_{3}=1 and aperture vectors:

a=(1,12,12),b=(0,32,32).a=\left(1,\frac{1}{2},-\frac{1}{2}\right),\;\;b=\left(0,-\frac{\sqrt{3}}{2},-\frac{\sqrt{3}}{2}\right).

In general, one can get stable setups when the absolute values of weights are close to each other.

The star transforms with weights of mixed algebraic signs satisfying the condition

i=1mci=0\sum_{i=1}^{m}c_{i}=0 (28)

play an important role in simultaneous reconstruction of scattering and absorption coefficients in single scattering tomography [9, 10, 15, 26]. Simultaneously satisfying the above condition and keeping the absolute values of weights as close to each other as possible, suggest choosing the weights (up to a proportionality coefficient and re-indexing) as follows:

c1==ck=1k,ck+1==c2k+1=1k+1,m=2k+1.c_{1}=\ldots=c_{k}=-\frac{1}{k},\;\;c_{k+1}=\ldots=c_{2k+1}=\frac{1}{k+1},\;\;m=2k+1. (29)

In the case of m=3m=3, this implies choosing the weights proportional to c1=c2=1c_{1}=c_{2}=1 and c3=2c_{3}=-2. Such a setup was analyzed in detail in [26].

We finish this section with generalizations of the statements of Section 5.1 to the case of transforms with non-uniform weights. For the stars with an even number of rays, both the result and its proof are almost identical. Namely,

Theorem 5

Consider the star transform 𝒮=i=1mci𝒳γi\mathcal{S}=\sum_{i=1}^{m}c_{i}\mathcal{X}_{\gamma_{i}}. If mm is even, 𝒮\mathcal{S} must contain a singular direction of Type 2.

In the case of star transforms with an odd number of rays, it should be clarified in what sense we want to generalize Part 2 of Theorem 3. If one is allowed to choose both the weights cic_{i} and the ray directions γi\gamma_{i} to claim existence of setups with and without Type 2 singularities, then the statement has a trivial proof by choosing uniform weights and referring to Theorem 3. Hence, we consider the following question. Are there configurations (i.e. ray directions γ1,,γm\gamma_{1},\ldots,\gamma_{m}) with/without Type 2 singularities for a specified set of weights c1,,cmc_{1},\ldots,c_{m}?

Theorem 6

Consider a star transform 𝒮=i=1mci𝒳γi\mathcal{S}=\sum_{i=1}^{m}c_{i}\mathcal{X}_{\gamma_{i}}, where m=2k+1m=2k+1.

(a) For any set of specified weights c1,cmc_{1}\ldots,c_{m}, there exist γ1,,γm\gamma_{1},\ldots,\gamma_{m} such that 𝒮\mathcal{S} contains singular directions of Type 2. In other words, for any set of weights there are configurations of 𝒮\mathcal{S} with unstable inversion.

(b) Let m=3m=3. For any set of specified weights c1,c2,c3c_{1},c_{2},c_{3}, there exist γ1,γ2,γ3\gamma_{1},\gamma_{2},\gamma_{3} such that 𝒮\mathcal{S} does not contain singular directions of Type 2. In other words, for any set of weights there are configurations of 𝒮\mathcal{S} with a stable inversion.

Proof of (a): Existence of configurations with Type 2 singularities.
Both examples provided in the proof of Theorem 3 work here with small modifications. If γi=γj\gamma_{i}=-\gamma_{j} for some indices ii and jj, then the direction normal to them is a singular direction of Type 2.

A more interesting example is the case when all γi\gamma_{i}’s belong to the same half-plane. Recall the corresponding setup from the proof of Theorem 3. Without loss of generality, let us assume that the c1c2>0c_{1}c_{2}>0, i.e. the weights c1c_{1} and c2c_{2} have the same sign. Since we consider the case of m=2k+1m=2k+1 for k1k\geq 1, the assumption above is just a matter of re-indexing the weights (if necessary).

Let ψ1=γ1\psi_{1}=\gamma_{1}^{\perp} and ψ2=γ2\psi_{2}=\gamma_{2}^{\perp} be vectors obtained by a clockwise rotation by π/2\pi/2 from γ1\gamma_{1} and γ2\gamma_{2} correspondingly. Then, it is easy to verify that P2(ψ1)P_{2}(\psi_{1}) and P2(ψ2)P_{2}(\psi_{2}) have different signs. Hence, by continuity of P2P_{2} there exists some ψ\psi s.t. P2(ψ)=0P_{2}(\psi)=0. \hfill\blacksquare

Before proceeding to the proof of the second statement of Theorem 6, we define a new class of planes in 3\mathbb{R}^{3} and discuss their properties.

Definition 7

We call a plane T3T\subseteq\mathbb{R}^{3} “spannable by star aperture vectors” or SSAV if T=span{a,b}T=\operatorname{span}\{a,b\}, where a,b3a,b\in\mathbb{R}^{3} are aperture vectors of some star with m=3m=3 rays. In particular this means that (ai,bi)(a_{i},b_{i}) are unit vectors in 2\mathbb{R}^{2} for i=1,2,3i=1,2,3.

Not every plane (containing the origin) in 3\mathbb{R}^{3} is SSAV. The following lemma gives a complete description of admissible normal vectors to an SSAV plane.

Lemma 5

A vector n=(n1,n2,n3)3n=(n_{1},n_{2},n_{3})\in\mathbb{R}^{3} is normal to an SSAV plane if and only if |n1|\left|n_{1}\right|, |n2|\left|n_{2}\right| and |n3|\left|n_{3}\right| correspond to side lengths of some triangle. We will denote the set of all such vectors by AA.

Proof. Let nn be normal to a plane T3T\in\mathbb{R}^{3} containing the origin. TT is SSAV if and only if there exists a star with aperture vectors a,ba,b such that n,a=n,b=0\langle n,a\rangle=\langle n,b\rangle=0. By definition of the aperture vectors, the last relation implies

n1γ1+n2γ2+n3γ3=0,n_{1}\gamma_{1}+n_{2}\gamma_{2}+n_{3}\gamma_{3}=0,

where γ1\gamma_{1}, γ2\gamma_{2} and γ3\gamma_{3} are the ray directions of the corresponding star. Since γi\gamma_{i}’s are unit vectors, the statement of the lemma follows from the above equality interpreting it as a sum of vectors tracing the perimeter of a triangle. \hfill\blacksquare

The set AA of admissible normals to an SSAV plane has an interesting geometric description due to the three triangle inequalities

|n1|+|n2||n3|,|n2|+|n3||n1|,|n1|+|n3||n2|.\left|n_{1}\right|+\left|n_{2}\right|\geq\left|n_{3}\right|,\quad\left|n_{2}\right|+\left|n_{3}\right|\geq\left|n_{1}\right|,\quad\left|n_{1}\right|+\left|n_{3}\right|\geq\left|n_{2}\right|.

In the octant n1>0n_{1}>0, n2>0n_{2}>0, n3>0n_{3}>0 it coincides with an infinite polyhedral cone with a vertex at the origin and infinite edges along vectors (0,1,1)(0,1,1), (1,0,1)(1,0,1) and (1,1,0)(1,1,0). We denote the portion of AA in this first octant by A+A^{+}. In all other octants AA coincides with a similar tetrahedron with the edges along appropriately signed versions of the same vectors.

Formula (26) shows that every plane with a normal vector nA+n\in A^{+} intersects e21(0)e_{2}^{-1}(0) only at the origin. Moreover, among the SSAV planes, those with a normal vector in A+A^{+} are the only ones that intersect e21(0)e_{2}^{-1}(0) only at the origin. Hence, the set A+A^{+} contains normals to SSAV planes that correspond (through the spanning aperture vectors) to star transforms with uniform weights c1=c2=c3=1c_{1}=c_{2}=c_{3}=1 and a stable inversion, i.e. without a singular direction of Type 2.

Let

W=diag(c1,c2,c3)W=\operatorname{diag}(c_{1},c_{2},c_{3})

denote the diagonal matrix of weights.

Lemma 6

For any set of specified weights c1,c2,c3>0c_{1},c_{2},c_{3}>0 there exist SSAV planes T1,T2T_{1},T_{2} such that T2=W1T1T_{2}=W^{-1}T_{1} and Tie21(0)={0}T_{i}\cap e_{2}^{-1}(0)=\{0\} for i=1,2i=1,2.

Proof. Let n(1),n(2)n^{(1)},n^{(2)} denote some vectors normal to the planes T1T_{1} and T2T_{2} correspondingly. The proof will follow immediately if we show that for any c1,c2,c3>0c_{1},c_{2},c_{3}>0 there exist n(1),n(2)A+n^{(1)},n^{(2)}\in A^{+} such that n(2)=Wn(1)n^{(2)}=Wn^{(1)}.

We first note that the problem is invariant with respect to multiplication of all weights c1,c2,c3>0c_{1},c_{2},c_{3}>0 by a positive constant, as well as their permutation. Hence, without loss of generality we will assume that 0<c1c2=1c30<c_{1}\leq c_{2}=1\leq c_{3}.

A choice of desired pair of vectors n(1),n(2)n^{(1)},n^{(2)} is given by

n(1)=(1,1,1/c3)A+,n(2)=Wn(1)=(c1,1,1)A+,n^{(1)}=\left(1,1,1/c_{3}\right)\in A^{+},\quad n^{(2)}=Wn^{(1)}=\left(c_{1},1,1\right)\in A^{+},

which ends the proof. \hfill\blacksquare

Proof of Theorem 6 (b): Configurations without Type 2 singularities.
Assume m=3m=3 and let c1,c2,c30c_{1},c_{2},c_{3}\neq 0 be specified weights. By Remark 2, each stable configuration of 𝒮\mathcal{S} with signed weights c1,c2,c3c_{1},c_{2},c_{3} corresponds to another stable configuration of 𝒮\mathcal{S} with positive weights |c1||c_{1}|, |c2||c_{2}|, |c3||c_{3}| and the other way around. Hence, without loss of generality we will prove our statement for c1,c2,c3>0c_{1},c_{2},c_{3}>0.

We will use our knowledge of existence of stable configurations for every star transform with uniform weights (marked by superscript (2) in the proof below) to show existence of stable configurations for star transforms with positive weights (marked by superscript (1) in the proof below).

By Lemma 6 there exist SSAV planes T1,T2T_{1},T_{2} such that T2=W1T1T_{2}=W^{-1}T_{1} and Tie21(0)={0}T_{i}\cap e_{2}^{-1}(0)=\{0\} for i=1,2i=1,2. Let a(1),b(1)a^{(1)},b^{(1)} and a(2),b(2)a^{(2)},b^{(2)} be some star aperture vector pairs spanning correspondingly T1T_{1} and T2T_{2}.

To prove the theorem, it is enough to show that for ray directions γ1(1),γ2(1),γ3(1)\gamma_{1}^{(1)},\gamma_{2}^{(1)},\gamma_{3}^{(1)} corresponding to the aperture vectors a(1),b(1)a^{(1)},b^{(1)} the following relation holds

P2(1)(ψ)c1ψ,γ2(1)ψ,γ3(1)+c2ψ,γ1(1)ψ,γ3(1)+c3ψ,γ1(1)ψ,γ2(1)0P_{2}^{(1)}(\psi)\doteq c_{1}\langle\psi,\gamma_{2}^{(1)}\rangle\langle\psi,\gamma_{3}^{(1)}\rangle+c_{2}\langle\psi,\gamma_{1}^{(1)}\rangle\langle\psi,\gamma_{3}^{(1)}\rangle+c_{3}\langle\psi,\gamma_{1}^{(1)}\rangle\langle\psi,\gamma_{2}^{(1)}\rangle\neq 0

for all ψ𝕊1\psi\in\mathbb{S}^{1}.

Due to the properties of selected planes and corresponding aperture vectors

P2(2)(ψ)e2(ra(2)+sb(2))0 for all ψ=(r,s)𝕊1.P_{2}^{(2)}(\psi)\doteq e_{2}\left(ra^{(2)}+sb^{(2)}\right)\neq 0\mbox{ for all }\psi=(r,s)\in\mathbb{S}^{1}.

At the same time for any ψ=(r,s)𝕊1\psi=(r,s)\in\mathbb{S}^{1} there exists ψ^=(r^,s^)𝕊1\hat{\psi}=(\hat{r},\hat{s})\in\mathbb{S}^{1} such that

P2(1)(ψ)=ce2(W1(ra(1)+sb(1)))=ke2(r^a(2)+s^b(2))=kP2(2)(ψ^)0,P_{2}^{(1)}(\psi)=c\;e_{2}\left(W^{-1}(ra^{(1)}+sb^{(1)})\right)=k\;e_{2}\left(\hat{r}a^{(2)}+\hat{s}b^{(2)}\right)=k\;P^{(2)}_{2}(\hat{\psi})\neq 0,

where c=c1c2c3c=c_{1}c_{2}c_{3} and kk is some positive constant. In the second equality above we use the fact that W1(ra(1)+sb(1))T2W^{-1}(ra^{(1)}+sb^{(1)})\in T_{2} and homogeneity of e2e_{2}. \hfill\blacksquare

The generalization of Theorem 6 (b) to the case of m=2k+1>3m=2k+1>3 is technically more complicated and we plan to address it in a future publication. So we finish this sections with the following.

Conjecture 1

Let m=2k+1m=2k+1. For any set of specified weights c1,,cmc_{1},\ldots,c_{m}, there exist γ1,,γm\gamma_{1},\ldots,\gamma_{m} such that 𝒮\mathcal{S} does not contain singular directions of Type 2. In other words, for any set of weights there are configurations of 𝒮\mathcal{S} with a stable inversion.

6 Zeros of em1(y1,,ym)e_{m-1}(y_{1},\ldots,y_{m}) for Odd mm.

In paper [7] published in 2006, the author formulated the following conjecture:

Conjecture 2

If rr is even then er1(0)e^{-1}_{r}(0) contains no real vector subspace of dimension rr.

Furthermore, it is stated there that one of the extreme cases is “the case em1(y1,,ym),m1(mod2)e_{m-1}(y_{1},...,y_{m}),\,m\equiv 1\;(\operatorname{mod}2), which becomes a task quite hard to tackle”.

Our proof of Theorem 4 includes a proof of the aforementioned extreme case. Namely,

Theorem 7

Let m=2k+1m=2k+1 for some kk\in\mathbb{N}. Then em11(0)e_{m-1}^{-1}(0) contains no real vector subspace of dimension m1m-1.

Indeed, let a,bma,b\in\mathbb{R}^{m} be the (linearly independent) aperture vectors of the regular star with m=2k+1m=2k+1 rays. It was shown that em1(ra+sb)=P2(ψ)0e_{m-1}(ra+sb)=P_{2}(\psi)\neq 0 for any ψ=(r,s)𝕊1\psi=(r,s)\in\mathbb{S}^{1}. The set C{ra+sb:(r,s)𝕊1}C\doteq\left\{ra+sb:\,(r,s)\in\mathbb{S}^{1}\right\} represents a closed contour around the origin in the 2-dimensional subspace TmT\subset\mathbb{R}^{m} spanned by the aperture vectors aa and bb. Due to homogeneity of em1(y)e_{m-1}(y), the fact that it has no zeros on CC, implies that em11(0)T={0}e^{-1}_{m-1}(0)\bigcap T=\{0\}. Since dim(T)=2\dim(T)=2, the zero set em11(0)e^{-1}_{m-1}(0) can not contain a real vector subspace of dimension m1m-1. \hfill\blacksquare

7 The Star Transform in Imaging

A known application of the star transform is related to imaging of a heterogeneous medium using single-scattered radiation [26]. To describe a simple model of such procedure, consider the following setup (see Figure 4).

Refer to caption
Figure 4: A simple setup of single-scattering tomography.

Let pp and qq be some points outside the imaged body. A source of radiation located at point pp emits a beam of particles into the body along the direction γi-\gamma_{i}. A receiver placed at point qq collimated along the direction γj\gamma_{j} measures the intensity of scattered radiation. Ignoring particles that scatter more than once and have very small energy at qq, each collimated emitter-receiver pair (p,q)(p,q) uniquely identifies the scattering location xx inside the body. That point xx corresponds to the vertex of the V-line with rays along γi,γj\gamma_{i},\gamma_{j} passing respectively through p,qp,q. The measured data ϕij(x)\phi_{ij}(x) can then be expressed as

ϕij(x)=𝒳if(x)+kij𝒳jf(x)+η(x),\phi_{ij}(x)=\mathcal{X}_{i}f(x)+k_{ij}\mathcal{X}_{j}f(x)+\eta{(x)}, (30)

where f(x)f(x) and η(x)\eta(x) represent correspondingly the attenuation and scattering coefficients at xx, while kijk_{ij} is a known constant. Relation (30) can be derived using single-scattering approximation to the radiative transport equation (see [9, 16, 26]).

In simple models of single scattering tomography it is customary to take kij=1k_{ij}=1, assuming that the medium has a unique (space dependent) attenuation coefficient. However, it is well known that the attenuation is energy dependent. Given the drop of energy of the particles after scattering, it is natural to consider different ff’s before and after scattering. For example, the primary type of scattering in X-rays is Compton scattering, where the energy loss depends on the scattering angle. If one collects data corresponding to two fixed directions γi,γj\gamma_{i},\gamma_{j}, then (under certain simplifying assumptions) it is reasonable to consider the attenuation coefficients at different energy levels (before and after the single scattering event) as ff and kijfk_{ij}f (e.g. see [16]).

Let us now return to formula (30). By varying the location of points p,qp,q and using finitely many fixed directions for γ\gamma’s (see Figure 4), one can generate a two-dimensional family of measured data corresponding to all possible locations of xx inside the body. If η(x)\eta(x) is known, then the recovery of f(x)f(x) boils down to inversion of a weighted V-line transform. In this case just two fixed directions γ1,γ2\gamma_{1},\gamma_{2} will suffice.

However, even in the simple model using kij=1k_{ij}=1, if η(x)\eta(x) is not known, the recovery of both attenuation and scattering coefficients from the same data set requires the use of the star transform with at least three distinct ray directions. The key idea here is to consider various linear combinations of ϕij(x)\phi_{ij}(x), which will eliminate η\eta and produce a weighted star transform of ff. For example, in a setup using three fixed directions γ1,γ2,γ3\gamma_{1},\gamma_{2},\gamma_{3} one can generate star transform data as follows:

ϕ12(x)ϕ23(x)=𝒳1f(x)+(k121)𝒳2f(x)+k23𝒳3f(x).\phi_{12}(x)-\phi_{23}(x)=\mathcal{X}_{1}f(x)+(k_{12}-1)\mathcal{X}_{2}f(x)+k_{23}\mathcal{X}_{3}f(x).

An inversion of the weighted star transform produces f(x)f(x), which then can be used to recover η(x)\eta(x) from (30).

The choice of a linear combination of ϕij(x)\phi_{ij}(x) eliminating η(x)\eta(x) is not unique, even when kij=1k_{ij}=1 for all i,ji,j. For example, let us assume that the existing hardware allows measurements along mm directions γ1,,γm\gamma_{1},\ldots,\gamma_{m}. To get an appropriate linear combination of data at each vertex xx and cancel the contribution from the η\eta term, one can choose a symmetric matrix ω\omega with zeros on the main diagonal and satisfying i,jωij=0\sum_{i,j}\omega_{ij}=0. Writing this linear combination, we get

i=1mj=1mωijϕij(x)=i=1m(j=1mωij)𝒳if(x)=i=1mci𝒳if(x),\sum_{i=1}^{m}\sum_{j=1}^{m}\omega_{ij}\phi_{ij}(x)=\sum_{i=1}^{m}\left(\sum_{j=1}^{m}\omega_{ij}\right)\mathcal{X}_{i}f(x)=\sum_{i=1}^{m}c_{i}\mathcal{X}_{i}f(x),

where

ci=j=1mωij.c_{i}=\sum_{j=1}^{m}\omega_{ij}.

After applying the inversion of the star transform and finding ff, we can recover the scattering term using (30).

It is easy to notice that when m>3m>3, for any predefined set of weights ci0c_{i}\neq 0, i=1,,mi=1,\ldots,m there are infinitely many choices of ω\omega satisfying the described requirements. In the case of m=3m=3, not all choices of c1,c2,c3c_{1},c_{2},c_{3} work, but there are still infinitely many admissible options.

We refer the interested reader to [16, 26] for more details on physical aspects and imaging settings of single-scattering optical and X-ray tomography.

Remark 3

Another potential imaging application of V-line transforms is emission tomography, where the goal is recovery of the distribution f(x)f(x) of a radioactive material inside an object. Here the measured data corresponding to scattered particles can be represented by the weighted V-line transforms of f(x)f(x), where the non-constant weight will depend on a spatially varying attenuation μ(x)\mu(x). To the best of our knowledge, no inversion formula is known for such V-line transforms with non-constant weights.

8 Numerical Simulations

In this section we present some examples of numerical reconstructions of the standard Shepp-Logan phantom (see Figure 5a) from its star transforms of various configurations. The reconstruction algorithm is based on Theorem 1 and the filtered backprojection algorithm of the Radon transform. The resolution of all images is 400x400 pixels. The weights ci=1c_{i}=1 for all ii.

In all simulations below we have generated the star transform 𝒮f\mathcal{S}f of ff depicted in Figure 5a by numerically computing the divergent beam transforms 𝒳γjf\mathcal{X}_{\gamma_{j}}f along given directions γj\gamma_{j} and adding them up. Since all 𝒳γjf\mathcal{X}_{\gamma_{j}}f’s and 𝒮f\mathcal{S}f have unbounded support (even for compactly supported ff), one has to truncate the data. In other words, numerically 𝒮f\mathcal{S}f is represented by a matrix AA, where each entry ai,ja_{i,j} corresponds to 𝒮f(xi,j)\mathcal{S}f(x_{i,j}) with vertices xi,jx_{i,j} uniformly sampled inside the square [1,1]×[1,1][-1,1]\times[-1,1], while 𝒮f\mathcal{S}f has unbounded support. Such truncation creates numerical errors when computing the Radon transform of (𝒮f)(ψ,t)\mathcal{R}(\mathcal{S}f)(\psi,{t}) along lines that pass through truncated “tail” of 𝒮f\mathcal{S}f. In the reconstructed images these errors appear in the form of artifacts at the edges of the unit square. Moreover, in unstable geometric configurations those errors get amplified along lines l(ψ,t)l(\psi,t) with normal unit vector ψ\psi corresponding to singular directions of Type 2 for 𝒮\mathcal{S}, due to the presence of q(ψ)q(\psi) in the inversion formula (15).

It is easy to notice that in the setup corresponding to an even number of rays (see Figure 5b) the reconstruction has severe artifacts propagating along lines with normals in the direction of Type 2 singularities. Namely, the directions of Type 2 singularities are ψ1=(cosπ6,sinπ6)\psi_{1}=(\cos\frac{\pi}{6},-\sin\frac{\pi}{6}) and ψ2=(cos5π6,sin5π6)\psi_{2}=(\cos\frac{5\pi}{6},\sin\frac{5\pi}{6}), and the severe artifacts are pronounced along lines parallel to the vector (cosπ3,sinπ3)(\cos\frac{\pi}{3},\sin\frac{\pi}{3}). Similar issues can be observed in Figure 7(a) corresponding to the configuration with odd number of rays and Type 2 singularities described in the proof of Part 2a of Theorem 3.

The setups with an odd number of rays that do not have singular directions of Type 2 are of much better quality and do not have such severe artifacts (see Figures 6(a), 6(b) and 7(b)).The milder artifacts close to the boundary of the unit square are almost entirely outside of the support of the image function (i.e. outside of the unit disc) and could have been cleared after the reconstruction, but we kept them for completeness of the experimental results.

Refer to caption
(a) Shepp-Logan phantom
used in all reconstructions
Refer to caption
(b) Reconstruction from 𝒮\mathcal{S} with
γ1=(1,0),γ2=(cos2π3,sin2π3)\gamma_{1}=(1,0),\gamma_{2}=(\cos\frac{2\pi}{3},\sin\frac{2\pi}{3})
Figure 5:
Refer to caption
(a) Reconstruction from regular 𝒮\mathcal{S}
with γ1=(1,0),γ2=(cos2π3,sin2π3),γ3=(cos4π3,sin4π3)\gamma_{1}=(1,0),\gamma_{2}=(\cos\frac{2\pi}{3},\sin\frac{2\pi}{3}),\\ \gamma_{3}=(\cos\frac{4\pi}{3},\sin\frac{4\pi}{3})
Refer to caption
(b) Reconstruction from 𝒮\mathcal{S} with
γ1=(cosπ20,sinπ20),γ2=(cos2π3,sin2π3),γ3=(cos4π3,sin4π3)\gamma_{1}=(\cos\frac{\pi}{20},\sin\frac{\pi}{20}),\gamma_{2}=(\cos\frac{2\pi}{3},\sin\frac{2\pi}{3}),\\ \gamma_{3}=(\cos\frac{4\pi}{3},\sin\frac{4\pi}{3})
Figure 6:
Refer to caption
(a) Reconstruction from 𝒮\mathcal{S} with
γ1=(1,0),γ2=(0,1),γ3=(cos3π4,sin3π4)\gamma_{1}=(1,0),\gamma_{2}=(0,1),\\ \gamma_{3}=(\cos\frac{3\pi}{4},\sin\frac{3\pi}{4})
Refer to caption
(b) Reconstruction from regular 𝒮\mathcal{S} with
γ1=(1,0)\gamma_{1}=(1,0), γ2=(cos2π5,sin2π5),\gamma_{2}=(\cos\frac{2\pi}{5},\sin\frac{2\pi}{5}),
γ3=(cos4π5,sin4π5),\gamma_{3}=(\cos\frac{4\pi}{5},\sin\frac{4\pi}{5}), γ4=(cos6π5,sin6π5),\gamma_{4}=(\cos\frac{6\pi}{5},\sin\frac{6\pi}{5}),
γ5=(cos8π5,sin8π5)\gamma_{5}=(\cos\frac{8\pi}{5},\sin\frac{8\pi}{5})
Figure 7:

9 Additional Remarks

  1. 1.

    The statement of a special case of Theorem 1 was published without proof in the abstract [4].

  2. 2.

    We did not aim to formulate the most general conditions under which Theorem 1 holds. The requirement of fCc(2)f\in C_{c}(\mathbb{R}^{2}) is sufficient, but not necessary. The proof will hold under weaker assumptions, e.g. when ff decays sufficiently fast at infinity.

  3. 3.

    In the case of m=1m=1, Theorem 1 establishes a relation between the divergent beam transform and the (ordinary) Radon transform. A different relation between these transforms was obtained in [14], where the divergent beam transform is integrated over all possible angles in Sn1S^{n-1} to generate the integral of the function over the whole space n\mathbb{R}^{n}. The latter can also be obtained by integrating the Radon transform over all hyperplanes normal to a given direction.

    Similarly, in the case of m=2m=2, our Theorem 1 establishes a relation between the V-line transform and the Radon transform. A different relation between these transforms was obtained in [21] using an overdetermined setup of the V-line transform. It is similar in the spirit to the formula for the divergent beam transform in [14], as it uses all possible angles for the branch directions of the V-line transform.

    Our formulas derived in Theorem 1 are essentially different from those in [14] and [21], as we do not require overdetermined data and use only a finite number of branch directions.

10 Acknowledgements

We would like to thank the anonymous reviewers for several insightful observations that allowed us to shorten some of the proofs and considerably improve the manuscript.

This work was partially funded by NSF grant DMS 1616564 and Simons Foundation grant 360357. The authors are thankful to Alessandro Conflitti for updates about the current status of his conjecture and useful references.

References

  • [1] G. Ambartsoumian, Inversion of the V-line Radon transform in a disc and its applications in imaging, Computers & Mathematics with Applications, 64, 3 (2012), pp 260–-265.
  • [2] G. Ambartsoumian, V-line and conical Radon transforms with applications in imaging, Chapter 7 in “The Radon Transform: The First 100 Years and Beyond”, edited by R. Ramlau and O. Scherzer, Radon Series on Computational and Applied Mathematics, De Gruyter, 2019.
  • [3] G. Ambartsoumian and M. J. Latifi Jebelli, The V-line transform with some generalizations and cone differentiation, Inverse Problems, 35 3 (2019), 034003.
  • [4] G. Ambartsoumian and M. J. Latifi Jebelli, Inversion of the star transform, an abstract in “Tomographic Inverse Problems: Theory and Applications”, edited by M. Burger, B. Hahn and E. T. Quinto, Oberwolfach Reports, EMS, 2019.
  • [5] G. Ambartsoumian and S. Moon, A series formula for inversion of the V-line Radon transform in a disc, Computers & Mathematics with Applications, 66 9 (2013), pp 1567–1572.
  • [6] G. Ambartsoumian and S. Roy, Numerical inversion of a broken ray transform arising in single scattering optical tomography, IEEE Transactions on Computational Imaging, 2 2 (2016), pp 166–173.
  • [7] A. Conflitti, Zeros of real symmetric polynomials, Applied Mathematics E-Notes, 6 (2006), pp 219–224.
  • [8] L. Florescu, V. Markel and J. Schotland, Inversion formulas for the broken ray Radon transform, Inverse Problems, 27 2 (2011) 025002.
  • [9] L. Florescu, J. Schotland and V. Markel, Single scattering optical tomography, Physical Review E, 79 3 (2009), 036607.
  • [10] L. Florescu, V. Markel and J. Schotland, Single scattering optical tomography: simultaneous reconstruction of scattering and absorption Physical Review E, 81 1 (2010), 016602.
  • [11] L. Florescu, V. Markel and J. Schotland, Nonreciprocal broken ray transforms with applications to fluorescence imaging, Inverse Problems, 34 9 (2018) 094002.
  • [12] R. Gouia-Zarrad and G. Ambartsoumian, Exact inversion of the conical Radon transform with a fixed opening angle, Inverse Problems, 30 4 (2014), 045007.
  • [13] M. Haltmeier, S.  Moon and D. Schiefeneder. Inversion of the attenuated V-line transform for SPECT with Compton cameras. IEEE Transactions on Computational Imaging, 3 4 (2017) pp 853–863.
  • [14] C. Hamaker, K. T. Smith, D. C. Solomon and S. l. Wagner, The divergent beam x-ray transform, Rocky Mountain J. Math., 10 1 (1980), pp 253–284.
  • [15] A. Katsevich and R. Krylov, Broken ray transform: inversion and a range condition, Inverse Problems, 29 7 (2013), 075008.
  • [16] R. Krylov and A. Katsevich, Inversion of the broken ray transform in the case of energy dependent attenuation, Physics in Medicine & Biology, 60 11 (2015), pp 4313–4334.
  • [17] I. G. Macdonald, Symmetric Functions and Hall Polynomials, 2-nd Edition, Clarendon Press, Oxford 1998.
  • [18] V. Maxim, M. Frandes and R. Prost Analytical inversion of the Compton transform using the full set of available projections, Inverse Problems, 25 9 (2009), 095001.
  • [19] M. K. Nguyen, T. T. Truong and P. Grangeat. Radon transforms on a class of cones with fixed axis direction Journal of Physics A: Mathematical and General, 38 37 (2005), 8003-8015.
  • [20] B. Sherson, Some results in single-scattering tomography, PhD Thesis, Oregon State University, 2015.
  • [21] F. Terzioglu, Some inversion formulas for the cone transform, Inverse Problems, 31 11 (2015), 115010.
  • [22] F. Terzioglu, P. Kuchment and L. Kunyansky, Compton camera imaging and the cone transform: a brief overview, Inverse Problems, 34 5 (2018), 054002.
  • [23] T. T. Truong and M. K. Nguyen, On new V-line Radon transforms in 2\mathbb{R}^{2} and their inversion, Journal of Physics A: Mathematical and Theoretical, 44 7 (2011), 075206.
  • [24] T. T. Truong and M. K. Nguyen, New properties of the V-line Radon transform and their imaging applications, Journal of Physics A: Mathematical and Theoretical, 48 40 (2015), 405204.
  • [25] M. R. Walker and J. A. O’Sullivan, The broken ray transform: additional properties and new inversion formula, Inverse Problems, 35 11 (2019), 115003.
  • [26] Z. Zhao, J. Schotland and V. Markel, Inversion of the star transform, Inverse Problems, 30 10 (2014), 105001.