Paper

Null space and resolution in dynamic computerized tomography

Published 11 January 2016 © 2016 IOP Publishing Ltd
, , Citation Bernadette N Hahn 2016 Inverse Problems 32 025006 DOI 10.1088/0266-5611/32/2/025006

0266-5611/32/2/025006

Abstract

One major challenge in computerized tomography is to image objects which change during the data acquisition and hence lead to inconsistent data sets. Motion artefacts in the reconstructions can be reduced by applying specially adapted algorithms which take the dynamic behaviour into account. Within this article, we analyse the achievable resolution in the dynamic setting in case of two-dimensional affine deformations. To this end, we characterize the null space of the operator describing the dynamic case, using its singular value decomposition and a necessary dynamic consistency condition. This shows that independent of any reconstruction method, the specimen's dynamics results in a loss of resolution compared to the stationary setting. Our theoretical results are illustrated at a numerical example.

Export citation and abstract BibTeX RIS

1. Introduction

The mathematical model of computerized tomography with stationary specimen is given by the Radon transform ${ \mathcal R }$ which maps integrable functions in ${{\mathbb{R}}}^{N}$ to their integrals over hyperplanes, i.e.

Equation (1)

where δ denotes the delta distribution. The searched-for x-ray attenuation coefficient f has to be recovered from the measured Radon data by solving the equation ${ \mathcal R }f=g.$ In [22], Radon proved the existence of a unique solution if the data g are known for all $({\boldsymbol{\theta }},s)\in {S}^{N-1}\times {\mathbb{R}}$ and $N=2,3.$ However, this condition cannot be realized in practice. If only a finite number of projections, ${{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p},$ is available, the uniqueness result does no longer hold [23], i.e. there is a non-trivial null space. Its elements, the so-called ghosts, are invisible for the considered directions, and hence, the reconstruction method has to prevent them from disturbing the computed images. Thus, to guarantee a reliable diagnostics in applications, it is essential to study the possible resolution carefully. Disregarding the resolution properties for the data acquisition could lead e.g. to false medical diagnoses when ghosts are mistaken as tumours.

In the case of stationary specimens, the singular value decomposition of the Radon transform is given in [4, 16, 18] for arbitrary dimensions and different weights. This provided the basis for the characterization of the elements in the null space [15, 17], revealing that they are highly oscillating functions, and therefore affect only the small details of the object. Especially, this provides a statement about the overall resolution of the imaging system. A respective singular value decomposition, null space characterizations and resolution properties for the x-ray transform, which in arbirtrary dimensions integrates a function along straight lines instead of hyperplanes, have been derived in [20].

In many applications, the assumption that the specimen is stationary during the measuring does not hold, e.g. in medical imaging due to moving organs or in non-destructive testing, e.g. in imaging driven liquid fronts. Since the scanning process is time-dependent due to the rotation of the radiation source, the dynamics of the object results in inconsistent data sets. Applying standard reconstruction techniques to motion-corrupted data leads in general to motion artefacts in the reconstructed images. Therefore, e.g. iterative [1, 13] and analytical [3, 5, 9, 14] reconstruction methods have been designed which aim to compensate for the object's motion.

Within this article, we study resolution properties in this dynamic setting with two-dimensional affine deformations. Based on our theoretical results for dynamic linear inverse problems with moderate deformations, we state the singular value decomposition of the operator describing the dynamic setting in section 3. Together with a necessary consistency condition on data for moving objects, this yields a characterization of the null space in case of finitely many projections. In section 4, we discuss the theoretical results with respect to the resolution properties. Especially, this shows that the temporal changes diminish the resolution compared to the classical stationary setting. This loss in resolution is also illustrated at a numerical example.

2. Mathematical basis

2.1. Preliminaries on the Radon transform

The 2D Radon transform (1) can be considered as mapping

with the cylinder $Z:= {S}^{1}\times {\mathbb{R}}.$ Without loss of generality, it is assumed that the investigated object described by the x-ray attenuation coefficient f is compactly supported in the unit ball ${V}_{1}(0)\subset {{\mathbb{R}}}^{2},$ [21].

The aim is to determine the unknown function f from the measured Radon data, i.e. to solve the inverse problem

A data set $g\in {L}_{2}(Z)$ is called consistent if it is in the range of ${ \mathcal R },$ denoted by $\mathrm{Im}({ \mathcal R }).$ The following Helgason–Ludwig-consistency conditions hold for the Radon transform, [6, 12].

Theorem 2.1. It holds $g\in \bar{\mathrm{Im}({ \mathcal R })}$ if and only if

  • (i)  
    $g({\boldsymbol{\theta }},s)=0$ for $| s| \geqslant 1,$
  • (ii)  
    g is even,
  • (iii)  
    ${\displaystyle \int }_{{\mathbb{R}}}{s}^{n}\;g({\boldsymbol{\theta }},s)\;{\rm{d}}s={p}_{n}({\boldsymbol{\theta }})$ with a trigonometric polynomial pn of degree $\leqslant n.$

In [16, 17], the singular value decomposition of the Radon transform was derived in arbitrary dimensions. Denote $w(s):= {(1-{s}^{2})}^{-1/2}.$

Theorem 2.2. The singular value decomposition of the Radon transform

is given by

where

and

with the Jacobi-polynomials ${P}_{n}^{(\alpha ,\beta )},$ the Chebychev-polynomials of the second kind Un and the spherical harmonics Yl of degree l in ${{\mathbb{R}}}^{2}.$

In practical applications, only a finite number of projections can be measured. In the stationary setting, it is known that the Radon transform with complete projections for finitely many directions has a nontrivial null space, so f is not uniquely determined [16]. Complete projections means that $g({{\boldsymbol{\theta }}}_{j},s)$ is known for $s\in {\mathbb{R}}.$ Thus, this is refered to as the semi-discrete case. Let

denote this null space and let $A:= \{{{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p}\}\subset {S}^{1}$ not be contained in an algebraic variety of degree $\lt p,$ i.e. there is no $q\in {{ \mathcal P }}_{l},$ $l\lt p,q\not\equiv 0$ with $q({\boldsymbol{\theta }})=0$ for all ${\boldsymbol{\theta }}\in A,$ see [16]. Here

Equation (2)

is the set of spherical harmonics of degree $\lt l$ which are even for l even, and odd for l odd. Please note that in the two-dimensional case, a set of p distinct source positions fulfils this condition, and that the angles of ${{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p}$ are not necessarily equispaced, see [21].

Then, for $f\in {{ \mathcal N }}_{p}({ \mathcal R }),$ it holds

Equation (3)

with a trigonometric polynomial qn of degree $\leqslant n,$ see [15, 16]. This means, for fixed ${\boldsymbol{\theta }}\in {S}^{1},$ the series expansion of ${ \mathcal R }f({\boldsymbol{\theta }},\cdot )$ in Chebychev-polynomials Un contains only polynomials of degree p and larger. An explicit characterization of the elements in the null space, the so-called ghosts, is given in [17], showing that they are highly oscillating functions by studying their frequency distribution. Especially, this yields that only details of size $2\pi /p$ and larger can be reliably reconstructed from Radon data measured from p directions.

2.2. The mathematical model of dynamic computerized tomography

Due to the rotation of the radiation source, the data acquisition in computerized tomography takes a certain amount of time. The position of the x-ray source is described by ${\boldsymbol{\theta }}\in {S}^{1},$ and therefore, each unit vector ${\boldsymbol{\theta }}\in {S}^{1}$ can be uniquely described by a time instance t, via the parametrization

where ϕ is the rotation angle of the source [8, 9].

Occasionally, throughout the paper, we have to depict a unit vector in terms of its angular phase. To keep the notation simple, ${\boldsymbol{\theta }}$ denotes the unit vector and θ its respective angle, so it holds the correlation ${\boldsymbol{\theta }}={(\mathrm{cos}\theta ,\mathrm{sin}\theta )}^{T}.$

In the following, let $f\in {L}_{2}({V}_{1}(0))$ describe the state of the investigated object at the beginning of the scanning. Its state at the moment of measuring the data ${g}_{{\boldsymbol{\theta }}}$ is then assumed to be given by $f\;\circ \;{{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}$ with diffeomorphic motion functions ${{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}\ :{{\mathbb{R}}}^{2}\to {{\mathbb{R}}}^{2}.$ This is a common model for the dynamic behaviour in tomography, see e.g. [2, 8, 14]. Within this setting, the mathematical model of dynamic computerized tomography is given by

A detailed derivation of the dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}}$ is given in [10] in the general context of linear inverse problems.

The task of reconstruction methods in the dynamic case is to determine both the motion functions and the reference function f from the measured data. In this paper, we want to study the effect of the dynamic behaviour on the achievable resolution, independent of a reconstruction procedure. Therefore, considering the exact motion functions ${{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}$ for our mathematical analysis is justified.

We now discuss some properties of the underlying motion model. Without loss of generality, we can make the assumption that the trajectory of a particle x,

Equation (4)

is smooth, since in applications, only a discrete number of projections is measured.

The classical Radon transform ${ \mathcal R }$ is a symmetric operator in the sense that

Equation (5)

see also theorem 2.1. In contrast, functions in the range of the dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}}$ are in general not even functions of the cylinder variables. However, the data acquisition scheme in 2D parallel scanning geometry makes use of the symmetry of the Radon transform (5), i.e. the data are collected for unit vectors ${\boldsymbol{\theta }}$ covering only the half sphere. The same acquisition process is applied in the dynamic case, i.e. we can assume ${{\rm{\Gamma }}}_{-{\boldsymbol{\theta }}}={{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}$ without affecting the measured data, [8]. This yields

Equation (6)

3. Properties of the dynamic imaging operator

3.1. A necessary consistency condition

Changes of the object during the scanning lead in general to inconsistent data sets, i.e. $g\notin \mathrm{Im}({ \mathcal R }).$ However, in the noise-free case, it is an element in the range of the dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}}.$ For $g\in \mathrm{Im}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ we prove a condition similar to the property in theorem 2.1, iii), which is then a necessary consistency condition for ${{ \mathcal R }}_{{\rm{\Gamma }}}.$ To keep the notation simple, we denote

Theorem 3.1. Let the inverse motion functions be given by

Equation (7)

Then, for $f\in {L}_{2}({V}_{1}(0)),$ it holds

Proof. Using the transformation of coordinates $x:= {{\rm{\Gamma }}}_{\theta }^{-1}z,$ we obtain

Using the representation (7) of the inverse motion functions ${{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}^{-1},$ it holds

Thus, the mapping $\theta \mapsto {\left({({{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}^{-1}z)}^{T}{\boldsymbol{\theta }}\right)}^{n}$ is an element in ${{ \mathcal T }}_{(L+1)n}.$ The components of the Jacobian $D{{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}^{-1}z$ are given by

Since the differentiation takes place with respect to the spatial coordinates only, we obtain

and thus

Altogether, it holds $p\in {{ \mathcal T }}_{(L+1)n+2L}.$

This result will play an important role within the characterization of the null space of the semi-discrete dynamic operator. Especially, it is not restricted to affine deformations and holds for all motion models satisfying (7). In applications, this condition (7) can be realized by choosing a suitable underlying motion model. Thus, it is not a restrictive condition.

3.2. Affine deformations

We want to study resolution properties in the dynamic setting at the case of affine deformations. Thus, throughout the rest of the paper, the motion functions are given by

with invertible matrix ${A}_{{\boldsymbol{\theta }}}\in {{\mathbb{R}}}^{2\times 2}$ and vector ${b}_{{\boldsymbol{\theta }}}\in {{\mathbb{R}}}^{2}.$ Due to the smoothness assumption on the trajectories ${\mathrm{tr}}_{x},$ (4), the entries of ${A}_{{\boldsymbol{\theta }}}$ and ${b}_{{\boldsymbol{\theta }}}$ are considered to be continuously differentiable. Further, we assume that the motion model satisfies

Equation (8)

where ${A}_{{\boldsymbol{\theta }}}^{-*}:= {\left({A}_{{\boldsymbol{\theta }}}^{*}\right)}^{-1}$ denotes the inverse of the adjoint matrix ${A}_{{\boldsymbol{\theta }}}^{*}$. This condition means that the dynamic behaviour does not lead to infinitesimally rigid integration curves, see [11], where this condition is analysed in more detail.

Theorem 3.2. Let the object's deformation be given by affine motion functions as described above. Then, the dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}}\ :{L}_{2}({V}_{1}(0))\to {L}_{2}(Z)$ is related to the classical Radon transform via

with

Equation (9)

and

Equation (10)

The proof of this representation can be found in [10].

In the following, we consider the case where the measured data cover directions distributed over the complete sphere. For our analytical framework, we therefore assume the condition

Otherwise, the dynamic behaviour would result in limited-data, where only certain singularities of the object are detectable from the data, which has to be taken into account while analysing the spatial resolution.

Lemma 3.3. The mappings

and

are, except on a measure-zero set, diffeomorphic functions. Especially, it holds

with $h({\boldsymbol{\theta }})$ stated in (8).

Proof. The mapping ${T}_{{\boldsymbol{\theta }}}$ is differentiable with respect to s. Since ${A}_{{\boldsymbol{\theta }}}$ is injective, it holds especially

Hence, ${T}_{{\boldsymbol{\theta }}}$ is a diffeomorphism.

To prove the property for the mapping T, we use again the isometric correlation between a unit vector and its phase angle. Especially, T corresponds to

In case of ${({A}_{{\boldsymbol{\theta }}}^{-*}{\boldsymbol{\theta }})}_{1}=0,$ it holds either

such that $\mu \left(\left\{\varphi \in [0,2\pi )\ :{({A}_{{\boldsymbol{\theta }}}^{-*}{\boldsymbol{\theta }})}_{1}=0\right\}\right)=0$ with Lebesgue-measure μ. In case of ${({A}_{{\boldsymbol{\theta }}}^{-*}{\boldsymbol{\theta }})}_{1}\ne 0,$ it holds

Thus, we obtain the representation

Using the chain rule and the property (8) yields

3.3. The singular value decomposition

The following theorem states the singular value decomposition of ${{ \mathcal R }}_{{\rm{\Gamma }}}$ with Γ describing affine deformations.

Theorem 3.4. Let ${w}_{{\rm{\Gamma }}}({\boldsymbol{\theta }},s):= | \mathrm{det}{A}_{{\boldsymbol{\theta }}}{| }^{2}\;| h({\boldsymbol{\theta }})| \;| \mathrm{det}{{DT}}_{{\boldsymbol{\theta }}}(s)| \;w({T}_{{\boldsymbol{\theta }}}(s))$ with the weight function $w(s)={(1-{s}^{2})}^{-1/2}$ of the Radon transform and $h({\boldsymbol{\theta }})$ as in (8).

The singular value decomposition of the operator

is given by

with

and

where ${I}_{{\boldsymbol{\theta }}}:= {T}_{{\boldsymbol{\theta }}}^{-1}([-1,1]).$ As in the singular value decomposition of the Radon transform, P denotes the Jacobi polynomials, Un the Chebychev-polynomials of the second kind and Yl the spherical harmonics.

We prove this statement in the appendix, using a result derived in [10] in the general context of dynamic linear inverse problems.

Remark. Since the singular values of ${{ \mathcal R }}_{{\rm{\Gamma }}}$ coincide with the ones of the classical Radon transform, this shows that the degree of the ill-posedness is not changed by a dynamic behaviour with affine motion functions. This was already shown in [8] in terms of sobolev space estimates for ${{ \mathcal R }}_{{\rm{\Gamma }}}.$

In addition, the singular value decomposition of ${{ \mathcal R }}_{{\rm{\Gamma }}}$ provides with the Picard criterion a sufficient consistency condition for $g\in {L}_{2}(Z,{w}_{{\rm{\Gamma }}})$ to be in $\mathrm{Im}({{ \mathcal R }}_{{\rm{\Gamma }}}).$

Based on the singular value decomposition, we deduce a representation of ${{ \mathcal R }}_{{\rm{\Gamma }}}f,$ which is later used to characterize the null space of the dynamic operator. Within the proof, the following lemma is required.

Lemma 3.5. For ${{ \mathcal R }}_{{\rm{\Gamma }}}f\in {L}_{2}(Z,{w}_{{\rm{\Gamma }}}),$ it holds

Proof. For reasons of simplicity, we shortly write ${\displaystyle \sum }_{m}{\displaystyle \sum }_{l}$ for the sum ${\displaystyle \sum }_{m}{\displaystyle \sum }_{| l| \leqslant m,l+m\in 2{\mathbb{Z}}}.$ Using the singular value decomposition of ${{ \mathcal R }}_{{\rm{\Gamma }}},$ along with the coordinate transform $\tilde{s}:= {T}_{{\boldsymbol{\theta }}}(s),$ it holds

Since $w{(s)}^{-1}=\sqrt{1-{s}^{2}},$ the assertion follows with the fact that the Chebyshev-polynomials form an orthonormal set on ${L}_{2}([-1,1],{w}^{-1}).$

For $f\in {L}_{2}({V}_{1}(0)),$ we denote in the following

Equation (11)

Theorem 3.6. Let ${{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}$ satisfy condition (7), i.e. ${{\rm{\Gamma }}}_{{\boldsymbol{\theta }}}^{-1}$ is given by trigonometric polynomials of degree $L\in {\mathbb{N}}.$ Then, for $f\in {L}_{2}({V}_{1}(0)),$ it holds

and ${q}_{n}^{{\rm{\Gamma }},f}\in {{ \mathcal T }}_{(L+1)n+2L}$ with $((L+1)n+2L+1)$ coefficients.

Proof. Together with the representation of the singular functions ${u}_{{nl}}^{{\rm{\Gamma }}}$ and lemma 3.5, the singular value decomposition yields

For fixed ${\boldsymbol{\theta }}\in {S}^{1},\;$ ${U}_{n}({T}_{{\boldsymbol{\theta }}}(s))$ is a polynomial in s of degree n. Thus, according to the consistency condition 3.1 for the dynamic operator, we have

Further, we obtain

With the symmetry ${{\rm{\Gamma }}}_{-{\boldsymbol{\theta }}}={{\rm{\Gamma }}}_{{\boldsymbol{\theta }}},$ it holds $T(-{\boldsymbol{\theta }})=-T({\boldsymbol{\theta }}),$ and, using the symmetry property of the Radon transform and the fact that Un is even (odd) for n even (odd), we obtain

This property yields the representation

with ${\alpha }_{l}^{n}=\left\{\begin{array}{ll}2l & \mathrm{if}\ n\ \mathrm{even},\\ 2l+1 & \mathrm{if}\ n\ \mathrm{odd}.\end{array}\right.$

4. The null space and resolution properties

4.1. The null space

In the following, we study the elements in the null space of the semi-discrete dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}},$

with p distinct source positions ${{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p}\in {S}^{1}.$ As in the static case, $\{{{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p}\}$ is then not contained in an algebraic variety of degree $\lt p,$ i.e. there is no $q\in {{ \mathcal P }}_{l},l\lt p,q\not\equiv 0$ with $q({{\boldsymbol{\theta }}}_{j})=0$ for all $j=1,...,p,$ see (2), [21].

Now, we first derive a series expansion for ${{ \mathcal R }}_{{\rm{\Gamma }}}f$ with $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ which holds independently of any sampling scheme. In a next step, the a priori information of the specific source positions is taken into account. Finally, based on these characterizations of ${{ \mathcal R }}_{{\rm{\Gamma }}}f,$ we state an explicit representation of the ghosts in the dynamic case.

Theorem 4.1. For $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ it holds

with ${q}_{n}^{{\rm{\Gamma }},f}$ as in (11) and ${q}_{n}^{{\rm{\Gamma }},f}({{\boldsymbol{\theta }}}_{j})=0$ for $j=1,...,p.$

Proof. For $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ it holds ${{ \mathcal R }}_{{\rm{\Gamma }}}f({{\boldsymbol{\theta }}}_{j},s)=0\ \mathrm{for}\ \mathrm{almost}\ \mathrm{all}\ s\in {\mathbb{R}}\ \mathrm{and}\ j\quad =\quad 1,...,p.$ Due to theorem 3.6, this is equivalent to

for almost all $s\in {\mathbb{R}}$ and $j=1,...,p.$

For fixed ${\boldsymbol{\theta }},$ the transformed Chebyshev-polynomials ${({U}_{n}({T}_{{\boldsymbol{\theta }}}(s)))}_{n}$ form an orthonormal system on ${L}_{2}\left({T}_{{\boldsymbol{\theta }}}^{-1}([-1,1]),| \mathrm{det}{{DT}}_{{\boldsymbol{\theta }}}(s)| \;w{({T}_{{\boldsymbol{\theta }}}(s))}^{-1}\right),$ which can be directly verified by substituting ${T}_{{\boldsymbol{\theta }}}(s).$ Hence, it holds

Equation (12)

Due to theorem 3.6, it is ${q}_{n}^{{\rm{\Gamma }},f}\in {{ \mathcal T }}_{(L+1)n+2L}$ with $(L+1)n\quad +\quad 2L+1$ coefficients. Hence, for each n, the above condition (12) corresponds to a system of p linear homogenous equations with $((L+1)n+2L+1)$ unknowns. If $(L+1)n\quad +\quad 2L\lt p,$ or equivalently

the matrix has full rank since $\{{{\boldsymbol{\theta }}}_{1},...,{{\boldsymbol{\theta }}}_{p}\}$ is not contained in an algebraic variety of degree $\lt p,$ and it follows ${q}_{n}^{{\rm{\Gamma }},f}\equiv 0.$

Since ${T}_{{\boldsymbol{\theta }}}(s)$ is a polynomial in s of degree 1, theorem 4.1 yields that ${{ \mathcal R }}_{{\rm{\Gamma }}}f({\boldsymbol{\theta }},\cdot )$ has a series expansion in polynomials of degree $\frac{p-2L}{L+1}$ and larger. However, in the static case, the series expansion of ${ \mathcal R }$ starts with polynomials of degree p, see (3). For $L\gt 0,$ i.e. the truly dynamic case, it holds the estimate

This comparison shows that, in the dynamic case, the series expansion of the imaging operator can contain lower order terms than in the stationary case.

Our provided series expansion, especially the lower bound, describes the worst-case scenario for any deformation of degree L and for any discretization of the projection space S1. By taking into account the effect of a given motion model on a specific sampling scheme, one can obtain more detailed information. Therefore, we denote

and $\tilde{p}:= | {\rm{\Xi }}| .$ The set Ξ comprises the directions from which the reference object is actually scanned, and especially, it holds $\tilde{p}\leqslant p.$ In applications, it is possible to determine the number $\tilde{p}$ directly from the data by detecting e.g. redundancies in the measured data set.

Theorem 4.2. Let $\tilde{p}=| {\rm{\Xi }}| $ as introduced above. Then, for $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ it holds

with ${q}_{n}^{{\rm{\Gamma }},f}$ as in (11) and ${q}_{n}^{{\rm{\Gamma }},f}({{\boldsymbol{\theta }}}_{j})=0$ for $j=1,...,p.$

Proof. For $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ theorem 4.1 yields the representation

where ${q}_{n}^{{\rm{\Gamma }},f}$ is the polynomial as in (11) with property

Equation (13)

Using lemma 3.5, ${q}_{n}^{{\rm{\Gamma }},f}$ has the representation

with ${c}_{{nl}}={\sigma }_{n}\langle f,{v}_{{nl}}\rangle {\pi }^{-1}\in {\mathbb{R}}.$ Since ${| \mathrm{det}{A}_{{{\boldsymbol{\theta }}}_{j}}| }^{-1}\ne 0$ for $j=1,...,p,$ (13) is equivalent to the condition

This corresponds to a system of p equations and $n+1$ unknowns. Since ${\xi }_{j}\in {\rm{\Xi }}$ with $| {\rm{\Xi }}| =\tilde{p},$ the respective matrix has rank $\mathrm{min}(\tilde{p},n+1).$ Thus, for $n\lt \tilde{p},$ we have $\sum {c}_{{nl}}{Y}_{l}(\xi )=0\ \forall \ \xi \in {S}^{1},$ and so

leading to the asserted representation of ${{ \mathcal R }}_{{\rm{\Gamma }}}f.$

Remark. Please note that theorem 4.2 and its proof hold even in the non-symmetric case, i.e. when measurements are taken over the complete sphere and when ${{\rm{\Gamma }}}_{-\theta }={{\rm{\Gamma }}}_{\theta }$ is not necessarily fulfilled for all θ, see (6).

From theorem 4.2 and the respective proof, we obtain the following representation of the functions in the null space.

Corollary 4.3. For $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}),$ it holds the series expansion

with coefficients ${c}_{{nl}}\in {\mathbb{R}}.$

Proof. Let $f(x)={\displaystyle \sum }_{n}{\displaystyle \sum }_{l}{c}_{{nl}}{v}_{{nl}}$ be the series expansion of f in the orthonormal set ${({v}_{{nl}})}_{{nl}},$ with ${c}_{{nl}}=\langle f,{v}_{{nl}}\rangle .$ According to theorem 4.2 and its proof, we have for $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}})$

where ${q}_{n}^{{\rm{\Gamma }},f}$ has the representation

Thus, with ${q}_{n}^{{\rm{\Gamma }},f}({\boldsymbol{\theta }})\equiv 0,$ it holds also

This yields the condition

for the coefficients in the series expansion of f. Since ${({Y}_{l})}_{l\in {\mathbb{N}}}$ is an orthonormal system on ${L}_{2}({S}^{1}),$ this yields

and hence the asserted representation of f.

4.2. Resolution

To interpret the previous results in terms of resolution, we characterize the elements of ${{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}})$ in the Fourier space.

Theorem 4.4. Let $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}}).$ Then, its Fourier transform has the representation

with the Bessel-functions Jn of the first kind.

Proof. With the series expansion of ${{ \mathcal R }}_{{\rm{\Gamma }}}f$ as given in theorem 4.2, it holds for $f\in {{ \mathcal N }}_{p}({{ \mathcal R }}_{{\rm{\Gamma }}})$

with the function

For the Fourier transform ${ \mathcal F }{H}_{{\boldsymbol{\theta }}},$ we obtain with the substitution $z:= {T}_{{\boldsymbol{\theta }}}(s)$ and the definition of ${T}_{{\boldsymbol{\theta }}},$ (10),

In the last step, we used that the Bessel-functions Jn fulfill

see formula 7.321, [7]. Altogether, we obtain

Equation (14)

The Fourier transform $\widehat{{{ \mathcal R }}_{{\rm{\Gamma }}}f}$ is related to the Fourier transform of f by the Fourier-slice-theorem

see [8]. It provides the representation of $\widehat{f}$ in polar coordinates

Using (14) yields

With the representation of ${q}_{n}^{{\rm{\Gamma }},f}$ according to lemma 3.5, we obtain immediately

with a ${q}_{n}\in \mathrm{span}\{{Y}_{l},| l| \leqslant n,l+n\in 2{\mathbb{Z}}\},{q}_{n}({\boldsymbol{\theta }})=0$ for ${\boldsymbol{\theta }}\in {\rm{\Xi }}.$

Now, we want to study the frequency distribution of these ghosts by comparing the energy of their Fourier transform restricted to a circle ${V}_{c}(0)$ around 0 of radius c with their total energy. Therefore, we denote $M:= \mathrm{max}\left\{\tilde{p},\frac{p-2L}{L+1}\right\}$ and

The asymptotic behaviour in the static case was derived in [17]. The proof therein is based on a representation of $f\in {{ \mathcal N }}_{p}({ \mathcal R })$ in Fourier space as

with ${q}_{n}\in \mathrm{span}\{{Y}_{l},| l| \leqslant n,l+n\in 2{\mathbb{Z}}\},{q}_{n}({{\boldsymbol{\theta }}}_{j})=0$ for $j=1,...,M$, $M\gt 0.$ Thus, with the same arguments as in [17], we obtain immediately the following behaviour.

Lemma 4.5. Let $\alpha ,\beta \in {\mathbb{R}}.$ It holds

This provides the following interpretation with respect to resolution. The spectrum of the ghosts is, at least asymptotically, concentrated outside a disk around zero with radius $\mathrm{max}\left\{\tilde{p},\frac{p-2L}{L+1}\right\}.$ Therefore, a reliable reconstruction is only possible for functions which are essentially M-band limited with $M=\mathrm{max}\left\{\tilde{p},\frac{p-2L}{L+1}\right\},$ i.e. for details corresponding to frequencies

In accordance to Shannon's sampling theorem, these are the details of size

and larger.

We compare this result in the dynamic case with the respective result from the stationary setting, where data from p directions are required to reliably reconstruct details of size $2\pi /p$ and larger. Since $\mathrm{max}\left\{\tilde{p},\frac{p-2L}{L+1}\right\}\leqslant p,$ i.e.

we obtain that, in the static case, smaller details can be reliably reconstructed from the same amount of data than in the dynamic setting. To guarantee a similar resolution in dynamic and static case, more data, i.e. a finer sampling, is required.

Our analysis also provides a worst case estimate on how many projections are required to guarantee a spatial resolution of $2\pi /b:$ Using $(p-2L)/(L+1)$ as lower bound for the resolution, i.e. without any knowledge about the sampling scheme, approximately $(L+1)b$ directions are required. If we want to guarantee the same resolution as in the static setting, i.e. for b = p, approximately $L+1$ times as many data have to be collected.

4.3. Numerical example

As a first example, we consider the specimen shown in figure 1, performing a rotation during the scanning with

Thus, the inverse motion functions are given by trigonometric polynomials of degree L = 3. The Radon data are collected from p = 136 projections uniformly distributed over S1, i.e.

To study the effect of different numbers of projections, the fixed number of 451 detector points is used for all numerical examples, independent of the number of projections. More precisely, we used the detector points

with q = 225. The worst-case estimate guarantees a spatial resolution of at least $2\pi /33.$ To get a sharper estimate, we take also the effect of the motion on the specific sampling scheme into account. With

we obtain

and hence, $\tilde{p}=| {\rm{\Xi }}| =34.$ Since $\mathrm{max}\left\{\tilde{p},\frac{p-2L}{L+1}\right\}=34,$ the achievable resolution in this dynamic setting is $2\pi /34.$ In comparison, in case of stationary data, the achievable resolution would be $2\pi /136.$ This illustrates that the resolution is actually significantly reduced by the dynamic behaviour of the object, which gets also obvious by comparing the reconstruction results from both dynamic and stationary data, see figures 2 and 3.

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Original phantom at the initial time of the scanning.

Standard image High-resolution image
Figure 2. Refer to the following caption and surrounding text.

Figure 2. Reconstruction from dynamic data with p = 136 projections.

Standard image High-resolution image
Figure 3. Refer to the following caption and surrounding text.

Figure 3. Reconstruction from static data with p = 136 projections.

Standard image High-resolution image
Figure 4. Refer to the following caption and surrounding text.

Figure 4. Reconstruction from dynamic data with p = 544 projections.

Standard image High-resolution image
Figure 5. Refer to the following caption and surrounding text.

Figure 5. Reconstruction from dynamic data with p = 139 projections fulfilling (16).

Standard image High-resolution image

For the reconstruction in the stationary case, the classical filtered backprojection algorithm was used with the filter

Equation (15)

where D is the Dawson integral

see [19]. To reconstruct the reference image from dynamic data, we applied the algorithm proposed in [8] which is exact for affine deformations and corresponds to a filtered backprojection method with a respectively modified version of the filter (15).

Next, we address the question how to achieve the same spatial resolution $2\pi /136.$ Using the lower estimate which is independent of the specific dynamics and sampling, approximately $(L+1)136=544$ projections are required to guarantee the desired resolution. With this number of source positions, we obtain

Hence, it is $\tilde{p}=| {\rm{\Xi }}| =136$ which proves that we obtain actually the desired resolution.

However, a comparable resolution could be obtained with far less than 544 projections, see also Figure 4 and 5. Consider e.g. p = 139, for which we obtain

Equation (16)

such that $\tilde{p}=| {\rm{\Xi }}| =139\gg \lceil \frac{p-2L}{L+1}\rceil=34,$ leading to an actual resolution of $2\pi /139,$ which is far better than the predicted resolution $2\pi /34$ by the worst-case estimate. Thus, this example illustrates that the estimate independent of Ξ only provides a coarse estimation on how many projections are required to obtain a desired resolution. Especially for high-order motion functions, a high number of projections is required to guarantee a desired resolution independent of the specific dynamic sampling scheme.

5. Conclusion

In this paper, we analysed the imaging operator of dynamic computerized tomography with 2D affine deformations with respect to achievable resolution. We showed that the dynamic behaviour of the object can lead to a loss in spatial resolution compared to the stationary case. Two predictions for the achievable resolution are given. The first provides a worst-case prediction, independent of a specific sampling scheme and motion model. The second one takes these information into account and hence provides an improved result. It corresponds to the best possible resolution which can be obtained by any reconstruction algorithm.

For applications, it is essential to know about this limitation in resolution due to the object's motion in order to interpret reconstructed images adequately. Our analysis results can also be used to choose or adapt suitable sampling schemes to image dynamic processes. For example, the number $\tilde{p}$ of actual source positions could be determined on the fly, i.e. already during the data acquisition, and so upcoming source positions could be chosen to increase the resolution.

Acknowledgement

The work of the author was supported by the Deutsche Forschungsgemeinschaft under grant LO 310/11-2.

Appendix

To prove the singular value decomposition of the dynamic operator ${{ \mathcal R }}_{{\rm{\Gamma }}},$ i.e. Theorem 3.4, we first recall a result from [10], which was derived in the context of dynamic linear inverse problems.

Analogously to the derivation of ${{ \mathcal R }}_{{\rm{\Gamma }}}$ in section 2.2, a respective dynamic operator ${{ \mathcal A }}_{{\rm{\Gamma }}}$ can be stated for each linear operator ${ \mathcal A }.$ Let ${ \mathcal A }\ :{L}_{2}({{\rm{\Omega }}}_{X})\to {L}_{2}({{\rm{\Omega }}}_{Y})$ describe an initial inverse problem with ${{\rm{\Omega }}}_{X},{{\rm{\Omega }}}_{Y}$ being compact subsets of ${{\mathbb{R}}}^{N}$ and ${{\mathbb{R}}}^{M},$ respectively. Then, its dynamic counterpart is given by

In [10], the following classificiation scheme for linear dynamic problems depending on the object's motion is provided: If the operator ${{ \mathcal A }}_{{\rm{\Gamma }}}$ is still related to the initial operator ${ \mathcal A },$ i.e. ${{ \mathcal A }}_{{\rm{\Gamma }}}={{ \mathcal V }}_{{\rm{\Gamma }}}{ \mathcal A }$ with a bijective transformation ${{ \mathcal V }}_{{\rm{\Gamma }}},$ the motion is called moderate, and strong otherwise. For operators ${{ \mathcal A }}_{{\rm{\Gamma }}}$ with moderate dynamics Γ, a singular value decomposition was derived in [10].

Theorem. Let ${ \mathcal A }\ :{L}_{2}({{\rm{\Omega }}}_{X})\to {L}_{2}({{\rm{\Omega }}}_{Y},w)$ with weight w and singular value decomposition ${({\sigma }_{n},{v}_{n},{u}_{n})}_{n\in {\mathbb{N}}}.$ Then, the singular value decomposition of

is given by

where ${w}_{{\rm{\Gamma }}}(y)={\left({{ \mathcal V }}_{{\rm{\Gamma }}}{\chi }_{{{\rm{\Omega }}}_{Y}}(y)\right)}^{-1}{{ \mathcal V }}_{{\rm{\Gamma }}}^{-*}w(y)$ with the characteristic function ${\chi }_{{{\rm{\Omega }}}_{Y}}$ of ${{\rm{\Omega }}}_{Y}.$

The Radon transform ${ \mathcal R }$ can be considered as mapping

since ${ \mathcal R }f({\boldsymbol{\theta }},s)=0$ for $| s| \gt 1.$ Thus, it fits into this general setting with ${{\rm{\Omega }}}_{X}={V}_{1}(0)$ and ${{\rm{\Omega }}}_{Y}={S}^{1}\times [-1,1].$

In theorem 3.2, it was shown that the operator ${{ \mathcal R }}_{{\rm{\Gamma }}}$ is related to the Radon transform ${ \mathcal R }$ via a bijective transformation ${{ \mathcal V }}_{{\rm{\Gamma }}}.$ Thus, affine deformations are moderate motion models with respect to the Radon transform, and we can use the afore mentioned general theorem to derive the singular value decomposition of ${{ \mathcal R }}_{{\rm{\Gamma }}}$ between suitably weighted spaces. Using the respective SVD of the Radon transform ${ \mathcal R }\ :{L}_{2}({V}_{1}(0))\to {L}_{2}({S}^{1}\times [-1,1],w),$ we obtain the singular system

for ${{ \mathcal R }}_{{\rm{\Gamma }}}\ :{L}_{2}({V}_{1}(0))\to {L}_{2}({S}^{1}\times {T}_{{\boldsymbol{\theta }}}^{-1}([-1,1]),{w}_{{\rm{\Gamma }}})$ with

and ${\sigma }_{n},{v}_{{nl}},{u}_{{nl}}$ as in theorem 2.2. With the representation of the transformation ${{ \mathcal V }}_{{\rm{\Gamma }}},$ (9), we obtain

and

for $s\in {T}_{{\boldsymbol{\theta }}}^{-1}([-1,1]).$ Since ${{ \mathcal R }}_{{\rm{\Gamma }}}f({\boldsymbol{\theta }},s)=0$ for $s\notin {T}_{{\boldsymbol{\theta }}}^{-1}([-1,1]),$ we obtain the stated singular value decomposition of ${{ \mathcal R }}_{{\rm{\Gamma }}}\ :{L}_{2}({V}_{1}(0))\to {L}_{2}({S}^{1}\times {\mathbb{R}},{w}_{{\rm{\Gamma }}}).$

Please wait… references are loading.
10.1088/0266-5611/32/2/025006