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

thanks: Address: 408 W. Simpson St., Tucson, AZ, 85701. Author to whom correspondence should be addressed.

Running title: Invertibility of ME X-ray transform

Invertibility of Multi-Energy X-ray Transform

Yijun Ding dingy@email.arizona.edu Wyant College of Optical Sciences, University of Arizona, Tucson, AZ, United States, 85719    Eric W. Clarkson clarkson@radiology.arizona.edu Department of Medical Imaging and Wyant College of Optical Sciences, University of Arizona, Tucson, AZ, United States, 85719    Amit Ashok ashoka@optics.arizona.edu Wyant College of Optical Sciences and Department of Electrical and Computer Engineering, Tucson, AZ, United States, 85719
Abstract

Purpose: The goal is to provide a sufficient condition for the invertibility of a multi-energy (ME) X-ray transform. The energy-dependent X-ray attenuation profiles can be represented by a set of coefficients using the Alvarez-Macovski (AM) method. An ME X-ray transform is a mapping from NN AM coefficients to NN noise-free energy-weighted measurements, where N2N\geq 2.

Methods: We apply a general invertibility theorem to prove the equivalence of global and local invertibility for an ME X-ray transform. We explore the global invertibility through testing whether the Jacobian of the mapping J(𝑨)J(\bm{A}) has zero values over the support of the mapping. The Jacobian of an arbitrary ME X-ray transform is an integration over all spectral measurements. A sufficient condition for J(𝑨)0J(\bm{A})\neq 0 for all 𝑨\bm{A} is that the integrand of J(𝑨)J(\bm{A}) is 0\geq 0 (or 0\leq 0) everywhere. Note that the trivial case of the integrand equals 0 everywhere is ignored. Using symmetry, we simplified the integrand of the Jacobian to three factors that are determined by the total attenuation, the basis functions, and the energy-weighting functions, respectively. The factor related to the total attenuation is always positive, hence the invertibility of the X-ray transform can be determined by testing the signs of the other two factors. Furthermore, we use the Cramér-Rao lower bound (CRLB) to characterize the noise-induced estimation uncertainty and provide a maximum-likelihood (ML) estimator.

Results: The factor related to the basis functions is always negative when the photoelectric/Compton/Rayleigh basis functions are used and K-edge materials are not considered. The sign of the energy-weighting factor depends on the system source spectra and the detector response functions. For four special types of X-ray detectors, the sign of this factor stays the same over the integration range. Therefore, when these four types of detectors are used for imaging non-K-edge materials, the ME X-ray transform is globally invertible. The same framework can be used to study an arbitrary ME X-ray imaging system, e,g, when K-edge materials are present. Furthermore, the ML estimator we presented is an unbiased, efficient estimator and can be used for a wide range of scenes.

Conclusions: We have provided a framework to study the invertibility of an arbitrary ME X-ray transform and proved the global invertibility for four types of systems.

invertibility, X-ray, multi-energy X-ray imaging, spectral X-ray imaging

I Introduction

Muti-energy (ME) X-ray imaging, also referred to as spectral or energy-selective X-ray imaging, has long been used to image the chemical composition of the object being scanned [1, 2, 3, 4, 5, 6]. In X-ray imaging, the chemical composition of a material is characterized by the energy dependence of the X-ray attenuation profile. As an X-ray attenuation profile can be represented as a linear combination of basis functions with known energy dependences, it can be summarized by a few energy-independent coefficients as in the Alvarez-Macovski method [1]. We refer to these coefficients as AM coefficients. Imaging AM coefficients requires multiple energy-weighted measurements, e.g. energy integration with varying source tube voltages or photo-counting with multiple energy bins. We refer to the mapping from the AM coefficients to the energy-weighted measurements an X-ray transform. The question whether an X-ray transform is invertible has only been explored recently for dual-energy (DE) [7, 8] and ME measurements [9]. The purpose of this work is to provide a sufficient condition for the invertibility of a general ME X-ray transform from a different perspective.

With the recent developments in detectors, ME X-ray imaging is becoming more tangible. DE X-ray imaging recovers two AM coefficients [1] that represent contributions from photoelectric absorption and Compton scattering to the linear attenuation profile, respectively. The contribution from Rayleigh scattering has been considered negligible or assumed to be captured by the other two AM coefficients in DE X-ray imaging [10, 11, 12]. However, it is difficult to predict the effect caused by ignoring the Rayleigh scattering term due to the nonlinear nature of the X-ray transform, especially for security- and industrial-screening applications where the materials of interest are not necessarily low-Z materials. With ME detectors, the AM coefficient corresponding to the Rayleigh scattering can be recovered. Furthermore, ME X-ray imaging systems can image materials containing K-edges in the spectral range used for imaging [5].

With broad-spectrum X-ray sources, measurements of many X-ray systems are naturally energy-weighted [13]. ME measurements can be acquired with varying source settings [14, 15] or with detectors with varying energy responses, such as sandwich detectors [16], counting and integrating X-ray (CIX) detectors [17] and multi-bin photon-counting detectors [18]. More specifically, the recent advancement in photon counting detectors with pulse-height analysis, which output signals in multiple energy levels, provides a paradigm shift in X-ray detector technology and is enabling many new applications [19, 20].

The invertibility of a transform is a fundamental question in inverse problems. The invertibility problem considers noise-free measurements and determines whether a unique solution exists. A system of NN linear equations of NN unknowns has a unique solution (as long as the forward matrix is invertible); this is not necessarily true for nonlinear transforms. Levine et. al. [7] demonstrated a case of DE X-ray imaging with non-unique solutions. Alvarez et. al. [8] has applied a two-dimensional global inverse theorem [21] to DE X-ray transforms. Bal et. al. [9] provided an invertibility criteria for an ME X-ray transform by placing strong orientation constraints on the Jacobian matrices and demonstrated the equivalence of global and local invertibility for some examples through numerical experiments. We apply a global inverse function theorem for a N-dimensional map and prove that, for an ME X-ray transform, local invertibility is equivalent to global invertibility. Our global invertibility criteria is local invertibility, which is a weaker sufficient condition than the criteria provided by Bal et. al. [9]. This is proved through topological properties of the definition region and inequalities of the X-ray transform. Furthermore, we provide a sufficient condition for the global invertibility by taking advantage of the symmetries in the expression of the Jacobian. With its simple expression, this condition can be applied to the design of ME X-ray imaging systems and detectors.

In this paper, we provide a framework to study the invertibility of an arbitrary ME X-ray transform and prove the invertibility for four special cases of energy-weighted detectors. Furthermore, we consider Poisson noise in the measurement data and present the Cramér-Rao lower bound (CRLB) on the estimation of AM coefficients. Lastly, we provide a fast maximum-likelihood (ML) algorithm for coefficients estimation and demonstrate its application in an X-ray reconstruction problem.

II Forward problem: ME X-ray transform

In the energy range 20—200 keV, which is commonly used for X-ray transmission imaging, the interaction between X-ray photons and the medium can be categorized into the following three processes: photoelectric absorption, Compton (incoherent) scattering and Rayleigh (coherent ) scattering. Correspondingly, the X-ray linear attenuation coefficient profiles can be represented accurately by a summation of NN terms as:

μ(E)=iNaifi(E)=𝒂𝒇(E),\mu(E)=\sum_{i}^{N}a_{i}f_{i}(E)={\bm{a}}\cdot{\bm{f}}(E), (1)

where each component of 𝒇{\bm{f}} is a function of energy EE, the coefficients 𝐚\mathbf{a} are determined by the material composition, and the NN terms include photo electric, Compton scattering, Rayleigh scattering and K-edges. Here ‘photo electric’ refers to the smooth energy dependence of the photo electric effect and ‘K-edges’ refers to the discontinuities in the energy dependence of the photo electric effect just above the binding energy of the K-shell electrons. We use this set of fi(E)f_{i}(E) functions as basis functions and the coefficients 𝒂\bm{a} as the AM coefficients.

For materials that do not contain K-edges in the energy range of interest, the number of basis functions needed is N=3N=3. Approximated expressions of photo electric and Rayleigh scattering term have been provided in Williamson et al. [22] by fitting to DLC-146 cross-section data [23] and the Klein-Nishina function [1] describes the Compton scattering term:

{f1(E)=c1E3.088,f2(E)=c2(1+αα2[2(1+α)1+2α1αln(1+2α)]+12αln(1+2α)1+3α(1+2α)2),f3(E)=c3E1.737,\begin{cases}\begin{split}f_{1}(E)&=c_{1}E^{-3.088},\\ f_{2}(E)&=c_{2}\left(\frac{1+\alpha}{\alpha^{2}}\left[\frac{2(1+\alpha)}{1+2\alpha}-\frac{1}{\alpha}\ln(1+2\alpha)\right]+\frac{1}{2\alpha}\ln(1+2\alpha)-\frac{1+3\alpha}{{(1+2\alpha)}^{2}}\right),\\ f_{3}(E)&=c_{3}E^{-1.737},\\ \end{split}\end{cases} (2)

where α=E/(510.975keV)\alpha=E/(510.975~{}keV), the subscript 1,2,3 refers to the photoelectric effect, the Compton scattering and the Rayleigh scattering, respectively, and cic_{i} are normalization factors so that fi(E)2=1\lVert f_{i}(E)\rVert_{2}=1. The normalized basis functions are presented in Figure 1 (left). The usefulness of these functions in representing attenuation coefficient profiles is well known. We generated attenuation profiles for 128 materials based on the NIST XCOM data [24]. As an example, the fitted attenuation profile and the XCOM data for water are presented in Figure 1 (right).

In a tomographic imaging or measurement system, the total attenuation τ(E)\tau(E) is the line integral of the X-ray attenuation coefficient μ(E)\mu(E) along the ray path

τ(E)=dlμ(E)=i=1NAifi(E)=𝑨𝒇(E),\tau(E)=\int\mathrm{dl}~{}\mu(E)=\sum_{i=1}^{N}A_{i}f_{i}(E)={\bm{A}}\cdot{\bm{f}}(E), (3)

where

Ai=dlai(𝑹)A_{i}=\int\mathrm{dl}~{}a_{i}(\bm{R}) (4)

is a sinogram of the ithi^{th} AM coefficient. For a parallel-beam system, 𝑨i(θ,ρ)\bm{A}_{i}(\theta,\rho) is the Radon transform of 𝒂i(𝑹)\bm{a}_{i}(\bm{R}), where θ\theta is the rotation angle of the ray path and ρ\rho is the position along the detector plane.

The object ai(𝑹)a_{i}(\bm{R}) can be reconstructed from the sinograms 𝑨i(θ,ρ)\bm{A}_{i}(\theta,\rho), and the line integrals 𝑨i(θ,ρ)\bm{A}_{i}(\theta,\rho) can be estimated from ME measurements of the corresponding ray path. Consider a ME X-ray imaging system producing MM energy-weighted measurements with a source photon budget I0I_{0} (total number of photons emitted by the source across the energy range of interest). To describe the mthm^{th} energy-weighted measurement, where m=1,,Mm=1,...,M, denote Dm(E)D_{m}(E) as the detector response and Sm(E)S_{m}(E) as the normalized source spectrum of the mthm^{th} measurement. For a given ray path, the mean signal of the mthm^{th} measurement can be described by

Im=I00dEDm(E)Sm(E)exp[𝑨𝒇(E)],=I00dEpm(E)exp[𝑨𝒇(E)],\begin{split}I_{m}&=I_{0}\int_{0}^{\infty}\mathrm{dE}~{}D_{m}(E)\,S_{m}(E)\,\exp\left[-{\bm{A}}\cdot{\bm{f}}(E)\right],\\ &=I_{0}\int_{0}^{\infty}\mathrm{dE}~{}p_{m}(E)\,\exp\left[-{\bm{A}}\cdot{\bm{f}}(E)\right],\\ \end{split} (5)

where pm(E)=Dm(E)Sm(E)p_{m}(E)=D_{m}(E)S_{m}(E) is the combined energy-weighting function. This equation can be used to describe many energy-weighted measurements, such as a photon-counting (PC) binning detector and an energy-integrating detector. In the most general case, the source spectra may vary across measurements and the combined weighting functions are arbitrary and can take on any real non-negative values at each energy. The basis functions 𝒇(E)\bm{f}(E) can contain components describing K-edges as well. Therefore, Equation (LABEL:eq:mean_mapping) describes a general ME X-ray transform. In the following sections, we study the invertibility of the mapping defined by this equation in the domain Ai0A_{i}\geq 0 for i=1,2,,Ni=1,2,...,N.

A special case of a ME detector is a CIX detector that counts the number of photons and integrates both the energy and the momentum of the photons (PC/EI/MI), providing measurements with detector response D1(E)=1D_{1}(E)=1, D2(E)ED_{2}(E)\propto E and D3(E)ED_{3}(E)\propto\sqrt{E} as shown in Figure 2(a). As a CIX PC/EI detector has been developed [17], it is reasonable to assume that it is feasible to build a CIX PC/EI/MI detector. A second special case is a binning detector where the weighting functions are arbitrary and non-overlapping as shown in Figure 2(b). Another special case is an ideal PC detector as illustrated in Figure 2(c), where the detector response of each bin can be considered as 𝑟𝑒𝑐𝑡\it rect functions and there may be overlaps between different bins. Binning detectors in real life tend to have non-overlapping bins. Here, for the comprehensiveness, we include detectors with overlapping bins. Another special case considers a slightly overlapping three-bin detector, where the overlap is introduced by the finite energy resolution of the detector. The detector response functions of such a detector are plotted in Figure 2(d).

III Invertibility

We explore the invertibility of the mapping from the AM coefficients 𝑨{\bm{A}} to the noise-free ME measurement data 𝑰{\bm{I}}. Suppose we have M=NM=N measurements. The coefficients 𝑨{\bm{A}} and the mean photon count 𝑰{\bm{I}} are both subsets of N-dimensional Euclidean space 𝑹N{\bm{R}}^{N}. We define the ME X-ray transform from 𝑨{\bm{A}} to 𝑰{\bm{I}} as 𝒳:M1M2\mathcal{X}:M_{1}\rightarrow M_{2}, where the domain of the mapping is 𝑨{\bm{A}} in M1M_{1} and the range of the mapping is 𝑰{\bm{I}} in M2M_{2}.

The Hadamard’s global inverse function theorem[25]: Let M1M_{1} and M2M_{2} be smooth, connected N-dimensional manifolds and let 𝒳:M1M2\mathcal{X}:M_{1}\rightarrow M_{2} be a C1C^{1} function. If (1) 𝒳\mathcal{X} is proper, (2) the Jacobian of 𝒳\mathcal{X} vanishes nowhere and (3) M2M_{2} is simply connected, then 𝒳\mathcal{X} is a homeomorphism. A homeomorphism is one-to-one and onto, which implies global invertibility, while non-vanishing Jacobian implies local invertibility.

In the following sections, we will use the Hadamard’s global inverse function theorem to prove the equivalence of global and local invertibility for an ME X-ray transform. We will first construct a simply-connected range M2M_{2}, then prove the mapping 𝒳:M1M2\mathcal{X}:M_{1}\rightarrow M_{2} is proper through inequality relations. Lastly, we will provide a simplified expression for the Jacobian determinant and a sufficient condition for the Jacobian to vanish nowhere.

III.1 Simply connected

We briefly summarize the property of the mapping 𝒳\mathcal{X}. The first-order derivative of the X-ray transform can be expressed as follows:

ImAi=I00dEpm(E)fi(E)exp[𝑨𝒇(E)].\frac{\partial I_{m}}{\partial A_{i}}=-I_{0}\int_{0}^{\infty}\mathrm{dE}~{}p_{m}(E)\,f_{i}(E)\exp\left[-{\bm{A}}\cdot{\bm{f}}(E)\right]. (6)

The first-order derivative exists and is continuous, therefore the mapping 𝒳\mathcal{X} is a C1C^{1} mapping. The values 𝒳(0)\mathcal{X}(0), which represents the mean signals of an air scan, are finite and equal to the maximum (mean) count values. We further define a normalization factor dm=0dEpm(E)d_{m}=\int_{0}^{\infty}\mathrm{dE}~{}{p_{m}\left(E\right)}. With this definition, the maximum mean count measured by the mthm^{th} detector is I0dmI_{0}d_{m}. As the magnitude of 𝑨\bm{A} approaches infinity, the counts approach zeros. We will use these properties and the assumption that the Jacobian is non-vanishing in 𝑹N\bm{R}^{N} to construct a simply connected range M2M_{2} with the corresponding domain M1M_{1} connected. Furthermore, we will justify that the interior of M1M_{1} and M2M_{2} are both smooth N-dimensional manifolds.

The ME X-ray transform, as defined in Equation (LABEL:eq:mean_mapping), has physical meaning when 𝑨\bm{A} is in the positive subspace of 𝑹N{\bm{R}}^{N}, denoted as 𝑷N{\bm{P}}^{N}, where Ai0A_{i}\geq 0 for all ii. However, the transform is mathematically valid over the domain 𝑹N{\bm{R}}^{N}. In order to construct a simply connected M2M_{2}, we will expand the domain of the mapping from 𝑷N{\bm{P}}^{N}.

Let U0=𝑷NU_{0}={\bm{P}}^{N} and V0V_{0} be the image of U0U_{0} under the mapping 𝒳\mathcal{X}. In V0V_{0}, the mean photon count 𝑰\bm{I} is bounded by 0<ImI0dm0<I_{m}\leq I_{0}d_{m}. Furthermore, as U0U_{0} is path connected and 𝒳\mathcal{X} is a continuous mapping, V0V_{0} is path connected [26](page 150). From every point 𝑰i\bm{I}_{i} in V0V_{0}, draw a straight line to the maximum-count point I0𝒅I_{0}\bm{d} and define this line with end points as ViV_{i}. Every point in ViV_{i} is bounded by IimImI0dmI_{im}\leq I_{m}\leq I_{0}d_{m}. Define M2M_{2} as the union of all ViV_{i}. As V0V_{0} and ViV_{i} are all path connected, M2M_{2} is path connected. The space M2M_{2} is simply connected if every closed curve in M2M_{2} can be contracted to a point [25]. Define a closed curve ϕ(s):[0,1]M2\phi(s):[0,1]\rightarrow M_{2}. We can contract ϕ(s)\phi(s) to the maximum-count point I0𝒅I_{0}\bm{d} through the following continuous function HH: [0,1]×[0,1]M2[0,1]\times[0,1]\rightarrow M_{2},

H(s,t)=tϕ(s)+(1t)I0𝒅.H(s,t)=t\phi(s)+(1-t)I_{0}\bm{d}. (7)

As M2M_{2} is path connected, the closed curve ϕ(s)\phi(s) can be contracted to any points in M2M_{2} [26] (page 332). Therefore, M2M_{2} is simply connected.

We assume that the Jacobian vanishes nowhere in 𝑹N{\bm{R}}^{N}. In other words, the mapping 𝒳\mathcal{X} is a local homeomorphism, which means that every point of 𝑨𝑹N\bm{A}\in{\bm{R}}^{N} has a neighborhood that is homeomorphic to an open subset in the range. For every straight line Vi(𝑰)V_{i}(\bm{I}), the corresponding preimage ui(𝑨)u_{i}(\bm{A}) in the domain 𝑹N{\bm{R}}^{N} can be constructed by successive local inverses 𝒳1(𝑰)\mathcal{X}^{-1}(\bm{I}). The maximum-count point I0𝒅I_{0}\bm{d} corresponds to only one point in U0U_{0}, and this point is the origin of the coefficient space. Therefore, the origin is one end point of all preimages. The point 𝑰i\bm{I}_{i} may have multiple local inverses, which we can index with subscript jj. The jthj^{th} local inverse introduces an inverse curve uiju_{ij}, where the jthj^{th} local inverse is the second end point of the corresponding preimage uiju_{ij}. Each uiju_{ij} is connected and connected to U0U_{0}. We define the union of all uij(𝑨)u_{ij}(\bm{A}) as UiU_{i}. Furthermore, we define M1M_{1} as the union of all UiU_{i}. M1M_{1} is connected and a superset of 𝑷N{\bm{P}}^{N}. The expansion of the range and domain for N=2N=2 is illustrated in Figure 3.

The interior points of M1M_{1} and M2M_{2} (excluding the boundaries) are both smooth N-dimensional manifolds, because they are open subsets of 𝑹N{\bm{R}}^{N} [27] (page 19). We will limit our proof to the interior of M1M_{1} and M2M_{2} and discuss the boundary points in Section VI.

III.2 Proper

We derive the bounds on the coefficients 𝑨{\bm{A}} for given measurement data 𝑰{\bm{I}}. Using Jensen’s inequality, we have

lnImdmI0>0dEpm(E)dm[𝑨𝒇(E)],\ln\frac{I_{m}}{d_{m}I_{0}}>\int_{0}^{\infty}\mathrm{dE}~{}\frac{p_{m}\left(E\right)}{d_{m}}\left[-{\bm{A}}\cdot{\bm{f}}\left(E\right)\right], (8)

where dmd_{m} has been defined previously and dmI0d_{m}I_{0} is the maximum count in the mthm^{th} measurement. These inequalities can be written as

𝑨𝒏m>ln[(dmI0)/Im]{\bm{A}}\cdot{\bm{n}}_{m}>\ln[(d_{m}I_{0})/I_{m}] (9)

where

𝒏m=0dEpm(E)𝒇(E)/dm{\bm{n}}_{m}=\int_{0}^{\infty}\mathrm{dE}~{}p_{m}\left(E\right){\bm{f}}\left(E\right)/d_{m} (10)

The vector 𝒏m{\bm{n}}_{m} has all non-negative components. Furthermore, the mean photon count ImI_{m} in M2M_{2}, is always less than or equal to the maximum count, dmI0d_{m}I_{0}. Therefore, the right hand side of Equation (9) is always larger than or equal to 0. Each of the inequalities in Equation (9) forces the vector 𝑨{\bm{A}} to be on the side that is opposite to the origin of the hyperplane defined by to 𝑨𝒏m=ln[Im/(dmI0)]{\bm{A}}\cdot{\bm{n}}_{m}=-\ln[I_{m}/(d_{m}I_{0})], as shown in Figure 4(a) for the case of N=2N=2.

We define the support of the weighting functions pm(E)p_{m}(E) as Ωm\Omega_{m}. Using the Schwarz inequality we have

Im2I02[ΩmdEpm2(E)][ΩmdEexp[2𝑨𝒇(E)]],I_{m}^{2}\leq I_{0}^{2}\left[\int_{\Omega_{m}}\mathrm{dE}~{}p_{m}^{2}\left(E\right)\right]\left[\int_{\Omega_{m}}\mathrm{dE}~{}\exp\left[-2{\bm{A}}\cdot{\bm{f}}\left(E\right)\right]\right], (11)

with equality if and only if pm(E)exp[𝑨𝒇(E)]p_{m}(E)\propto\exp\left[-{\bm{A}}\cdot{\bm{f}}\left(E\right)\right]. In many occasions, the equality condition is not attainable. For example, when the three basis functions given in Equation (2) are used, exp[𝑨𝒇(E)]\exp\left[-{\bm{A}}\cdot{\bm{f}}\left(E\right)\right] is not proportional to the pm(E)p_{m}(E) of the detectors illustrated in FIG. 2(b)-(d). If we define

γm=Im2I02[ΩmdEpm2(E)]1,\gamma_{m}=\frac{I_{m}^{2}}{I_{0}^{2}}\left[\int_{\Omega_{m}}~{}\mathrm{dE}~{}p_{m}^{2}\left(E\right)\right]^{-1}, (12)

then we have

γmΩmdEexp[2𝑨𝒇(E)].\gamma_{m}\leq\int_{\Omega_{m}}\mathrm{dE}~{}\exp\left[-2{\bm{A}}\cdot{\bm{f}}\left(E\right)\right]. (13)

Assume that the length |Ωm|\left|\Omega_{m}\right| of each support set is finite. Replacing the integrand with its maximum possible value gives

γmexp{2minEΩm[𝑨𝒇(E)]}|Ωm|\gamma_{m}\leq\exp\left\{-2\min_{E\in\Omega_{m}}\left[{\bm{A}}\cdot{\bm{f}}\left(E\right)\right]\right\}\left|\Omega_{m}\right| (14)

Therefore we have another set of inequalities

minEΩm[𝑨𝒇(E)]12ln(|Ωm|γm)\min_{E\in\Omega_{m}}\left[{\bm{A}}\cdot{\bm{f}}\left(E\right)\right]\leq\frac{1}{2}\ln\left(\frac{\left|\Omega_{m}\right|}{\gamma_{m}}\right) (15)

Now we may choose an energy EmE_{m} such that

𝑨𝒇(Em)ln(|Ωm|γm){\bm{A}}\cdot{\bm{f}}\left(E_{m}\right)\leq\ln\left(\sqrt{\frac{\left|\Omega_{m}\right|}{\gamma_{m}}}\right) (16)

The right-hand side satisfies ln(|Ωm|/γm)ln(dmI0/Im)0\ln\left(\sqrt{\left|\Omega_{m}\right|/\gamma_{m}}\right)\geq\ln(d_{m}I_{0}/I_{m})\geq 0 through the Schwarz inequality as follows:

dm2=[ΩmdEpm(E)]2|Ωm|[ΩmdEpm2(E)].d_{m}^{2}=\left[\int_{\Omega_{m}}\mathrm{dE}~{}p_{m}\left(E\right)\right]^{2}\leq|\Omega_{m}|\left[\int_{\Omega_{m}}\mathrm{dE}~{}p_{m}^{2}\left(E\right)\right]. (17)

Each of the inequalities in Equation (16) forces the vector 𝑨{\bm{A}} to be on the same side of the corresponding hyperplane as the origin.

Therefore, for given mean photon count 𝑰{\bm{I}}, where dmI0Im>0d_{m}I_{0}\geq I_{m}>0, the inequalities defined by Equation (9) and (16) force 𝑨{\bm{A}} to be in a bounded set defined by the first set of hyperplanes and the second set of hyperplanes. A typical picture of this scenario for N=2N=2 is shown in Figure 4(a). Note that for a physical measurement, the corresponding coefficients 𝑨\bm{A} are further bounded by the coordinate planes. Here we focus on demonstrating that 𝑨\bm{A} is bounded for a given 𝑰\bm{I} even without the positivity constraints on 𝑨\bm{A}.

Now we show that the mapping 𝒳:M1M2\mathcal{X}:M_{1}\rightarrow M_{2} is a proper mapping. If we have a compact set CC in the data space M2M_{2}, where all of the data vectors are located, then there are maximum and minimum values for each ImI_{m} over all 𝑰{\bm{I}} in CC. The maximum value for ImI_{m} determines the hyperplane 𝑨𝒏m=ln[(dmI0)/Im]{\bm{A}}\cdot{\bm{n}}_{m}=\ln[(d_{m}I_{0})/I_{m}] that is closet to the origin. The minimum value for ImI_{m} determines the hyperplane 𝑨𝒇(Em)=ln(|Ωm|/γm){\bm{A}}\cdot{\bm{f}}\left(E_{m}\right)=\ln\left(\sqrt{\left|\Omega_{m}\right|/\gamma_{m}}\right) that is furthest away from the origin. Therefore, the set of 𝑨{\bm{A}} that are mapped into CC are contained in a region bounded by these two sets of hyperplanes. This bounded region together with its boundary form a closed and bounded set in N\mathbb{R}^{N}, hence a compact set. As the map 𝒳(𝑨){\mathcal{X}}\left({\bm{A}}\right) is continuous, the set of 𝑨{\bm{A}} that are mapped into the closed set CC is closed. The set of 𝑨{\bm{A}} that are mapped into CC is a closed subset of a compact set. This set is therefore also compact. As a result, the mapping 𝒳(𝑨){\mathcal{X}}({\bm{A}}) is proper.

III.3 Jacobian

The Jacobian of the mapping is J(𝑨)=|det(𝑨𝑰)|J({\bm{A}})=|\det(\nabla_{{\bm{A}}}{\bm{I}})|, where |||\cdot| represents the absolute value and det()\det(\cdot) is the determinant of a matrix. The matrix inside the determinant is

𝑨𝑰=[I1A1I1A2I1ANI2A1I2A2I2ANIMA1IMA2IMAN].\nabla_{{\bm{A}}}{\bm{I}}=\begin{bmatrix}\frac{\partial I_{1}}{\partial A_{1}}&\frac{\partial I_{1}}{\partial A_{2}}&...&\frac{\partial I_{1}}{\partial A_{N}}\\ \frac{\partial I_{2}}{\partial A_{1}}&\frac{\partial I_{2}}{\partial A_{2}}&...&\frac{\partial I_{2}}{\partial A_{N}}\\ ...&...&...&...\\ \frac{\partial I_{M}}{\partial A_{1}}&\frac{\partial I_{M}}{\partial A_{2}}&...&\frac{\partial I_{M}}{\partial A_{N}}\end{bmatrix}. (18)

As M=NM=N, 𝑨𝑰\nabla_{{\bm{A}}}{\bm{I}} is a square matrix. Defining the set Ω=Ω1×Ω2×Ωm\Omega=\Omega_{1}\times\Omega_{2}...\times\Omega_{m}, the determinant of a square matrix can be expressed in Leibniz formula as

det(𝑨𝑰)=𝝈sgn(𝝈)m=1MImAσm,=(I0)M𝝈sgn(𝝈)m=1MΩmdEmpm(Em)fσm(Em)exp[𝑨𝒇(Em)]=(I0)MΩdM𝐄{m=1Mpm(Em)e𝑨𝒇(Em)}𝝈sgn(𝝈)m=1Mfσm(Em),\begin{split}\det(\nabla_{\bm{A}}{\bm{I}})&=\sum_{\bm{\sigma}}\text{sgn}(\bm{\sigma})\prod_{m=1}^{M}\frac{\partial I_{m}}{\partial A_{\sigma_{m}}},\\ &=(-I_{0})^{M}\sum_{\bm{\sigma}}\text{sgn}(\bm{\sigma})\prod_{m=1}^{M}\int_{\Omega_{m}}\mathrm{dE_{m}}~{}p_{m}(E_{m})\,f_{\sigma_{m}}(E_{m})\exp\left[-{\bm{A}}\cdot{\bm{f}}(E_{m})\right]\\ &=(-I_{0})^{M}\int_{\Omega}\mathrm{d^{M}{\bf E}}~{}\left\{\prod_{m=1}^{M}p_{m}(E_{m})e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}\right\}\sum_{\bm{\sigma}}\text{sgn}(\bm{\sigma})\prod_{m=1}^{M}f_{\sigma_{m}}(E_{m}),\\ \end{split} (19)

where the sum is computed over all permutations 𝝈\bm{\sigma} of the set {1,2,,M}\{1,2,...,M\}, and the sign of the permutation 𝝈\bm{\sigma}, sgn(𝝈)\text{sgn}(\bm{\sigma}), is +1 or -1 for even or odd permutations, respectively. Invoking the Leibniz formula again, we can simplify the Jacobian to:

det(𝑨𝑰)=(I0)MΩdM𝐄{m=1Mpm(Em)e𝑨𝒇(Em)}det[F(𝑬)],=(I0)MΩdM𝐄{m=1Mfm(Em)e𝑨𝒇(Em)}det[P(𝑬)],\begin{split}\det(\nabla_{\bm{A}}{\bm{I}})&=(-I_{0})^{M}\int_{\Omega}\mathrm{d^{M}{\bf E}}~{}\left\{\prod_{m=1}^{M}p_{m}(E_{m})e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}\right\}\det[F({\bm{E}})],\\ &=(-I_{0})^{M}\int_{\Omega}\mathrm{d^{M}{\bf E}}~{}\left\{\prod_{m=1}^{M}f_{m}(E_{m})e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}\right\}\det[P({\bm{E}})],\\ \end{split} (20)

where the second line follows a similar derivation as Equation (19) by swapping the subscripts of p(E)p(E) and f(E)f(E). The matrix FF as a function of 𝑬=(E1,E2,,EM){\bm{E}}=(E_{1},E_{2},...,E_{M}) is

F(𝑬)=[f1(E1)f2(E1)fN(E1)f1(E2)f2(E2)fN(E2)f1(EM)f2(EM)fN(EM)],F({\bm{E}})=\begin{bmatrix}f_{1}(E_{1})&f_{2}(E_{1})&...&f_{N}(E_{1})\\ f_{1}(E_{2})&f_{2}(E_{2})&...&f_{N}(E_{2})\\ ...&...&...&...\\ f_{1}(E_{M})&f_{2}(E_{M})&...&f_{N}(E_{M})\end{bmatrix}, (21)

and the matrix PP as a function of 𝑬{\bm{E}} is

P(𝑬)=[p1(E1)p2(E1)pN(E1)p1(E2)p2(E2)pN(E2)p1(EM)p2(EM)pN(EM)].P({\bm{E}})=\begin{bmatrix}p_{1}(E_{1})&p_{2}(E_{1})&...&p_{N}(E_{1})\\ p_{1}(E_{2})&p_{2}(E_{2})&...&p_{N}(E_{2})\\ ...&...&...&...\\ p_{1}(E_{M})&p_{2}(E_{M})&...&p_{N}(E_{M})\end{bmatrix}. (22)

The integrand in Equation (20) has several interesting symmetry properties. The factor m=1Me𝑨𝒇(Em)=e𝑨m=1M𝒇(Em)\prod_{m=1}^{M}e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}=e^{-{\bm{A}}\cdot{\sum_{m=1}^{M}\bm{f}}(E_{m})} has mirror symmetry about all hyperplanes Ei=EjE_{i}=E_{j} for i,j{1,2,..,M}i,j\in\{1,2,..,M\}. The other factor, det[F(𝑬)]\det[F(\bm{E})], has sign-switching mirror symmetry about the same hyperplanes, which can be described mathematically as:

det[F(E1,E2,,EM)]=sgn(𝝈)det[F(Eσ1,Eσ2,,EσM)].\det[F({E_{1},E_{2},...,E_{M}})]=\text{sgn}(\bm{\sigma})\det[F(E_{\sigma_{1}},E_{\sigma_{2}},...,E_{\sigma_{M}})]. (23)

A sign-switching mirror symmetry means that, when we switch the positions of two coordinates (odd permutation), the sign of the function changes but the absolute value of the function is preserved. For example, with two coordinates E1E_{1} and E2E_{2}, det[F(E1,E2)]=f1(E1)f2(E2)f2(E1)f1(E2)=det[F(E2,E1)]\det[F({E_{1},E_{2}})]=f_{1}(E_{1})f_{2}(E_{2})-f_{2}(E_{1})f_{1}(E_{2})=-\det[F({E_{2},E_{1}})]. To illustrate the sign-switching mirror symmetry, we plotted det[F(E1,E2)]\det[F(E_{1},E_{2})] for the case when f1(E)f_{1}(E) and f2(E)f_{2}(E) are both Gaussian functions in Figure 4(b).

Now we can divide the space occupied by Ω\Omega into M!M! subspaces with hyperplanes Ei=EjE_{i}=E_{j} for i,j{1,2,..,M}i,j\in\{1,2,..,M\}. One of the subspace has property E1<E2<<EME_{1}<E_{2}<...<E_{M} and we define this subspace as Ω1,2M\Omega_{1,2...M}. For every point (E1,E2,,EM)(E_{1},E_{2},...,E_{M}) in the subspace Ω1,2M\Omega_{1,2...M}, there is a corresponding point (Eσ1,Eσ2,,EσM)(E_{\sigma_{1}},E_{\sigma_{2}},...,E_{\sigma_{M}}) in each of the remaining subspaces. Applying the sign-switching mirror symmetry of the determinant, we can further simplify the Jacobian to:

det(𝑨𝑰)=(I0)MΩ12..MdM𝐄{m=1Me𝑨𝒇(Em)}det[F(𝐄)]𝝈sgn(𝝈)m=1Mpm(Eσm),=(I0)MΩ12..MdM𝐄{m=1Me𝑨𝒇(Em)}det[F(𝑬)]det[P(𝑬)].\begin{split}\det(\nabla_{\bm{A}}{\bm{I}})&=(-I_{0})^{M}\int_{\Omega_{12..M}}\mathrm{d^{M}{\bf E}}~{}\left\{\prod_{m=1}^{M}e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}\right\}\det[F({\bf E})]\sum_{\bm{\sigma}}\text{sgn}(\bm{\sigma})\prod_{m=1}^{M}p_{m}(E_{\sigma_{m}}),\\ &=(-I_{0})^{M}\int_{\Omega_{12..M}}\mathrm{d^{M}{\bf E}}~{}\left\{\prod_{m=1}^{M}e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}\right\}\det[F({\bm{E}})]\det[P(\bm{E})].\\ \end{split} (24)

The symbol MM emphasizes that the integration is over all spectral measurements. The three factors of the integrand, m=1Me𝑨𝒇(Em)\prod_{m=1}^{M}e^{-{\bm{A}}\cdot{\bm{f}}(E_{m})}, det[F(𝑬)]\det[F({\bm{E}})] and det[P(𝑬)]\det[P(\bm{E})] depend on the total attenuation, the basis functions and the energy-weighting functions, respectively.

When the Jacobian is non-vanishing everywhere in 𝑹N{\bm{R}}^{N}, we can construct the domain M1M_{1} and the range M2M_{2} and prove that the mapping 𝒳:M1M2\mathcal{X}:M_{1}\rightarrow M_{2} is globally invertible. As a result, the ME X-ray transform defined on the domain PNP^{N}, which is a subset of M1M_{1}, is globally invertible. Therefore, we have proved the equivalence of global invertibility with local invertibility for an ME X-ray transform. This equivalence holds when K-edge basis functions are considered.

III.4 A sufficient condition for invertibility

A sufficient but not necessary condition for J(𝑨)0J(\bm{A})\neq 0 is that the integrand of J(𝑨)J(\bm{A}), which is given in Equation (24), has the same sign over the subspace S12NS_{12...N} and has non-zero values. The first factor in the integrand, which depends solely on the total attenuation, is always positive. If we ignore the trivial case that det[F(𝑬)]det[P(𝑬)]=0\det[F({\bm{E}})]\det[P(\bm{E})]=0 everywhere, a sufficient condition for the invertibility of the ME X-ray transform is det[F(E)]det[P(E)]0\det[F({\bm{E}})]\det[P(\bm{E})]\leq 0 (or 0\geq 0) for all E\bm{E} in S12NS_{12...N} . As the sign of the integrand does not depend on 𝑨\bm{A}, if the sufficient condition is satisfied, the Jacobian is non-vanishing for all 𝑨\bm{A} in 𝑹N{\bm{R}}^{N}.

When the photoelectric/Compton/Rayleigh basis functions are used, the basis-function determinant det[F(𝑬)]\det[F({\bm{E}})] is always negative in the subspace Ω123\Omega_{123}. This set of three basis functions is sufficient to describe an object when the materials of interest have no K-edges in the energy range used for imaging, e.g. soft tissue and bone. The proof of det[F(𝑬)]<0\det[F({\bm{E}})]<0 for all 𝑬\bm{E} that satisfy E1<E2<E3E_{1}<E_{2}<E_{3} is provided in Appendix A. When K-edge materials are considered, the values of det[F(𝑬)]\det[F({\bm{E}})] can be calculated numerically and the positive regions of det[F(𝑬)]\det[F({\bm{E}})] can be avoided by adjusting the detector sensitivity or source spectrum in pm(E)p_{m}(E).

Now we apply the sufficient condition for invertibility to the DE scenario that has non-unique solutions discussed by Levine [7]. For DE X-ray imaging, a set of two basis functions, photoelectric and Compton, can be used. The basis-function determinant is always positive in the subspace S12S_{12}. Levine assumed the same detector response for the two measurements, hence the sign of det[P(𝑬)]\det[P({\bm{E}})] is the same with the sign of det[S(𝑬)]\det[S({\bm{E}})], which is the source-spectra determinant. Similar to the definition in Equation (22), the matrix S(𝑬)S({\bm{E}}) can be written as

S(E1,E2)=[S1(E1)S2(E1)S1(E2)S2(E2)],S(E_{1},E_{2})=\begin{bmatrix}S_{1}(E_{1})&S_{2}(E_{1})\\ S_{1}(E_{2})&S_{2}(E_{2})\\ \end{bmatrix}, (25)

where (E1,E2)(E_{1},E_{2}) can be any combinations of two energies. With these, det[S(𝑬)]=S1(E1)S2(E2)S1(E2)S2(E1)\det[S({\bm{E}})]=S_{1}(E_{1})S_{2}(E_{2})-S_{1}(E_{2})S_{2}(E_{1}). Levine assumed that both source spectra S1(E)S_{1}(E) and S2(E)S_{2}(E) are not zero only at three energy points (30, 60, 100) keV. Hence, det[S(𝑬)]\det[S({\bm{E}})] is not zero only when (E1,E2)(E_{1},E_{2}) are combinations of the set {30,60,100}\{30,60,100\} keV. Within the subspace S12S_{12}, where E1E_{1} is always less than E2E_{2}, det[S(𝑬)]\det[S({\bm{E}})] is non-vanishing only at three points (E1,E2)=(30,60)(E_{1},E_{2})=(30,60), (30,100)(30,100), and (60,100)(60,100) keV. Given the two source spectra as S1(E)=(1,1,1)S_{1}(E)=(1,1,1) and S2(E)=(0.93,1.71,0.30)S_{2}(E)=(0.93,1.71,0.30) at E=(30,60,100)E=(30,60,100) keV, respectively. The values of det[S(𝑬)]\det[S({\bm{E}})] at (E1,E2)=(30,60)(E_{1},E_{2})=(30,60), (30,100)(30,100), and (60,100)(60,100) keV are 0.78, -0.63 and -1.41, respectively. Therefore, det[S(𝑬)]\det[S({\bm{E}})] does not have constant sign over the subspace S12NS_{12...N}, hence the invertibility of the DE X-ray transform for the proposed scenario is not guaranteed. Note that this analysis only shows that the existence of non-unique solutions is possible, it does not prove their existence.

We then consider the sign of det[P(𝑬)]\det[P({\bm{E}})] for the four types of detectors illustrated in Figure 2. Assuming the source spectrum S(E)S(E) is same for different mm, the weighting-function determinant det[P(𝑬)]\det[P({\bm{E}})] has the same sign with the detector-response determinant det[D(𝑬)]\det[D({\bm{E}})], where the (i,j)(i,j) element of matrix D(𝑬)D({\bm{E}}) is the detector response of the ithi^{th} measurement at EjE_{j} denoted as Di(Ej)D_{i}(E_{j}). The sign of det[D(𝑬)]\det[D({\bm{E}})] for the four types of detectors are studied in Appendix A and the main results are presented as follows.

  1. (a)

    An CIX-PC/EI/MI detector: det[D(𝑬)]<0\det[D({\bm{E}})]<0 for all 𝑬Ω123\bm{E}\in\Omega_{123}.

  2. (b)

    A three bin detector, where the three bins are not overlapping and the energy-response functions are arbitrary: det[D(𝑬)]0\det[D({\bm{E}})]\geq 0 for all 𝑬Ω123\bm{E}\in\Omega_{123}.

  3. (c)

    A three bin PC detector with rect-response functions and possible overlaps between bins: if Bin 1 and Bin 3 has no overlap, the lower edges of the three bins satisfy l1<l2<l3l_{1}<l_{2}<l_{3} and the upper edges of the three bins satisfy u1<u2,<u3u_{1}<u_{2},<u_{3}, the determinant det[D(𝑬)]0\det[D({\bm{E}})]\geq 0 for all 𝑬Ω123\bm{E}\in\Omega_{123}.

  4. (d)

    A non-overlapping three bin PC detector with finite energy resolution: if the energy resolution of the detector can be modeled by a narrow truncated function (for mathematical description see Appendix A) and there is no overlap between Bin 1 and Bin 3, det[D(𝑬)]0\det[D({\bm{E}})]\geq 0 in Ω123\Omega_{123}.

In conclusion, the mapping 𝒳:𝑨𝑰\mathcal{X}:{\bm{A}}\rightarrow{\bm{I}} is globally invertible for these four types of detectors when measuring attenuation profiles without K-edges. For arbitrary detectors or systems with varying source spectra, the values of det[P(𝑬)]\det[P({\bm{E}})] can be calculated numerically. One can alway maintain det[F(𝑬)]det[P(𝑬)]0\det[F({\bm{E}})]\det[P({\bm{E}})]\leq 0 for any EE in S12NS_{12...N} by adjusting the energy-response function or the bin boundaries of the detectors.

IV Estimation uncertainties for Poisson data

From a practical point of view, it is also crucial to consider the uncertainty in the estimation under the presence of noise. In this section, we consider PC detectors with non-overlapping bins. If only inherent quantum noise is considered, the data of the mthm^{th} measurement at a given ray path, gmg_{m}, is a Poisson random variable with mean equals to ImI_{m},

gm(𝑨)=𝒫oiss(Im(𝑨)),g_{m}(\bm{A})={\mathcal{P}oiss}(I_{m}(\bm{A})), (26)

where Im(𝑨)I_{m}(\bm{A}) is the mean photon count of the mthm^{th} measurement given in Equation (LABEL:eq:mean_mapping). Combining all MM measurements, we get the measurement data 𝒈{\bm{g}}. The probability density function of the data 𝒈{\bm{g}} given the AM coefficient 𝑨\bm{A} along the ray path is

pr(𝒈|𝑨)=m=1MIm(𝑨)gmeIm(𝑨)gm!\mathrm{pr}{({\bm{g}}|{\bm{A}})}=\prod_{m=1}^{M}\frac{I_{m}({\bm{A}})^{g_{m}}e^{-I_{m}({\bm{A}})}}{g_{m}!} (27)

The log-likelihood function of the AM coefficients 𝑨\bm{A} is

L(𝐀|𝒈)=lnpr(𝒈|𝑨)=m=1MgmlnIm(𝐀)Im(𝑨)lngm!.L({\bf A}|\bm{g})=\ln{\mathrm{pr}({{\bm{g}}|{\bm{A}}})}=\sum_{m=1}^{M}g_{m}\ln{I_{m}({\bf A}})-I_{m}(\bm{A})-\ln{g_{m}!}. (28)

The first derivative of the log-likelihood function is

LAi(𝑨)=m=1MgmIm(𝑨)Im(𝑨)ImAi(𝑨)=m=1MgmIm(𝑨)Im(𝑨)[𝑨𝑰]mi.\frac{\partial L}{\partial A_{i}}(\bm{A})=\sum_{m=1}^{M}\frac{g_{m}-I_{m}(\bm{A})}{I_{m}(\bm{A})}\frac{\partial I_{m}}{\partial A_{i}}(\bm{A})=\sum_{m=1}^{M}\frac{g_{m}-I_{m}(\bm{A})}{I_{m}(\bm{A})}[\nabla_{\bm{A}}{\bm{I}}]_{mi}. (29)

The Hessian, or second derivative, of the log-likelihood function is given by

[𝑨2L]ij=2LAiAj=m=1MgmImIm2ImAiAjgmIm2ImAiImAj.[\nabla^{2}_{\bm{A}}L]_{ij}=\frac{\partial^{2}L}{\partial A_{i}\partial A_{j}}=\sum_{m=1}^{M}\frac{g_{m}-I_{m}}{I_{m}}\frac{\partial^{2}I_{m}}{\partial A_{i}\partial A_{j}}-\frac{g_{m}}{I_{m}^{2}}\frac{\partial I_{m}}{\partial A_{i}}\frac{\partial I_{m}}{\partial A_{j}}. (30)

The components of the Fisher information matrix are

FIMij(𝑨)=2LAiAj𝒈|𝑨=dMgpr(𝒈|𝑨)m=1M[gmImIm2ImAiAjImAiImAjgmIm2]=m=1M1ImImAiImAj.\begin{split}\mathrm{FIM}_{ij}({\bm{A}})&=-\left<\frac{\partial^{2}L}{\partial A_{i}\partial A_{j}}\right>_{{\bm{g}}|{\bm{A}}}\\ &=-\int\mathrm{d^{M}g}~{}\mathrm{pr}({\bm{g}}|{\bm{A}})\sum_{m=1}^{M}\left[\frac{g_{m}-I_{m}}{I_{m}}\frac{\partial^{2}I_{m}}{\partial A_{i}\partial A_{j}}-\frac{\partial I_{m}}{\partial A_{i}}\frac{\partial I_{m}}{\partial A_{j}}\frac{g_{m}}{I_{m}^{2}}\right]\\ &=\sum_{m=1}^{M}\frac{1}{I_{m}}\frac{\partial I_{m}}{\partial A_{i}}\frac{\partial I_{m}}{\partial A_{j}}.\end{split} (31)

Therefore, the Fisher information matrix is

FIM(𝑨)=(𝑨𝑰)TΛ1(𝑨𝑰),\mathrm{FIM}({\bm{A}})=(\nabla_{\bm{A}}{\bm{I}})^{T}\Lambda^{-1}(\nabla_{\bm{A}}{\bm{I}}), (32)

where Λ\Lambda is a diagonal matrix with the mthm^{th} diagonal element equals to ImI_{m}.

The Crámer-Rao bounds [28, 29] characterize the limit on the estimation uncertainties induced by noise. It states that for an unbiased estimate of the ithi^{th} parameter, its variance must be at least as large as the ithi^{th} diagonal element of the inverse of the Fisher information matrix. Mathematically, the Crámer-Rao lower bounds are

Var(A^iAi)[FIM1]ii=[(𝑨𝑰)1Λ(𝑨𝑰)1,T]ii,\text{Var}(\hat{A}_{i}-A_{i})\geq[\mathrm{FIM}^{-1}]_{ii}=[(\nabla_{\bm{A}}{\bm{I}})^{-1}\Lambda(\nabla_{\bm{A}}{\bm{I}})^{-1,T}]_{ii}, (33)

where the symbol A^\hat{A} indicates an estimate of AA. Note that the uncertainty in the estimation is inversely related with the source photon budget I0I_{0}, which agrees with our intuition. Also note that the uncertainty of an unbiased estimation is inversely related with the Jacobian J(𝑨)=|det(𝑨𝑰)|J(\bm{A})=|\det(\nabla_{\bm{A}}{\bm{I}})|. If the Jacobian J(𝑨)J(\bm{A}) is close to zero, the estimation uncertainty is close to infinity and the coefficients can not be estimated accurately in practice.

V Estimation algorithm and illustrative results

In this section we develop a maximum-likelihood (ML) algorithm for Poisson data. The goal of the algorithm is to estimate AM-coefficients 𝑨\bm{A} from noisy data 𝒈\bm{g}. The assumption for the algorithm is that both Equations (LABEL:eq:mean_mapping) and  (27) are valid. First, consider LL as a function of the mean signal 𝑰\bm{I}. The first derivation of this function is

LIi=giIi1.\frac{\partial L}{\partial I_{i}}=\frac{g_{i}}{I_{i}}-1. (34)

Hence, the point 𝑰=𝒈\bm{I}=\bm{g} is a critical point. The Hessian of the function L(𝑰|𝒈)L(\bm{I}|\bm{g}) is

[𝑰2L]ij=2LIiIj=giIi2σij[\nabla^{2}_{\bm{I}}L]_{ij}=\frac{\partial^{2}L}{\partial I_{i}\partial I_{j}}=-\frac{g_{i}}{I_{i}^{2}}\sigma_{ij} (35)

where σij=1\sigma_{ij}=1 when i=ji=j, and σij=0\sigma_{ij}=0 otherwise. This Hessian is a diagonal matrix with all negative elements when gm>0g_{m}>0. Therefore, the function L(𝑰|𝒈)L(\bm{I}|\bm{g}) is a concave function and the critical point at 𝑰=𝒈\bm{I}=\bm{g} is the global maximum for LL. When the mapping 𝒳:𝑨𝑰\mathcal{X}:{\bm{A}}\rightarrow{\bm{I}} is invertible and 𝒈\bm{g} is within the range of the mapping, the maximum value of LL corresponds to a point 𝑨g\bm{A}_{g} that satisfies 𝑰(𝑨g)=𝒈\bm{I}(\bm{A}_{g})=\bm{g}.

We then consider LL as a function of the AM coefficients 𝑨\bm{A}. When the matrix 𝑨𝑰\nabla_{\bm{A}}{\bm{I}} is invertible, which is true when the X-ray transform is invertible, 𝑨g\bm{A}_{g} is a critical point of the likelihood function L(𝑨|𝒈)L(\bm{A}|\bm{g}), as L(𝑨g|𝒈)L(\bm{A}_{g}|\bm{g}) is the maximum likelihood value. Furthermore, when 𝑨\bm{A} is located within the region defined by Im(𝑨)gmI_{m}(\bm{A})\geq g_{m}, the likelihood function L(𝑨|𝒈)L(\bm{A}|\bm{g}) is a concave function of 𝑨\bm{A}. An ML algorithm can be developed based on Newton’s method [30, 31] with iterations described as

𝑨k+1=𝑨k+tkΔ𝑨k,andΔ𝑨k=[𝑨2L(𝑨k)]1𝑨L(𝑨k),\bm{A}_{k+1}=\bm{A}_{k}+t_{k}*\Delta\bm{A}_{k},\quad\quad{\text{a}nd}\quad\quad\Delta\bm{A}_{k}=-[\nabla^{2}_{\bm{A}}L(\bm{A}_{k})]^{-1}\nabla_{\bm{A}}L(\bm{A}_{k}), (36)

where 𝑨k\bm{A}_{k} is the attenuation coefficients at iteration kk and tkt_{k} is the step size chosen with an Armijo-type (or back-tracking) line search to enforce sufficient increase in LL and the negativeness of the Hessian 𝑨2L\nabla^{2}_{\bm{A}}L. Note that the enforcement of negative-definiteness of the Hessian is important, as the algorithm may not converge otherwise. The likelihood function is convex in the region 0𝑨𝑨g0\leq\bm{A}\leq\bm{A}_{g}, hence the algorithm works the best when the initialization point 𝑨0\bm{A}_{0} is less than 𝑨g\bm{A}_{g}. Furthermore, during the iterations, the Hessian may become rank-deficient due to numerical accuracy. In this case, a gradient-descent step can be used instead of the Newton step. For example, in our implementation, we used an initialization of 𝑨0=(0,0,0)\bm{A}_{0}=(0,0,0) and back-tracking parameters α=0.1\alpha=0.1 and β=0.1\beta=0.1 (α\alpha and β\beta are defined as in [30]), and the convergence of the algorithm required less than 15 iteration steps for every case we have tested in the following sections.

V.1 AM coefficients estimation uncertainties

To demonstrate the applications of this maximum-likelihood algorithm, we considered an ideal three-bin PC detector with non-overlapping rect response functions, as shown in Fig 2(c) but with equal heights and no overlapping. The source is operated at 160 kVp and generates a broad X-ray spectrum [32]. The material attenuation profiles are extracted from the NIST XCOM data. X-ray attenuation was simulated according to Beer’s law. X-ray scattering and detector imperfections are not considered. The data were simulated at source photon budget I0=107I_{0}=10^{7} with Poisson noise.

We simulated a single X-ray path with different lengths of water as the attenuating media. The energy-bin boundaries of the detector were set at [30, 75, 100, 160] keV. The length of water ranged from 1 cm to 28 cm. For each length, 1000 sets of noisy data were generated, and the AM coefficients were estimated for each set of data. We calculated the mean and the variance of the estimated coefficients and compared the estimation uncertainty to the Crámer-Rao lower bound. The results are presented in Figure 5.

To check if the algorithm works for materials that are very different from water, we changed the material in the X-ray path to iron and the detector bin boundaries to [30, 110, 140, 160] keV. The bin boundaries were changed so that there are sufficient number of photons collected in all three bins. The length of iron ranged from 0.1 cm to 3 cm. The mean and variance of the 1000 repeated estimations at each length are presented in Figure 6.

In both scenarios, the mean of the estimates (black line) matches well with the true coefficient (red circle). The slight deviation between the true coefficient and the mean estimation at high attenuation region can be attributed to sampling error in Monte-Carlo simulation. The standard deviation of the estimates (purple area) is almost perfectly aligned with the Crámer-Rao lower bound. These results demonstrate that our estimation algorithm is unbiased and efficient. Furthermore, the AM coefficient corresponding to the Rayleigh scattering is estimable and the uncertainty in A^rs\hat{A}_{rs} is comparable to the uncertainty in A^pe\hat{A}_{pe}, which corresponds to the photoelectric effect. However, the A^pe\hat{A}_{pe} and A^rs\hat{A}_{rs} are anti-correlated, which is demonstrated by a negative value of the element [FIM1]13[\mathrm{FIM}^{-1}]_{13}. This correlation is probably due to the fact that the basis functions fpe(E)f_{pe}(E) and frs(E)f_{rs}(E) have similar shapes. This correlation may partially explain why A^pe\hat{A}_{pe} and A^rs\hat{A}_{rs} have more variance than A^cs\hat{A}_{cs}, as shown in Figure 5. As a result of the correlation, the estimated total attenuation τ^(E)=A^pefpe(E)+A^csfcs(E)+A^rsfrs(E)\hat{\tau}(E)=\hat{A}_{pe}f_{pe}(E)+\hat{A}_{cs}f_{cs}(E)+\hat{A}_{rs}f_{rs}(E), which is the ultimate physical quantity we are interest in, is not very noisy, as will be shown in the reconstruction results of the phantom.

V.2 Phantom reconstruction

We further applied the ML estimation algorithm for an image reconstruction. The reconstruction problem in X-ray computed tomography (CT) is to estimate the distribution 𝒂(𝑹)\bm{a}(\bm{R}) from the estimated line integrals 𝑨^\hat{\bm{A}} for each ray path. The AM coefficients 𝒂\bm{a} at each location 𝑹\bm{R} correspond to an attenuation profile μ(E)\mu(E). For a two-dimensional scene, the object μ(E,𝑹)\mu(E,\bm{R}) and the reconstruction μ^(E,𝑹)\hat{\mu}(E,\bm{R}) are both three-dimensional data cubes. To present the reconstruction result, we plot μ(E,𝑹)\mu(E,\bm{R}) and μ^(E,𝑹)\hat{\mu}(E,\bm{R}) at an arbitrary energy.

We simulate a two-dimensional fan-beam CT system (62 fan-beam angle) with 360 views and 245 detectors. The field of view is 256x256 pixels with a pixel pitch of 1—1.5 mm. The same source, detector, and material database described in the previous section are used. X-ray attenuation is simulated according to Beer’s law, while scattering and detector imperfections are not considered. The detector energy bin boundaries are [30, 75, 100, 160] keV. The source photon budget for each beam path is 10710^{7}. The AM coefficients 𝑨\bm{A} are estimated for each beam path and the object represented by 𝒂(𝑹)\bm{a}(\bm{R}) is reconstructed from 𝑨^\hat{\bm{A}} using a filtered-back projection (FBP) algorithm.

The first phantom reconstructed is a circular water phantom of diameter about 30 cm with pixel-pitch 0.15 cm. This phantom and its reconstruction are plotted at E=75.5E=75.5 keV in Figure 7. The reconstruction matches well with the object. As shown in the center-line plot in the right panel of Figure 7, the reconstruction has no cupping artifacts, which is typically associated with beam hardening.

A second phantom is a multi-material resolution phantom design inspired by Gong et.al. [33]. The length of the phantom is 20 cm with pixel pitch around 0.1 cm. The phantom, as shown in Figure 8 (left), is a Delrin block with 25 circular inserts in five rows and five columns. In each row, the inserts are made from the same material; in each column, the inserts have the same diameter. From top to bottom, the five materials for the inserts are water, polyvinyl chloride, cast magnesium, acrylic and methanol. The diameter of the inserts are from 0.6 to 1.8 cm with 0.3 cm step. The reconstruction of the resolution phantom are plotted at E=75.5E=75.5 keV in Figure 8. In the plots, different shades of grey represent different materials. The reconstruction results show that the ML estimation algorithm works for a broad range of AM coefficients

Each reconstruction, which calls the ML estimation algorithm 88,200 times, takes approximately 130 seconds on a desktop with a quad-core central processing unit (CPU). The reconstruction can be further sped up using a graphic processing unit (GPU).

VI Discussion

In our proof of invertibility, we focused on the interior points of M1M_{1} and M2M_{2} and proved that, for an ME X-ray transform, the global invertibility is equivalent to local invertibility for 𝑨\bm{A} in the interior of M1M_{1} and 𝑰\bm{I} in the interior of M2M_{2}. This equivalence can be extended to the boundaries of M1M_{1} and M2M_{2} by invoking Theorem 6 in Sandberg et. al. [34]. As mentioned in the introduction, Bal et. al. [9] has also provided a sufficient condition for the invertibility of ME X-ray transform. Their sufficient condition is that the Jacobian is a P-matrix in a rectangle in 𝑹N\bm{R}^{N}. P-matrix, which is a concept related to the preservation of orientation, requires the matrix and a few sub-matrices to be all positive. For the definition of P-matrix, please refer to Bal et. al. [9] or Gale and Nikaido [35]. Bal et. al. studied the invertibility of different ME X-ray systems by numerically calculating the Jacobian matrix on a grid of 𝑨\bm{A} values for each system. Based on numerical simulations, they have suggested that an ME X-ray system may become invertible as soon as the mapping is locally invertible in the rectangle. Our sufficient condition for global invertibility is that the Jacobian is non-vanishing for all points in M1M_{1}. In comparison to Bal et. al., our sufficient condition is weaker (better), but the domain where the Jacobian matrix needs to be checked is larger. For ease of computation, we also provide a sufficient condition for non-vanishing Jacobian, which requires the integrand of the Jacobian to have constant sign over all energy combinations in S12NS_{12...N}. From a practical point of view, the latter condition is significantly easier to use, as (1) the sign of the Jacobian integrand does not depend on 𝑨\bm{A}, and (2) the properties of the basis functions and the detector response functions can be studied separately.

Invertibility only requires that the Jacobian J(𝑨)0J(\bm{A})\neq 0 for all coefficients 𝑨\bm{A}. Nonetheless, a smaller J(𝑨)J(\bm{A}) leads to a worse-conditioned inverse problem and hence more uncertainty in the estimation as discussed in Section IV. Take the binning detector depicted in Figure 2 (c) as an example, when Bin 1 and Bin 3 do not overlap, the system is invertible for N=3N=3 (when imaging materials with no K-edges). However, the overlap between bins would result in a reduction in the Jacobian and hence more uncertainty in the coefficient estimation, which has also been observed in other works [36, 37]. One can employ the CRLB to optimize bin boundaries for a given set of AM coefficients 𝑨\bm{A}. As the CRLB varies with 𝑨\bm{A}, the optimum bin boundaries depend on the prior distribution of the objects. An optimum energy-weighting strategy that is not object dependent has been proposed by Wang et. al. [38]. Their strategy is to set the weights pm(E)p_{m}(E) same as the attenuation basis functions fi(E)f_{i}(E). They have proved that this measurement strategy provided a sufficient statistic to the X-ray spectral flux. From Equation (24), we can prove that this strategy is globally invertible, because det[F(𝑬)]det[P(𝑬)]={det[F(𝑬)]}20\det[F(\bm{E})]\det[P(\bm{E})]=\{\det[F(\bm{E})]\}^{2}\geq 0, for all 𝑬\bm{E}.

Conventional DECT systems reconstruct the effective atomic number (ZeZ_{e}) and the electron density (ρ\rho) [39] from two energy-weighted measurements. However, ZeZ_{e} and ρ\rho may not capture all of the information about material composition measurable from attenuation-based X-ray systems. Based on principal component analysis (PCA), Bornefalk et. al. [40] have suggested that the intrinsic dimensionality of the attenuation profiles of low-Z materials in the XCOM data base is four. Midgley et. al. [41] also showed similar degrees of freedom in the parameterization of the X-ray linear attenuation profiles. However, whether these intrinsic dimensions are accessible or not is still up to debate [42, 43]. There is potential value in collecting ME X-ray data, but the benefits may depend on the task of the imaging system and the experimental setup.

We used a set of basis functions that describe photoelectric, Compton scattering and Rayleigh scattering, because we wanted to investigate whether the Rayleigh coefficient is estimable or not. Rayleigh scattering has often been ignored in DE imaging due to its small contribution in the X-ray attenuation profile. Our results show that the Rayleigh component, ArsA_{rs}, is solvable and the uncertainty in its estimation is comparable to that of the photoelectric coefficient. However, we did not specifically study how important ArsA_{rs} is for the task of material discrimination. Other basis functions that are based on materials of interest (such as water and bone) or on PCA [44, 40, 45, 46] can be used as well. As pointed out by Alvarez et. al. [8], the choice of a particular basis set does not affect the invertibility. The uncertainty in the estimated attenuation profile should not be affected by the choice of the basis set either.

We demonstrated a two-step reconstruction algorithm that consists of an ML estimation of the AM coefficients and the FBP reconstruction. Many work have been done in DECT and MECT reconstruction. Reconstruction algorithms are currently available in three main flavors: object-domain based [47, 48], projection-domain based [49, 50], and one-step statistical algorithms [51, 52, 53, 54, 55] that estimate 𝒂^(𝑹)\hat{\bm{a}}(\bm{R}) from the raw data directly. Our ML estimation algorithm was designed for Poisson likelihood and ideal ME X-ray transform where Equation (LABEL:eq:mean_mapping) is valid. When the Poisson likelihood model or the ideal forward model are not accurate, the estimation algorithm needs modification.

Our ML estimation algorithm is almost unbiased and achieves the Crámer-Rao lower bound. The ML estimation is an established paradigm for nonlinear estimation tasks. At high signal-to-noise ratio (SNR), ML estimates are asymptotically unbiased and efficient (achieves Crámer-Rao lower bound). At low SNR (e.g. short exposure time), however, the ML estimates tend to be skewed and the variance is often larger than the Cramer-Rao bound [56]. The reason why our estimator is efficient is probably that our simulation was carried out in the high SNR regime. In our experiment, the smallest photon count collected in an energy bin is about 300 photons, which still has a relatively high SNR. To analyze a realistic system, one need to first identify if the system is operating at low SNR regime. If that is the case, instead of using the Crámer-Rao bound to characterize the variance of the ML estimates, one can apply other measures such as χpdfML2\chi^{2}_{pdf-ML}-isocontours [57] to describe the distribution of the ML estimates.

There are several limitations in our work. Firstly, the physical process considered by the mapping 𝒳:𝑨𝑰\mathcal{X}:{\bm{A}}\rightarrow{\bm{I}} includes only the attenuation of the X-ray photons, which follows Beer’s law, but not signals due to scattered radiation or background radiation. Although scattered radiation and background radiation can be significantly mitigated through anti-scatter grids [58], those signals should be characterized and accounted for, as they may affect the invertibility of a realistic system. Secondly, effects that limit detector performance, such as charge sharing [59, 60], charge trapping [61, 60] and pulse pileup, are ignored. Such details should be considered in the system model when the framework is applied to a specific imaging system. For a given realistic detector response function, the invertibility of the system can be studied by calculating det[P(𝑬)]\det[P({\bm{E}})] numerically over the subspace Ω12N\Omega_{12…N}. In this case, det[P(𝑬)]0\det[P({\bm{E}})]\leq 0 (or 0\geq 0) over Ω12N\Omega_{12…N} may not be guaranteed, hence the invertibility is not guaranteed. However, one can apply the invertibility framework in the optimization of detector designs. Lastly, we assumed the energy-weighting functions, including the source spectrum and the detector energy response functions, are known exactly. In reality, one can measure the energy-weighting functions experimentally [62] within some uncertainty.

VII Conclusion

We have provided a sufficient condition for the global invertibility of an ME X-ray transform for attenuation-based X-ray imaging. The ME X-ray transform is defined as the mapping from NN (N2N\geq 2) AM coefficients to NN energy-weighted noise-free measurements. The invertibility of this transform depends greatly on the weighting schemes used in the measurements. Considering scenes with no K-edge materials, we represented the X-ray attenuation profiles with N=3N=3 AM coefficients and proved the global invertibility of the transform for four commonly used weighting schemes. The same framework can be used to examine the invertibility of an arbitrary ME X-ray system, such as a system with non-ideal detectors, a system with multiple source emission spectra, and scenes with K-edge materials. This mathematical framework can be applied broadly in the design of X-ray detectors and systems.

We also considered Poisson noise in the measurement data and presented the CRLB on the estimation uncertainty. Furthermore, we presented an ML estimation algorithm and applied the algorithm to estimate AM coefficients for varying lengths of water and varying lengths of iron. The results have shown that the coefficient corresponding to Rayleigh scattering is estimable. Last but not least, we demonstrated the application of the ML estimator in reconstruction. Two phantoms imaged through a simulated fan-beam CT with ideal three-energy discriminating photon-counting detectors were reconstructed. The reconstructed images match well with the objects and are free of the ‘cupping artifacts’ induced by beam hardening.

Acknowledgments

Dr. Clarkson acknowledges the support of NIH R01- EB000803 and P41- EB002035.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Data Availability Statement

The data and code that support the findings of this study are available from the corresponding author upon reasonable request.

References

References

  • [1] Alvarez Robert E, Macovski Albert. Energy-selective reconstructions in X-ray computerised tomography Physics in Medicine & Biology. 1976;21:733.
  • [2] Lehmann LA, Alvarez RE, Macovski Aetal, et al. Generalized image combinations in dual kVp digital radiography Medical Physics. 1981;8:659–667.
  • [3] Barnes Gary T, Sones Richard A, Tesic Mike M, Morgan Douglas R, Sanders John N. Detector for dual-energy digital radiography. Radiology. 1985;156:537–540.
  • [4] Cardinal H Neale, Fenster Aaron. Theoretical optimization of a split septaless Xenon ionization detector for dual-energy chest radiography Medical Physics. 1988;15:167–180.
  • [5] Roessl E, Proksa R. K-edge imaging in X-ray computed tomography using multi-bin photon counting detectors Physics in Medicine & Biology. 2007;52:4679.
  • [6] Fredenberg Erik. Spectral and dual-energy X-ray imaging for medical applications Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2018;878:74–87.
  • [7] Levine Zachary H. Nonuniqueness in dual-energy CT Medical Physics. 2017;44:e202–e206.
  • [8] Alvarez Robert E. Invertibility of the dual energy X-ray data transform Medical Physics. 2019;46:93–103.
  • [9] Bal Guillaume, Terzioglu Fatma. Uniqueness criteria in multi-energy CT Inverse Problems. 2020;36:065006.
  • [10] Sukovle P, Clinthorne NH. Basis material decomposition using triple-energy X-ray computed tomography in IMTC/99. Proceedings of the 16th IEEE Instrumentation and Measurement Technology Conference (Cat. No. 99CH36309);3:1615–1618IEEE 1999.
  • [11] Lehmann L, Alvarez R, Macovski A, et al. Energy-selective reconstructions in X-ray computerized tomography Medical Physics. 1981;8:659–667.
  • [12] Macovski A, Alvarez RE, Chan JL-H, Stonestrom JP, Zatz LM. Energy dependent reconstruction in X-ray computerized tomography Computers in Biology and Medicine. 1976;6:325–336.
  • [13] Tapiovaara Markku J, Wagner R. SNR and DQE analysis of broad spectrum X-ray imaging Physics in Medicine & Biology. 1985;30:519.
  • [14] Flohr Thomas G, McCollough Cynthia H, Bruder Herbert, et al. First performance evaluation of a dual-source CT (DSCT) system European Radiology. 2006;16:256–268.
  • [15] Zhang Da, Li Xinhua, Liu Bob. Objective characterization of GE discovery CT750 HD scanner: gemstone spectral imaging mode Medical Physics. 2011;38:1178–1188.
  • [16] Altman A, Carmi R. TU-E-210A-03: a double-layer detector, dual-energy CT—principles, advantages and applications Medical Physics. 2009;36:2750–2750.
  • [17] Kraft Edgar, Fischer Peter, Karagounis Michael, et al. Counting and integrating readout for direct conversion X-ray imaging: Concept, realization and first prototype measurements IEEE Transactions on Nuclear Science. 2007;54:383–390.
  • [18] Karg J, Niederlöhner D, Giersch J, Anton G. Using the Medipix2 detector for energy weighting Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment. 2005;546:306–311.
  • [19] Taguchi Katsuyuki, Iwanczyk Jan S. Vision 20/20: Single photon counting X-ray detectors in medical imaging Medical Physics. 2013;40.
  • [20] Fredette Nathaniel R, Kavuri Amar, Das Mini. Multi-step material decomposition for spectral computed tomography Physics in Medicine & Biology. 2019;64:145001.
  • [21] Fulks Watson. Advanced calculus: an introduction to analysis. John Wiley & Sons 1978.
  • [22] Williamson Jeffrey F, Li Sicong, Devic Slobodan, Whiting Bruce R, Lerma Fritz A. On two-parameter models of photon cross sections: Application to dual-energy CT imaging Medical Physics. 2006;33:4115–4129.
  • [23] Trubey D. K.. HUGI VI notes documentation for the DLC-146 code package. Radiaton Safety Information Computational Center 1989.
  • [24] Berger MJ, Hubbell JH, Seltzer SM, et al. XCOM: Photon Cross Sections Database. NIST Standard Reference Database 8 (XGAM) http://physics.nist.gov/PhysRefData/Xcom/Text/XCOM.html. 1998.
  • [25] Krantz Steven G, Parks Harold R. The implicit function theorem: history, theory, and applications. Springer Science & Business Media 2012.
  • [26] Munkres James Raymond. Topology;2. Prentice Hall 2000.
  • [27] Lee John. Introduction to Smooth Manifolds. Springer 2012.
  • [28] Cramer Harald. Mathematical methods of statistics. Princeton University Press 1946.
  • [29] Rao C Radhakrishna. Information and the accuracy attainable in the estimation of statistical parameters Bulletin of the Calcutta Mathematical Society. 1945;37:81–89.
  • [30] Boyd Stephen, Boyd Stephen P, Vandenberghe Lieven. Convex optimization. Cambridge university press 2004.
  • [31] Nocedal Jorge, Wright Stephen. Numerical optimization. Springer Science & Business Media 2006.
  • [32] Ding Yijun, Ashok Amit. X-ray measurement model and information-theoretic metric incorporating material variability with energy correlations in Anomaly Detection and Imaging with X-rays (ADIX) IV;10999:109990JInternational Society for Optics and Photonics 2019.
  • [33] Gong Qian, Stoian Razvan-Ionut, Coccarelli David S, Greenberg Joel A, Vera Esteban, Gehm Michael E. Rapid simulation of X-ray transmission imaging for baggage inspection via GPU-based ray-tracing Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms. 2018;415:100–109.
  • [34] Sandberg I. Global inverse function theorems IEEE Transactions on Circuits and Systems. 1980;27:998–1004.
  • [35] Gale David, Nikaido Hukukane. The Jacobian matrix and global univalence of mappings Mathematische Annalen. 1965;159:81–93.
  • [36] Ren Yan, Xie Huiqiao, Long Wenting, Yang Xiaofeng, Tang Xiangyang. On the conditioning of spectral channelization (energy binning) and its impact on multi-material decomposition based spectral imaging in photon-counting CT IEEE Transactions on Biomedical Engineering. 2021.
  • [37] Tang Xiangyang, Ren Yan. On the conditioning of basis materials and its impact on multi-material decomposition based spectral imaging in photon-counting CT Medical Physics. 2021.
  • [38] Wang Adam S, Pelc Norbert J. Sufficient statistics as a generalization of binning in spectral X-ray imaging IEEE transactions on medical imaging. 2010;30:84–93.
  • [39] Martz Harry E, Seetho Issac M, Champley Kyle E, Smith Jerel A, Azevedo Stephen G. CT dual-energy decomposition into X-ray signatures ρe\rho_{e} and ZeZ_{e} in Anomaly Detection and Imaging with X-Rays (ADIX);9847:98470DInternational Society for Optics and Photonics 2016.
  • [40] Bornefalk Hans. XCOM intrinsic dimensionality for low-Z elements at diagnostic energies Medical Physics. 2012;39:654–657.
  • [41] Midgley Stewart Michael. A parameterization scheme for the x-ray linear attenuation coefficient and energy absorption coefficient Physics in Medicine & Biology. 2004;49:307.
  • [42] Midgley Stewart Michael. Materials analysis using x-ray linear attenuation coefficient measurements at four photon energies Physics in Medicine & Biology. 2005;50:4139.
  • [43] Levine Zachary H, Peskin Adele P, Holmgren Andrew D, Garboczi Edward J. Preliminary X-ray CT investigation to link Hounsfield unit measurements with the International System of Units (SI) Plos one. 2018;13:e0208820.
  • [44] Alvarez Robert E. Dimensionality and noise in energy selective X-ray imaging Medical Physics. 2013;40:111909.
  • [45] Weaver John B, Huddleston Alan L. Attenuation coefficients of body tissues using principal-components analysis Medical Physics. 1985;12:40–45.
  • [46] Xie Huiqiao, Ren Yan, Long Wenting, Yang Xiaofeng, Tang Xiangyang. Principal component analysis in projection and image domains Another form of spectral imaging in photon-counting CT IEEE Transactions on Biomedical Engineering. 2020.
  • [47] Brooks Rodney A. A quantitative theory of the Hounsfield unit and its application to dual energy scanning. Journal of computer assisted tomography. 1977;1:487–493.
  • [48] Maaß Clemens, Baer Matthias, Kachelrieß Marc. Image-based dual energy CT using optimized precorrection functions: A practical new approach of material decomposition in image domain Medical physics. 2009;36:3818–3829.
  • [49] Abascal Juan FPJ, Ducros Nicolas, Peyrin Françoise. Nonlinear material decomposition using a regularized iterative scheme based on the Bregman distance Inverse Problems. 2018;34:124003.
  • [50] Wu Dufan, Zhang Li, Zhu Xiaohua, Xu Xiaofei, Wang Sen. A weighted polynomial based material decomposition method for spectral x-ray CT imaging Physics in Medicine & Biology. 2016;61:3749.
  • [51] Mory Cyril, Sixou Bruno, Si-Mohamed Salim, Boussel Loic, Rit Simon. Comparison of five one-step reconstruction algorithms for spectral CT Physics in Medicine & Biology. 2018;63:235001.
  • [52] Mechlem Korbinian, Ehn Sebastian, Sellerer Thorsten, et al. Joint statistical iterative material image reconstruction for spectral computed tomography using a semi-empirical forward model IEEE transactions on medical imaging. 2017;37:68–80.
  • [53] Long Yong, Fessler Jeffrey A. Multi-material decomposition using statistical image reconstruction for spectral CT IEEE transactions on medical imaging. 2014;33:1614–1626.
  • [54] Barber Rina Foygel, Sidky Emil Y, Schmidt Taly Gilat, Pan Xiaochuan. An algorithm for constrained one-step inversion of spectral CT data Physics in Medicine & Biology. 2016;61:3784.
  • [55] Kazantsev Daniil, Jørgensen Jakob S, Andersen Martin S, Lionheart William RB, Lee Peter D, Withers Philip J. Joint image reconstruction method with correlative multi-channel prior for x-ray spectral computed tomography Inverse Problems. 2018;34:064001.
  • [56] Mueller Stefan P, Kijewski Marie Foley, Kappeler Christian, Moore Stephen C. Estimation performance at low SNR: predictions of the Barankin bound in Medical Imaging 1995: Physics of Medical Imaging;2432:157–166International Society for Optics and Photonics 1995.
  • [57] Müller Stefan P, Abbey Craig K, Rybicki Frank J, Moore Stephen C, Kijewski Marie Foley. Measures of performance in nonlinear estimation tasks: prediction of estimation performance at low signal-to-noise ratio Physics in Medicine & Biology. 2005;50:3697.
  • [58] Tang C-M, Stier E, Fischer K, Guckel H. Anti-scattering X-ray grid Microsystem Technologies. 1998;4:187–192.
  • [59] Shikhaliev Polad M, Fritz Shannon G, Chapman John W. Photon counting multienergy X-ray imaging: Effect of the characteristic X rays on detector performance Medical Physics. 2009;36:5107–5119.
  • [60] Xu Cheng, Danielsson Mats, Bornefalk Hans. Evaluation of energy loss and charge sharing in cadmium telluride detectors for photon-counting computed tomography IEEE Transactions on Nuclear Science. 2011;58:614–625.
  • [61] Knoll Glenn F. Radiation detection and measurement. John Wiley & Sons 2010.
  • [62] Ha Wooseok, Sidky Emil Y, Barber Rina Foygel, Schmidt Taly Gilat, Pan Xiaochuan. Estimating the spectrum in computed tomography via Kullback–Leibler divergence constrained optimization Medical physics. 2019;46:81–92.

Figures

Refer to caption
Figure 1: Shape of fi(E)f_{i}(E) for i=1,2,3i=1,2,3 (left) and fitted attenuation profile of water (right).
Refer to caption
Figure 2: Energy weighting functions for four special cases: (a) CIX-PC/EI/MI, (b)non-overlapping bins with arbitrary response, (c) three rect bins with ideal energy resolution, and (d) a slightly overlapping three-bin detector, where the overlap is introduced by the finite energy resolution of the detector.
Refer to caption
Figure 3: Constructing a simply connected range M2M_{2} by expanding the domain of the map 𝒳\mathcal{X}: (a) The initial domain U0U_{0} and range V0V_{0}. (b) ViV_{i}, defined by points 𝑰i\bm{I}_{i} and 𝑰0𝒅\bm{I}_{0}\bm{d}, and its corresponding preimages uiju_{ij}. UiU_{i} is defined as the union of all uiju_{ij}. (c) The expanded domain M1M_{1} and range M2M_{2}. (d) Every closed curve ϕ(s)\phi(s) (blue loop) can be shrunk down to the point I0𝒅I_{0}\bm{d} through the function H(s,t)H(s,t). Color varying from blue to green represents tt from 11 to 0. The black-dashed line is the trajectory of a point ss in the loop ϕ\phi as tt decreases from 1 to 0.
Refer to caption
Figure 4: (a) For given noise-free measurement data 𝑰\bm{I}, the vector 𝑨\bm{A} is bounded in the area indicated in grey. Each energy-weighted measurement generates a pair of red and blue hyperplanes, which bound the vector 𝑨\bm{A}. (b) Illustration of the sign-switched mirror symmetry in function det[F(E1,E2)\det[F(E_{1},E_{2})] along line E1=E2E_{1}=E_{2}.
Refer to caption
Figure 5: Uncertainties in the three estimated photoelectric/Compton-scattering/Rayleigh-scattering coefficients for different lengths of water.
Refer to caption
Figure 6: Uncertainties in the three estimated photoelectric/Compton-scattering/Rayleigh-scattering coefficients for different lengths of iron.
Refer to caption
Figure 7: Reconstruction of the circular phantom. The object μ(E,𝑹)\mu(E,\bm{R}) (left), the reconstruction μ^(E,𝑹)\hat{\mu}(E,\bm{R}) (middle) and the center-line plots (right) are all presented at E=75.5E=75.5 keV.
Refer to caption
Figure 8: Reconstruction of the resolution phantom. The object μ(E,𝑹)\mu(E,\bm{R}) (left), the reconstruction μ^(E,𝑹)\hat{\mu}(E,\bm{R}) (middle) and the center-line plots (right) are all presented at E=75.5E=75.5 keV.