Fact-checked by Grok 2 weeks ago

Radon transform

The Radon transform is a fundamental in that maps a multidimensional to the set of its integrals over a family of hyperplanes, enabling the reconstruction of the original from these projections. Introduced by Austrian mathematician in 1917, it provides an exact mathematical framework for inverting such line integrals, laying the groundwork for modern imaging techniques. Mathematically, for a f defined on \mathbb{R}^n, the Radon transform Rf at a specified by a unit normal vector \theta \in S^{n-1} and signed distance s \in \mathbb{R} is given by Rf(\theta, s) = \int_{\mathbb{R}^n} f(x) \delta(s - x \cdot \theta) \, dx, where \delta is the , effectively computing the of f along the \{x : x \cdot \theta = s\}. This is linear, satisfying such as superposition (i.e., R(af + bg) = aRf + bRg for scalars a, b and functions f, g) and rotational symmetry, with analogs to convolution theorems and Plancherel's formula in . The transform is invertible under suitable conditions, such as when f is compactly supported and sufficiently smooth, allowing recovery via filtered back-projection methods involving the , which links the 1D of projections to slices of the 2D of f. Beyond pure mathematics, the Radon transform is pivotal in applied fields, particularly computed tomography (CT) in , where it models the of X-rays along lines to reconstruct cross-sectional images of the human body from non-invasive scans. Its principles extend to emission (e.g., and SPECT), geophysical imaging, and even planetary , such as deriving surface elevations from polar echoes. Since the 1970s, advancements in inversion algorithms have made it indispensable for real-time diagnostics, with ongoing research exploring generalizations like the cone-beam transform for helical scanning.

Introduction

Overview

The Radon transform is a fundamental mathematical tool that captures the essence of a by its values over straight lines in two dimensions or hyperplanes in higher dimensions, producing a set of projections akin to shadow silhouettes from multiple viewpoints. This process intuitively reveals how the function's mass or density is distributed along those directions, much like how beams traverse an object to measure cumulative opacity without direct access to its interior. At the heart of image reconstruction techniques, the Radon transform models the data collection in computed tomography (CT) scans, where projections from various angles enable the recovery of cross-sectional images of the human body, revolutionizing medical diagnostics by allowing precise visualization of tissues and organs. Formally introduced by Austrian mathematician Johann Radon in his 1917 paper, the transform builds on earlier related ideas in integral representations. Its applications extend to geophysics, where it aids in seismic signal processing to delineate underground formations from reflection data, and to pure mathematics, bridging integral geometry—concerned with invariants under group actions on manifolds—and harmonic analysis for studying function decompositions and operator properties.

Historical Development

The origins of the Radon transform can be traced to early developments in geometry, with precursors involving s over curves and surfaces. In 1904, employed s on to study bodies of constant width, laying groundwork for later geometric methods. This was extended by Paul Funk in 1911, who introduced what is now known as the Funk transform, computing mean values of functions on the two-dimensional along to enable . The transform was formally defined by in in his seminal paper "Über die Bestimmung von Funktionen durch ihre Integralwerte längs gewisser Mannigfaltigkeiten," where he generalized the concept to integrals over hyperplanes in n-dimensional and provided an explicit inversion formula for reconstructing the original function. Radon's work built directly on Funk's spherical integrals but shifted focus to flat manifolds, establishing the mathematical framework for what would later become essential in multidimensional problems. However, early formulations, including Radon's, offered primarily theoretical inversion methods that were not yet adapted for practical numerical computation or discrete data. Significant advancements occurred in the mid-20th century through applications to medical imaging, particularly computed tomography (CT). In the 1960s, Allan M. Cormack developed reconstruction algorithms based on Radon-like projections, publishing key results in 1963 and 1964 that addressed the inversion problem for X-ray attenuation data using series expansions and Fourier techniques. Independently, engineer Godfrey N. Hounsfield constructed the first practical CT scanner in 1971 at EMI Laboratories, implementing filtered backprojection for image reconstruction from projection data. Their combined theoretical and engineering contributions revolutionized diagnostic imaging, earning them the 1979 Nobel Prize in Physiology or Medicine for the development of computer-assisted tomography. This era marked the transition from abstract mathematics to real-world application, though inversion methods remained challenged by noise and limited data until refinements in the 1970s provided more stable and complete solutions for practical use. In the , the Radon transform saw expansions into and numerical domains, driven by advances in and . A pivotal contribution was the 1987 introduction of the Radon transform by Gregory Beylkin and colleagues, which enabled exact inversion for finite datasets on uniform grids, facilitating efficient computation in image processing and seismic analysis. These adaptations, including fast algorithms for limited-angle projections, integrated the transform into pipelines, broadening its use in fields like and geophysical imaging while addressing computational efficiency for large-scale data.

Mathematical Definition and Properties

Formal Definition

The Radon transform, originally introduced by Johann Radon in 1917, provides a mathematical framework for integrating a function over hyperplanes in Euclidean space. For a compactly supported continuous function f: \mathbb{R}^n \to \mathbb{R}, the Radon transform Rf is defined as Rf(\theta, s) = \int_{H(\theta, s)} f(x) \, d\mu(x), where H(\theta, s) = \{ x \in \mathbb{R}^n \mid x \cdot \theta = s \} denotes the hyperplane orthogonal to the unit vector \theta \in S^{n-1} at signed distance s \in \mathbb{R} from the origin, and d\mu(x) is the Euclidean surface measure induced on this hyperplane. This parametrization identifies hyperplanes via their normal direction \theta on the unit sphere and offset s. In the specific case of n=2, the Radon transform reduces to line integrals, often expressed in polar coordinates for clarity. Let \theta = (\cos \phi, \sin \phi) with \phi \in [0, \pi) and s \in \mathbb{R}; then Rf(\phi, s) = \int_{-\infty}^{\infty} f(s \cos \phi - t \sin \phi, s \sin \phi + t \cos \phi) \, dt, where the integration variable t parametrizes the line to \theta. This form aligns with the by setting the direction as \theta^\perp = (-\sin \phi, \cos \phi). The definition extends naturally to the space of Schwartz functions \mathcal{S}(\mathbb{R}^n), which are smooth and rapidly decreasing, ensuring Rf \in \mathcal{S}(S^{n-1} \times \mathbb{R}) and preserving key analytic properties under the transform. Further, by duality, it applies to tempered distributions, where Rf is interpreted via test function integration, facilitating applications in generalized function theory.

Basic Properties

The Radon transform R, defined as the operator that maps a f on \mathbb{R}^n to its over hyperplanes parameterized by \theta \in S^{n-1} and signed distance s \in \mathbb{R}, is . Specifically, for scalars a, b \in \mathbb{R} and f, g, it satisfies R(af + bg)(\theta, s) = a Rf(\theta, s) + b Rg(\theta, s). This follows directly from the definition of the transform. The transform is continuous in suitable function spaces. On the Schwartz space \mathcal{S}(\mathbb{R}^n) of rapidly decreasing smooth functions, R maps to \mathcal{S}(S^{n-1} \times \mathbb{R}), preserving the rapid decay and smoothness properties, which ensures continuity in the Fréchet topology of these spaces. It extends continuously to Sobolev spaces H^s(\mathbb{R}^n) for appropriate s, mapping to weighted Sobolev spaces on the parameter domain. A key property is the support theorem due to Helgason, which relates the support of f to that of Rf. If f \in C^\infty(\mathbb{R}^n) is rapidly decreasing and Rf(\theta, s) = 0 for all |s| > r and all \theta \in S^{n-1}, then f(x) = 0 for all |x| > r; the converse holds without the rapid decay assumption. This theorem enables recovery of the support of f from its projections alone. If f has compact support in a of radius A, then Rf(\theta, s) = 0 for |s| > A. Additionally, Rf is an even function on the S^{n-1} \times \mathbb{R}, satisfying Rf(-\theta, -s) = Rf(\theta, s). The Radon transform is covariant under the actions of and , reflecting its geometric origins. For an () U, R(Uf)(U\theta, s) = Rf(\theta, s). For a f_t(x) = f(x - t), Rf_t(\theta, s) = Rf(\theta, s - \theta \cdot t), which implies a convolution structure along each direction: R(f * g)(\theta, s) = \int_\mathbb{R} Rf(\theta, t) Rg(\theta, s - t) \, dt. These properties hold uniformly in n \geq 2. In (n=2), projections are over lines; in (n=3), over planes.

Connections to Other Transforms

Relation to the Fourier Transform

One of the most fundamental connections between the Radon transform and the is provided by the , also known as the . This theorem establishes that the one-dimensional of the Radon transform of a f along a projection at angle \theta, evaluated at frequency \sigma, equals the n-dimensional of f evaluated at the point \sigma \theta in Fourier space. In mathematical terms, for a function f: \mathbb{R}^n \to \mathbb{R} and its Radon transform Rf(\theta, s), the theorem states that \mathcal{F}(Rf)(\theta, \sigma) = \hat{f}(\sigma \theta), where \mathcal{F} denotes the one-dimensional with respect to the radial variable s, and \hat{f} is the n-dimensional of f. This relation holds under suitable assumptions on f, such as membership in L^1(\mathbb{R}^n) or the , ensuring integrability. The theorem implies a direct mapping from projection data in the Radon domain to slices in the Fourier domain of the original function, facilitating representation in polar coordinates where the radial coordinate corresponds to the frequency \sigma and the angular coordinate to \theta. This polar sampling structure is particularly advantageous for computational efficiency, as it allows the use of the (FFT) to interpolate between Cartesian and polar grids, enabling rapid evaluation of the from projection data without full matrix inversion. A sketch of the proof for the two-dimensional case proceeds by direct computation. Consider the Radon transform p_\theta(r) = \int_{-\infty}^\infty f(x \cos \theta + y \sin \theta, -x \sin \theta + y \cos \theta) \, dy. The one-dimensional Fourier transform is P_\theta(\rho) = \int_{-\infty}^\infty p_\theta(r) e^{-j 2\pi \rho r} \, dr = \int_{-\infty}^\infty \int_{-\infty}^\infty f(x, y) e^{-j 2\pi \rho (x \cos \theta + y \sin \theta)} \, dx \, dy, where the interchange of integrals is justified by Fubini's theorem under the integrability assumptions on f. The right-hand side is precisely the two-dimensional Fourier transform of f evaluated at (\rho \cos \theta, \rho \sin \theta). The n-dimensional case follows analogously by coordinate rotation and integration over the hyperplane orthogonal to \theta.

Dual Radon Transform

The dual Radon transform, also known as the backprojection operator, is the operator R^* of the Radon transform R with respect to the L^2 inner product. For a g defined on the space of hyperplanes, typically parameterized by direction \theta \in S^{n-1} and signed distance s \in \mathbb{R}, the dual Radon transform is given by R^* g(x) = \int_{S^{n-1}} g(\theta, x \cdot \theta) \, d\theta, where the integral is taken with respect to the standard surface measure on the unit S^{n-1}, and x \in \mathbb{R}^n. This operator maps functions on the cylinder S^{n-1} \times \mathbb{R} to functions on \mathbb{R}^n, effectively integrating the data g over all hyperplanes passing through the point x. The adjoint property establishes that R^* is the formal L^2-adjoint of R, satisfying \langle R f, g \rangle_{L^2(S^{n-1} \times \mathbb{R})} = \langle f, R^* g \rangle_{L^2(\mathbb{R}^n)} for suitable functions f \in L^2(\mathbb{R}^n) and g \in L^2(S^{n-1} \times \mathbb{R}), where the inner products are appropriately weighted to account for the invariant measures on the spaces. This duality is fundamental in the analysis of the Radon transform, as it preserves the inner product structure and facilitates the study of inversion problems through operator compositions. The normal operator R^* R, formed by composing the dual with the forward transform, acts on functions in \mathbb{R}^n and has a smoothing effect by integrating the projections over all directions. Specifically, for f \in L^2(\mathbb{R}^n), (R^* R f)(x) = \int_{S^{n-1}} (R f)(\theta, x \cdot \theta) \, d\theta, which represents the integrated value of the projections of f along all hyperplanes containing x. In the case, R^* R is an elliptic of order -1, which attenuates high-frequency components and smooths the input function, adding roughly $1/2 in the Sobolev . This smoothing property arises because R^* R integrates local information globally over directions, reducing singularities while preserving the overall scale of the function. Geometrically, the dual Radon transform can be interpreted as a superposition of projections: for each fixed point x, R^* g(x) accumulates the values of g from all lines (or hyperplanes) passing through x, effectively smearing or backprojecting the line integrals onto the space. This interpretation underscores its role in algorithms, where it reverses the line-averaging effect of the forward Radon transform by redistributing projection data uniformly along intersecting lines.

Intertwining Property

The relation between the backprojection operator R^* and the \mathcal{F} provides an analytic connection between projection data and frequency domain representations, manifesting through the , where the 1D Fourier transform along projections aligns with slices of the nD of the original function, and the backprojection integrates these consistently across directions. More precisely, the of the backprojection R^* g of a function g on the space (angles and s) yields a radial in space. For the case, \mathcal{F}(R^* g)(\xi) = \int_{S^1} \hat{g}(\theta, \theta \cdot \xi) \, d\theta, where \hat{g}(\theta, \sigma) denotes the 1D of g(\theta, \cdot) with respect to the offset variable at \sigma. This shows that at each \xi \in \mathbb{R}^2, the value is the integral over the projected frequencies along the unit circle S^1. A key consequence of this property is its role in deriving inversion formulas via filtered projections. The radial integration introduces a smoothing effect equivalent to multiplication by $1/|\xi| in Fourier space (arising from the measure on of radius |\xi|), which is counteracted by applying a ramp with Fourier multiplier |\xi| to the projections before backprojection. For the 2D case with backprojection over [0, \pi), this yields the standard filtered backprojection inversion f(x) = \frac{1}{2} \int_0^\pi \left( Rf(\cdot, \theta) * h \right) (x \cdot \theta) \, d\theta, where \hat{h}(\sigma) = |\sigma|. The proof relies on the homogeneity of the operators (both the backprojection and preserve scaling degrees) and the rotational invariance of the Radon setup. Starting from the definition R^* g(x) = \int_{S^1} g(\theta, \theta \cdot x) \, d\theta, the is computed via \mathcal{F}(R^* g)(\xi) = \int_{\mathbb{R}^2} R^* g(x) e^{-i \xi \cdot x} \, dx. Applying Fubini's theorem to interchange the integrals over space and directions, and substituting the 1D Fourier inversion for each fixed \theta, reduces the expression to the radial integral over the projected frequencies.

Inversion and Reconstruction

Direct Inversion Formulas

The direct inversion of the Radon transform provides explicit analytical expressions to recover the original function f from its Radon transform Rf, assuming f is sufficiently and compactly supported, and that complete data over all directions and offsets are available. These formulas originated with Radon's foundational work, where he derived inversion expressions using integral operators along hyperplanes. In two dimensions, the classical inversion formula expresses f(x) in terms of a integral involving second derivatives of the Radon transform: f(x) = \frac{1}{4\pi^2} \int_{S^1} \int_{\mathbb{R}} \frac{\partial^2}{\partial s^2} Rf(\theta, s) \cdot \frac{1}{x \cdot \theta - s} \, ds \, d\theta, where the inner integral is understood in the sense to handle the singularity. This form, equivalent to Radon's original expression, relies on the Fourier slice theorem to relate projections to the of f. An equivalent and computationally practical variant is the filtered backprojection formula, which incorporates a ramp filter in the followed by backprojection: f(x) = \frac{1}{4\pi^2} \int_{S^1} \mathcal{H} \left\{ |\sigma| \widehat{Rf}(\theta, \sigma) \right\} (x \cdot \theta) \, d\theta, where \widehat{Rf}(\theta, \sigma) denotes the one-dimensional of Rf(\theta, \cdot) with respect to s at \sigma, and \mathcal{H} is the applied along the projection direction. This , widely used in , stems from inverting the after applying the central slice theorem. For the general n-dimensional case, inversion involves higher-order derivatives or generalized Hilbert transforms over hyperspheres. One standard expression uses a principal value integral with a kernel adjusted for dimension: f(x) = c_n \int_{S^{n-1}} \pv \int_{\mathbb{R}^{n-1}} \frac{\partial^{n-1}}{\partial s^{n-1}} Rf(\theta, s) \cdot \frac{1}{|x_\perp - s|^{n-1}} \, ds \, d\theta, where c_n = (-1)^{(n-1)/2} / (2^{n-1} (n-1)! \pi^{(n-1)/2} \Gamma((n-1)/2)) for odd n, x_\perp is the component of x orthogonal to \theta, and the principal value ensures convergence for smooth f. Helgason generalized Radon's approach to manifolds, providing these formulas via representation theory on symmetric spaces. These direct methods assume ideal conditions, such as infinite data resolution, for exact reconstruction.

Ill-Posedness and Stability

The inversion of the Radon transform is a classic example of an ill-posed problem in the sense of due to the lack of continuous dependence on the data, despite existence and uniqueness holding in appropriate function spaces such as compactly supported smooth functions. Sequences of functions converging in appropriate norms can have Radon transforms converging to zero without the original functions converging, demonstrating discontinuity of the inverse operator. This ill-posedness manifests through severe amplification of high-frequency components during inversion, arising from the smoothing nature of the forward Radon . The singular value decomposition reveals that the singular values decay as \sigma(\omega) \approx \sqrt{2\pi / |\omega|}, or equivalently like $1/|\sigma|^{1/2} where \sigma denotes the variable, leading to unbounded amplification factors for high frequencies in the inverse . In discrete settings, such as pixelated data, the Picard condition provides a for the existence and uniqueness of to the discretized equations. This condition requires that the expansion coefficients of the noisy data in the singular vector basis decay faster than the singular values themselves; violation results in , as small perturbations can dominate the due to the gradual decay of singular values without a spectral gap. To stabilize direct inversion formulas amid , regularization techniques such as Tikhonov regularization and truncated (TSVD) are employed. Tikhonov regularization minimizes a penalized least-squares functional \|Rf - g\|^2 + \alpha \|Lf\|^2, where L is a smoothing operator (often the or a ), introducing a \alpha > 0 to balance fidelity and solution , thereby bounding the effectively. TSVD, in contrast, discards singular values below a , directly mitigating high-frequency amplification while preserving low-frequency structure. For noisy data g^\delta = g + \delta, where \delta denotes the level, bounds for regularized reconstructions typically as O(\delta^{1/2}) in appropriate norms, reflecting the mild ill-posedness degree induced by the singular value decay. The of the discrete Radon operator, defined as the ratio of the largest to smallest retained , grows like O(N^{1/2}) with grid size N, quantifying the inherent sensitivity and guiding threshold choices in regularization.

Iterative Reconstruction Methods

Iterative reconstruction methods address the inversion of the by solving discretized linear systems arising from data, particularly when dealing with incomplete, noisy, or sparse measurements. These approaches iteratively refine an estimate of the underlying f to satisfy the projection constraints Rf \approx g, where R is the discrete Radon operator and g represents the measured projections. Unlike direct inversion formulas, iterative methods incorporate models for noise and data inconsistencies, enabling robust reconstruction in practical scenarios such as . Algebraic reconstruction techniques (ART), based on the , solve the from ray sums by sequentially projecting the current estimate onto hyperplanes defined by each equation. The , originally proposed for approximate solutions to linear systems, updates the image estimate f^{(k)} along each ray i as f^{(k+1)} = f^{(k)} + [\lambda](/page/Lambda) \frac{g_i - \langle r_i, f^{(k)} \rangle}{\|r_i\|^2} r_i, where r_i is the ray and [\lambda](/page/Lambda) is a relaxation . In , ART was adapted by , , and Herman to reconstruct three-dimensional objects from limited projections, iterating through all rays in a sequential manner to minimize the residual error. This method converges to the exact solution for consistent data and exhibits linear convergence rates under certain conditions. For probabilistic models in emission tomography, such as (), the - () maximizes the likelihood of observed under a noise model. Shepp and Vardi formulated the maximum-likelihood (MLEM) update as f^{(k+1)}_j = \frac{f^{(k)}_j}{\sum_i r_{ij} } \sum_i \frac{r_{ij} g_i }{\sum_l r_{il} f^{(k)}_l }, where the summation is over pixels j and i, effectively alternating between (forward ) and maximization ( with ) steps. This approach accounts for the statistical nature of counts, improving image quality in low-count regimes compared to algebraic methods. Convergence of MLEM is monotonic in the likelihood but typically requires many iterations for practical accuracy. Gradient descent methods minimize a least-squares objective \|Rf - g\|^2 + \lambda \|Lf\|^2, where L is a regularization operator (e.g., discrete Laplacian for ), using updates f^{(k+1)} = f^{(k)} - \alpha (R^T (Rf^{(k)} - g) + \lambda L^T L f^{(k)} ). These iterative schemes, often implemented via conjugate acceleration, handle ill-conditioned systems from the Radon transform by incorporating priors to stabilize reconstruction. In sparse data scenarios, such as limited-angle , regularization prevents artifacts that plague direct methods. Convergence guarantees for these methods vary: the classical Kaczmarz/ART converges linearly for consistent systems with rate depending on the of the rows, while randomized variants achieve exponential convergence in expectation. EM converges to a of the likelihood, with rates analyzed via majorization-minimization properties. per iteration is typically O(N^3) for an N \times N with O(N) projections, dominated by forward and backprojection operations, though ordered-subset accelerations can reduce this effectively. Iterative methods excel over inversion for sparse by flexibly incorporating constraints and models, yielding higher fidelity reconstructions with fewer projections.

Applications and Extensions

In Imaging and Tomography

The Radon transform plays a central role in computed (CT) by modeling the acquisition of projections, which are line integrals through the imaged object. In parallel-beam , projections are collected along at various , directly approximating the continuous Radon transform for reconstruction. Fan-beam , commonly used in clinical CT scanners for efficiency, diverges beams from a , requiring rebinning algorithms to convert into parallel-beam equivalents before applying Radon-based inversion. Discrete approximations of the Radon transform, such as the 2-D discrete Radon transform (DRT), enable numerical reconstruction from pixelated images, converging to the continuous case as sampling density increases, with relative errors scaling as O(1/N) for N×N grids. A landmark application occurred in Godfrey Hounsfield's 1971 prototype, which produced the first clinical brain images by processing 180 parallel projections over 180 degrees, each with 160 ray measurements, and using Radon transform principles for backprojection reconstruction, taking 4.5 minutes per slice. This device revolutionized by enabling non-invasive visualization of soft tissues without invasive procedures. Beyond , the Radon transform supports projection reconstruction in (MRI), where radial sampling approximates projections that are inverted via filtered backprojection to form images, particularly useful for motion-robust sequences. In ultrasound imaging, plane-wave transmissions generate data mappable to the Radon domain, allowing high-frame-rate reconstruction through interpolation and inverse Radon transform, achieving contrast levels comparable to delay-and-sum methods (e.g., 8.5 dB) while reducing computation. in employs the Radon transform for from tilt-series projections, but limited-angle acquisition—typically restricted to 60–70 degrees due to sample thickness—poses ill-posedness, addressed via regularization to mitigate missing wedge artifacts. Practical challenges include motion from patient respiration or heartbeat, which distort projections and cause blurring or streaking in Radon-inverted images, often mitigated by gating or fewer projections to shorten times. Dose reduction addresses concerns in by leveraging with Radon priors, such as minimization on local projections, enabling accurate interior reconstruction from undersampled data (e.g., 1160 projections for sheep ) while suppressing noise. Resolution in Radon-based tomography is limited by projection sampling density; in the angular domain, Nyquist sampling requires at least πN angles for an N-pixel object to avoid aliasing, with undersampling causing edge blurring and artifacts proportional to the angular gap.

In Algebraic Geometry

In algebraic geometry, the Radon transform is defined over projective spaces by integrating sections of sheaves or functions along linear subspaces, with the parameter space given by the Grassmannian of k-planes in \mathbb{P}^n. For the case of lines, it maps a function on \mathbb{P}^n (or more generally, a constructible sheaf) to its integrals over all lines, parameterized by the Grassmannian \mathrm{Gr}(2, n+1). This setup generalizes the classical Euclidean Radon transform to a projective setting, preserving key analytic properties such as injectivity under suitable conditions. A primary application arises in integral geometry for computing MacPherson's Chern classes of singular algebraic , which extend the classical Chern classes to the singular case via a on constructible functions. The topological Radon transform, which combines and operations on sheaves, ensures compatibility with this transformation, allowing the Chern-Mather class of the variety of a projective to be expressed in terms of the original variety's class. Moreover, through integral , the of a variety is recoverable from the Radon integrals over generic linear projections, providing a tool to define these classes globally without resolving singularities. The Guillemin-Sternberg framework analyzes the Radon transform as an elliptic Fourier integral operator associated to a canonical relation, enabling microlocal inversion and range characterization. In the algebraic geometry context, this corresponds to Fourier-Mukai transforms on flag varieties, where both serve as integral functors between derived categories of coherent sheaves or D-modules, intertwining geometric structures with cohomological data. On flag varieties, the Radon transform establishes equivalences between categories of twisted D-modules, acting as intertwining functors that relate representations of semisimple Lie groups to geometric objects via Bernstein's localization theorem. For partial flag varieties, these transforms preserve quasi-equivariance and provide explicit isomorphisms between blocks of category \mathcal{O}, linking integral geometry to representation theory. An illustrative example occurs in toric varieties, where projections onto coordinate subspaces allow reconstruction of algebraic cycles using Radon inversion, leveraging the fan's combinatorial data to recover cycle classes from their Euler characteristics.

Generalizations to Higher Dimensions

The Radon transform generalizes to higher dimensions in \mathbb{R}^n (n > 3) by integrating a function over (n-1)-dimensional hyperplanes, defined as g(\theta, s) = \int_{\{x \in \mathbb{R}^n : x \cdot \theta = s\}} f(x) \, dm(x), where \theta \in S^{n-1}, s \in \mathbb{R}, and dm(x) is the Euclidean volume element on the hyperplane. This extension preserves key properties like injectivity for smooth functions of compact support, but requires moment conditions for inversion, such as the zeroth-order Gelfand-Graev-Helgason-Ludwig condition ensuring the integrals vanish outside certain regions. Inversion formulas, such as those by Mader and Helgason, involve derivatives of back-projected data: for even n, f(x) = A_0 \frac{\partial^n}{\partial t^n} F_0(x, t) \big|_{t=0}, where A_0 = (-1)^{(n-2)/2} \pi (n-2)! / \sigma_{n-2} and F_0 integrates the data with a logarithmic kernel; similar expressions hold for odd n using a sign function kernel. On manifolds, the Radon transform adapts to spaces like spheres or compact groups by integrating over submanifolds with invariant measures. For the sphere S^n, the spherical Radon transform averages functions over great spheres or caps, using the rotationally invariant , which extends the classical case while respecting the manifold's geometry. On compact groups G, such as SU(2) or SO(3), the transform integrates over closed geodesics parametrized by the bi-invariant metric, yielding (Rf)(\gamma) = \int_0^1 f(\gamma(t)) \, dt for periodic geodesics \gamma, with injectivity holding for smooth functions if G has rank at least 2 (e.g., excluding S^1 and S^3). These versions leverage group representation theory for analysis, ensuring the transform intertwines with Fourier-like decompositions on the manifold. Weighted and generalized variants modify the integration measure or submanifold type. The X-ray transform, a line integral version in \mathbb{R}^n, computes I_f(\ell) = \int_\ell f(x) \, ds(x) along lines \ell with arc-length ds, contrasting the full hyperplane integrals of the standard Radon by focusing on endpoints or rays, which is injective under visibility conditions but requires different inversion techniques like filtered backprojection adapted to lower codimensions. Weighted forms incorporate densities, such as attenuated X-ray transforms I_f^\mu(\ell) = \int_\ell f(x) e^{-\int_0^s \mu(\gamma(t)) dt} \, ds, generalizing to non-uniform media while preserving duality with the Radon via exponential factors. In abstract settings, the Radon transform appears in as a map between point and subspace incidence structures, where for a combinatorial geometry over a of zero, the transform's equals the number of points for \geq 2, enabling reconstruction via linear over flats. It can also be viewed functorially in , associating sheaves or representations to incidence relations in geometric categories, though explicit inversions remain tied to the underlying poset . Inversion in non-Euclidean spaces, such as , faces challenges due to overdetermined data and incomplete coverage; for quarter-spherical transforms on S^n, exact formulas exist via and error functions, but numerical stability suffers from ill-posedness in limited angular spans, often requiring regularization for practical .

References

  1. [1]
    Radon Transform -- from Wolfram MathWorld
    The Radon transform is an integral transform used to reconstruct images, like in medical CT scans, and also for mapping polar regions.
  2. [2]
    [PDF] 69, 1917
    SITZUNG VOM 30. APRIL 1917. Über die Bestimmung von Funktionen durch ihre. Integralwerte längs gewisser Mannigfaltigkeiten. ―.
  3. [3]
    [PDF] THE RADON TRANSFORM JOSHUA BENJAMIN III
    The Radon transform for Cn is known as the Penrose transform and it is related to integral geometry (the modern approach to integral geometry was largely ...
  4. [4]
    The Radon Transform and Medical Imaging
    This book surveys the main mathematical ideas and techniques behind some well-established imaging modalities such as X-ray CT and emission tomography.
  5. [5]
    [PDF] The Radon Transform and the Mathematics of Medical Imaging
    The Radon transform offers a means of determining the total density of a certain function f along a given line l. This line l is determined by an angle θ ...
  6. [6]
    Computed Tomography - Medical Imaging Systems - NCBI Bookshelf
    Turning the function values f(x, y) into line integral values p(θ,s) is known as the Radon transform in 2-D. The aim of CT reconstruction is the computation of ...
  7. [7]
    The first 100 years of the Radon transform - IOPscience
    Jul 11, 2018 · The first 100 years of the Radon transform · 1. Johann Radon and his fundamental paper on the Radon transform · 2. On the special issue.
  8. [8]
    [PDF] Applications of the Radon transform - VTechWorks
    Sep 22, 2004 · The first application of the Radon transform focuses on enhancing diffraction patterns in surface recorded seismic data by filtering primary ...
  9. [9]
    Integral Geometry and Radon Transforms - SpringerLink
    In stockIn this text, integral geometry deals with Radon's problem of representing a function on a manifold in terms of its integrals over certain ...
  10. [10]
    [PDF] The Funk–Radon transform and its generalizations
    Johann Radon, 1917: Über die Bestimmung von Funktionen durch ihre Integralwerte längs gewisser Mannigfaltigkeiten. Wanted to reconstruct a function f : R2 ...
  11. [11]
    [PDF] 100 Years of the Radon Transform - RICAM
    Mar 27, 2017 · Jo- hann Radon is well-known for his ground-breaking achievements in Mathematics, such as the Radon- transformation, the Radon-numbers, the ...
  12. [12]
    The Nobel Prize in Physiology or Medicine 1979 - NobelPrize.org
    The Nobel Prize in Physiology or Medicine 1979 was awarded jointly to Allan M. Cormack and Godfrey N. Hounsfield for the development of computer assisted ...Missing: Radon transform
  13. [13]
    [PDF] The ART of tomography In 1979, the Nobel Prize for Medicine and ...
    Cormack, then a Professor at Tufts University, in Massachusetts, developed algorithms based on the. Fourier transform to create an image from X-ray data.
  14. [14]
    [PDF] Discrete Radon Transform - Applied Mathematics
    Abstract-This paper describes the discrete Radon transform (DRT) and the exact inversion algorithm for it. Similar to the discrete Fourier.
  15. [15]
    On the Invertibility of the Discrete Radon Transform - SIAM.org
    The Radon transform is a useful device for analyzing multidimensional data. It is closely connected to what has become known as “projection pursuit.Missing: post | Show results with:post
  16. [16]
    The Radon Transform | SpringerLink
    Bibliographic Information · Book Title: The Radon Transform · Authors: Sigurdur Helgason · Series Title: Progress in Mathematics · Publisher: Springer New York, NY.
  17. [17]
    [PDF] The Radon transform - Homepages of UvA/FNWI staff
    Dec 18, 2014 · The Radon transformed is named after the Austrian mathematician Johann Radon, who studied it in a paper that appeared in 1917. ... Reprint of the ...
  18. [18]
  19. [19]
    [PDF] Application: Tomography - Purdue Engineering
    May 1, 2002 · • The Radon transform of the function f(x, y) is defined as pθ(r) ... Proof of the Fourier Slice Theorem. • By definition pθ(r) ...
  20. [20]
    Sigurdur Helgason - Integral Geometry and Radon Transforms
    This work, initiated by Cormack and Hounsfield and earning them a Nobel Prize, has greatly increased interest in Radon transform theory. The support theorem ...
  21. [21]
    [PDF] Support theorems for the Radon transform and Cramér-Wold ... - arXiv
    Feb 29, 2008 · The dual Radon transform R∗ is defined for ψ ∈ C0(Sd−1×R) by. R∗ψ(x) = ZSd−1 ψ(ω, x · ω)dω. (2.4). If ψ is even, ψ(ω, p) = ψ(−ω,−p) ...
  22. [22]
    [PDF] The Radon Transform in the Plane - Assets - Cambridge University ...
    1.3 The Normal Operator and Singularities. 1.3.1 Normal Operator. We will now proceed to study the normal operator R∗R of the Radon transform, where the ...
  23. [23]
    The Mathematics of Computerized Tomography
    The mathematics of computerized tomography. I. Tomography-Data processing. 2 ... second equivalent definition of the Radon transform in terms of the one-.
  24. [24]
    The Radon Transform - Book - SpringerLink
    The Radon Transform on Rn Pages 1-75 A Duality in Integral Geometry. Generalized Radon Transforms and Orbital Integrals Pages 76-92
  25. [25]
    On the determination of functions from their integral values along ...
    Apr 4, 2018 · PDF | This document contains a facsimile of the original paper on the Radon transform, "Über die Bestimmung von Funktionen durch ihre ...
  26. [26]
    On the Problem of Well-Posedness for the Radon Transform
    About this paper. Cite this paper. Hertle, A. (1981). On the Problem of Well-Posedness for the Radon Transform. In: Herman, G.T., Natterer, F. (eds) ...
  27. [27]
    [PDF] February 18, 2021 1 Outline 2 Ill-posedness of the inverse problem
    We have thus provided a quantitative statement about the ill-possedness of inverting the Radon transform. We can express this in the language of linear algebra.
  28. [28]
    [PDF] The Discrete Picard Condition for Discrete Ill-Posed Problems
    We illustrate the importance of this condition theoretically as well as experimentally. AMS subject classification: 65F30, 65F20. Key words: Ill-posed problems.Missing: Radon | Show results with:Radon
  29. [29]
    Tikhonov-Phillips Regularization of the Radon Transform
    With the help of the singular value decomposition of the Radon transform the Tikhonov-Phillips regularization with suitable Sobolev norm is analytically ...
  30. [30]
    Error analysis for filtered back projection reconstructions in Besov ...
    Error estimates for noisy data. We now consider the case of noisy Radon data. To this end, we follow a deterministic approach and assume that the Radon data ...<|control11|><|separator|>
  31. [31]
    Basics of iterative reconstruction methods in computed tomography
    This review provides an overview of the underlying basic principles of iterative image reconstruction methods currently available for and applied in CT imaging
  32. [32]
    Approximate solution of systems of linear equations†
    Approximate solution of systems of linear equationst. S. KACZMARZ. Although the approximate solution of the equation f(x) = 0 with an un- known is shown with ...
  33. [33]
    Algebraic Reconstruction Techniques (ART) for three-dimensional ...
    We give a new method for direct reconstruction of three-dimensional objects from a few electron micrographs taken at angles which need not exceed a range of 60 ...
  34. [34]
    The standard forms and convergence theory of the Kaczmarz ...
    Dec 15, 2023 · As a famous image reconstruction method in computerized tomography, the Kaczmarz method is simple and easy to implement, but its convergence ...
  35. [35]
  36. [36]
    [PDF] Iterative Methods in regularized tomographic reconstruction
    Mar 14, 2018 · shows that the Radon transform is a shift-invariant operator, so its least-squares inverse can be efficiently computed – at least in the ...
  37. [37]
    [PDF] Reconstruction of CT Images by Iterative Least Squares Methods ...
    In this paper, we propose image-reconstruction methods using non-negative matrix factorization (NMF) [5, 6] and iterative least squares methods. As the solution ...
  38. [38]
    [PDF] CT Reconstruction From Parallel and Fan-Beam Projections by a 2 ...
    In addition, it is shown in [5] that the discrete Radon transform (DRT) converges to the con- tinuous Radon transform as the discretization step goes to zero.
  39. [39]
    Sir Godfrey Hounsfield - PMC - NIH
    Short abstract. Engineer who invented computed tomography and won the Nobel prize for medicine. Sir Godfrey Hounsfield invented the computed tomographic scanner ...
  40. [40]
    Image Characteristics Quizzes
    First, the backprojection operator is not the inverse of the Radon transform (technically, it is the adjoint of the radon transform). Furthermore, only a ...
  41. [41]
    A Radon diffraction theorem for plane wave ultrasound imaging
    Feb 9, 2023 · A Radon diffraction theorem for plane wave ultrasound imaging ... Radon transform might as well be applied to image data in post-processing.
  42. [42]
  43. [43]
    [PDF] Tomographic Image Reconstruction - AAPM
    A typical slice image and its Radon transform are shown below: Figure 2. Shepp-Logan phantom and its Radon transform. The task of tomographic reconstruction ...
  44. [44]
    Compressed Sensing Based Interior Tomography - PMC
    Based on the compressed sensing theory, here we prove that if an object under reconstruction is essentially piecewise constant, a local ROI can be exactly and ...
  45. [45]
    The Radon transform with finitely many angles - IOPscience
    Aug 29, 2023 · We study the Radon transform in the plane in parallel geometry possibly undersampled in the angular variable. We study resolution, aliasing artifacts, and edge ...
  46. [46]
    radon transforms on higher rank grassmannians - Project Euclid
    We define a Radon transform R from functions Gr(&, A?), the Grassmannian of projective A:-planes in CP" to functions on Gr(/, n). If / e C°°(Gr(A:,n)).
  47. [47]
    On topological Radon transformations - ScienceDirect.com
    Aug 28, 1997 · We show that the Chern-Schwartz-MacPherson transformation makes the topological Radon transform of constructible functions compatible with a ...
  48. [48]
    Radon transforms for quasi-equivariant D-modules on generalized ...
    Nov 14, 1999 · Abstract: In this paper we deal with Radon transforms for generalized flag manifolds in the framework of quasi-equivariant D-modules.Missing: Guillemin- Sternberg Fourier- Mukai
  49. [49]
    Radon transforms of twisted D-modules on partial flag varieties - arXiv
    Jan 10, 2017 · In this paper we study intertwining functors (Radon transforms) for twisted D-modules on partial flag varieties and their relation to the representations of ...Missing: algebraic geometry
  50. [50]
    None
    ### Summary of Hyperplane Radon Transform in R^n, Mader-Helgason Inversion Formulas, and Moment Conditions
  51. [51]
    The Radon Transform | SpringerLink
    Free delivery 14-day returnsAug 1, 1999 · The Radon transform is an important topic in integral geometry which deals with the problem of expressing a function on a manifold in terms of its integrals ...
  52. [52]
  53. [53]
    Generalizations and Variants of the Radon Transform (Chapter 5)
    There are two main ways of generalizing the Radon transform. One can add a weight to the hyperplane integral or one can integrate over subvarieties that are not ...
  54. [54]
    [PDF] The weighted X-ray transform and applications - UCSB Math
    We review both attenuated X-ray transforms and. X-ray transforms with general weights, in particular the matrix version, with emphasis on the approaches using ...
  55. [55]
    The radon transforms of a combinatorial geometry, I - ScienceDirect
    Over a field of characteristic zero, the rank of the point-copoint incidence matrix of a combinatorial geometry of rank ⩾ 2 equals the number of points.
  56. [56]
    Inversion formulas for quarter-spherical Radon transforms
    ### Summary of Inversion Challenges for Spherical Radon Transform in Non-Euclidean Spaces