Fact-checked by Grok 2 weeks ago

Edge detection

Edge detection is a fundamental technique in and that identifies and locates boundaries within an image by detecting significant local changes in , typically arising from discontinuities in the image's brightness or its first . These edges often correspond to physical phenomena such as object outlines, surface texture variations, depth discontinuities, or illumination changes. As a core preprocessing step, edge detection enables higher-level image analysis tasks, including feature extraction, , segmentation, and registration. It plays a pivotal role in diverse applications such as for tumor boundary identification, autonomous vehicle navigation for obstacle detection, 3D scene reconstruction, motion , and image enhancement or compression. Challenges in edge detection include handling , which can produce false edges, and achieving precise localization amid varying or complex scenes. Most edge detection algorithms follow a structured pipeline involving noise reduction through smoothing (often via Gaussian filters), enhancement of intensity gradients, and thresholding to pinpoint edge pixels. First-order derivative methods, which approximate the image gradient, include the Sobel operator—employing 3x3 convolution kernels to emphasize horizontal and vertical changes—and the Prewitt operator, a similar gradient-based approach. The Canny edge detector, introduced in 1986, stands out for its optimality, combining Gaussian smoothing, gradient computation, non-maximum suppression, and hysteresis thresholding to achieve high signal-to-noise ratios, accurate edge positioning, and reduced multiple responses. Second-order methods, like the Laplacian of Gaussian, apply Gaussian filtering followed by the Laplacian operator to detect zero-crossings indicative of edges. Contemporary advancements extend these with multi-scale to capture edges at varying resolutions, energy-based models for local feature emphasis, or of for noise-robust detection in real-world scenarios. Despite progress, no universal algorithm excels across all contexts, as performance depends on image characteristics and downstream needs.

Fundamentals

Motivations

Edge detection is fundamental in image processing and because it identifies boundaries between regions in an image that differ in properties such as , color, or , thereby highlighting significant structural changes. These boundaries, often corresponding to object outlines or material transitions, provide essential low-level cues for higher-level interpretation of visual data. The primary role of edge detection lies in feature extraction, which supports core tasks including , , and shape analysis. By isolating these boundaries, algorithms can simplify complex images into representations that facilitate subsequent processing, such as matching detected edges to known models for identification or partitioning scenes into meaningful regions. This process is pivotal in enabling machines to interpret visual scenes akin to human perception, reducing while preserving critical information. Historically, edge detection emerged in the 1960s as part of early efforts to analyze scenes through , with pioneering work by Larry Roberts on interpreting line drawings from images. A seminal theoretical foundation was laid by David Marr's vision framework in , which posited edges as key primitives for building hierarchical representations of visual information, from raw primal sketches to object-centered descriptions. This theory underscored the motivation to mimic biological vision systems for robust scene understanding. In practical applications, edge detection is indispensable across domains. In , it enables precise delineation of tumor boundaries in MRI or scans, aiding and planning by isolating pathological regions from healthy tissue. For autonomous vehicles, it supports lane detection by extracting road markings from camera feeds, enhancing navigation safety and path correction in real-time. Similarly, in , edge detection separates objects from backgrounds in visual inputs, facilitating tasks like obstacle avoidance, grasping, and environmental mapping in mobile systems.

Edge Properties

Edges in digital images are characterized by discontinuities in intensity, manifesting in various forms depending on the underlying scene and imaging process. Ideal step edges represent abrupt changes in pixel intensity between two homogeneous regions, such as at sharp object boundaries. Ramp edges, in contrast, exhibit a gradual transition in intensity over a finite distance, often resulting from the smoothing effects on step edges. Roof edges feature symmetric intensity peaks, where the brightness rises and falls sharply, typically occurring at line discontinuities or specular highlights on surfaces. Key properties of edges include their , strength, and thickness. Orientation describes the perpendicular to the edge, derived from the gradient vector's angle, which indicates the local alignment. Strength quantifies the edge's prominence through the , measuring the rate of intensity change across the . Thickness refers to the spatial extent of the intensity transition, which can vary from a single in ideal cases to broader profiles influenced by factors like or preprocessing. Imaging conditions significantly alter edge characteristics, particularly through blur mechanisms. Defocus blur, arising from optical imperfections, spreads edge transitions, converting sharp step edges into wider ramp profiles. , caused by relative movement between the camera and scene during exposure, similarly elongates edges along the direction of motion, increasing their effective thickness and reducing sharpness. The quality of detected edges is evaluated using metrics that assess fidelity to true boundaries. Localization accuracy measures how precisely the detected edge position aligns with the actual discontinuity, often aiming for sub-pixel through techniques. Multiple responses occur when a single true edge produces several nearby detections, typically due to thick profiles, leading to redundant outputs. False detections, or erroneous edge identifications, arise from or , inflating the error rate in edge maps. These criteria, emphasizing low false alarm rates, precise positioning, and singular responses per edge, guide the design of effective edge detectors.

Simple Edge Model

In the simplest mathematical representation of an edge in image processing, a one-dimensional (1D) edge is modeled as an step discontinuity in . This is captured by the unit step function u(x), defined as u(x) = 0 for x < 0 and u(x) = 1 for x \geq 0. The profile across such an edge is then expressed as f(x) = I_0 + A \cdot u(x), where I_0 represents the background level on one side of the edge, and A denotes the contrast or amplitude of the change. This model assumes a perfect, instantaneous transition, serving as a foundational idealization for understanding edge detection principles. The derivatives of this profile highlight key properties: the first derivative is f'(x) = A \cdot \delta(x), where \delta(x) is the , producing a sharp peak at the edge location. The second derivative follows as f''(x) = A \cdot \delta'(x), the derivative of the , which exhibits a positive lobe followed by a negative lobe, crossing zero at the edge. Extending this to two dimensions (2D), an edge is conceptualized as a line or curve in the image plane where the intensity changes abruptly in the direction perpendicular to the edge, manifesting as a gradient vector normal to the edge tangent. This perpendicular gradient captures the directional nature of the intensity transition in the image function I(x, y).

Challenges

One of the primary challenges in edge detection arises from noise sensitivity, particularly , which can introduce false edges or attenuate the strength of genuine ones by corrupting the intensity gradients essential for identification. This issue is exacerbated in real-world images where noise from sensor limitations or environmental factors reduces the signal-to-noise ratio, leading to erroneous detections that mimic true boundaries. Illumination variations and shadows pose another significant difficulty, as uneven lighting creates intensity discontinuities that generate spurious edges unrelated to object boundaries. Shadows, in particular, block light and produce strong, artifactual contrasts that traditional gradient-based methods often misinterpret as valid edges, complicating accurate localization in scenes like outdoor environments or complex indoor settings. Blur effects, stemming from optical imperfections or motion during capture, further degrade edge sharpness by broadening transitions and diluting gradient magnitudes, making precise detection harder. Such defocus or motion blur convolves the ideal step edge model with a point-spread function, shifting detected positions away from true inflections and increasing delocalization errors in practical applications. In textured regions or areas with occlusions, edge detection faces ambiguities that result in fragmented or incomplete contours, as fine-scale patterns mimic edges while partial overlaps obscure underlying boundaries. These challenges arise because texture-induced variations blend with true discontinuities, and occlusions introduce T-junctions or depth cues that fragment responses, deviating from the simple edge model's assumption of isolated, step-like changes. A fundamental trade-off exists between detection accuracy (minimizing false negatives and positives) and localization precision, where efforts to suppress noise for reliable detection often widen smoothing kernels, thereby reducing sub-pixel accuracy in edge positioning. This uncertainty principle, inherent to derivative-based approaches, means optimizing one criterion typically compromises the other, limiting overall performance in noisy or complex scenes.

Theoretical Foundations

Gradient-Based Detection

Gradient-based edge detection relies on the principle that edges correspond to abrupt changes in image intensity, manifesting as local maxima in the magnitude of the first-order gradient. For a continuous image function I(x, y), the gradient is the vector \nabla I = \left( \frac{\partial I}{\partial x}, \frac{\partial I}{\partial y} \right), and its magnitude is computed as |\nabla I| = \sqrt{ \left( \frac{\partial I}{\partial x} \right)^2 + \left( \frac{\partial I}{\partial y} \right)^2 }, where the magnitude quantifies edge strength and the direction \theta = \tan^{-1} \left( \frac{\partial I / \partial y}{\partial I / \partial x} \right) indicates edge orientation. This approach assumes edges align with the direction of maximum intensity variation, enabling localization by seeking peaks in |\nabla I|. In discrete digital images, partial derivatives are approximated using finite differences to handle pixelation. The forward difference estimates \frac{\partial I}{\partial x} \approx I(x+1, y) - I(x, y), the backward difference uses \frac{\partial I}{\partial x} \approx I(x, y) - I(x-1, y), and the central difference provides a more accurate symmetric approximation \frac{\partial I}{\partial x} \approx \frac{I(x+1, y) - I(x-1, y)}{2}, with analogous forms for the y-direction. These can be realized through convolution with compact kernels, distinguishing between isotropic filters, which treat all directions uniformly, and separable filters, which independently convolve horizontal and vertical components for efficiency. An early isotropic example is the Roberts cross operator, employing 2x2 diagonal kernels to approximate gradients as G_x = I(x, y) - I(x+1, y+1) and G_y = I(x+1, y) - I(x, y+1), with magnitude |G| = |G_x| + |G_y|; introduced by Lawrence Roberts in 1963 for three-dimensional solid perception, it offers simple implementation but is sensitive to noise due to its small support. To produce binary edge maps from continuous gradient magnitudes, thresholding is applied to classify pixels as edge or non-edge based on a predefined value, suppressing weak responses while retaining strong ones. Hysteresis thresholding refines this conceptually by employing dual thresholds—a high threshold for definitive strong edges and a low threshold for potential weak edges—where weak edges are retained only if connected to strong ones via , thereby suppressing isolated noise while preserving continuous edge structures. Modern gradient enhancements address limitations in detecting anisotropic edges, where gradient direction varies locally, by incorporating the structure tensor—a 2x2 symmetric matrix J = \begin{pmatrix} \langle I_x^2 \rangle & \langle I_x I_y \rangle \\ \langle I_x I_y \rangle & \langle I_y^2 \rangle \end{pmatrix}, formed from averaged outer products of image derivatives over a Gaussian-weighted neighborhood. Its eigenvalues and eigenvectors reveal edge coherence and orientation, improving robustness to noise and enabling better handling of textured or curved edges compared to scalar magnitudes. This tensor-based approach, building on early formulations, has become influential in refining first-order detection for complex scenes.

Second-Order Derivative Methods

Second-order derivative methods for edge detection rely on the second spatial derivative of the image intensity function to identify edges at points of inflection, where the intensity profile changes curvature, typically detected as zero-crossings in the derivative response. Unlike first-order approaches that highlight edge strength via gradient magnitude, these methods emphasize precise localization by capturing the transition from positive to negative concavity in the intensity surface. The core operator in these methods is the Laplacian, defined as the sum of the second partial derivatives along the x and y axes: \nabla^2 I = \frac{\partial^2 I}{\partial x^2} + \frac{\partial^2 I}{\partial y^2}. Edges correspond to the zero-crossings of this operator applied to the image, where the second derivative passes through zero, indicating a change in the sign of the intensity curvature. However, the Laplacian is highly sensitive to noise, as second derivatives amplify high-frequency components, leading to numerous false edges. To mitigate noise while preserving edge localization, the Laplacian is often convolved with a Gaussian smoothing kernel, yielding the Laplacian of Gaussian (LoG) filter. The LoG combines multi-scale analysis by scaling the Gaussian with parameter \sigma, which controls the trade-off between localization accuracy and noise suppression; larger \sigma values detect broader edges but smooth finer details. The 2D LoG kernel is given by: g(x,y) = -\frac{1}{\pi \sigma^4} \left(1 - \frac{x^2 + y^2}{2\sigma^2}\right) \exp\left(-\frac{x^2 + y^2}{2\sigma^2}\right) Zero-crossings in the LoG response then delineate edges across scales. In discrete implementations, the LoG is frequently approximated by the Difference of Gaussians (DoG), which subtracts two Gaussian filters of differing scales: one low-pass filtered version from another with a slightly larger \sigma. This approximation is computationally efficient, as it avoids direct computation of second derivatives, and closely mimics the LoG's bandpass characteristics for edge enhancement. These methods offer precise subpixel edge localization due to the sharpness of zero-crossings, making them suitable for applications requiring accurate boundary detection. However, they can produce multiple responses (such as echoes or halos) to a single edge, particularly at step-like discontinuities, complicating edge linking and increasing false positives in noisy images. The theoretical foundations of second-order derivative methods, including the , were proposed by in their seminal 1980 work on computational vision.

Phase Congruency Principles

Phase congruency was originally proposed by in 1988 as a phase-dependent energy model for feature detection in human vision. It represents a fundamental principle in edge detection, positing that image features, such as edges, occur at locations where the Fourier components of the signal are maximally aligned in phase. This concept originates from the observation that discontinuities in an image, like step edges, cause the phases of sinusoidal components across multiple scales and orientations to converge, maximizing their constructive interference. Unlike traditional intensity-based methods, phase congruency treats edges as points of high phase alignment rather than abrupt changes in amplitude, providing a more biologically plausible model inspired by human visual processing. The measure of phase congruency is defined as PC(x) = \frac{ \sum_n A_n \left[ \cos(\phi_n - \bar{\phi}) - \left| \sin(\phi_n - \bar{\phi}) \right| \right] - T }{ \sum_n A_n + \epsilon }, where A_n denotes the amplitude of the n-th Fourier component, \phi_n its phase, \bar{\phi} the moment-weighted average phase angle, T a noise compensation threshold, and \epsilon a small constant to avoid division by zero (with the term in brackets set to zero if negative). This formulation normalizes the aligned phase contributions against the total amplitude, yielding a dimensionless value between 0 and 1, with higher values indicating stronger feature significance. Edges are then identified at local maxima of this measure across scales. A key advantage of phase congruency is its invariance to variations in illumination and contrast, as it depends solely on phase relationships rather than absolute intensity gradients. In contrast to gradient-based detectors, which can produce false edges under uneven lighting, phase congruency maintains consistent detection thresholds regardless of global brightness changes, making it robust for real-world images. This property arises because phase alignment is preserved under monotonic transformations of pixel values. Peter Kovesi advanced the practical implementation of phase congruency in 1999, employing log-Gabor filters to perform multi-scale and orientation-selective analysis. Log-Gabor filters, which have Gaussian amplitude spectra on a logarithmic frequency scale, allow efficient capture of a wide range of frequencies without the limitations of traditional Gabor filters, such as poor representation of high frequencies. By convolving the image with filter banks at multiple scales (e.g., wavelengths of 3, 6, 12, and 24 pixels) and orientations (e.g., 0°, 30°, 60°, 90°, 120°, 150°), Kovesi's approach computes phase congruency robustly, followed by non-maximum suppression to localize edges precisely. This method has become a standard for phase-based feature detection due to its computational efficiency and accuracy.

Classical Algorithms

Sobel, Prewitt, and Scharr Operators

The , introduced in 1968 by and , approximates the gradient of an image intensity function using two 3×3 discrete convolution kernels, one for the horizontal direction G_x and one for the vertical direction G_y. The kernels incorporate central weighting to provide some smoothing in the direction perpendicular to the edge, reducing noise sensitivity compared to simpler differencing. Specifically, the horizontal kernel is given by G_x = \frac{1}{8} \begin{bmatrix} -1 & 0 & 1 \\ -2 & 0 & 2 \\ -1 & 0 & 1 \end{bmatrix}, and the vertical kernel is the transpose, G_y = \frac{1}{8} \begin{bmatrix} -1 & -2 & -1 \\ 0 & 0 & 0 \\ 1 & 2 & 1 \end{bmatrix}. The Prewitt operator, developed in 1970 by , employs similar 3×3 kernels but with uniform weighting, omitting the diagonal emphasis present in the Sobel operator, which results in less smoothing along edges. The horizontal kernel is G_x = \frac{1}{3} \begin{bmatrix} -1 & 0 & 1 \\ -1 & 0 & 1 \\ -1 & 0 & 1 \end{bmatrix}, and the vertical kernel is the transpose, G_y = \frac{1}{3} \begin{bmatrix} -1 & -1 & -1 \\ 0 & 0 & 0 \\ 1 & 1 & 1 \end{bmatrix}. The Scharr operator, proposed in 2000 by , refines these approximations with kernels optimized for better rotational invariance and isotropy, particularly in the Fourier domain, by minimizing weighted mean squared angular error. Its horizontal kernel is G_x = \frac{1}{32} \begin{bmatrix} -3 & 0 & 3 \\ -10 & 0 & 10 \\ -3 & 0 & 3 \end{bmatrix}, and the vertical kernel is the transpose, G_y = \frac{1}{32} \begin{bmatrix} -3 & -10 & -3 \\ 0 & 0 & 0 \\ 3 & 10 & 3 \end{bmatrix}. Compared to the , the exhibits reduced angular bias, providing more accurate gradient direction estimates across various orientations, while the offers the simplest computation but with comparable isotropy limitations to Sobel. All three operators are susceptible to noise amplification due to their differentiator nature, often requiring pre-smoothing with a for robust performance. In practice, these operators compute the gradient components by convolving the image with G_x and G_y, yielding the edge magnitude as \sqrt{G_x^2 + G_y^2} and direction as \atan2(G_y, G_x) to produce edge maps highlighting intensity transitions.

Canny Edge Detector

The , introduced in 1986, represents a seminal multi-stage algorithm for first-order edge detection that optimizes performance through a combination of noise reduction, gradient analysis, and edge linking. Developed by , it addresses key challenges in edge detection by prioritizing three main criteria: low error rate in detecting true edges while minimizing false positives, good localization to ensure detected edges are as close as possible to the actual edge positions, and a single response per edge to avoid multiple detections along the same boundary. These criteria were derived from a computational framework that models edges as step discontinuities in image intensity, allowing the algorithm to perform robustly on real-world images with varying noise levels. The algorithm proceeds in five primary stages to achieve these goals. First, the input image is convolved with a to smooth out noise, controlled by a standard deviation parameter σ that balances noise suppression with edge preservation; larger σ values reduce noise more effectively but can blur fine edges. Second, the intensity gradient is computed to identify potential edges, typically using finite difference operators like the for approximating the first derivatives in the x and y directions, yielding gradient magnitude and direction maps. Third, non-maximum suppression thins the edges by retaining only the local maxima in the gradient magnitude along the direction perpendicular to the edge, ensuring single-pixel-wide responses. Fourth, a double-thresholding process applies low (T1) and high (T2) thresholds to the suppressed gradient magnitudes, classifying pixels as strong edges (above T2), weak edges (between T1 and T2), or non-edges (below T1). Finally, edge tracking by hysteresis connects weak edges to strong ones, suppressing isolated weak edges to form continuous contours while eliminating noise-induced fragments. Parameter tuning is central to the Canny detector's effectiveness, with σ influencing the scale of detected edges and the T1/T2 ratio (often around 2:1 to 3:1) controlling sensitivity to faint edges; optimal values are typically selected empirically based on image characteristics to minimize localization error. The algorithm's strengths lie in its noise robustness, achieved through Gaussian smoothing and hysteresis, which outperform simpler gradient-based methods in noisy environments by reducing false edges while preserving true ones. However, it exhibits weaknesses in parameter sensitivity, where suboptimal σ or thresholds can lead to missed edges or excessive fragmentation, and its computational complexity makes it slower than basic operators for real-time applications. To mitigate threshold sensitivity, adaptive thresholding variants have been proposed, which dynamically adjust T1 and T2 based on local image statistics like gradient histograms or entropy measures, improving performance on images with non-uniform illumination without manual tuning. For instance, one such approach employs bilateral filtering prior to gradient computation and adapts thresholds regionally to enhance edge continuity in varied lighting conditions. These extensions maintain the core hysteresis mechanism but extend applicability to complex scenes, though they increase computational demands.

Marr-Hildreth Algorithm

The Marr-Hildreth algorithm represents a foundational approach in edge detection, leveraging second-order derivatives to identify boundaries in images through zero-crossing analysis. Developed within the framework of computational vision, it processes images by applying the (LoG) filter, which integrates Gaussian smoothing to reduce noise while computing the Laplacian to highlight intensity transitions. This method detects edges as points where the filtered intensity crosses zero, indicating a change in the sign of the second derivative. The algorithm's design emphasizes multi-scale analysis to capture edges robustly across varying resolutions, distinguishing it as a key contribution to early computer vision theory. The core process begins with convolving the input image with the at multiple spatial scales, determined by the standard deviation σ of the Gaussian component. At each scale, zero-crossings are located by examining the convolved image for sign changes between adjacent pixels, often using a 3x3 neighborhood to verify transitions along principal directions. These isolated zero-crossings are then linked across scales to form continuous edge segments: starting from the coarsest scale to outline major structures, subsequent finer scales refine details and resolve fragmented edges. This hierarchical linking mitigates multiple responses to the same edge by propagating connectivity and selecting the most consistent paths. The , as a second-order derivative method, enables this by approximating the after , providing a noise-tolerant basis for zero-crossing detection. Scale selection in the Marr-Hildreth algorithm is guided by the need to match the filter's response to the actual scale of intensity changes in the image, using the smallest σ where the edge appears step-like for accurate localization. Processing begins at the largest scale to detect broad features, then progresses to smaller scales for localization, ensuring that coarse-to-fine integration avoids over-segmentation from noise-induced responses. This multi-channel approach, involving several σ values (e.g., increasing by factors of √2), constructs a scale-space representation that aligns with natural image properties. Theoretically, the algorithm forms a critical component of David Marr's primal sketch, the initial stage in his hierarchical model of visual processing, which posits that edges represent localized intensity discontinuities essential for higher-level scene understanding. Introduced in Marr's 1980 paper with Ellen Hildreth, it draws from neurophysiological insights into human vision, aiming to replicate how the visual cortex detects boundaries through derivative computations. In Marr's broader 1982 vision theory, the primal sketch aggregates such edges into symbolic descriptions for object recognition. Despite its elegance, the Marr-Hildreth algorithm exhibits limitations, particularly its sensitivity to noise at finer scales, which can produce spurious zero-crossings or halo artifacts around true edges. It often requires post-processing, such as curve fitting or hysteresis thresholding, to achieve edge closure and eliminate disconnected segments. For implementation, zero-crossings are validated by thresholding their strength—typically the absolute value of the Laplacian or the gradient magnitude at the crossing—against a fraction (e.g., 10-20%) of the image's maximum Laplacian response, ensuring only significant edges are retained.

Kovalevsky Method

The Kovalevsky method represents a topological approach to edge detection, treating the digital image as a 2D manifold composed of pixels arranged in 8-connected regions. Edges are identified through boundary extraction, where the boundaries separate distinct regions such as objects and backgrounds, leveraging the inherent connectivity of the pixel grid without relying on intensity gradients or derivative computations. This principle draws from digital geometry and cellular topology, enabling the delineation of closed boundaries that preserve the topological integrity of the image structures. The algorithm operates via iterative region growing initiated from seed pixels, which are selected as representatives of interior regions. Starting from these seeds, neighboring pixels are examined using simple adjacency rules to classify each pixel as interior (connected to the seed via a path of similar pixels), exterior (not connected), or boundary (pixels adjacent to both interior and exterior regions). This process expands regions progressively, merging compatible areas while extracting boundary pixels along the interfaces, resulting in a set of boundary curves that can be encoded for further analysis. The method avoids explicit thresholding by relying solely on connectivity criteria, making it suitable for binary images where regions are predefined by pixel labels. Key advantages include the absence of gradient calculations or noise-sensitive thresholding, allowing the method to handle thick edges naturally by resolving them into precise boundaries during region expansion. It demonstrates robustness to noise, as minor perturbations do not disrupt the connectivity-based classification, and it performs efficiently in industrial inspection tasks where reliable boundary detection is critical. Developed by Vladimir Kovalevsky in 1980, the approach was initially applied to binary image processing for shape analysis in automated systems.

Advanced and Modern Techniques

Subpixel Localization

Subpixel localization refines the positions of edges detected at the integer-pixel level to achieve fractional-pixel accuracy, enabling higher precision in image analysis tasks where discrete grid limitations would otherwise introduce quantization errors. This process typically operates on the output of initial edge detectors, such as gradient-based operators, by modeling the local intensity variation across the edge profile. One common approach involves interpolating the gradient magnitude profile perpendicular to the detected edge direction. For instance, quadratic interpolation fits a parabola to the gradient magnitudes at the peak pixel and its immediate neighbors along the edge normal, solving for the subpixel offset that maximizes the fitted curve. This method assumes a smooth, unimodal response in the gradient magnitude and can yield accuracies on the order of 0.1 pixels under moderate noise conditions. Moment-based techniques treat the edge as a step function and compute the centroid of the local intensity distribution to estimate its position. By calculating the first-order spatial moment (centroid) and zero-order moment (total intensity change) within a small window around the edge pixel, the subpixel location is determined as the ratio of these moments, providing robustness to noise through integration over the neighborhood. This approach, invariant to uniform scaling and translation in intensity, achieves subpixel precision by fitting the moments to an ideal ramp edge model. The Taylor expansion method approximates the true edge position by expanding the intensity or gradient function around the integer-pixel maximum. Assuming the edge response f(x) reaches its extremum near a discrete sample x_0, the subpixel offset \delta x is given by: \delta x = -\frac{f'(x_0)}{f''(x_0)} where f'(x_0) and f''(x_0) are the first and second derivatives estimated from neighboring pixels. This second-order approximation corrects for the discrete sampling offset and is computationally efficient for 1D profiles extracted from 2D edges. Iterative methods, such as the Lyvers approach, enhance accuracy by modeling the local image patch as a planar facet and iteratively refining the 1D edge profile along the estimated normal direction. Starting from an initial gradient-based edge point, the algorithm computes moments of the projected intensity line to update the position, orientation, and contrast until convergence, often requiring only a few iterations for subpixel results. This facet-based fitting accounts for slight tilts in the edge plane, improving localization in textured regions. These techniques find application in high-precision domains like stereo vision, where subpixel edge correspondences enhance depth estimation accuracy, and metrology, where they support micron-level measurements of manufactured parts by refining boundary positions beyond camera resolution limits.

Edge Thinning and Linking

Edge thinning and linking constitute essential post-processing steps in edge detection pipelines, aimed at refining binary edge maps into thin, continuous curves suitable for higher-level analysis such as object recognition or boundary tracing. Thinning reduces thick edge regions to single-pixel-wide skeletons while preserving the overall topology and connectivity of the original structures, whereas linking bridges gaps between fragmented edge segments to form coherent contours. These techniques address common artifacts like edge thickening from gradient approximations and discontinuities arising from noise or thresholding variability, ultimately facilitating applications in vectorization and shape representation. Thinning algorithms typically employ iterative morphological erosion, where boundary pixels are successively removed based on local neighborhood configurations to achieve a one-pixel-thick representation without altering the object's fundamental shape or branching points. A seminal approach is the , which operates in parallel sub-iterations: the first targets southeast and northwest boundary points for deletion, while the second addresses northeast and southwest ones, using simple 3x3 neighborhood rules to ensure topology preservation—such as maintaining the number of connected components and avoiding spurious breaks. This method, proposed in 1984, is particularly efficient for binary images and has been widely adopted due to its simplicity and guaranteed single-pixel skeletons for elongated patterns. These iterative processes continue until no further pixels can be safely removed, yielding skeletons that capture medial axes essential for feature extraction. The historical roots of edge thinning trace back to 1970s skeletonization techniques, which formalized parallel reduction operations on digital pictures to extract curve-like representations while conserving topological properties like Euler numbers. Early characterizations, such as those by Rosenfeld in 1975, defined conditions for topology-preserving deletions in 8-connected grids, laying the groundwork for subsequent parallel algorithms that avoid excessive erosion or distortion. Edge linking reconnects disjoint segments by evaluating potential bridges based on spatial proximity, directional continuity, and gradient consistency, often modeling edges as graphs where pixels or short segments serve as nodes. Common methods include tracking algorithms that propagate from seed points along predicted directions derived from local gradient orientations, filling small gaps if the angle deviation remains below a threshold (e.g., 10-20 degrees) and intensity gradients align. Graph-based approaches, such as union-find structures, efficiently merge nearby segments by treating edge pixels as sets and unioning them when proximity and collinearity criteria are met, achieving near-linear time complexity for large images. For instance, linear-time union-find variants tailored for image processing resolve equivalences during region growing, adaptable to edge contours by prioritizing merges along consistent curvatures. Gap filling further relies on criteria like curvature matching—where interpolated paths minimize bends beyond local edge tangents—or gradient direction consistency, ensuring linked segments form smooth, plausible boundaries without introducing artifacts. These refinement steps mitigate fragmentation induced by detection thresholds, producing cleaner edges that enhance accuracy in downstream tasks like vectorization for CAD or GIS applications. In the Canny detector, hysteresis thresholding provides a basic form of linking by connecting weak edges to strong ones, but dedicated thinning and linking extend this to more robust contour completion.

Phase Stretch Transform

The Phase Stretch Transform (PST) is a physics-inspired computational method for edge detection that amplifies edges by stretching the phase of an image's analytic representation, effectively converting the theoretically infinite energy at ideal edges into finite, detectable energy. Developed to emulate the dispersive effects of light propagation through a nonlinear medium, PST enhances high-frequency components associated with intensity transitions while suppressing noise in lower frequencies. This approach draws loosely from principles by leveraging phase information but introduces a nonlinear warping to better isolate features in complex images. The process begins with computing the analytic image of the input intensity image, typically via convolution with a kernel to extract the local phase φ at each pixel. A phase stretching operation is then applied using tunable parameters S (stretch strength) and W (warp parameter) that control the degree of nonlinearity and sensitivity to edges. The phase stretching is applied in the frequency domain as \phi(\omega) = S \ln\left(1 + \frac{|\omega|}{W}\right), where ω is the radial frequency; this logarithmic transformation warps the phase to amplify abrupt changes. The resulting stretched phase map is thresholded, often with an adaptive strategy based on local statistics, and may undergo morphological operations like thinning to produce a clean binary edge map. PST offers advantages in real-time processing due to its efficient frequency-domain implementation via fast Fourier transforms, making it suitable for dynamic scenes where traditional gradient-based detectors like may struggle with motion blur or varying illumination. It was introduced and patented by Asghari and Jalali in 2015. Applications include video surveillance for detecting moving objects in cluttered environments and biomedical imaging, such as enhancing edges in histology slides or MRI scans to improve feature segmentation without excessive preprocessing.

Learning-Based Methods

Learning-based methods for edge detection represent a paradigm shift from handcrafted filters to data-driven approaches, leveraging machine learning to learn edge representations directly from annotated images. Early efforts employed supervised classifiers, such as logistic regression models trained on local features like brightness, color, and texture gradients, to predict boundary probabilities at pixel locations. These methods improved upon classical operators by incorporating contextual cues but were limited by shallow architectures and manual feature engineering. The advent of deep learning marked a significant evolution, with the first convolutional neural network (CNN) specifically designed for edge detection, DeepEdge, introduced in 2015 as a multi-scale bifurcated network that fuses low-level contours with high-level object detection features for top-down contour prediction. Subsequent CNN-based models advanced this foundation by addressing multi-scale integration and supervision strategies. The Holistically-Nested Edge Detection (HED) framework, proposed in 2015, utilized a fully convolutional network with side outputs at multiple layers, enabling end-to-end image-to-image prediction of edges while fusing hierarchical features through a weighted fusion layer. Building on this, the Richer Convolutional Features (RCF) model from 2017 extended VGG16 architecture to capture richer multi-level features, employing a novel convolutional neural network that sorts and fuses all convolutional outputs into a single edge probability map, achieving superior performance on natural images with diverse scales. Training these models typically involves pixel-wise supervision, where each pixel is labeled as edge or non-edge using datasets like , which comprises 500 natural images with human-annotated boundaries divided into training, validation, and test sets. Loss functions often combine binary cross-entropy for classification with edge-specific terms, such as class-balanced variants in to handle imbalanced edge/non-edge pixels, or multi-scale supervision in to emphasize boundary precision across resolutions. Recent developments have focused on efficiency and novel architectures, exemplified by PiDiNet in 2021, which introduces pixel difference convolutions to integrate traditional edge operators into lightweight CNNs, enabling real-time detection at over 100 FPS on while rivaling human-level accuracy without pretraining on large datasets like . Transformer-based approaches, such as from 2022, further enhance global context capture by employing in a two-stage framework, where a transformer encoder processes multi-scale features followed by a decoder for crisp boundary refinement, demonstrating improved handling of long-range dependencies in complex scenes. These methods excel in managing intricate textures, occlusions, and varying lighting conditions in real-world images, outperforming classical algorithms on benchmarks like BSDS500 with ODS F-scores exceeding 0.80 for models like RCF and PiDiNet. However, they require large volumes of annotated data for training, which is labor-intensive to curate, and incur high computational costs during inference, particularly for transformer variants, limiting deployment on resource-constrained devices. Recent developments as of 2024 include efficient transformer-based models such as EFED, which achieves an ODS F-score of 0.824 on BSDS500, enhancing real-time applicability through optimized architectures.

Implementation Aspects

Thresholding Strategies

Thresholding strategies in edge detection involve applying criteria to the output of gradient operators, such as the magnitude of intensity changes, to classify pixels as edge or non-edge, thereby producing a binary edge map. These methods balance the detection of true edges against noise suppression, often applied after computing gradient magnitudes from first-order derivative filters. The choice of threshold significantly impacts edge localization accuracy and robustness to variations in image conditions like illumination or noise. Global thresholding employs a single fixed or automatically determined value across the entire image to binarize the gradient magnitude map. A common automatic approach is the , which selects the threshold by minimizing the intra-class variance of the histogram of gradient magnitudes, effectively separating strong edges from background. This technique, originally proposed for grayscale image segmentation, has been widely adopted in edge detection for its computational efficiency and lack of need for manual parameter tuning, achieving optimal separation in bimodal histograms typical of edge responses. Local or adaptive thresholding addresses non-uniform illumination by computing region-specific thresholds, often based on local image statistics such as the mean intensity plus a multiple of the standard deviation within a neighborhood. For instance, methods using local variance estimate thresholds that adapt to spatial variations, improving edge detection in shadowed or textured regions compared to global approaches, though at higher computational cost. These strategies enhance robustness in real-world images with varying lighting, as demonstrated in applications to medical and remote sensing imagery. Hysteresis thresholding uses dual thresholds to distinguish strong edges from potential weak ones, reducing false positives while preserving continuity. Pixels exceeding the high threshold are classified as definite edges, while those between the low and high thresholds are tentatively marked as weak edges and retained only if connected to strong edges via . This approach mitigates the effects of noise and gaps in edges, leading to more coherent edge maps than single-threshold methods. Introduced as a key refinement in optimal edge detection frameworks, hysteresis improves overall edge tracing in complex scenes. Edge-specific thresholding incorporates additional criteria beyond magnitude, such as gradient direction consistency, to refine edge classification. For example, candidate edge pixels are validated if neighboring gradient directions align within a small angular range, suppressing spurious responses from texture or noise. This directional filtering enhances precision by focusing on coherent edge structures, particularly useful in gradient-based detectors. Quantitative evaluation of thresholding strategies often relies on precision-recall trade-offs, where precision measures the fraction of detected edges that are true positives, and recall captures the fraction of actual edges identified. The F-measure, defined as the harmonic mean of precision and recall, provides a balanced metric for comparing methods, with optimal thresholds maximizing this value on benchmark datasets. Such evaluations highlight hysteresis and adaptive methods' superiority in practical edge detection performance.

Connectivity and Post-Processing

After initial edge detection, connectivity analysis groups detected pixels into coherent components to form continuous edge structures. In binary edge maps, pixels are typically considered connected if they share adjacency, with two primary models: 4-connectivity, where neighbors are limited to horizontal and vertical directions, and 8-connectivity, which additionally includes diagonal neighbors for more flexible linking in diagonal edges. The choice impacts the topology of resulting components, as 8-connectivity can merge structures that 4-connectivity separates, potentially reducing fragmented edges but risking over-connection in noisy images. To preserve the topological integrity of these components during processing, the Euler number serves as a key invariant metric, calculated as the number of connected components minus the number of holes in the binary image. This scalar value remains unchanged under continuous deformations and is used to validate that post-processing operations, such as linking or smoothing, do not alter the fundamental genus or connectivity holes of edge structures. For instance, algorithms monitor the Euler number to ensure topology preservation when bridging gaps or removing artifacts, preventing unintended mergers or splits that could distort object boundaries. Gap bridging addresses discontinuities in detected edges by predicting and linking nearby segments based on local edge directions and orientations. Predictive methods, such as adapted for curve tracking, model edge chains as dynamic trajectories, estimating missing pixels by propagating position and velocity from prior segments while accounting for noise-induced gaps. This approach excels in reconstructing smooth curves from fragmented edgels, iteratively refining links to minimize prediction errors and maintain directional continuity. Noise removal in post-processing employs morphological operations to eliminate spurious elements without disrupting valid edges. Spur pruning iteratively removes short protrusions or branches from edge skeletons using erosion with linear structuring elements, targeting endpoints that extend less than a predefined length. Similarly, small segment deletion identifies and erases isolated or brief edge fragments below a threshold length, often after initial thinning to isolate noise from main contours. These operations enhance edge coherence by focusing on shape-based filtering rather than intensity alone. Edge validation assesses the quality of linked chains through scoring mechanisms that evaluate geometric properties. Segments are scored by their total length, favoring longer chains as more reliable representations of object boundaries; curvature smoothness is quantified via second-order derivatives or approximation errors, penalizing abrupt changes indicative of artifacts. Closure detection fits polygons to candidate chains, measuring fit quality with metrics like the integral of squared distances from points to the polygon, retaining only those achieving low error for complete loops. For global optimization of edge chains, graph cut methods model the problem as an energy minimization on a graph where nodes represent edge pixels and edges encode linking costs based on proximity, direction, and smoothness. The min-cut algorithm partitions the graph to select optimal chains, balancing local affinities with global coherence to resolve ambiguities in noisy or occluded scenes. This yields gap-free, topologically consistent structures superior to greedy local linking.

Example Code Implementations

Example implementations of edge detection post-processing can be found in popular libraries. For thresholding, the with hysteresis is available in . Below is a Python example using to apply with default thresholds:
python
import cv2
import numpy as np

# Load image
image = cv2.imread('input_image.jpg', cv2.IMREAD_GRAYSCALE)

# Apply Gaussian blur to reduce noise
blurred = cv2.GaussianBlur(image, (5, 5), 0)

# Apply [Canny](/page/Canny) edge detection with hysteresis thresholds (low=50, high=150)
edges = cv2.Canny(blurred, 50, 150)

# Display or save the result
cv2.imshow('Edges', edges)
cv2.waitKey(0)
cv2.destroyAllWindows()
This code performs smoothing and thresholding; further post-processing like connectivity can be added using cv2.connectedComponents for 4- or 8-connectivity. For morphological noise removal, SciPy provides tools for erosion and dilation. An example for spur pruning via thinning:
python
from skimage import io, morphology
import numpy as np

# Load binary edge image
edges = io.imread('binary_edges.png', as_gray=True)
edges = (edges > 0).astype(bool)

# Thin the edges
skeleton = [morphology](/page/Morphology).skeletonize(edges)

# Remove small objects (small segments)
skeleton = [morphology](/page/Morphology).remove_small_objects(skeleton, min_size=20)

io.imsave('cleaned_skeleton.png', skeleton.astype(np.uint8))
This uses skeletonization followed by removal of small objects to delete short segments.

References

  1. [1]
    [PDF] Chapter 5. Edge Detection
    This chapter covers only the detection and localization of edges. Basic concepts in edge detection will be discussed. Several common edge detectors will be used ...
  2. [2]
    [PDF] Edge Detection Techniques - An Overview - Hal-Inria
    In computer vision, edge detection is a process which attempts to capture the signi cant prop- erties of objects in the image. These properties include ...
  3. [3]
    [PDF] Methods of Image Edge Detection: A Review - Hilaris Publisher
    Jul 30, 2015 · Edge detection is a basic and important tool in the main areas of image processing such as feature detection and feature extraction [1].
  4. [4]
  5. [5]
    [PDF] Edge Detection 1 Low Level Vision - Cornell: Computer Science
    Edge detection identifies where image intensity changes quickly, finding intensity boundaries, and is related to the gradient of image intensity.
  6. [6]
    [PDF] Lecture #06: Edge Detection - Stanford Computer Vision Lab
    The detection of edges provides meaningful semantic information that facilitate the understanding of an image. This can help analyzing the shape of elements, ...
  7. [7]
    Theory of edge detection | Proceedings of the Royal Society of ...
    A theory of edge detection is presented. The analysis proceeds in two parts. (1) Intensity changes, which occur in a natural image over a wide range of scales, ...Missing: 1982 | Show results with:1982
  8. [8]
    The Beginnings of Computer Vision - Zbigatron
    Jul 13, 2018 · The Early History of Computer Vision · This post, therefore, will focus on seminal developments in computer vision between the 60s and early 80s.The Summer Vision Project · Continued Work With Line... · Computer Vision Gathers...<|separator|>
  9. [9]
    [PDF] David Marr - Vision - Cvut
    David Marr's 'Vision' is a computational investigation into the human representation and processing of visual information.
  10. [10]
    Predicting Source and Age of Brain Tumor Using Canny Edge ... - NIH
    The proposed system uses Canny Edge Detection algorithm and threshold technique for locating the source of tumor, spread area and age of the tumor cells in the ...
  11. [11]
    Analysis of Edge Detection for Road Lanes through Hardware ...
    One of the significant aspects of self-driving vehicles is the detection of lanes through Edge Detection to correct the vehicle from inadvertent road departure.
  12. [12]
    Application of Edge Detection Algorithm Based on Morphology in ...
    To a mobile robot vision system, it is of vital importance to separate objects from the background. This needs us firstly to extract edges from the images ...
  13. [13]
    [PDF] Edge detection
    Ramp edge: a step edge where the intensity change is not instantaneous but occur over a finite distance. Ridge edge: the image intensity abruptly changes ...<|control11|><|separator|>
  14. [14]
    [PDF] Edge Detection - RPI ECSE
    An edge indicates discontinuity in image intensity function. Edge detection amounts to locating each edge pixel (also referred to as edgel) and determining its ...
  15. [15]
    [PDF] A Computational Approach to Edge Detection
    Nov 6, 1986 · Index Terms-Edge detection, feature extraction, image processing, machine vision, multiscale image analysis. I. INTRODUCTION. EDGE detectors of ...
  16. [16]
    [PDF] Image Features Edge Model (1D) 1-D edge detection Edges
    Feb 19, 2005 · • An ideal edge can be modeled as an step. ⎩. ⎨. ⎧. ≤. <. = 0 if. 0 if ... • Think of an image as function. • Computing derivative of 2D ...
  17. [17]
    [PDF] Edge Detection
    Let the signal S have a step edge of height H at location x0, and let the noise at each pixel be Gaussian, independent and identically distributed (IID).
  18. [18]
  19. [19]
    [PDF] Local Scale Control for Edge Detection and Blur Estimation
    Local scale control is shown to be important for the estimation of blur in complex images, where the potential for interference between nearby edges of very ...
  20. [20]
    [PDF] Local Detection of Occlusion Boundaries in Video
    Improvement is most apparent for edges bordering textured or cluttered regions, which are precisely those edges at which occlusion is most likely to be visible.
  21. [21]
    [PDF] Statistical Edge Detection: Learning and Evaluating Edge Cues.
    Dec 26, 2002 · Statistical edge detection has fewer false negatives in the textured regions and is also better at detecting edges with are partially defined by ...
  22. [22]
    None
    ### Summary of Gradient Operator, First Derivatives, Finite Differences, and Roberts Operator for Edge Detection
  23. [23]
    [PDF] Adopted from pdflib image sample (C) - DSpace@MIT
    MACHINE PERCEPTION OF THREE-DIMENSIONAL SOLIDS by. LAWRENCE GILMAN ROBERTS. S. B., Massachusetts Institute of Technology. (1961). M.S., Massachusetts Institute ...Missing: detection | Show results with:detection
  24. [24]
    [PDF] Edge and Junction Detection with an Improved Structure Tensor
    In this paper we propose three improvements to the structure tensor approach that address the aforementioned problems: we use a higher sampling rate to avoid ...
  25. [25]
    [PDF] Notes on Corner Detection
    Jul 13, 2000 · cal feature analysis goes by many names: Plessey-Harris operator, Förstner corner detector, orientation tensor, local structure tensor, etc.
  26. [26]
  27. [27]
    [PDF] A Review of Classic Edge Detectors - IPOL Journal
    Jun 16, 2015 · Abstract. In this paper some of the classic alternatives for edge detection in digital images are studied. The main idea of edge detection ...<|separator|>
  28. [28]
    Spatial Filters - Laplacian/Laplacian of Gaussian
    The Laplacian of an image highlights regions of rapid intensity change and is therefore often used for edge detection (see zero crossing edge detectors).
  29. [29]
    [PDF] Image Features from Phase Congruency
    They recognize that step edges and bars have specific local phase properties that can be detected using filters in quadrature; however, they do not connect the.
  30. [30]
    [PDF] Phase Congruency Detects Corners and Edges - Peter Kovesi
    This paper describes a new corner and edge detector developed from the phase con- gruency model of feature detection. The new operator uses the principal.
  31. [31]
    (PDF) An Isotropic 3x3 Image Gradient Operator - ResearchGate
    A 3x3 Isotropic Gradient Operator for Image Processing, presented at the Stanford Artificial Intelligence Project (SAIL) in 1968.
  32. [32]
    [PDF] Object enhancement and extraction - UTK-EECS
    Object enhancement, extraction, characterization and recognition are closely linked aspects of visual perception, and cognizance of their inter- play as well as ...
  33. [33]
    Optimale Operatoren in der Digitalen Bildverarbeitung - heiDOK
    Abstract. Eine neuartige Methode zur optimalen Wahl von Filteroperatoren wird vorgestellt. Dabei können sowohl einzelne Filter als auch Filterfamilien mit ...
  34. [34]
  35. [35]
    A New Method of Edge Detection - SpringerLink
    Dec 11, 2018 · A connected subset of a boundary is called a boundary curve . For the notion of connectedness, please refer to Kovalevsky (1989). We consider ...Missing: Kovalevsky | Show results with:Kovalevsky
  36. [36]
    [PDF] an Implementation of the Canny/Devernay Algorithm - IPOL Journal
    This sub-pixel edge detector uses Canny's ideas, Devernay's post-processing for sub-pixel accuracy, and a modification to improve the accuracy.
  37. [37]
    6.4. Subpixel Localization of Local Structure
    In this section we look at sub pixel accurate localization methods for contour lines (isophotes), edges, lines and X-corners (saddle points).
  38. [38]
    Comprehensive investigation of subpixel edge detection schemes in ...
    Aug 9, 2025 · In this paper, we review several sub-pixel edge detection schemes and compare them with respect to two figures of merit - resolution accuracy ...
  39. [39]
    Edge linking by sequential search - ScienceDirect
    In this paper we develop a linking algorithm for the combination of edge elements enhanced by an optimal filter. The linking algorithm (LINK) is based on ...Missing: seminal | Show results with:seminal
  40. [40]
    [PDF] Two linear time Union-Find strategies for image processing - Hal-Inria
    In this Paper we consider two different variants of Union-Find that solve region growing and then give generalizations of them. For both we give an algorithm ...
  41. [41]
    [PDF] Learning to Detect Natural Image Boundaries Using Local ...
    Abstract—The goal of this work is to accurately detect and localize boundaries in natural scenes using local image measurements.
  42. [42]
    A Multi-Scale Bifurcated Deep Network for Top-Down Contour ...
    Dec 2, 2014 · DeepEdge uses a multi-scale deep network with a bifurcated sub-network to detect contours by exploiting object-related features, achieving ...
  43. [43]
    [1504.06375] Holistically-Nested Edge Detection - arXiv
    Apr 24, 2015 · Our proposed method, holistically-nested edge detection (HED), performs image-to-image prediction by means of a deep learning model that leverages fully ...
  44. [44]
    [1612.02103] Richer Convolutional Features for Edge Detection - arXiv
    Dec 7, 2016 · In this paper, we propose an accurate edge detector using richer convolutional features (RCF). Since objects in nature images have various scales and aspect ...
  45. [45]
    The Berkeley Segmentation Dataset and Benchmark
    New: The BSDS500, an extended version of the BSDS300 that includes 200 fresh test images, is now available here. The goal of this work is to provide an ...
  46. [46]
    [2108.07009] Pixel Difference Networks for Efficient Edge Detection
    Aug 16, 2021 · We propose a simple, lightweight yet effective architecture named Pixel Difference Network (PiDiNet) for efficient edge detection.
  47. [47]
    [2203.08566] EDTER: Edge Detection with Transformer - arXiv
    Mar 16, 2022 · We propose a novel transformer-based edge detector, \emph{Edge Detection TransformER (EDTER)}, to extract clear and crisp object boundaries and meaningful ...Missing: 2021 | Show results with:2021
  48. [48]
    The connected-component labeling problem: A review of state-of-the ...
    Moreover, because objects with 8-connectivity are more complicated than those with 4-connectivity, we will only consider 8-connectivity for objects. For ...Missing: seminal | Show results with:seminal
  49. [49]
    A Review for the Euler Number Computing Problem - MDPI
    Oct 25, 2023 · Among the features, the Euler number, which can be obtained by subtracting the number of holes from the number of connected components ( ...
  50. [50]
    [PDF] Topology-Preserving Downsampling of Binary Images
    We present a novel discrete optimization-based approach to generate downsampled versions of binary images that are guaranteed to have the same topology as the ...
  51. [51]
    Computing the Euler Number of a Binary Image Based on a Vertex ...
    We describe a method to compute the Euler number of a binary digital image based on a codification of contour pixels of the image's shapes.Missing: preservation | Show results with:preservation
  52. [52]
    [PDF] Curve Finder Combining Perceptual Grouping and a Kalman Like ...
    Abstract. We present an algorithm that extracts curves from a set of edgels within a specific class in a decreasing or- der of their “length”.
  53. [53]
    [PDF] Circle extraction via least squares and the Kalman filter - ePrints Soton
    The Kalman filter algorithm also uses the discrepancy in edge direction information to correct parameter estimates and can produce estimates in a point by ...
  54. [54]
    (PDF) Implementation of Morphological Operation through Intercept ...
    The main goal of thinning is to degrade particular foreground pixels from original input image. A pruning focused on removing short spurs present in an image.
  55. [55]
    [PDF] EDGE DETECTION USING MORPHOLOGICAL AMOEBAS IN ...
    In this paper, we propose morphological edge detectors with amoebas which consider the image contour variations to adapt their shape. Note that the proposed ...
  56. [56]
    [PDF] Segmentation and Classi cation of Edges Using Minimum ...
    Abstract. This article presents a method for segmenting and classifying edges using minimum description length (MDL) approximation with automatically ...
  57. [57]
    [PDF] Techniques for Assessing Polygonal Approximations of Curves
    In this paper we shall restrict ourself to approaches which require the polygon vertices to lie on the curve. Thus, given a curve C = fxi; yigN i=1 the goal is ...
  58. [58]
    [PDF] Graph Cuts in Vision and Graphics: Theories and Applications
    Graph cuts are combinatorial min-cut algorithms used for geometric interpretation and energy minimization in vision and graphics, and are a multidimensional  ...
  59. [59]
    [PDF] An Alternative Graph Cut Algorithm for Morphological Edge Detection
    The graph cuts algorithm produces the changes in edges on each step, which are more deformed edges starts from the beginning. Global form of final edges.Missing: chains | Show results with:chains