Fact-checked by Grok 2 weeks ago

Walsh function

Walsh functions are a complete set of pairwise defined on the interval [0, 1], each consisting of a square waveform that takes only the values +1 or -1 (except at points of discontinuity, where the value is 0), and remaining constant over dyadic subintervals of [0, 1]. These functions form a basis for the L^2[0, 1], allowing the representation of any as an infinite analogous to the but using binary-valued basis elements instead of sinusoids. Introduced by American mathematician Joseph L. Walsh in his 1923 paper "A Closed Set of Normal ," they are normalized such that \int_0^1 \phi_n(x) \phi_m(x) \, dx = \delta_{nm}, where \delta_{nm} is the , and each function \phi_n(x) exhibits exactly n sign changes within (0, 1). The functions are typically ordered by sequency, a measure of the number of sign changes (zero-crossings) per , which serves as a discrete counterpart to frequency in . They can be generated recursively or via the rows of Hadamard matrices, which are square matrices with entries \pm 1 satisfying H H^T = N I for an N \times N matrix H and identity I. Walsh functions possess properties such as (no non-zero integrable is orthogonal to all of them) and uniform boundedness, making them suitable for expansions of continuous functions that converge uniformly when grouped by sequency order. Unlike the continuous basis, their discontinuous, binary nature aligns well with digital implementations, facilitating efficient computation through the Walsh-Hadamard transform (WHT), which requires only additions and subtractions, akin to the (FFT) but without multiplications. In applications, Walsh functions underpin the WHT for tasks, including power spectrum estimation, filtering of speech and medical signals, , and , where their enables decomposition into sequency-ordered components with computational complexity O(N \log N) for N-point transforms. They are integral to systems in wireless communications, where finite-length Walsh codes (derived from Hadamard matrices) provide orthogonal spreading sequences to separate simultaneous user signals on the downlink, as standardized in mobile networks like IS-95 and CDMA2000. Additional uses include reducing in antenna arrays and designing logic circuits or schemes in .

Definition and History

Definition

Walsh functions constitute a complete orthogonal system in the L^2[0,1], comprising square-wave functions that take values \pm 1 on the interval [0,1], with discontinuities occurring solely at dyadic rational points of the form m/2^k for integers m and k \geq 0. These functions form an for L^2[0,1], enabling the expansion of any as an infinite series of Walsh functions. The Walsh functions are constructed as products of Rademacher functions, which provide the fundamental building blocks. The k-th Rademacher function is defined as r_k(x) = \sgn(\sin(2^{k+1} \pi x)) for k = 0, 1, 2, \dots and x \in [0,1), where \sgn denotes the sign function, yielding periodic square waves with increasing frequencies that alternate between +1 and -1. For a nonnegative integer n with binary expansion n = \sum_{j=0}^\infty n_j 2^j where n_j \in \{0,1\}, the Walsh function in Paley ordering (also known as dyadic or natural ordering) is given by \wal(n, x) = \prod_{j=0}^\infty r_j(x)^{n_j} = (-1)^{\sum_{j=0}^\infty n_j x_j}, with the second form using the binary digits x_j of x \in [0,1); this ordering follows the natural binary representation of the index n. In contrast, sequency ordering (also called Hadamard ordering) arranges the Walsh functions by their sequency, defined as the number of sign changes (zero crossings) in the interval [0,1), analogous to frequency in Fourier analysis. This ordering is obtained by reordering the Paley-ordered functions or rows of the via bit-reversal permutation. Examples of the first few sequency-ordered Walsh functions include:
  • \wal(0, x) = 1 (sequency 0),
  • \wal(1, x) = \sgn(\sin(2\pi x)) (sequency 1),
  • \wal(2, x) = \sgn(\sin(2\pi x)) \cdot \sgn(\sin(4\pi x)) (sequency 2),
  • \wal(3, x) = \sgn(\sin(4\pi x)) (sequency 3).
The Walsh transform of a function f \in L^2[0,1] is defined by the coefficients \hat{c}_k = \int_0^1 f(x) \wal(k, x) \, dx, \quad k = 0, 1, 2, \dots, which represent the projections onto the Walsh basis and allow reconstruction via the series \sum_k \hat{c}_k \wal(k, x). The set \{\wal(n, x)\}_{n=0}^\infty is normalized such that \int_0^1 \wal(m, x) \wal(n, x) \, dx = \delta_{mn}, forming an , and by the completeness theorem, the span of these functions is dense in L^2[0,1].

Historical Development

The Walsh functions were first introduced by Joseph L. Walsh in his 1923 paper, where he constructed a complete orthogonal system of square-wave functions on the interval [0,1] primarily for applications in interpolation theory. This work built upon the earlier independent discovery of Rademacher functions by Hans Rademacher in 1922, which served as fundamental building blocks consisting of simple ±1 square waves that exhibit orthogonality properties. In the 1930s, Raymond E. A. C. Paley advanced the understanding of these functions by establishing a connection between the Walsh and Rademacher systems, leading to the development of the Paley ordering, which arranges the functions based on their structure. The term "sequency," analogous to in and reflecting the number of zero crossings, was introduced by Heinrich F. Harmuth in 1969. Despite these mathematical foundations, the functions remained relatively obscure until the , when pioneers in digital computing and electronics, such as Heinrich F. Harmuth, popularized them for tasks, recognizing their computational efficiency in binary systems. It was during this period that the functions became widely known as "Walsh functions," distinguishing them from other orthogonal bases. Harmuth's contributions, including early applications in communications engineering, alongside works by researchers like K. G. Beauchamp, spurred the first international symposia on Walsh functions starting in 1969. The 100th anniversary of Walsh's introduction in prompted renewed scholarly attention, coinciding with the 50th anniversary of his death, and featured commemorative publications highlighting the evolution of analysis as a distinct field parallel to classical . These efforts underscored the functions' enduring legacy in bridging and applied sciences.

Mathematical Properties

Core Properties

Walsh functions possess several fundamental properties that establish them as an orthonormal basis for the Hilbert space L^2[0,1]. Central to their utility is : for distinct nonnegative integers m and n, \int_0^1 \mathrm{wal}(m,x) \mathrm{wal}(n,x) \, dx = \delta_{mn}, where \delta_{mn} denotes the Kronecker delta (equal to 1 if m = n and 0 otherwise), and the integral equals 1 when m = n due to normalization. This relation holds because Walsh functions can be expressed as finite products of Rademacher functions r_k(x) = \mathrm{sgn}(\sin(2^k \pi x)), which are themselves orthogonal over [0,1]. A proof sketch relies on the binary expansions of the indices m and n: the product \mathrm{wal}(m,x) \mathrm{wal}(n,x) corresponds to a Walsh function whose index reflects the bitwise XOR of the binary representations of m and n, and the integral vanishes unless this index is zero, leveraging the independence of the Rademacher components in their binary-induced structure. Complementing is the of the Walsh system in L^2[0,1], ensuring that any f admits a expansion f(x) = \sum_{k=0}^\infty \hat{c}_k \mathrm{wal}(k,x) in the L^2 sense, where the coefficients are given by \hat{c}_k = \int_0^1 f(x) \mathrm{wal}(k,x) \, dx. This implies that if \int_0^1 f(x) \mathrm{wal}(k,x) \, dx = 0 for all k, then f = 0 . As a consequence, applies: \|f\|_2^2 = \int_0^1 |f(x)|^2 \, dx = \sum_{k=0}^\infty |\hat{c}_k|^2, preserving the energy of the function in the Walsh domain and facilitating norm-equivalent transformations akin to those in Fourier analysis. From a group-theoretic viewpoint, Walsh functions serve as the characters of the dyadic group \mathbb{Z}_2^\mathbb{N}, consisting of infinite binary sequences under componentwise addition modulo 2, compactified with the product topology. Each Walsh function \mathrm{wal}(k,x) corresponds to the character evaluating the k-th binary digit sum modulo 2 for the binary expansion of x, yielding \pm 1. This identification embeds Walsh analysis within the representation theory of compact abelian groups, where the Walsh transform acts as the Fourier transform, enabling harmonic analysis on dyadic structures. An algebraic hallmark is the closure of the Walsh system under pointwise : for any m, n \geq 0, \mathrm{wal}(m,x) \cdot \mathrm{wal}(n,x) = \mathrm{wal}(m \oplus n, x), where \oplus denotes bitwise XOR, which aligns with the binary index structure. This closure implies that the Walsh functions form a group under (with the constant 1 as and each its own ), supporting efficient algebraic operations in expansions and convolutions. Structurally, all Walsh functions are piecewise constant, taking values \pm 1 on intervals of the form [j/2^l, (j+1)/2^l) for integers j, l \geq 0, with discontinuities solely at dyadic rationals (where they are conventionally set to 0). In sequency ordering—arranged by the number of sign changes across [0,1)—higher-indexed functions exhibit more rapid oscillations, with the k-th function changing sign exactly k times, mirroring frequency escalation in a binary-dyadic sense.

Comparison to Trigonometric Functions

Walsh functions and the of both constitute complete orthonormal bases for the L^2[0,1], enabling the representation of any on the unit interval as an infinite . Often characterized as a "square-wave" analog to the basis, Walsh functions replace the continuous oscillations of sines and cosines with binary-valued (±1) step functions, providing a discrete, digital-friendly alternative for . In contrast to the smooth, infinitely differentiable, and periodic nature of , Walsh functions are piecewise constant with abrupt transitions exclusively at dyadic rational points, aligning naturally with binary subdivisions of the interval. This discontinuous, rectangular support renders Walsh functions less prone to issues in discrete sampling scenarios, where trigonometric bases may introduce errors due to their . The Walsh transform is computed via the , achieving O(n \log n) complexity for n data points, matching the efficiency of the (FFT) but relying solely on additions and subtractions without the FFT's multiplicative twiddle factors—yielding practical speedups for binary or low-precision data on specialized hardware. Walsh series further avoid the inherent in Fourier approximations near jumps, eliminating overshoot and ringing artifacts. Regarding approximation properties, Walsh series converge in the mean-square sense to the original function, akin to , but exhibit enhanced rates for functions discontinuous at dyadic points, where the basis aligns with the jumps. For instance, the Walsh expansion of a at x = 1/2 achieves exact finite-term representation, bypassing the slow, oscillatory convergence and persistent Gibbs overshoot observed in its counterpart.

Generalizations

Walsh-Ferleger Systems

The Walsh-Ferleger system refers to an for expansions in the L²(μ), where μ is an arbitrary on the [0,1], extending the classical Walsh system beyond its dyadic structure tied to the . This generalization allows for the representation of square-integrable functions with respect to non-uniform or irregular measures, facilitating in probabilistic settings where the underlying distribution deviates from uniform sampling. The construction relies on generalized Rademacher functions tailored to the measure μ. For a μ with density ρ ∈ L¹([0,1]), ρ > 0, and ∫₀¹ ρ(x) dx = 1, the system begins with martingale differences derived from the Rademacher functions r_n(x) = sign(sin(2^{n+1} π x)). These are normalized as φ_n = (r_n - E_μ^n r_n) / √Var_μ(r_n), where E_μ^n denotes with respect to the σ-algebra generated by μ, ensuring zero mean and unit variance under μ. The Walsh-Ferleger functions ψ_m are then defined as finite products ψ_m = ∏k φ{m_k}^k, where m = ∑ m_k 2^k with m_k ∈ {0,1}, forming a complete system in L²(μ) for non-atomic μ. A fundamental result is that the system {ψ_m} is orthonormal in L²(μ), satisfying ⟨ψ_m, ψ_n⟩μ = ∫{[0,1]} ψ_m(x) ψ_n(x) dμ(x) = δ_{mn}, and complete provided μ has no atoms, meaning μ({x}) = 0 for all x ∈ [0,1]. This and hold under these mild conditions on μ, enabling the Parseval identity for series expansions and supporting applications such as irregular sampling, where μ captures non-equispaced data distributions. In contrast to the standard Walsh system, which assumes the dyadic partitioning implicit in the and is limited to uniform expansions, the Walsh-Ferleger system accommodates arbitrary supports, including fractal or singular continuous measures that lack density but remain non-atomic. The classical Walsh functions arise as the special case when μ is the on [0,1]. This framework was developed in the context of probabilistic during the late 20th century.

Vilenkin System

The Vilenkin system was introduced by Nikolai Yakovlevich Vilenkin in to extend beyond dyadic structures to more general compact abelian groups of zero dimension. These groups, often called Vilenkin groups, are constructed as the infinite of finite cyclic groups \mathbb{Z}/p\mathbb{Z} for a fixed prime p \geq 2, endowed with the , making them topologically isomorphic to the additive group of p-adic integers \mathbb{Z}_p. This framework allows for a natural generalization of on such spaces, where the Vilenkin functions serve as the characters. The Vilenkin functions \{\chi_m\}_{m \in \mathbb{N}_0} are defined as the continuous characters of the Vilenkin group G_p, expressed using the p-adic expansions of elements. Specifically, for m = \sum_{k=0}^\infty a_k p^k with digits $0 \leq a_k < p and x = \sum_{k=0}^\infty b_k p^{-(k+1)} with digits $0 \leq b_k < p, \chi_m(x) = \exp\left(2\pi i \sum_{k=0}^\infty \frac{a_k b_k}{p^{k+1}}\right). When p=2, this reduces to the classical . The system forms an orthonormal basis for L^2(G_p, \mu), where \mu is the normalized on G_p. The associated Vilenkin-Fourier transform of a function f \in L^1(G_p) is given by \hat{f}(m) = \int_{G_p} f(x) \overline{\chi_m(x)} \, d\mu(x), with the inversion formula recovering f for f \in L^1(G_p) \cap L^2(G_p). This transform generalizes the and preserves key analytic properties, such as Plancherel's theorem: \|f\|_{L^2}^2 = \sum_{m} |\hat{f}(m)|^2. In , Vilenkin functions facilitate the analysis of for sequences in p-adic settings, such as \mathbb{Z}_p. By expanding the empirical measure of a sequence in Vilenkin- series, the discrepancy—measuring deviation from uniformity—can be bounded using the decay of Fourier coefficients, analogous to Weyl's in the classical case. For instance, ergodic transformations on G_p preserve when their Vilenkin spectra satisfy certain mixing conditions.

Nonlinear Phase Extensions

Nonlinear Walsh functions represent an extension of the traditional Walsh basis by incorporating phase functions φ(n, x) that deviate from the strict dyadic linearity of standard constructions, enabling more flexible representations of signals with varying content. A prototypical form is given by \mathrm{wal}_\phi(n, x) = \sgn(\sin(2\pi \phi(n, x))), where the φ(n, x) is chosen to be nonlinear, such as a monotonic increasing function that introduces discontinuities at non- points, often described as "curved" in the of signal . This generalization preserves the square-wave nature of Walsh functions while allowing for adaptive adjustments to the and shapes of transitions. The construction begins with generalizing the Rademacher functions, which form the building blocks of Walsh functions, through nonlinear mappings applied to their phase arguments. Specifically, instead of the exponential dyadic scaling in standard Rademacher functions r_n(x) = \sgn(\sin(2^n \pi x)), a nonlinear phase φ(n, x) is substituted, ensuring the resulting set remains orthogonal under conditions like monotonicity of φ to maintain inner product zero-crossings. Walsh-like functions are then formed as products or linear combinations of these generalized Rademacher components, often verified computationally for small orders to confirm orthogonality. For instance, sets of length 32 have been generated via exhaustive search in binary spaces, yielding bases without the zero-crossing constraints of linear Walsh sequences. Key properties include partial orthogonality for finite sets, where the basis functions are mutually orthogonal over [0,1) with respect to the Lebesgue measure, and tolerance to nonzero means, unlike the balanced standard Walsh functions. These extensions exhibit enhanced correlation characteristics, with lower sidelobes in autocorrelation compared to linear counterparts, facilitating better resolution in expansions. They also support adaptive bases, where φ can be tuned based on signal features to optimize representation efficiency. Compared to linear-phase Walsh functions, nonlinear variants offer superior approximation capabilities for non-stationary signals, such as chirp-like or frequency-modulated waveforms, by aligning discontinuities more closely with the signal's instantaneous variations, thus reducing the number of terms needed for accurate . This is particularly evident in scenarios requiring localized shifts, where linear bases suffer from fixed partitioning. Development of these extensions gained traction in the within nonlinear , with foundational contributions focusing on relaxing phase linearity to broaden applicability in adaptive signal . Seminal work by Akansu and Poluri in 2007 introduced practical Walsh-like nonlinear phase bases through search-based methods, demonstrating their and benefits.

Applications

In Signal Processing

The Walsh-Hadamard transform (WHT) serves as the discrete counterpart to continuous Walsh functions, enabling the analysis of finite-length sequences in by decomposing them into orthogonal Walsh basis functions. This transform is particularly suited for binary or square-wave signals, where it provides a sequency-domain representation analogous to frequency in , facilitating efficient spectral examination without trigonometric computations. In practical applications, the WHT is employed for of binary signals, such as those in digital modulation schemes, where it reveals sequency content to identify dominant components and support filtering operations. It also plays a role in , leveraging Hadamard matrices—special cases of Walsh matrices—to encode blocks with reduced , as demonstrated in early probe imagery systems that achieved efficient data transmission through sequency-ordered transformations. Additionally, the WHT aids in switching circuits and adaptive filtering, where its allows separation of signal from interference in real-time hardware environments, such as digital logic designs. A key advantage of the WHT lies in its computational efficiency, relying solely on integer additions and subtractions rather than multiplications, which enables implementation via a butterfly algorithm structurally similar to the (FFT), achieving O(N log N) complexity for sequences of length N. This property made it ideal for real-time processing in 1960s-era sequency analyzers, hardware devices developed by researchers like H.F. Harmuth for on-the-fly signal decomposition in resource-constrained settings. The of Walsh functions underpins this efficiency, allowing invertible transformations with minimal numerical overhead. As an illustrative example, watermarking techniques embed proprietary data by modifying selected WHT coefficients of an or signal; for instance, a hybrid scheme integrates WHT with to insert watermarks into mid-sequency bands, ensuring robustness against common attacks like or while preserving perceptual quality. In modern wireless communications, Walsh functions generate orthogonal spreading codes akin to those in (CDMA) systems, where 64-chip Walsh-Hadamard sequences distinguish user channels and mitigate interference in multi-user environments, as seen in OFDM-CDMA hybrids for enhanced .

In Coding Theory and Statistics

In coding theory, Walsh functions form the basis for Hadamard-Walsh codes, which are linear error-correcting codes used for detecting and correcting errors in noisy channels. These codes are constructed from Walsh-Hadamard matrices, where the rows represent codewords with equal , enabling a relative of 1/2 that supports correction of up to 1/4 of the errors in a codeword. A prominent example is the simplex code, obtained as a shortened version of the from Walsh matrices, which has parameters [2^m - 1, m, 2^{m-1}] and exhibits constant nonzero codeword weight, making it suitable for applications requiring equidistant codewords. Such codes were employed by in the late for space communications, including the Mariner missions to Mars in 1969 and 1971, where reliable error correction was essential over deep-space channels with high noise levels. Walsh spectrum analysis, derived from the Walsh transform, plays a key role in evaluating the nonlinearity of Boolean functions, a critical property for designing secure cryptographic primitives like S-boxes in block ciphers. The nonlinearity of a Boolean function is quantified as its minimum Hamming distance to any affine function, computed via the maximum absolute value in the Walsh spectrum, ensuring resistance to linear cryptanalysis attacks. In recent developments, the Walsh transforms of symmetric and rotation-symmetric Boolean functions—common in cryptographic constructions—have been shown to satisfy linear recurrence relations with integer coefficients, facilitating efficient computation and analysis of their spectral properties for enhanced security assessments. In , Walsh-Fourier methods leverage the of Walsh functions for nonparametric approaches to and , particularly with or . For instance, Ott and Kronmal (1976) developed procedures for multivariate by expanding probability densities in a Walsh series, enabling discriminant analysis and through the transform's ability to capture interactions in orthogonal components. An illustrative application is dyadic ANOVA, where the Walsh basis decomposes effects in 2^n designs into main effects and interactions analogous to a modulo 2, as noted by Good (1958), providing a framework for estimating variances in experimental settings with factors.

References

  1. [1]
    [PDF] A Closed Set of Normal Orthogonal Functions
    It is the object ot the present paper to study a certain new closed set of functions ... J. L. WALSH: Normal Orthogonal Functions. S24.1(x) cannot differ from F(.
  2. [2]
    Walsh Function -- from Wolfram MathWorld
    The Walsh functions consist of trains of square pulses (with the allowed states being -1 and 1) such that transitions may only occur at fixed intervals of a ...
  3. [3]
    Walsh-Hadamard Transform - MATLAB & Simulink - MathWorks
    The Walsh-Hadamard transform is a non-sinusoidal, orthogonal transformation technique that decomposes a signal into a set of basis functions.
  4. [4]
    Walsh Function - an overview | ScienceDirect Topics
    Walsh functions are defined as rows of Hadamard matrices that serve as a complete orthogonal system, enabling representation of numeric vectors in spaces of ...<|control11|><|separator|>
  5. [5]
    Discrete Walsh-Hadamard Transform - MATLAB & Simulink Example
    WHTs are used in many different applications, such as power spectrum analysis, filtering, processing speech and medical signals, multiplexing and coding in ...Introduction · Walsh (or Hadamard) Functions · Walsh-Transform Applications
  6. [6]
    Role of Walsh Codes and pseudorandom noise sequences in CDMA
    Abstract: Walsh Codes are a set of orthonormal codes while PN Sequences are statistically random but deterministically formed binary sequences.
  7. [7]
    Walsh Functions - an overview | ScienceDirect Topics
    The Walsh function is an orthogonal function and can be used as the basis for a continuous or discrete transform.<|control11|><|separator|>
  8. [8]
    [PDF] arXiv:2304.06003v2 [math.CA] 13 Jun 2023
    Jun 13, 2023 · We introduce some concepts of Walsh-Fourier analysis. The Rademacher func- tions are defined as. rk(x) := (−1)xk (x ∈ G, k ∈ N). The Walsh- ...
  9. [9]
    [PDF] Walsh-Fourier Analysis and Its Statistical Applications
    These articles deal with some of the mathematical properties of Walsh functions. The 1950 paper establishes the generalized Walsh functions by extending the ...
  10. [10]
    [PDF] On the Walsh Functions Author(s): N. J. Fine Source
    On the Walsh Functions. Author(s): N. J. Fine. Source: Transactions of the American Mathematical Society, Vol. 65, No. 3 (May, 1949), pp. 372-414. Published ...Missing: dyadic group 1950
  11. [11]
    [PDF] 72. Completeness and basis properties of sets of special functions
    THEOREM The Walsh system forms a CON sequence in. L2(0,1). Proof (i) ... (ii) Completeness We shall verify Parseval's theorem directly. Green's ...Missing: scholarly | Show results with:scholarly
  12. [12]
  13. [13]
    Walsh Functions - Yale Math
    The basis corresponding to a fixed row provides a windowed Walsh transform. The discrete Haar wavelet basis is obtained by choosing the second block in each row ...
  14. [14]
    [PDF] in search of the optimal walsh-hadamard transform - SPIRAL Project
    Fast al- gorithms for computing the WHT are similar to the fast Fourier transform (FFT) and its variants [6]. The only difference is that there are no twiddle ...
  15. [15]
    [PDF] Robustifying the Sparse Walsh-Hadamard Transform without ...
    The WHT may be computed using N samples and N log N operations via a recursive algorithm [4], [5] analogous to the Fast Fourier Transform (FFT). However, these ...
  16. [16]
    [PDF] On Walsh Functions with Respect to Weights 1 Ferenc Schipp
    Among others, the Walsh system and its generalizations can be originated in this way. ... measure we take a probability measure, generated by a positive weight ...
  17. [17]
  18. [18]
    Characterizing four-body indistinguishability via symmetries
    By equation (3) we now obtained an (anti)symmetric state, describing four identical bosons (fermions). ... Walsh functions $w(j,\mathcal{S})$ for all modes j ∈ {1 ...
  19. [19]
    Uniformly distributed sequences of p-adic integers, II - math - arXiv
    Sep 30, 2002 · The paper describes ergodic (with respect to the Haar measure) functions in the class of all functions, which are defined on (and take values in) ...
  20. [20]
    [PDF] Walsh-Like Nonlinear Phase Orthogonal Codes for Direct Sequence ...
    In this paper, we relaxed linear phase requirements and new sets of Walsh-like nonlinear phase binary orthogonal user codes (transforms) are obtained for ...Missing: original | Show results with:original
  21. [21]
    Walsh-Hadamard Transform | SpringerLink
    The WHT has gained prominence in various digital signal processing applications, since it can essentially be computed using additions and subtractions only.
  22. [22]
    [1607.01039] Practical Tera-scale Walsh-Hadamard Transform - arXiv
    Jun 28, 2016 · WHT is popular in a variety of applications in image and video coding, speech processing, data compression, digital logic design, communications ...
  23. [23]
  24. [24]
    An FPGA implementation of Walsh-Hadamard transforms for signal ...
    These transforms are important in many signal processing applications including speech compression, filtering and coding.
  25. [25]
    ADAPTIVE DIGITAL FILTERS USING THE WALSH-HADAMARD ...
    The authors discuss the role of orthogonal transformation in adaptive noise canceling and demonstrate the use of the Walsh-Hadamard transform (WHT) for ...Missing: switching | Show results with:switching
  26. [26]
  27. [27]
    A Hybrid Watermarking Scheme Using Walsh Hadamard Transform ...
    In this work, a robust hybridized medical images watermarking approach based on Walsh Hadamard Transform (WHT ) and singular value decomposition (S VD) is ...
  28. [28]