Fact-checked by Grok 2 weeks ago

K-space

In physics, particularly and , k-space (also known as reciprocal space) is a conceptual mathematical domain that represents the spatial frequencies or wave vectors (k) of plane waves, serving as the counterpart to real (direct) space where positions are described by coordinate vectors (r). The wave vector k has components (k_x, k_y, k_z) with units of inverse length (e.g., cm⁻¹), quantifying the modulation or periodicity of waves such that k = 2π/λ, where λ is the . This space is independent of any specific crystal structure initially but becomes structured by the when applied to periodic systems, defined by basis vectors (b_i) satisfying b_i · a_j = 2π δ_{ij}, where a_j are the real-space primitive lattice vectors. The in k-space consists of discrete points (G) given by integer combinations of the basis vectors, G = m₁ b₁ + m₂ b₂ + m₃ b₃ (with integers m_i), forming a periodic array that mirrors the of the direct lattice but with inverted dimensions. These points correspond to the possible vectors in experiments, where constructive (Bragg condition) occurs when the change in Δk equals a vector G, enabling the determination of crystal structures via , , or . In , k-space is essential for analyzing electronic band structures, as the energy eigenvalues of electrons in a crystal are periodic functions E(k) within the first —a primitive cell in reciprocal space bounded by planes perpendicular to vectors and passing through their midpoints. Beyond physics, the k-space formalism extends to fields like (MRI), where it denotes the raw data matrix of components acquired during scans, which is Fourier-transformed to yield the final image; the central region of k-space encodes low-frequency contrast information, while peripheral areas capture high-frequency details like edges. This shared mathematical foundation highlights k-space's versatility in representing periodic or modulated phenomena across disciplines, from —where it equates to momentum space via p = ℏk—to .

Mathematical Foundations

Definition and Spatial Frequencies

K-space, also referred to as reciprocal space, is the conjugate domain to real position space, where spatial structures are represented in terms of their wave-like components rather than direct positional coordinates. Points in k-space are specified by wavevectors \mathbf{k}, which characterize plane waves, with the magnitude |\mathbf{k}| = 2\pi / \lambda, where \lambda denotes the of the associated wave. This representation transforms descriptions of physical fields or functions from their spatial extent in real space to a domain emphasizing oscillatory behavior and periodicity. In k-space, the coordinates correspond to spatial frequencies that encode the scale of variations in the original real-space features. Low k values, associated with longer wavelengths, capture coarse, slowly varying structures, such as overall shapes or broad gradients, while high k values, linked to shorter wavelengths, encode fine details and rapid oscillations, like edges or textures. This duality allows for the decomposition of complex signals into constituent frequencies, facilitating analysis of how different scales contribute to the whole. K-space is generally three-dimensional in physics applications, spanned by the orthogonal components k_x, k_y, and k_z, reflecting the full nature of wave propagation in volumetric space. In contexts like , it may be treated as two- or three-dimensional depending on the dimensionality of the data. The space itself is linear and isotropic under standard metrics, meaning distances and directions are uniform unless modified by specific physical constraints, such as material symmetries. The foundational ideas underlying k-space trace back to Joseph Fourier's 1822 development of Fourier analysis in his work on heat conduction, which introduced the decomposition of functions into sinusoidal components. The term "k-space" gained prominence in the 20th century through wave mechanics in quantum and solid-state physics, where it describes momentum distributions via \mathbf{p} = \hbar \mathbf{k}. The Fourier transform provides the key mathematical bridge mapping representations between real space and k-space.

Relation to Fourier Transform

In physics and mathematics, k-space, or reciprocal space, represents the Fourier transform of a function defined in real space, providing a decomposition into spatial frequencies that capture variations across different scales. For a continuous function f(\mathbf{r}) in three-dimensional real space, its k-space representation F(\mathbf{k}) is given by the Fourier transform: F(\mathbf{k}) = \int f(\mathbf{r}) e^{-i \mathbf{k} \cdot \mathbf{r}} \, d^3\mathbf{r}, where \mathbf{k} is the wavevector and the integral extends over all space. This formulation generalizes the one-dimensional case, where the transform relates position x to wavenumber k, and extends naturally to higher dimensions for volumetric data in fields like and . The inverse Fourier transform recovers the original function: f(\mathbf{r}) = \frac{1}{(2\pi)^3} \int F(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} \, d^3\mathbf{k}, ensuring that k-space encodes all information about f(\mathbf{r}) without loss. In computational contexts, such as numerical simulations of periodic systems, the continuous integrals are approximated by the (DFT). For a sampled grid of N points in one dimension, the DFT takes the form: F_k = \sum_{j=0}^{N-1} f_j e^{-2\pi i j k / N}, with the inverse involving a factor of $1/N and the opposite sign in the exponent; this extends to multidimensional grids for efficient computation via the fast Fourier transform (FFT) algorithm, commonly used in materials simulations to map real-space densities to reciprocal-space structure factors. Key properties arise from this Fourier relation. Parseval's theorem ensures energy conservation between domains: \int |f(\mathbf{r})|^2 \, d^3\mathbf{r} = \frac{1}{(2\pi)^3} \int |F(\mathbf{k})|^2 \, d^3\mathbf{k}, preserving the L² norm across the transform (conventions may adjust the $2\pi factor). The convolution theorem states that convolution in real space corresponds to multiplication in k-space: the Fourier transform of f \ast g is F(\mathbf{k}) \cdot G(\mathbf{k}), facilitating efficient operations like filtering or correlation in reciprocal space. This relation derives from the completeness of plane waves as a basis for function expansion. Any square-integrable function f(\mathbf{r}) can be expressed as a superposition of plane waves: f(\mathbf{r}) = \frac{1}{(2\pi)^3} \int F(\mathbf{k}) e^{i \mathbf{k} \cdot \mathbf{r}} \, d^3\mathbf{k}. To find the coefficients F(\mathbf{k}), project f(\mathbf{r}) onto the complex conjugate basis e^{-i \mathbf{k} \cdot \mathbf{r}} by integrating: multiply both sides by e^{-i \mathbf{k}' \cdot \mathbf{r}} and integrate over \mathbf{r}, yielding \int f(\mathbf{r}) e^{-i \mathbf{k}' \cdot \mathbf{r}} \, d^3\mathbf{r} = F(\mathbf{k}') due to orthogonality \int e^{i (\mathbf{k} - \mathbf{k}') \cdot \mathbf{r}} \, d^3\mathbf{r} = (2\pi)^3 \delta(\mathbf{k} - \mathbf{k}'). Thus, k-space amplitudes F(\mathbf{k}) quantify contributions from each spatial frequency, fully representing variations in f(\mathbf{r}).

K-space in Physics

Wavevectors and Momentum Space

In wave mechanics, the wavevector \mathbf{k} is defined as \mathbf{k} = \frac{2\pi}{\lambda} \hat{\mathbf{n}}, where \lambda is the and \hat{\mathbf{n}} is the in the direction of wave propagation. This vector encodes both the of the wave, with magnitude k = 2\pi / \lambda, and its direction of propagation for plane waves of the form e^{i \mathbf{k} \cdot \mathbf{r}}. In quantum mechanics, k-space is equivalent to momentum space through the de Broglie relation \mathbf{p} = \hbar \mathbf{k}, where \hbar is the reduced Planck's constant, making k-space a scaled version of momentum space. The time-independent for a , -\frac{\hbar^2}{2m} \nabla^2 \psi(\mathbf{r}) = E \psi(\mathbf{r}), transforms via the into an algebraic form in k-space: E = \frac{\hbar^2 k^2}{2m}, highlighting the direct correspondence between wavevectors and momentum eigenstates. This provides the mathematical tool to represent wavefunctions in k-space, where position-space convolutions become multiplications. Plane wave solutions in k-space are given by \psi_{\mathbf{k}}(\mathbf{r}) = \frac{1}{\sqrt{V}} e^{i \mathbf{k} \cdot \mathbf{r}}, normalized over a finite volume V to ensure \int_V |\psi_{\mathbf{k}}(\mathbf{r})|^2 d^3\mathbf{r} = 1. These form a complete basis, satisfying the completeness relation \sum_{\mathbf{k}} |\mathbf{k}\rangle \langle \mathbf{k}| = I in the discrete case for a large , or \int \frac{d^3\mathbf{k}}{(2\pi)^3} |\mathbf{k}\rangle \langle \mathbf{k}| = I in the limit, allowing any wavefunction to be expanded as a superposition of s. For free particles in a large volume, the states in k-space exhibit a density, with the number of states in a differential volume d^3\mathbf{k} given by \frac{V}{(2\pi)^3} d^3\mathbf{k}, reflecting the even distribution of allowed wavevectors without gaps in the . This underscores the dense sampling of states available to free particles, essential for understanding quantum superpositions and evolution.

Reciprocal Lattice Construction

In periodic crystal structures, k-space is discretized by the , which provides a basis for representing wavevectors in momentum space while respecting the translational symmetry of the real-space lattice. The vectors are defined such that they satisfy the orthogonality condition with the direct lattice vectors, ensuring that plane waves with wavevectors differing by a vector are equivalent due to the periodicity. The reciprocal lattice vectors \mathbf{b}_i (for i = 1, 2, 3) are constructed from the primitive direct vectors \mathbf{a}_1, \mathbf{a}_2, \mathbf{a}_3 using the formula \mathbf{b}_i = 2\pi \frac{\mathbf{a}_j \times \mathbf{a}_k}{\mathbf{a}_i \cdot (\mathbf{a}_j \times \mathbf{a}_k)}, where the indices j, k are cyclic permutations of i (i.e., $1 \to 2 \to 3). This definition guarantees that for any direct lattice vector \mathbf{R} = n_1 \mathbf{a}_1 + n_2 \mathbf{a}_2 + n_3 \mathbf{a}_3 (with integers n_l) and any vector \mathbf{G} = m_1 \mathbf{b}_1 + m_2 \mathbf{b}_2 + m_3 \mathbf{b}_3 (with integers m_l), the relation \mathbf{G} \cdot \mathbf{R} = 2\pi times an integer holds, implying e^{i \mathbf{G} \cdot \mathbf{R}} = 1. Consequently, the of a e^{i (\mathbf{k} + \mathbf{G}) \cdot \mathbf{r}} remains unchanged under translations by \mathbf{R}, making \mathbf{k} and \mathbf{k} + \mathbf{G} physically indistinguishable in the periodic system. To construct the reciprocal lattice, one begins with the primitive cell in real space, defined by the direct vectors \mathbf{a}_i, and generates the in k-space via the above formula for \mathbf{b}_i. The reciprocal lattice points \mathbf{G} form a that tiles k-space periodically. The volume of the primitive cell in reciprocal space V^* = |\mathbf{b}_1 \cdot (\mathbf{b}_2 \times \mathbf{b}_3)| is related to the volume of the real-space primitive cell V = |\mathbf{a}_1 \cdot (\mathbf{a}_2 \times \mathbf{a}_3)| by V^* = (2\pi)^3 / V, reflecting the inverse scaling between the two spaces. This volume relation ensures that the density of k-points in the Brillouin zone matches the density of real-space unit cells. The first Brillouin zone is the fundamental domain of the reciprocal lattice, defined as the Wigner-Seitz primitive cell centered at the origin \mathbf{k} = 0. It consists of all points in k-space closer to the origin than to any other reciprocal lattice point, constructed by drawing perpendicular bisectors (Bragg planes) between the origin and neighboring \mathbf{G} points and taking the enclosed polyhedron. The volume of the first Brillouin zone is (2\pi)^3 / V, equal to that of the reciprocal primitive cell, and its boundaries correspond to planes where wavevectors \mathbf{k} and \mathbf{k} + \mathbf{G} (for some \mathbf{G} \neq 0) produce indistinguishable scattering due to the crystal periodicity. Wavevectors are thus restricted to this zone, with periodicity folding higher zones back into it. This structure underpins , which states that the eigenfunctions of the in a periodic potential can be written as \psi_{\mathbf{k}}(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u_{\mathbf{k}}(\mathbf{r}), where u_{\mathbf{k}}(\mathbf{r}) is periodic with the lattice periodicity (u_{\mathbf{k}}(\mathbf{r} + \mathbf{R}) = u_{\mathbf{k}}(\mathbf{r})) and \mathbf{k} lies within the first .

Applications in Solid-State Physics

Electronic Band Structures

In , electronic band structures map the allowed energy levels of electrons in crystalline materials as a function of their wavevector k in reciprocal space, denoted as E_n(\mathbf{k}), where n labels the band index. This dependence arises from solving the for electrons in a periodic potential, with E_n(\mathbf{k}) exhibiting periodicity matching the due to . The serves as the fundamental domain in k-space for plotting these structures, capturing the essential symmetries without redundancy. Band structures are commonly computed using models like the nearly free electron approximation or the tight-binding method. In the nearly free electron model, the weak periodic potential perturbs the parabolic dispersion of free electrons, E(\mathbf{k}) = \frac{\hbar^2 k^2}{2m}, resulting in energy gaps at the Brillouin zone boundaries where Bragg reflection causes strong backscattering of electron waves, forbidding certain energies and forming band gaps. The tight-binding model, conversely, starts from localized atomic orbitals on lattice sites, with band formation occurring through overlap and hybridization of these orbitals, yielding dispersive bands whose width depends on hopping integrals between neighboring sites. Visualization of band structures typically involves plotting E_n(\mathbf{k}) along high-symmetry paths in the , revealing energy contours or surfaces defined by E(\mathbf{k}) = constant. The , an isosurface at the E_F, delineates occupied from unoccupied states in k-space and profoundly influences electronic properties; in metals, it intersects partially filled bands, enabling high conductivity, while in semiconductors, it lies in a gap between . A key example is , a with an indirect bandgap of approximately 1.1 eV, where the valence band maximum occurs at the Γ point and the conduction band minimum near the X point along the Δ direction, necessitating assistance for optical transitions. The density of states g(E), which quantifies available electron states per unit energy, integrates over k-space as g(E) = \frac{V}{(2\pi)^3} \int \delta(E - E(\mathbf{k})) \, d^3\mathbf{k}, where V is the crystal volume, providing insight into thermodynamic and transport properties by weighting band dispersions. In semiconductors and metals, k-space filling up to E_F governs conductivity: full bands in insulators or semiconductors yield zero conductivity at zero temperature, whereas partial filling in metals allows free carrier motion. The effective mass m^*, approximating electron dynamics near band extrema, is derived from band curvature as m^* = \hbar^2 / \left( \frac{d^2 E}{dk^2} \right), often anisotropic and much smaller than the free electron mass in semiconductors, enhancing mobility.

Phonon Dispersion Relations

In , phonons represent the quantized normal modes of collective lattice vibrations in crystalline solids, where each mode corresponds to a discrete excitation energy \hbar \omega_j(\mathbf{k}) for branch j and wavevector \mathbf{k} in . These relations \omega_j(\mathbf{k}) describe how the vibrational frequencies vary with \mathbf{k}, typically plotted as curves along high-symmetry paths within the , the primitive cell of the that delineates the unique range of phonon wavevectors. The construction of phonon dispersion relations relies on solving the classical for atomic displacements under the harmonic approximation, assuming small-amplitude and quadratic . For a with a basis of multiple atoms, the resulting dynamical D_{\alpha\beta}(\mathbf{k}), a $3p \times 3p (where p is the number of atoms per and \alpha, \beta denote Cartesian components), is formed from the of interatomic force constants. The eigenvalues of this yield \omega^2_j(\mathbf{k}) for each branch j, while the eigenvectors describe the mode patterns. To handle finite , Born-von Kármán are imposed, treating the as a to ensure translational invariance and discrete \mathbf{k}-points. In monatomic crystals, there are three acoustic branches where frequencies vanish at the center (\mathbf{k} = 0), with linear \omega \approx v |\mathbf{k}| at long wavelengths, corresponding to at speed v. For polyatomic crystals, additional optical branches emerge, featuring non-zero frequencies at \mathbf{k} = 0 due to relative motion between sublattices of differing masses or charges, often appearing relatively flat near the zone center as the modes become localized. Phonon dispersion relations underpin thermal properties, particularly the lattice specific heat C_V, derived from the phonon g(\omega) obtained by integrating over \mathbf{k}-space in the . The simplifies this by approximating all acoustic branches with linear dispersion \omega = v k up to a cutoff Debye frequency \omega_D, yielding C_V \propto T^3 at low temperatures and matching classical Dulong-Petit behavior at high temperatures.

K-space in Magnetic Resonance Imaging

Data Acquisition Process

In magnetic resonance imaging (MRI), the data acquisition process involves collecting raw radiofrequency signals from excited nuclei and encoding them into k-space, a representation of spatial frequencies that captures the object's magnetization distribution. This encoding is achieved through the application of linear gradients, which impose position-dependent phase shifts on , effectively sampling the of the image at specific k-space locations. The trajectory through k-space during acquisition determines how spatial is mapped, with the position vector \mathbf{k}(t) = \frac{\gamma}{2\pi} \int_0^t \mathbf{G}(\tau) \, d\tau, where \gamma is the and \mathbf{G}(t) is the time-varying . Spatial encoding in k-space relies on two primary gradient mechanisms: frequency encoding and phase encoding. Frequency encoding, typically along the x-direction, occurs during the signal readout period when a readout G_x is applied, causing at different positions to precess at distinct frequencies proportional to their location; this linearly traverses k-space along the k_x dimension as the signal is sampled. Phase encoding, usually along the y-direction, involves brief pulses of a phase-encoding G_y with incrementally varying amplitudes applied before readout; each pulse imparts a position-dependent shift that selects a specific line in k_y, allowing discrete sampling across the phase-encoding direction. These orthogonal encodings fill a two-dimensional k-space grid, with slice selection achieved via a third along z during . In conventional spin-echo or gradient-echo sequences, k-space is filled using a Cartesian , where one phase-encoding line is acquired per time () in a line-by-line manner, progressing systematically from one edge of k-space to the other; this approach ensures uniform sampling but requires multiple excitations, limiting speed. For faster acquisition, echo-planar imaging (EPI) employs rapid oscillations of the readout to traverse multiple k-space lines in a zigzag pattern after a single excitation, enabling whole-plane coverage in milliseconds and facilitating applications like functional MRI. Non-Cartesian trajectories, such as spirals or radial projections, further accelerate filling by continuously modulating gradients to trace efficient paths through k-space, often starting from the center and spiraling outward or projecting from the origin. The central region of k-space, corresponding to low spatial frequencies, encodes broad image features and , while the peripheral regions at high spatial frequencies capture fine details and edges; the order of filling influences image appearance, with sequences prioritizing the center (e.g., centric ordering) enhancing at the expense of if acquisition is truncated. at high k-values risks artifacts or reduced , as fewer samples fail to adequately represent sharp boundaries, though this can be mitigated in faster sequences by design trade-offs. The conceptual foundations of k-space acquisition trace back to Fourier transform techniques in nuclear magnetic resonance (NMR) spectroscopy developed in the 1970s, which enabled efficient of frequency-domain data. Practical implementation in emerged with the introduction of gradient-based localization in the mid-1970s, as demonstrated in early two-dimensional methods. The k-space formalism was first patented in 1981 and independently formalized in 1983, providing a unified framework for analyzing and designing acquisition trajectories.

Image Reconstruction Techniques

In (MRI), the standard method for reconstructing images from fully sampled k-space data involves applying a two-dimensional () or three-dimensional () inverse (FFT) to convert the frequency-domain data into the spatial domain. This process assumes uniform Cartesian sampling of k-space, where the inverse FFT directly yields the and images without additional corrections, providing a computationally efficient reconstruction that has been the cornerstone of clinical MRI since the technique's early development. To enhance image quality or interpolate for display purposes, zero-filling—inserting zeros into the undersampled portions of k-space before the inverse FFT—can be employed, which effectively increases the apparent matrix size and reduces pixelation but does not add new information. Advanced reconstruction techniques address the limitations of full sampling by enabling faster scans through of k-space, thereby reducing acquisition time while mitigating associated artifacts. Parallel imaging methods, such as Sensitivity Encoding (), exploit the spatial sensitivity profiles of multi-coil receiver arrays to unfold aliased images resulting from undersampled k-space data, achieving acceleration factors of 2–8 depending on coil geometry and noise levels. Similarly, Generalized Autocalibrating Partially Parallel Acquisitions () synthesizes missing k-space lines from acquired data using kernel-based across coils, allowing robust reconstruction even with non-Cartesian trajectories and reducing sensitivity to coil calibration errors. further advances by enforcing image sparsity in a transform domain (e.g., ), combined with incoherent sampling patterns in k-space and iterative optimization to recover high-fidelity images from as little as 20–30% of the data, particularly effective for dynamic or high-resolution applications. More recently, (DL) techniques have revolutionized MRI reconstruction by using neural networks to learn mappings from undersampled k-space or intermediate images to high-quality reconstructions. These methods, including convolutional neural networks (CNNs) for direct inversion, unrolled networks that incorporate physics-based priors, and generative models, enable acceleration factors beyond traditional limits, improve artifact suppression, and enhance robustness to variations in data. As of 2025, DL-based reconstruction is increasingly integrated into clinical scanners for faster imaging protocols. Undersampling and truncation in k-space introduce specific artifacts that reconstruction algorithms must address to preserve diagnostic quality. Gibbs ringing, or truncation artifact, manifests as oscillatory overshoots and undershoots near high-contrast edges due to abrupt cutoff of high spatial frequencies in k-space, with the ringing typically 9% of the edge height for a rectangular truncation window. Motion artifacts arise from inconsistent k-space filling, where patient movement between phase-encoding steps causes phase inconsistencies that propagate as blurring or ghosting across the image, often mitigated by navigator echoes or prospective correction during acquisition. , or wraparound, occurs when violates the , folding peripheral signals into the field of view; this is resolved in advanced methods through regularization techniques, such as total variation minimization in , which enforce smoothness and sparsity to suppress replicas. The and in reconstructed MRI images are fundamentally tied to k-space coverage. High spatial frequencies at the periphery of k-space determine fine details and resolution, with the pixel size approximated as \Delta x \sim 1 / \Delta k, where \Delta k is the maximum extent of k-space sampling in the frequency-encoding direction; insufficient peripheral coverage thus leads to blurring. Conversely, low spatial frequencies near the k-space center encode overall image and , as they represent the dominant signal ; perturbations here, such as from motion, disproportionately affect differentiation. Partial methods exploit the Hermitian of k-space to reconstruct images from asymmetrically sampled data, typically acquiring 60–80% of the full grid and estimating the remainder via phase correction or , thereby shortening scan times by 20–40% without significant resolution loss.

References

  1. [1]
    [PDF] Reciprocal Space Fourier Transforms
    k is a wave-number and has units that are reciprocal to x: x -> cm k -> 2π/cm. So while x describes a position in space, k describes a spatial modulation.
  2. [2]
    [PDF] Lecture 3 - Diffraction and Reciprocal Lattice
    k = (kx, ky, kz ) is a vector in reciprocal space. • Reciprocal space is defined independent of any crystal! •The reciprocal lattice is the set of Fourier ...
  3. [3]
    [PDF] Band Structures and the Meaning of the Wave Vector k - UCSB MRL
    The energies of the bands are calculated in “k-space” or sometimes called “momentum space”. ... The Brillouin Zone exists in reciprocal space and is directly ...
  4. [4]
    k-space - Questions and Answers ​in MRI
    k-space is an array of numbers representing spatial frequencies in the MR image. The common pictorial representation of k-space as a "galaxy" perpetuates the ...
  5. [5]
    K-Space - an overview | ScienceDirect Topics
    K space is defined as a type of space characterized as a Banach lattice, which is an ordered linear space where the norm is compatible with the lattice ...
  6. [6]
    [PDF] Chapter 2 Reciprocal Lattice
    A reciprocal lattice is defined with reference to a particular Bravais lattice. • A set of vectors satisfying is called a reciprocal lattice only if the set ...<|control11|><|separator|>
  7. [7]
    Electromagnetic waves - Richard Fitzpatrick
    \begin{displaymath} \lambda = \frac{2\pi}{, (440). (see Fig. 35) ... $c$ is related to the magnitude of the wave-vector and the angular wave frequency via ...
  8. [8]
    [PDF] Chapter 4: Frequency Domain and Fourier Transforms
    On the other hand if we had a way to represent the “coarse” (or low frequency) information separately from the “fine” (or high frequency) information, we could.Missing: physics | Show results with:physics
  9. [9]
    [PDF] 12 Scattering in three dimensions
    ... space with coordi- nates (kx,ky,kz). The points form a cubic lattice with lattice spacing 2π/L. Thus the volume of k–space per lattice point is (2π/L)3, and ...
  10. [10]
    [PDF] Reciprocal Space and Brillouin Zones in Two and Three Dimensions
    We do this by constructing a region in reciprocal space that surrounds k = 0 such that all k-points enclosed are closer to k = 0 than to any other reciprocal ...
  11. [11]
    Joseph Fourier (1768 - 1830) - Biography - MacTutor
    Joseph Fourier studied the mathematical theory of heat conduction. He established the partial differential equation governing heat diffusion and solved it by ...
  12. [12]
    K-Space formulation of MRI - Stan's Hub
    Nov 23, 2007 · This article reviews in an easily accessible way the application of the universal concept of K-space to magnetic resonance imaging (MRI) and ...Missing: primary | Show results with:primary<|control11|><|separator|>
  13. [13]
    [PDF] Unfolding and effective bandstructure calculations as discrete real
    Mar 7, 2016 · Here we show that a discrete real-space description, based on discrete Fourier transforms, is fully general. Furthermore, such an approach ...
  14. [14]
    [PDF] Convolution Fourier Convolution
    One of the most central results of Fourier Theory is the convolution theorem (also called the Wiener-Khitchine theorem. ℑ{ f ⊗ g}= F(k)⋅ G(k) where, f(x) ⇔ F ...
  15. [15]
    [PDF] Fourier Transform - UCLA Department of Mathematics
    Very broadly speaking, the Fourier transform is a systematic way to decompose. “generic” functions into a superposition of “symmetric” functions.
  16. [16]
    [PDF] CHM 305 - Lecture 2 - The Classical Wave Equation
    where k ≡ 2π/λ is the wave vector (the number of radians of oscillation the wave undergoes per unit distance, representing a sort of “spatial frequency”), and ω ...
  17. [17]
    [PDF] Momentum vs. Wavevector
    ∴ k = ħ p p=ħk. Wavevector definition: Important commutators: Eigenvalue equation: Relation to momentum eigenstates: kp pck ħ= = Normalization. Closure.
  18. [18]
    Planck/de Broglie relation - MIT
    ... Broglie wavelength of a particle is given by \lambda_\dB = \frac hp. From these two relations we find that p = h λ = ℏ k p = \frac h\lambda = \hbar k p=λh​=ℏk.
  19. [19]
    [PDF] Chapter 8. The Momentum Representation
    Oct 10, 2013 · We converted the differential Schrödinger equation for (x|ψ) into an integral equation for the Fourier transform of the wave function. Often ...
  20. [20]
    [PDF] The Fourier Transform and Free Particle Wave Functions
    The Fourier transform decomposes a spatial function into Fourier modes, and the inverse transform reconstructs the original function. f(x) and A(k) are a  ...
  21. [21]
    [PDF] Lecture 6: Time Evolution and the Schrödinger Equation
    Feb 26, 2013 · The Schrödinger equation is a ... so the solution for all time is a traveling wave i(kx−ωt) ψ(x, t; E) = e disregarding normalization.<|control11|><|separator|>
  22. [22]
    [PDF] Quantum ergodicity: fundamentals and applications. - Physics
    Nov 30, 2013 · To get the last result we inserted the resolution of identity. I = ∫ dk. 2πh. |k⟩⟨k| inside the matrix element appearing in the integral. In ...
  23. [23]
    [PDF] Density of States Derivation
    The k space volume taken up by each allowed state is 3 / x y z. LLL π . The reciprocal is the state density in k space (# of states per volume in k space),. 3.
  24. [24]
    [PDF] 4 Crystal Lattices - AMOLF
    Since any direct lattice vector R has this property. (again by (5.2)), all direct lattice vectors are in the lattice reciprocal to the reciprocal lattice.
  25. [25]
    Brillouin Zone - an overview | ScienceDirect Topics
    The first Brillouin zone is defined as the Wigner–Seitz primitive cell of the reciprocal lattice. Thus, it is the set of points in the reciprocal space that is ...
  26. [26]
    [PDF] About the Quantum Mechanics of the Electrons in Crystal Lattices
    Abstract: This article aims to review Felix Bloch theorem of electron motion in a crystal lattice through his seminal paper that has also ...
  27. [27]
    [PDF] Lecture 4 - Nearly Free Electron Model
    Basic Assumptions: - crystal structure is periodic. - periodicity leads to formation of energy bands (allowed energy levels). - energy bands are separated ...
  28. [28]
    [PDF] Energy bands - Rutgers Physics
    We see that at the zone edge, the scattering is so strong that the reflected wave has the same amplitude as the incident wave. The electron is represented there ...
  29. [29]
    Nearly free electron model - Open Solid State Notes
    In this lecture, we will analyze how electrons behave in solids using the nearly-free electron model. ... We first find the band structure using the nearly free ...Missing: Ashcroft Mermin
  30. [30]
    Tight-binding methods - ScienceDirect.com
    Tight-binding theory becomes an independent, and in that sense ab initio, method for studying virtually all of the properties of surfaces.Missing: original | Show results with:original
  31. [31]
    Fermi Energy and Fermi Surface - Engineering LibreTexts
    Sep 7, 2021 · For each partially filled band there will be a surface in k-space separating the occupied from the unoccupied levels. The set of all such ...Fermi Function · Fermi Distribution Function to... · Fermi Surface
  32. [32]
    [2404.05634] Why does silicon have an indirect band gap? - arXiv
    Apr 8, 2024 · We show that its indirect gap arises from a competition between first and second nearest-neighbor bonds-where second nearest-neighbor interactions pull the ...Missing: point | Show results with:point
  33. [33]
    [PDF] Density of States: - Georgia Institute of Technology
    Dec 17, 2005 · The density of states function describes the number of states that are available in a system and is essential for determining the carrier ...
  34. [34]
    Effective Mass Approximation - an overview | ScienceDirect Topics
    The effective mass of an electron in semiconductors is typically one or two orders of magnitude smaller than that of the free electron.
  35. [35]
    [PDF] 4. Phonons
    Figure 57: Phonon dispersion relation for a diatomic chain. We make the ansatz u2n = Ae i!t2ikna and u2n+1 = B e i!t2ikna.Missing: seminal | Show results with:seminal
  36. [36]
    Phonon Dispersion - an overview | ScienceDirect Topics
    The phonon dispersion relations are defined as the k dependence of the frequencies, ω k , j , of the normal modes for all branches j and selected directions in ...Missing: seminal | Show results with:seminal
  37. [37]
    Dynamical Theory of Crystal Lattices - Max Born; Kun Huang
    The book begins with a general discussion of the statistical mechanics of ideal lattices, leading to the electric polarizability and to the scattering of light.
  38. [38]
  39. [39]
    Location of Spatial Frequencies - k-space - MRI Questions
    As an overview, the center of k-space contains low spatial frequency information, determining overall image contrast, brightness, and general shapes. The ...
  40. [40]
    Image Formation by Induced Local Interactions - Nature
    a new way of seeing. In 1973, Paul Lauterbur described an imaging technique that removed the usual resolution limits due to the ...
  41. [41]
    US4307343A - Moving gradient zeugmatography - Google Patents
    Nuclear magnetic resonance phenomena are employed to generate a two dimensional image of a thin planar slice through a body under investigation.
  42. [42]
    MODEL-BASED IMAGE RECONSTRUCTION FOR MRI - PMC
    The inverse fast Fourier transform (FFT) has served the MR community very well as the conventional image reconstruction method for k-space data with full ...Missing: seminal | Show results with:seminal
  43. [43]
    SENSE: sensitivity encoding for fast MRI - PubMed
    Abstract. New theoretical and practical concepts are presented for considerably enhancing the performance of magnetic resonance imaging (MRI) by means of arrays ...
  44. [44]
    Generalized autocalibrating partially parallel acquisitions (GRAPPA)
    In this study, a novel partially parallel acquisition (PPA) method is presented which can be used to accelerate image acquisition using an RF coil array for ...
  45. [45]
    Sparse MRI: The application of compressed sensing for rapid MR ...
    Sparse MRI: The application of compressed sensing for rapid MR imaging. Magn Reson Med. 2007 Dec;58(6):1182-95. doi: 10.1002/mrm.21391. Authors. Michael Lustig ...
  46. [46]
    Removal of Partial Fourier-Induced Gibbs (RPG) Ringing artifacts in ...
    Theory and Methods: Gibbs ringing of fully sampled data, leading to oscillations around tissue boundaries, is caused by the symmetric truncation of k-space.
  47. [47]
    Body MR Imaging: Artifacts, k-Space, and Solutions - PubMed Central
    Motion artifacts can be overcome through respiratory synchronization, alternative k-space sampling schemes, and parallel imaging. Aliasing and truncation ...Missing: inconsistent | Show results with:inconsistent
  48. [48]
    Parallel MR Imaging - PMC - NIH
    While it is possible to accelerate MRI data acquisition by collecting fewer phase-encoding lines in k-space, the resulting spatial aliasing must be removed ...
  49. [49]
    spatial frequency, image resolution and contrast | e-MRI - IMAIOS
    Most MR image information (contrast and general shape) is contained in the center of k-space. Low-spatial-frequency data have the highest amplitude.Missing: central | Show results with:central
  50. [50]
    Partial Fourier reconstruction for three-dimensional gradient echo ...
    Partial Fourier (PF) methods take advantage of data symmetry to allow for either faster image acquisition or increased image resolution.