Fact-checked by Grok 2 weeks ago

Dispersion relation

In wave physics, a dispersion relation is a mathematical that relates the \omega of a to its number k, encapsulating the propagation characteristics of waves in a physical medium. This relation is derived from the governing s of the system, such as the for mechanical waves or Maxwell's equations for electromagnetic waves, and it determines how different components evolve over space and time. For instance, in vacuum, the dispersion relation for electromagnetic waves is linear and given by \omega = c k, where c is the speed of light, implying non-dispersive propagation where all wavelengths travel at the same speed. The dispersion relation distinguishes between phase velocity v_p = \omega / k, which describes the speed of individual wave crests, and group velocity v_g = d\omega / dk, which governs the propagation of wave packets and the transport of energy or information. In dispersive media, where the relation is nonlinear (e.g., \omega^2 = c^2 k^2 + \omega_p^2 for electromagnetic waves in a plasma, with \omega_p as the plasma frequency), different frequencies travel at different speeds, causing wave packets to spread out over time—a phenomenon central to signal distortion in optics and communications. Linear dispersion relations, by contrast, yield constant velocities and no spreading, as seen in ideal sound waves where \omega = v k and v is the speed of sound. Dispersion relations find broad applications across physics, including acoustics (e.g., propagation in air or solids), fluid dynamics (e.g., surface on with \omega^2 = g k \tanh(k h), where g is and h is depth), and solid-state physics (e.g., dispersion in , influencing and lattice vibrations)./08%3A_Waves/8.02%3A_The_Dispersion_Relation) In , they underpin de Broglie relations for matter , linking particle p = \hbar k to E = \hbar \omega. These relations also reveal phenomena like band gaps in periodic structures and instabilities in plasmas, making them essential for analyzing wave phenomena in engineering and systems.

Basic Concepts

Definition and Terminology

A dispersion relation describes the functional dependence of the angular frequency \omega on the wave number k for waves propagating in a medium, typically expressed as \omega = \omega(k). This relation arises in the study of linear wave equations, where small-amplitude approximations allow the superposition of plane wave solutions. The wave number k is defined as k = 2\pi / \lambda, where \lambda is the wavelength, representing the spatial periodicity of the wave. Similarly, the angular frequency \omega is given by \omega = 2\pi f, with f denoting the ordinary frequency, capturing the temporal oscillation rate in radians per second. The fundamental form of a plane wave solution is e^{i(kx - \omega t)}, where x is the position and t is time; this complex exponential encodes both the oscillatory and propagating nature of the wave. In dispersive media, the dispersion relation \omega(k) is nonlinear in k, leading to a frequency-dependent propagation speed that causes wave packets to spread over time. Conversely, non-dispersive propagation occurs when \omega(k) is linear, such as \omega = c k for some constant c, resulting in all frequency components traveling at the same constant speed without distortion. Key velocities associated with the dispersion relation include the phase velocity v_p = \omega / k, which describes the speed of constant-phase surfaces, and the v_g = d\omega / dk, which indicates the propagation speed of the packet . These concepts are central to understanding wave behavior in contexts like acoustics, , and , always under the framework of linear, small-amplitude waves.

Phase and Group Velocities

The v_p of a monochromatic is defined as the velocity at which a surface of constant propagates through the medium. For a wave described by the dispersion relation \omega = \omega(k), where \omega is the angular frequency and k is the wavenumber, the is derived from the phase factor \phi = kz - \omega t in the wave expression e^{i(kz - \omega t)}. Setting the total constant for a moving point, \frac{dz}{dt} = \frac{\omega}{k}, yields v_p = \frac{\omega}{k}. This represents the speed at which individual crests or troughs of the advance, though it does not necessarily correspond to the propagation of energy or information. The group velocity v_g, in contrast, describes the velocity of the overall envelope of a wave packet formed by the superposition of waves with wavenumbers centered around some k_0. To derive it, consider the dispersion relation expanded via around k_0: \omega(k) \approx \omega(k_0) + \left. \frac{d\omega}{dk} \right|_{k=k_0} (k - k_0) + \frac{1}{2} \left. \frac{d^2\omega}{dk^2} \right|_{k=k_0} (k - k_0)^2 + \cdots The wave packet is then a product of a e^{i(k_0 z - \omega(k_0) t)} and an modulated by the spread in k. The propagates at v_g = \left. \frac{d\omega}{dk} \right|_{k=k_0}, as this term determines the shift in the phase synchronization across the packet. Physically, v_g corresponds to the velocity of transport in the wave, since the follows the in linear media. In non-dispersive media, where \omega(k) = v k for constant v, the phase and group velocities coincide: v_p = v_g = v, and wave packets maintain their shape without broadening. However, in dispersive media, where \frac{d^2 \omega}{dk^2} \neq 0, v_p \neq v_g, causing the wave packet to spread over time as different frequency components travel at varying speeds; this pulse broadening limits in applications like optical communications. In dispersive systems, the signal velocity—the speed at which information or a detectable front propagates—must respect and cannot exceed the in . shows this equals the at the point of stationary phase, ensuring no superluminal despite possible anomalous values of v_p or v_g in certain ranges.

Non-Dispersive Waves

Electromagnetic Waves in Vacuum

In , electromagnetic waves arise as solutions to in free space, which in the absence of charges and currents take the form: \nabla \cdot \mathbf{E} = 0, \quad \nabla \cdot \mathbf{B} = 0, \nabla \times \mathbf{E} = -\frac{\partial \mathbf{B}}{\partial t}, \quad \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \frac{\partial \mathbf{E}}{\partial t}, where \mathbf{E} is the electric field, \mathbf{B} is the magnetic field, \mu_0 is the vacuum permeability, and \epsilon_0 is the vacuum permittivity. To derive the wave equation, take the curl of Faraday's law \nabla \times \mathbf{E} = -\partial \mathbf{B}/\partial t: \nabla \times (\nabla \times \mathbf{E}) = -\frac{\partial}{\partial t} (\nabla \times \mathbf{B}). Substitute Ampère's law with Maxwell's correction \nabla \times \mathbf{B} = \mu_0 \epsilon_0 \partial \mathbf{E}/\partial t, yielding \nabla \times (\nabla \times \mathbf{E}) = -\mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}. Using the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E} and \nabla \cdot \mathbf{E} = 0, this simplifies to the wave equation \nabla^2 \mathbf{E} = \mu_0 \epsilon_0 \frac{\partial^2 \mathbf{E}}{\partial t^2}, or equivalently, \frac{\partial^2 \mathbf{E}}{\partial t^2} = c^2 \nabla^2 \mathbf{E}, where c = 1/\sqrt{\mu_0 \epsilon_0} is the speed of light in vacuum. A similar wave equation holds for \mathbf{B}. Assuming plane-wave solutions of the form \mathbf{E} = \mathbf{E}_0 e^{i(\mathbf{k} \cdot \mathbf{r} - \omega t)}, substitution into the wave equation yields the dispersion relation \omega = c |\mathbf{k}| for transverse electromagnetic waves, where \mathbf{k} is the wave vector and \omega is the angular frequency. This linear relation indicates that the phase velocity v_p = \omega / |\mathbf{k}| = c and the group velocity v_g = d\omega / d|\mathbf{k}| = c are both constant and equal to the speed of light. The constancy of both velocities implies no dispersion in vacuum: all frequencies propagate at the same speed c, so wave packets do not broaden over distance. Electromagnetic waves in are transverse, as the condition \nabla \cdot \mathbf{E} = 0 for plane waves requires \mathbf{k} \cdot \mathbf{E}_0 = 0, meaning the is to the . The of the wave is defined by the orientation of \mathbf{E}_0 in the plane to \mathbf{k}, which can be linear, circular, or elliptical.

Uniform Waves on a String

The transverse displacement y(x, t) of small-amplitude waves propagating along an idealized infinite string under constant satisfies the one-dimensional derived from Newton's second applied to a small string element. The net transverse force on the element arises from the difference in the vertical components of at its ends, leading to \frac{\partial^2 y}{\partial t^2} = v^2 \frac{\partial^2 y}{\partial x^2}, where the wave speed v = \sqrt{T / \mu}, with T the constant and \mu the constant linear . This equation assumes small slopes (|\partial y / \partial x| \ll 1) to linearize the tension components, transverse motion only, and no external forces or material damping. For plane wave solutions of the form y(x, t) = A \cos(kx - \omega t + \phi), substitution into the wave equation yields the linear dispersion relation \omega(k) = v |k|, where \omega is the and k is the . This relation is independent of , characteristic of non-dispersive where all wave components travel at the same speed. The phase velocity v_p = \omega / k = v and the group velocity v_g = d\omega / dk = v are both equal to the wave speed and constant, independent of frequency or wavenumber. Consequently, wave packets maintain their shape without spreading during propagation. In the case of a finite-length string fixed at both ends, boundary conditions y(0, t) = y(L, t) = 0 quantize the allowed wavenumbers to k_n = n \pi / L (for positive integer n), resulting in standing wave modes with frequencies \omega_n = v k_n. However, the focus here remains on traveling waves in the infinite string limit, where the dispersion relation governs free propagation without boundary-induced quantization. The idealized model neglects dispersion arising from bending stiffness, which introduces a higher-order restoring force proportional to the fourth spatial of displacement, causing to increase with frequency and leading to inharmonic partials. Similarly, nonlinear effects, such as those from large-amplitude vibrations altering tension, are excluded, as they would frequency-dependently modify propagation.

Dispersive Waves in Classical Media

Optical Waves in Dispersive Materials

In dispersive optical materials, such as and , the propagation of electromagnetic is characterized by a frequency-dependent refractive n(\omega), leading to the dispersion relation k = \frac{\omega n(\omega)}{c}, where k is the wave number, \omega is the , and c is the in vacuum. This relation emerges from for plane in isotropic with a frequency-dependent dielectric \epsilon(\omega) = n^2(\omega), reflecting the resonant interaction of with bound electrons in the . The refractive index n(\omega) is commonly modeled using the empirical Sellmeier equation, which approximates n^2(\lambda) = 1 + \sum_i \frac{B_i \lambda^2}{\lambda^2 - C_i}, where \lambda = 2\pi c / \omega is the vacuum , and the coefficients B_i and C_i (related to oscillator strengths and wavelengths) are determined from experimental measurements for specific materials. This form provides accurate predictions over broad ranges away from strong absorptions and was originally proposed by Wolfgang Sellmeier to explain anomalous color sequences in spectra. Materials exhibit normal dispersion in spectral regions distant from electronic absorption bands, where \frac{dn}{d\omega} > 0, so higher frequencies propagate more slowly than lower ones, as seen in visible light through glass. In contrast, anomalous dispersion occurs near ultraviolet or resonances, where \frac{dn}{d\omega} < 0, causing the refractive index to decrease with increasing frequency and potentially leading to superluminal phase velocities, though energy transport remains subluminal. These behaviors arise from the Lorentz oscillator model of atomic polarizability. A practical consequence of optical dispersion is chromatic dispersion in waveguides like silica optical fibers, where varying group velocities cause temporal broadening of light pulses, limiting high-speed data transmission over long distances. The group velocity is expressed as v_g = \frac{c}{n + \omega \frac{dn}{d\omega}}, quantifying how pulse envelopes propagate; in normal dispersion regimes typical for telecommunications wavelengths around 1550 nm, v_g < c/n, resulting in positive dispersion parameter D > 0 that spreads pulses. The real and imaginary parts of the \epsilon(\omega) = \epsilon_r(\omega) + i \epsilon_i(\omega), which determine and respectively, are interconnected by the Kramers-Kronig relations, principal-value integrals ensuring the causal response of materials to electromagnetic fields. These relations, derived from the analyticity of the dielectric function in the complex frequency plane, were independently formulated by Hendrik Kramers and Ralph Kronig in the late 1920s.

Deep Water Waves

Deep water waves refer to small-amplitude gravity waves propagating on the surface of an inviscid, incompressible occupying a domain of infinite depth, where the water depth h greatly exceeds the \lambda such that kh \gg 1, with k = 2\pi/\lambda being the . These are governed by the linearized Euler equations, assuming irrotational , which allows the introduction of a \phi satisfying \nabla^2 \phi = 0. The is two-dimensional, with propagating in the x-direction and vertical coordinate z increasing upward from the mean surface at z=0. To derive the dispersion relation, assume a monochromatic solution of the form \phi(x,z,t) = \mathrm{Re} \{ \hat{\phi}(z) e^{i(kx - \omega t)} \}, where \omega is the . Substituting into yields \hat{\phi}(z) = A e^{kz} for the deep water case, as the exponentially decaying solution ensures boundedness as z \to -\infty. The linearized kinematic boundary at the z=0 relates the vertical to the surface \eta, giving \partial_t \eta = \partial_z \phi. The dynamic boundary , enforcing constant pressure at the surface, linearizes to \partial_t \phi + g \eta = 0 at z=0. Combining these yields the dispersion relation \omega^2 = g k in the deep water limit. The phase velocity is v_p = \omega / k = \sqrt{g / k} = \sqrt{g \lambda / (2\pi)}, which increases with , meaning longer propagate faster than shorter ones. The , representing the speed of , is v_g = d\omega / dk = \frac{1}{2} v_p = \frac{1}{2} \sqrt{g / k}. This causes wave packets to spread out over time, with longer components advancing ahead of shorter ones, which explains the evolution of wind-generated wave fields into organized swell patterns where dominant longer separate from trailing shorter . This approximation holds for small-amplitude waves where the surface displacement is much less than the wavelength (a \ll \lambda) and under the deep water condition kh \gg 1, typically valid for ocean waves with periods less than about 20 seconds. Particle orbits are circular near the surface with radius decaying exponentially with depth as a e^{k z}, confining motion to the upper layer.

Acoustic Phonons in Solids

In crystalline solids, acoustic phonons arise from the collective of atoms in the , representing quantized modes of sound waves that propagate through the material. These modes are derived from the harmonic approximation of lattice dynamics, where atoms are treated as point masses connected by springs. Phonons, as the of these vibrations, provide a quantum mechanical description of the classical lattice oscillations. The simplest model for these is the one-dimensional monatomic , consisting of atoms of m separated by equilibrium distance a, interacting via nearest-neighbor harmonic springs with force constant K. The equation of motion for the u_n of the n-th atom is given by
m \ddot{u}_n = K (u_{n+1} + u_{n-1} - 2 u_n).
Assuming plane-wave solutions u_n = A e^{i(k n a - \omega t)}, this yields the dispersion relation
\omega(k) = 2 \sqrt{\frac{K}{m}} \left| \sin\left( \frac{k a}{2} \right) \right|,
where k is the wavevector in the first [-\pi/a, \pi/a]. This relation was foundational in the development of lattice dynamics.
The acoustic branch corresponds to this single dispersion curve, where all atoms move in phase at long wavelengths. For small k (long wavelengths, k a \ll 1), the relation linearizes to \omega \approx v_s k, with sound speed v_s = a \sqrt{K/m}, mimicking classical sound waves in a continuum. Near the Brillouin zone edge (k \approx \pi/a), the dispersion becomes nonlinear, with \omega flattening due to the periodic lattice structure, leading to standing-wave-like behavior. The group velocity, v_g = d\omega/dk, equals v_s at the zone center (k = 0) for propagating waves but approaches zero at the zone edges, where modes resemble stationary vibrations. In three dimensions, the model extends to a cubic with atoms having of , resulting in longitudinal acoustic () modes—where displacements are to the wavevector—and two transverse acoustic () modes—where displacements are . The remains linear at low k, with \omega = v_l k for LA modes (longitudinal speed v_l) and \omega = v_t k for TA modes (transverse speed v_t), derived from the constants of the material. The approximates this low-frequency behavior by assuming a linear \omega = v k up to a cutoff wavevector, treating the solid as a continuum of acoustic modes to simplify calculations of thermodynamic properties; here, v is an effective speed averaging v_l and v_t. This approximation, introduced by Debye, captures the essential physics for low-energy excitations. Acoustic phonons play a central role in sound propagation, as their linear dispersion at long wavelengths directly corresponds to the transmission of elastic waves through the solid at speeds v_l and v_t. Additionally, they dominate thermal conductivity in insulators and semiconductors, where heat transport occurs via phonon diffusion; the group velocity determines the rate of energy carry, while scattering from defects or anharmonicity limits the mean free path, as described by the Boltzmann transport equation for phonons.

Dispersive Waves in Quantum Systems

De Broglie Relations for Matter Waves

In 1924, Louis de Broglie proposed the hypothesis that particles of matter, such as electrons, exhibit wave-like properties, extending the wave-particle duality observed for light to all matter. This idea posits that every moving particle is associated with a periodic phenomenon, characterized by a wavelength \lambda and frequency f related to its momentum p and total energy E via the de Broglie relations: \lambda = h / p and f = E / h, where h is Planck's constant. In terms of angular frequency \omega = 2\pi f and wave number k = 2\pi / \lambda, these become \omega = E / \hbar and k = p / \hbar, with \hbar = h / 2\pi. For a , these relations yield the dispersion relation \omega(k) by substituting the energy-momentum relation into \omega = E / \hbar. In the non-relativistic limit, the is E = p^2 / 2m, where m is the particle , leading to p = \hbar k and thus \omega(k) = \hbar k^2 / 2m. This parabolic dispersion relation indicates that the wave is dispersive, as the v_p = \omega / k = \hbar k / 2m = p / 2m, which is half the particle v = p / m. However, the v_g = d\omega / dk = \hbar k / m = p / m matches the classical velocity of the particle, representing the propagation speed of the wave packet envelope that corresponds to the particle's motion. De Broglie's hypothesis was initially formulated in a relativistic context, where the total energy is E = \sqrt{(pc)^2 + (mc^2)^2}, with c the . Substituting p = \hbar k gives the relativistic dispersion relation \omega(k) = \sqrt{(\hbar c k)^2 + (m c^2)^2} / \hbar. In this case, the v_p = \omega / k = c^2 / v > c is superluminal, while the group velocity remains v_g = d\omega / dk = v < c, consistent with the particle's subluminal speed and ensuring no information travels faster than light. These relations form the foundation for describing matter waves in quantum mechanics, highlighting the dispersive nature inherent to massive particles.

Electron Dispersion in Band Structures

In solid-state physics, the dispersion relation for electrons in a crystal lattice describes how the energy E_n(\mathbf{k}) of an electron in the n-th energy band depends on its crystal momentum \mathbf{k}, which is confined to the first due to the periodic potential of the lattice. This relation extends the free-particle dispersion E = \frac{\hbar^2 k^2}{2m} by incorporating the effects of the lattice, leading to the formation of energy bands separated by bandgaps. The provides the foundational framework for this description, stating that the eigenfunctions of the Schrödinger equation in a periodic potential V(\mathbf{r}) = V(\mathbf{r} + \mathbf{R}), where \mathbf{R} is a , can be written as \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}), with u_{n\mathbf{k}}(\mathbf{r}) periodic and having the lattice periodicity. This form implies that the energy eigenvalues E_n(\mathbf{k}) are periodic in reciprocal space, E_n(\mathbf{k} + \mathbf{G}) = E_n(\mathbf{k}) for reciprocal \mathbf{G}, resulting in distinct energy bands labeled by the band index n. The approximates the relation by treating the periodic potential as a weak to the free-electron gas. In this approach, the unperturbed is parabolic, E(\mathbf{k}) \approx \frac{\hbar^2 k^2}{2m}, but the weak potential introduces Bragg scattering at the boundaries, where \mathbf{k} = \frac{\mathbf{G}}{2}, causing energy gaps to open due to avoided crossings between degenerate plane waves. The resulting near these boundaries shows a modified by the , approximately E(\mathbf{k}) \approx \frac{\hbar^2 k^2}{2m^*} + higher-order , where the effective m^* accounts for the influence. This model serves as a zero-order approximation extending the de Broglie relation for free particles, but it incorporates periodicity to explain formation. The effective mass m^* quantifies the curvature of the energy bands and is defined as m^* = \frac{\hbar^2}{\frac{d^2 E}{dk^2}} along a principal direction in \mathbf{k}-space, reflecting how the lattice modifies the electron's response to external fields compared to its bare mass m. In regions of high band curvature, m^* is small, leading to enhanced mobility, while flat bands imply large m^* and reduced response. The band structure divides into valence bands, filled with electrons at absolute zero, and conduction bands, which are empty; the bandgap E_g is the energy difference between the top of the valence band and the bottom of the conduction band. Bandgaps are classified as direct if the conduction band minimum and valence band maximum occur at the same \mathbf{k} (e.g., in GaAs), allowing efficient momentum-conserving optical transitions, or indirect if they occur at different \mathbf{k} (e.g., in Si), requiring phonon assistance for transitions. These dispersion features have key implications for electron transport in semiconductors. The group velocity, which determines the electron's propagation speed, is given by \mathbf{v}_g = \frac{1}{\hbar} \nabla_{\mathbf{k}} E(\mathbf{k}), analogous to the free-particle case but curved by the band structure, enabling directional transport in anisotropic materials. Electron mobility \mu = \frac{e \tau}{m^*}, where \tau is the relaxation time and e the charge, is inversely proportional to m^*, making low-effective-mass semiconductors like InSb suitable for high-speed devices, while indirect bandgaps in materials like silicon favor non-radiative recombination but enable efficient doping for transistors.

Mathematical Formulations

General Derivation from Wave Equations

The derivation of dispersion relations typically begins with a linear partial differential equation (PDE) governing wave propagation, such as the general second-order form \frac{\partial^2 \psi}{\partial t^2} = L\left(\frac{\partial}{\partial x}\right) \psi, where L is a linear spatial differential operator with constant coefficients. To solve this, one assumes a ansatz of the form \psi(x, t) = A e^{i(kx - \omega t)}, where A is a constant amplitude, k is the wave number, and \omega is the angular frequency. Substituting this into the PDE replaces time derivatives with -\omega^2 \psi and spatial derivatives with powers of ik, yielding the characteristic equation -\omega^2 + D(k) = 0, where D(k) is the Fourier symbol of the operator L, obtained by evaluating L at ik. This algebraic relation \omega = \omega(k) constitutes the dispersion relation, linking the frequency to the wave number and determining the wave's propagation characteristics. For illustration, consider the Klein-Gordon for a relativistic , (\partial_t^2 - c^2 \partial_x^2 + m^2 c^4 / \hbar^2) \psi = 0. Applying the yields the dispersion relation \omega^2 = c^2 k^2 + m^2 c^4 / \hbar^2, where the highlights how the mass term introduces dispersion by making \omega(k) nonlinear in k. The focus remains on the substitution process, which systematically extracts D(k) = c^2 k^2 + m^2 c^4 / \hbar^2 from the operator, emphasizing the general applicability across linear PDEs rather than the physics of the specific . The resulting normal modes correspond to solutions where \omega(k) is real-valued for all k, indicating propagating waves without decay or growth. Real \omega(k) ensures stability in the sense of bounded oscillatory solutions, as imaginary parts would imply exponential amplification or damping, which are analyzed separately for stability assessments. In non-dispersive cases, such as the standard wave equation, \omega(k) = c |k| yields constant phase speed, but the general framework reveals dispersion when D(k) is nonlinear. For multi-component systems, such as those described by vector fields or coupled equations (e.g., a chain of coupled oscillators), the extends to \psi_j(x, t) = A_j e^{i(kx - \omega t)} for components j = 1, \dots, n. This leads to a eigenvalue problem, where the dispersion relation emerges from the \det(-\omega^2 I + D(k)) = 0, with D(k) now a incorporating terms. In a one-dimensional lattice of N coupled oscillators, for instance, the normal modes satisfy a dispersion relation like \omega^2 = 4 \kappa / m \sin^2(k a / 2), derived via this approach, illustrating how interactions produce band-like spectra of propagating modes. This generalization underscores the method's versatility for systems beyond scalar waves.

Linearization and Approximations

In nonlinear wave systems, dispersion relations are frequently derived by the governing partial differential equations around an state, assuming small-amplitude perturbations. This process entails expanding the nonlinear terms in a and retaining only the linear contributions, as higher-order terms become negligible for sufficiently small deviations from . The resulting linear PDE admits plane-wave solutions of the form \exp[i(kx - \omega t)], leading to a dispersion relation \omega(k) that relates the \omega to the k. A representative example is the nonlinear Schrödinger equation (NLSE), i \hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \frac{\partial^2 \psi}{\partial x^2} + V |\psi|^2 \psi, which describes wave envelopes in nonlinear media such as optical fibers or Bose-Einstein condensates. For weak nonlinearity, where the coefficient V is small, the equation approximates the linear Schrödinger equation i \hbar \frac{\partial \psi}{\partial t} = -\frac{\hbar^2}{2m} \frac{\partial^2 \psi}{\partial x^2}, with the parabolic dispersion relation \omega = \frac{\hbar k^2}{2m}. This linear limit captures the essential dispersive spreading of the wave packet without self-phase modulation effects from the cubic nonlinearity. To incorporate weak nonlinear effects beyond the , is employed to compute corrections to the dispersion relation. The solution is expanded as \psi = \psi_0 + \epsilon \psi_1 + \cdots, where \epsilon quantifies the nonlinearity strength, and substitution into the full nonlinear PDE yields successive orders. At , this produces dispersion corrections of the form \delta \omega(k) = \epsilon f(k), where f(k) depends on the specific nonlinearity, enabling quantitative predictions of frequency shifts for mildly nonlinear . The validity of linearization and these perturbative corrections holds primarily in the small-k regime or long-wavelength limit, where linear propagation dominates over nonlinear interactions, and for amplitudes below thresholds that trigger instabilities like modulational instability. Breakdown occurs when nonlinear terms comparable to linear ones lead to phenomena such as wave steepening or soliton formation, invalidating the plane-wave assumption. For initial value problems governed by linear dispersive PDEs, facilitates extraction of the dispersion relation by decomposing the initial condition into a continuum of plane waves via the \hat{u}(k,0). Each mode then evolves independently as \hat{u}(k,t) = \hat{u}(k,0) \exp(-i \omega(k) t), so the temporal spectrum at fixed k reveals \omega(k) through phase accumulation or content analysis. This approach underpins numerical and analytical studies of evolution in dispersive media.

Historical Development

Early Concepts in Optics and Waves

The concept of in wave originated in the through optical experiments that revealed the separation of white light into colors when passing through . In 1672, conducted seminal experiments, observing that refracted by a produced a of colors, with each color deviating by a different angle; this indicated that the speed of light in the medium varied with color or wavelength, laying the groundwork for understanding frequency-dependent . Newton's findings, detailed in his letter to the Royal Society and later expanded in Opticks (1704), challenged earlier views of light as a homogeneous entity and suggested inherent material properties influencing wave speeds. By the early , the wave theory of , advanced by in Traité de la Lumière (1690) and refined by in works from 1815 to 1819, modeled as vibrations in an . In this , early formulations assumed a linear dispersion relation, where the was independent of , implying non-dispersive with constant speed in the ; however, these models began to incorporate effects to explain observed color separation in prisms. Fresnel's contributions, including his explanation of and , highlighted linear wave relations but increasingly accounted for ether perturbations in dense media, setting the stage for quantitative dispersion models. Empirical approaches to quantify optical dispersion emerged in the mid-19th century, focusing on the wavelength dependence of the refractive index. In 1836, Augustin-Louis Cauchy proposed a simple empirical formula relating the refractive index n to wavelength \lambda, capturing the increase in n for shorter wavelengths observed in transparent materials like glass. This was refined in 1871 by Wolfgang Sellmeier, who introduced a physically motivated formula incorporating resonance effects from atomic vibrations, providing better fits to experimental data across broader spectral ranges and influencing subsequent studies of material optics. In the 1880s, Lord Rayleigh advanced the theoretical understanding of dispersive by distinguishing between —the speed of individual wave crests—and —the speed of the overall or energy —in where velocity depends on . Rayleigh's , building on earlier wave theories, showed that in dispersive systems like light in glass or , these velocities differ, with determining signal ; his work in The of () and subsequent papers clarified these effects for both acoustic and optical contexts. Concurrently, 19th-century studies of by George and ( ) identified the dispersive character of , where increases with , leading to spreading of wave packets over . 's 1847 paper on oscillatory motions and 's 1870s investigations into tidal and ocean demonstrated how longer outpace shorter ones, providing classical examples of dispersion beyond .

Quantum and Solid-State Contributions

In 1924, proposed the hypothesis of wave-particle duality for matter, extending the concept of waves associated with particles beyond photons to all matter. This introduced the fundamental dispersion relations linking the \omega to the E via \omega = E / \hbar and the wave vector k to the momentum p via k = p / \hbar, where \hbar is the reduced Planck's constant. These relations provided the foundational framework for describing matter waves in , predicting wave-like interference for particles such as electrons. Building on de Broglie's ideas, Erwin Schrödinger developed the wave equation in 1926, which for a free particle yields a parabolic dispersion relation E = \frac{\hbar^2 k^2}{2m}, or equivalently \omega = \frac{\hbar k^2}{2m}, where m is the particle mass. This non-relativistic form describes how the energy of a free quantum particle disperses quadratically with wave vector, contrasting with the linear dispersion of light and highlighting the diffusive propagation of matter waves. The Schrödinger equation thus formalized the quantum description of dispersion for isolated particles, enabling solutions for bound states and scattering problems. In 1928, extended these to electrons in periodic potentials, such as , through his stating that functions can be expressed as modulated by periodic functions, \psi(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u(\mathbf{r}) with u(\mathbf{r} + \mathbf{R}) = u(\mathbf{r}) for vectors \mathbf{R}. This led to the emergence of energy band structures, where the relation E(\mathbf{k}) becomes periodic in reciprocal space, allowing forbidden energy gaps and explaining the electrical properties of solids. work marked a pivotal quantum of in condensed , shifting from particles to collective effects. During the , advanced this framework by introducing Brillouin zones in , which delineate regions where wave vectors \mathbf{k} lead to distinct energy bands due to Bragg reflections in the . These zones, first detailed in his of in metals, complement by providing a geometric of allowed states and highlighting discontinuities at zone boundaries. Concurrently, the evolved from the classical model established by and von Kármán in , which treated atomic displacements as coupled harmonic oscillators in a crystal . Their quantum extension in the 1920s, applying Schrödinger's equation to these modes, quantized the vibrations into phonons—bosonic quasiparticles with relations \omega(\mathbf{q}) derived from the dynamical matrix, enabling the description of thermal and acoustic properties in solids. Following World War II, dispersion relations found critical applications in solid-state physics, particularly in semiconductors, where band structures dictate carrier transport. In the 1950s, William Shockley's theory of p-n junctions and transistors relied on effective mass approximations from near-parabolic dispersion bands E(\mathbf{k}) \approx E_g + \frac{\hbar^2 k^2}{2m^*}, where m^* is the effective mass and E_g the bandgap, to model minority carrier injection and amplification. This work underpinned the development of transistor technology, revolutionizing electronics by leveraging quantum dispersion for device design.

References

  1. [1]
    [PDF] MIT 8.03SC Fall 2016 Textbook: The Physics of Waves
    Waves are everywhere. Everything waves. There are familiar, everyday sorts of waves in water, ropes and springs. There are less visible but equally ...
  2. [2]
    Dispersion Relation - an overview | ScienceDirect Topics
    Dispersion relation is defined as a mathematical expression that describes the relationship between the frequency and wavelength of waves in a given medium, ...
  3. [3]
    Dispersion relation - MIT
    A linear dispersion relation results in a constant phase velocity, while a nonlinear relation results in a phase velocity that is a function of frequency.Missing: definition | Show results with:definition<|control11|><|separator|>
  4. [4]
    [PDF] Lecture 11: Wavepackets and dispersion
    Recall the dispersion relation is defined as the relationship between the frequency and the wavenumber: ω(k). For non-dispersive systems, like most of what ...Missing: physics | Show results with:physics
  5. [5]
    [PDF] Wave Motion 1 - Duke Physics
    The wave number k is related to λ by k = 2π /λ . • The angular frequency is ω = 2π f . For a wave moving in the +x direction, x and t must appear in the ...
  6. [6]
    The Feynman Lectures on Physics Vol. I Ch. 29: Interference - Caltech
    It is oscillatory at the angular frequency ω. The angular frequency ω can be defined as the rate of change of phase with time (radians per second).
  7. [7]
    Waves in a Dispersive Medium - Graduate Program in Acoustics
    May 28, 2008 · In a non-dispersive medium all of the different frequency components travel at the same speed so the wave function doesn't change at all as it ...Missing: definition | Show results with:definition
  8. [8]
    [PDF] SIO 210: Waves
    Nov 14, 2019 · Non-dispersive waves*: group velocity equals phase velocity. Dispersive waves: group velocity not equal to phase velocity. *Non-dispersive: ω ...
  9. [9]
    Pulse Propagation
    Sinusoidal waves that satisfy nonlinear dispersion relations, such as (9.6) or (9.7), are known as dispersive waves, as opposed to waves that satisfy linear ...
  10. [10]
    Group Velocity
    The phase velocity is the propagation velocity of an individual wave maximum, whereas the group velocity is the propagation velocity of an interference peak.<|control11|><|separator|>
  11. [11]
    [PDF] 6.055J / 2.038J The Art of Approximation in Science and Engineering
    Jan 14, 2008 · So the group velocity is one-half of the phase velocity, as the result for power-law dispersion relation predicts. Within a wave train, the ...
  12. [12]
    [PDF] Maxwell's Equations
    Mar 15, 2013 · As far as the wave equation is concerned, the dispersion relation is the only constraint that needs to be satisfied in order for Eq. (18) to be ...
  13. [13]
    Faster-than-Light Laser Pulses?
    Jan 27, 2002 · In vacuum, all frequencies of light have a phase velocity of c, but in a transparent "dispersive" medium like glass or water, a given wave ...<|control11|><|separator|>
  14. [14]
    [PDF] Electromagnetic waves
    Electromagnetic waves traveling in a vacuum obey a perfect wave equation and have no dispersion. Page 4. 4. 4. Harmonic traveling wave aka sinusoidal waves.
  15. [15]
    [PDF] Lecture 14: Polarization
    \E0\. \. = c\. \B0\. \. (3) and that k · E0 = 0, k · B0 =0, E0 · B0 = 0. (4). In other words, the polarization vector of the electric field, the polarization ...Missing: transverse | Show results with:transverse
  16. [16]
    [PDF] Derivation of the Wave Equation
    Derivation of the Wave Equation. In these notes we apply Newton's law to an elastic string, concluding that small amplitude transverse vibrations of the ...
  17. [17]
    Standing Waves in Finite Continuous Medium
    The only simple solution of the wave equation, (6.1), that has stationary nodes and anti-nodes is a standing wave of the general form.
  18. [18]
    [PDF] Ideal Vibrating String - CCRMA
    Mar 12, 2008 · Recursion typically started by assuming zero past displacement: y(n, m)=0,n = −1,0,∀m. • Higher order wave equations yield more terms of the.
  19. [19]
    [PDF] Chapter 9: Electromagnetic Waves - MIT OpenCourseWare
    May 9, 2011 · The physical significance of (9.2. 1) is divided into two parts: Eo tells us the polarization, amplitude, and absolute phase of the wave at the ...
  20. [20]
    Zur Erklärung der abnormen Farbenfolge im Spectrum einiger ...
    ADS. Zur Erklärung der abnormen Farbenfolge im Spectrum einiger Substanzen. Sellmeier. Abstract. Publication: Annalen der Physik. Pub Date: 1871 ... harvard.edu.
  21. [21]
    refractive index, Sellmeier equation, dispersion formula - RP Photonics
    Sellmeier equations are very useful, as they make it possible to describe fairly accurately the refractive index in a wide wavelength range.
  22. [22]
    Anomalous Dispersion - Ocean Optics Web Book
    Oct 14, 2021 · This reversal of “colors” from that seen in normal life at visible wavelengths is called “anomalous dispersion.”
  23. [23]
    Chromatic Dispersion - RP Photonics
    Chromatic dispersion is the frequency dependence of phase velocity in a medium. It also affects the group velocities of light pulses.What is Chromatic Dispersion? · Modern Parameters for...
  24. [24]
    Group Velocity Dispersion - RP Photonics
    Group velocity dispersion is the frequency dependence of group velocity, or the derivative of inverse group velocity with respect to angular frequency.
  25. [25]
    Gravity Waves in Deep Water - Richard Fitzpatrick
    In deep water, where depth exceeds wavelength, gravity waves' phase velocity is proportional to the square root of wavelength, and group velocity is half the ...
  26. [26]
    [PDF] Gravity waves on water - UMD Physics
    Gravity waves on water arise from gravity, with surface tension neglected for longer wavelengths. Their speed is proportional to √ gλ or √ gh.<|control11|><|separator|>
  27. [27]
    [PDF] PROCESSES - Benoit Cushman-Roisin
    Nov 30, 2012 · In the case of deep-water gravity waves, the group travels half as fast as crests ... Derive dispersion relation (4.71) for inertia-gravity waves.
  28. [28]
    [PDF] Lecture 2: Concepts from Linear Theory - WHOI GFD
    which implies that shallow water waves are approximately non-dispersive and all travel with the same speed at a similar depth. This makes waves in shallow ...
  29. [29]
    [PDF] 4. Phonons - DAMTP
    The lower ! part of the dispersion relation is called the acoustic branch. The upper !+ part is called the optical branch. To understand where these names come ...
  30. [30]
    Born, M. and von Karman, T. (1912) Uber Schwingungen ... - Scirp.org.
    This article emphasizes that the Einstein and Debye models of specific heats of solids are correlated more tightly than currently acknowledged.
  31. [31]
    [PDF] Debye-1912.pdf
    Für die Zwecke diesesParagraphen genügt es,zuwissen,daB F eine für jeden Korper aus seiner Dichte und seinen elastischen. Konstanten berechenbare.
  32. [32]
    [PDF] Thermal conductivity Review: phonons as particles - Vishik Lab
    As these phonons diffuse to the cooler side of a solid, they will scatter many times and the value of the thermal conductivity coefficient will depend on how ...
  33. [33]
    [PDF] Louis de Broglie - Nobel Lecture
    Thus to describe the properties of matter as well as those of light, waves and corpuscles have to be referred to at one and the same time. The electron can ...Missing: primary | Show results with:primary
  34. [34]
    [PDF] Matter Waves - INFN
    The dispersion relation for de Broglie waves can be obtained as a function of k by substituting p ⫽ h/␭ ⫽បk into Equation 5.23. This gives. (5.24).
  35. [35]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.Missing: source | Show results with:source
  36. [36]
  37. [37]
    [PDF] Dispersion relations, stability and linearization 1 ... - Arizona Math
    Plugging either (1) or. (2) into the equation yields an algebraic relationship of the form ω = ω(k) or σ = σ(k), called the dispersion relation. It ...
  38. [38]
    [PDF] DERIVATION AND ANALYSIS OF SOME WAVE EQUATIONS
    In Section 4.2 we will do this for transverse waves on a tight string, and for Maxwell's equations describing electromagnetic waves. In both of these cases ...
  39. [39]
    [PDF] 129 Lecture Notes - Relativistic Quantum Mechanics 1 Need for ...
    This equation is called Klein–Gordon equation. You can find plane-wave ... that this equation gives Einstein's dispersion relation E2 = p2c2 + m2c4 like.
  40. [40]
    [PDF] Chapter 6. From Oscillations to Waves - Academic Commons
    In this chapter, the discussion of oscillations is extended to systems with two and more degrees of freedom. This extension naturally leads to another key ...
  41. [41]
    [PDF] 04 Linear Chain of Coupled Oscillators - DigitalCommons@USU
    We will find such a relationship in each instance of wave phenomena.* For reasons we shall discuss later, such a relation is called a dispersion relation.Missing: multi- component
  42. [42]
    Nonlinear Schrodinger systems: continuous and discrete
    Aug 15, 2008 · The Nonlinear Schrodinger (NLS) equation is a prototypical dispersive nonlinear partial differential equation (PDE) that has been derived in ...<|separator|>
  43. [43]
    [PDF] Derivation of Nonlinear Schrödinger Equation - arXiv
    Apr 1, 2011 · In addition, we prove the coefficient of nonlinear term is very small, i.e., the nonlinearity of Schrödinger equation is weak. We think ...
  44. [44]
    Anomalous Correlators in Nonlinear Dispersive Wave Systems
    May 26, 2020 · In its essence, the classical wave turbulence theory is a perturbation expansion in the amplitude a k of the nonlinearity, yielding, at the ...
  45. [45]
    [PDF] Perturbation theory for the Nonlinear Schrödinger Equation... - arXiv
    It is also argued that the same mecha- nism results in delocalization for the model (1.1) with sufficiently large β, while, for weak nonlinearity, localization ...
  46. [46]
    [PDF] Chapter 10 - Signals and Fourier Analysis - MIT OpenCourseWare
    We work out in some detail an example of the scattering of a wave packet. iv. We discuss the dispersion relation for electromagnetic waves in more detail and ...Missing: initial | Show results with:initial
  47. [47]
    How Isaac Newton's experiments revealed the mystery of light
    Nov 11, 2023 · Newton begins his 1672 account by relating his surprise that the colored spectrum produced by his prism was rectangular in shape rather than ...
  48. [48]
    The Science of Color - Smithsonian Libraries
    Opticks, one of the great works in the history of science, documents Newton's discoveries from his experiments passing light through a prism. He identified the ...
  49. [49]
    [PDF] THE WAVE-THEORY OFLIGHT - hlevkin
    ... Huygens, Newton, Young, and Fresnel, that each has in turn been considered the founder of the modern science of optics. What justification there is for each ...
  50. [50]
    Refraction—as the light is bent - Book chapter - IOPscience
    These are used to fit the data and find intermediate values by interpolation. The Cauchy formula, described by Augustin-Louis Cauchy (1789–1857) in 1836 , is.
  51. [51]
    (INVITED) Methods for determining the refractive indices and thermo ...
    2.2. Sellmeier equation ... The Sellmeier equation, first derived in 1871 [60], is one of the most widely used representations of refractive index dispersion.
  52. [52]
    Kelvin's method of stationary phase? - ScienceDirect.com
    Kelvin added a footnote to Eq. (8): “This is the group-velocity according to Lord Rayleigh's generalisation of Prof. Stokes's original result”. Later, Larmor, ...
  53. [53]
    [PDF] recherches sur la théorie des quanta - UB
    La théorie du rayonnement et les quanta (1er Congrès Solvay,. 1911), publiée par P. LANGEVIN et M. DE BROGLIE. M. PLANCK, Theorie der Wärmestrahlung, J.-A.
  54. [54]
    [PDF] 1926-Schrodinger.pdf
    The theory which is reported in the following pages is based on the very interesting and fundamental researches of L. de Broglie¹ on what he called "phase-waves ...Missing: citation | Show results with:citation
  55. [55]
    [PDF] Über die Quantenmechanik der Elektronen in Kristallgittern
    Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses durch ein zun~chst streng dreifaeh periodisches Kraftfeld schematisieren. Unter.Missing: Quantentheorie | Show results with:Quantentheorie
  56. [56]
    [PDF] Léon Brillouin and the Brillouin Zone - Physics Courses
    Contents of the paper (Brillouin, 1930). The paper studies selective reflection (Bragg reflection) and its relationship with the anomalies of propagation of ...Missing: citation | Show results with:citation
  57. [57]
    BSTJ 28: 3. July 1949: The Theory of p-n Junctions ... - Internet Archive
    Jan 19, 2013 · The Theory of p-n Junctions in Semiconductors and p-n Junction Transistors. (Shockley, W.) Bell System Technical Journal, 28: 3. July 1949 pp 435-489.