Fact-checked by Grok 2 weeks ago

Phonon

A phonon is representing the quantized collective vibrational modes of atoms in , analogous to as the quantum of electromagnetic . These excitations arise from the periodic arrangement of atoms in solids, where causes oscillations that propagate as , with each phonon carrying a discrete amount of energy given by E = \hbar \omega, where \hbar is the reduced Planck's constant and \omega is the angular frequency. As bosons obeying Bose-Einstein statistics, phonons can occupy the same quantum state, enabling phenomena like Bose-Einstein condensation in certain systems. In , phonons play a central role in understanding , electrical, and of materials. They dominate heat transport in insulators and dielectrics through phonon-phonon , which limits thermal conductivity, as described by the Boltzmann transport equation for phonons. Phonons also mediate electron-phonon interactions, which are crucial for conventional , where pairing of electrons via phonon exchange leads to zero-resistance states below critical temperatures, as established in . Additionally, phonons influence at low temperatures, following the , which predicts a T^3 dependence to the of phonon states. Phonons exhibit dispersion relations, where depends on wavevector, leading to acoustic and optical branches in multi-atom cells; acoustic phonons contribute to propagation, while optical phonons are involved in absorption. Recent advances, including topological phonons and anharmonic effects, have expanded their study to and quantum technologies, highlighting their role beyond classical .

Fundamentals

Definition and Basic Concepts

In solid-state physics, phonons are quasiparticles that represent the collective excitations resulting from the vibrational motions of atoms arranged in a crystal . These excitations arise from the ordered structure of atoms in solids, where small displacements from positions propagate as through the material. Unlike fundamental particles such as electrons, phonons are not elementary but emerge as effective descriptions of many-body interactions; they are bosonic quasiparticles, meaning they follow Bose-Einstein statistics and can occupy the same . The energy of a phonon is given by E = \hbar \omega, where \hbar is the reduced Planck's constant and \omega is the of the associated vibrational mode. A consisting of N atoms possesses $3N due to the three-dimensional motion of each , corresponding to $3N independent normal modes of vibration and thus $3N possible phonon modes. This finite number of modes contrasts with the infinite in free space but provides a complete basis for describing . Phonons bear a close to photons, the quantized excitations of electromagnetic , in that both are massless bosons mediating interactions—photons for electromagnetic forces and phonons for mechanical forces in solids—though phonons are confined to the . The foundational description of these vibrations begins with the harmonic approximation, where the for the takes the form H = \sum_i \left( \frac{p_i^2}{2m} + \frac{1}{2} k x_i^2 \right), with p_i and x_i as the and of the i-th , m the , and k the effective spring constant; this facilitates subsequent quantization into phonon states.

Historical Development

The concept of normal modes in continuous elastic media, foundational to later phonon theory, emerged in the late 19th century through the work of Lord Rayleigh, who analyzed vibrations in solids and fluids as superpositions of independent oscillatory modes. In 1907, proposed a quantum mechanical model for the specific heat of solids, treating the atoms as independent harmonic oscillators with quantized energy levels, thereby resolving the classical Dulong-Petit law's failure at low temperatures by introducing discrete vibrational quanta. This approach marked an early application of quantum ideas to lattice vibrations, though it assumed a single frequency for all oscillators. Peter Debye refined Einstein's model in 1912 by adopting a continuum approximation for the lattice, assuming a linear dispersion relation for acoustic modes up to a cutoff frequency known as the Debye frequency ω_D, which better captured the low-temperature specific heat behavior through a density of states proportional to ω². Independently in 1912, Max Born and Theodor von Kármán developed a discrete model of lattice dynamics, representing the crystal as a finite chain of atoms connected by springs and imposing periodic boundary conditions to simulate an infinite lattice, enabling the calculation of normal modes without surface effects. In the 1920s, advanced the understanding of lattice vibrations by deriving dispersion relations for phonons in periodic structures, introducing the concept of Brillouin zones in reciprocal space, where zone folding arises from the periodicity, leading to band gaps in the phonon spectrum. In 1930, Soviet physicist introduced the concept of phonons as quasiparticles representing the quantized modes of lattice vibrations. The name "phonon" was suggested by Yakov Frenkel in 1932. During the 1930s, and contributed to the framework by developing techniques, which interpreted lattice vibrations as bosonic fields and provided a many-body operator formalism for phonons, bridging classical normal modes to quantum . In 1950, Herbert Fröhlich introduced the concept, describing an in a polar as a dressed by phonon cloud due to electron-phonon coupling, which quantified the renormalization of and in ionic crystals.

Classical Lattice Dynamics

One-Dimensional Lattice Model

The one-dimensional lattice model serves as an introductory framework for classical lattice dynamics, modeling vibrations in a linear chain of identical atoms. Consider a monatomic chain where each atom has mass m and is connected to its nearest neighbors by harmonic springs of spring constant \kappa, with equilibrium spacing a between atoms. The displacement of the n-th atom from its equilibrium position is u_n(t), assuming longitudinal vibrations along the chain direction..pdf) The equations of motion for the atoms are derived from Newton's second law, considering the restoring forces from the adjacent springs. For the n-th atom, the net force is \kappa (u_{n+1} - u_n) + \kappa (u_{n-1} - u_n) = \kappa (u_{n+1} + u_{n-1} - 2u_n), leading to m \ddot{u}_n = \kappa (u_{n+1} + u_{n-1} - 2u_n). This second-order differential equation describes the coupled harmonic oscillations of the chain..pdf) To solve these equations, assume normal mode solutions of the form u_n(t) = A e^{i(qna - \omega t)}, where q is the wave vector and \omega is the angular frequency. Substituting this ansatz into the equation of motion yields the dispersion relation \omega(q) = 2 \sqrt{\frac{\kappa}{m}} \left| \sin\left( \frac{qa}{2} \right) \right|. This relation shows that the frequency \omega depends on q, with a maximum at the zone boundary and linear behavior near q = 0. The phase velocity is v_p = \omega / q, while the group velocity, representing energy propagation, is v_g = d\omega / dq. For small q, v_g \approx a \sqrt{\kappa / m}..pdf) For a finite chain of N atoms, are imposed via the Born-von Kármán approach, requiring u_{n+N} = u_n. This discretizes the allowed wave vectors as q = 2\pi k / (Na), where k = 0, 1, \dots, N-1, confining q to the first from -\pi/a to \pi/a. In the long-wavelength limit (qa \ll 1), the simplifies to \omega \approx c q, where c = a \sqrt{\kappa / m} is the sound speed, approximating continuum ..pdf)

Three-Dimensional Lattice Vibrations

In three-dimensional crystals, lattice vibrations are analyzed by extending the one-dimensional model to a with a basis consisting of p atoms per primitive . The position of the \kappa-th atom in the l-th is denoted by \mathbf{R}_l + \boldsymbol{\tau}_\kappa, where \mathbf{R}_l is the Bravais lattice vector and \boldsymbol{\tau}_\kappa specifies the position of atom \kappa within the cell. This structure accommodates the complexity of real crystals, such as those with multiple atom types or positions, allowing for vector displacements \mathbf{u}_l(\kappa, t) in Cartesian coordinates \alpha, \beta = x, y, z. The classical treatment assumes small oscillations, with the expanded to second order in displacements using interatomic force constants \Phi_{\alpha\beta}(\mathbf{R}_l - \mathbf{R}_{l'}, \kappa, \kappa'). The for the displacements are m_\kappa \ddot{u}_{\alpha l}(\kappa, t) = \sum_{\beta, l', \kappa'} \Phi_{\alpha\beta}(\mathbf{R}_l - \mathbf{R}_{l'}, \kappa, \kappa') u_{\beta l'}(\kappa', t), where m_\kappa is the mass of atom \kappa. Assuming normal mode solutions of the form u_{\alpha l}(\kappa, t) = e_{\alpha}(\kappa|\mathbf{q}) \exp[i (\mathbf{q} \cdot \mathbf{R}_l - \omega t)], with wavevector \mathbf{q} in the first , the system decouples into a Fourier-transformed form. This leads to the dynamical matrix, which encapsulates the force constants in reciprocal space: D_{\alpha\beta}(\mathbf{q}, \kappa, \kappa') = \frac{1}{\sqrt{m_\kappa m_{\kappa'}}} \sum_{\mathbf{R}} \Phi_{\alpha\beta}(\mathbf{R}, \kappa, \kappa') e^{i \mathbf{q} \cdot \mathbf{R}}, where the sum runs over all Bravais lattice vectors \mathbf{R}. This formulation, originating from the early lattice dynamics models, enables the computation of vibrational frequencies for arbitrary crystal structures. The normal modes are obtained by solving the secular eigenvalue equation \omega^2 e_{\alpha}(\kappa|\mathbf{q}) = \sum_{\beta, \kappa'} D_{\alpha\beta}(\mathbf{q}, \kappa, \kappa') e_{\beta}(\kappa'|\mathbf{q}), a $3p \times 3p matrix problem that yields $3p eigenvalues \omega_j^2(\mathbf{q}) (frequencies squared) and corresponding eigenvectors \mathbf{e}(\kappa|\mathbf{q}, j) (polarization vectors) for each \mathbf{q}. The $3p branches consist of three acoustic branches (one longitudinal and two transverse) and $3(p-1) optical branches, reflecting the degrees of freedom: three translational per atom. The polarization vectors describe the relative displacements of atoms in the mode, normalized such that \sum_{\kappa, \alpha} m_\kappa |e_{\alpha}(\kappa|\mathbf{q}, j)|^2 = 1. This eigenvalue approach, central to classical lattice dynamics, was formalized in the foundational cyclic boundary condition models for finite crystals. In the long-wavelength limit (\mathbf{q} \to 0), the dynamical matrix simplifies, revealing distinct polarizations. For acoustic branches, \omega(\mathbf{q}) \approx v |\mathbf{q}|, where v is the , and the modes decouple into one longitudinal (displacements parallel to \mathbf{q}) and two transverse (perpendicular) polarizations, assuming cubic or isotropic media. The sound speeds are determined by the Christoffel , derived from the \mathbf{q} = 0 dynamical matrix: \rho v_i^2 \hat{e}_\alpha = C_{\alpha\beta\gamma\delta} \hat{n}_\beta \hat{n}_\delta \hat{e}_\gamma, where \rho is the mass density, C_{\alpha\beta\gamma\delta} are the elastic constants, \hat{n} is the unit propagation direction, and \hat{e} the unit polarization. This relates macroscopic elasticity to microscopic vibrations, with the three eigenvalues giving the squared speeds for the polarizations. To bridge from simpler models, consider the generalization of the one-dimensional diatomic chain, where scalar displacements are replaced by , allowing via force constants in all directions and enabling transverse modes absent in 1D. In , this results in six branches for a diatomic basis (p=2): three acoustic and three optical, with polarizations mixing depending on the and \mathbf{q} direction. This vector extension captures realistic effects like in sound propagation and mode degeneracy lifting in non-cubic lattices.

Quantum Mechanical Treatment

Quantization of Lattice Modes

In the quantum mechanical treatment of lattice vibrations, the classical normal modes derived from the three-dimensional dynamics are quantized by associating each mode with an independent . This approach transforms the continuous classical displacements into discrete energy levels, where the quanta of vibration are known as phonons. The process begins with the classical for the , which, after into , separates into uncoupled oscillators for each wavevector \mathbf{q} and polarization branch s. Quantization proceeds by promoting the classical coordinate and to operators satisfying the canonical commutation relations, analogous to the single-particle in . The for a single labeled by \mathbf{q} and s takes the form \hat{H}_{\mathbf{q}s} = \hbar \omega_{\mathbf{q}s} \left( \hat{a}_{\mathbf{q}s}^\dagger \hat{a}_{\mathbf{q}s} + \frac{1}{2} \right), where \omega_{\mathbf{q}s} is the mode frequency, and \hat{a}_{\mathbf{q}s}^\dagger and \hat{a}_{\mathbf{q}s} are the , respectively. These operators are introduced via the of the classical mode amplitudes, replacing the classical energy \frac{1}{2} m \dot{Q}^2 + \frac{1}{2} m \omega^2 Q^2 (with normal coordinate Q) by operator expressions that satisfy [\hat{Q}, \hat{P}] = i\hbar, where \hat{P} is the conjugate . The full is then the sum over all modes: \hat{H} = \sum_{\mathbf{q},s} \hat{H}_{\mathbf{q}s}. This quantization ensures that the energy of each mode is , with eigenvalues \left( n_{\mathbf{q}s} + \frac{1}{2} \right) \hbar \omega_{\mathbf{q}s}, where n_{\mathbf{q}s} = 0, 1, 2, \dots represents the number of phonons in that mode. The atomic displacement operator at lattice site \mathbf{R} of basis atom \kappa is expressed in terms of these operators as \hat{u}_\kappa(\mathbf{R}) = \sum_{\mathbf{q},s} \sqrt{\frac{\hbar}{2 N m_\kappa \omega_{\mathbf{q}s}}} \, \mathbf{e}_{\kappa s}(\mathbf{q}) \left( \hat{a}_{\mathbf{q}s} + \hat{a}_{-\mathbf{q}s}^\dagger \right) e^{i \mathbf{q} \cdot \mathbf{R}}, where N is the number of unit cells, m_\kappa is the mass of atom \kappa, and \mathbf{e}_{\kappa s}(\mathbf{q}) is the polarization vector for branch s. The conjugate momentum operator has a similar form: \hat{p}_\kappa(\mathbf{R}) = i \sum_{\mathbf{q},s} \sqrt{\frac{\hbar m_\kappa \omega_{\mathbf{q}s}}{2 N}} \, \mathbf{e}_{\kappa s}(\mathbf{q}) \left( \hat{a}_{\mathbf{q}s}^\dagger - \hat{a}_{-\mathbf{q}s} \right) e^{i \mathbf{q} \cdot \mathbf{R}}. These expressions ensure the reality of the displacement field and incorporate the Hermitian conjugate terms to maintain physical consistency. The commutation relations [\hat{a}_{\mathbf{q}s}, \hat{a}_{\mathbf{q}'s'}^\dagger] = \delta_{\mathbf{q}\mathbf{q}'} \delta_{ss'}, with all other commutators vanishing, follow from the bosonic nature of the modes, confirming that phonons obey Bose-Einstein statistics. A key consequence of this quantization is the , the ground-state energy of the system when no phonons are excited (n_{\mathbf{q}s} = 0 for all modes), given by \frac{1}{2} \sum_{\mathbf{q},s} \hbar \omega_{\mathbf{q}s}. For a crystal with N primitive cells and three acoustic branches (or more generally $3r modes for r atoms per cell), this sums to approximately \frac{3N \hbar \bar{\omega}}{2}, where \bar{\omega} is an average frequency. This non-zero ground-state energy implies residual lattice vibrations even at temperature, known as zero-point motion, which contributes to phenomena like and elastic constants. The bosonic commutation relations further imply that multiple phonons can occupy the same mode without restriction, unlike fermions.

Phonon Operators and Second Quantization

In the quantum mechanical treatment of lattice vibrations, provides a powerful framework for describing phonons as bosonic quasiparticles in a many-body system, extending the picture to field theory. This approach treats the displacement of the lattice as a quantum field, allowing for the construction of operators that create and annihilate phonons while naturally incorporating their bosonic statistics. The formalism is particularly suited for handling multi-phonon states and interactions in extended systems. The phonon field operator, often denoted as the displacement field ψ(r, t), is expressed as a sum over wavevectors q and branch indices s (for acoustic or optical modes): \psi(\mathbf{r}, t) = \sum_{\mathbf{q} s} \sqrt{\frac{\hbar}{2 \rho \omega_{\mathbf{q} s} V}} \, \mathbf{e}_s(\mathbf{q}) \left( a_{\mathbf{q} s} e^{i (\mathbf{q} \cdot \mathbf{r} - \omega_{\mathbf{q} s} t)} + \mathrm{h.c.} \right), where ρ is the mass density of the crystal, V is the volume, ω_{q s} is the frequency of the mode, e_s(q) is the polarization vector, a_{q s} and a_{q s}^† are the annihilation and creation operators satisfying [a_{q s}, a_{q' s'}^†] = δ_{q q'} δ_{s s'}, and h.c. denotes the Hermitian conjugate. This expression quantizes the classical displacement field, promoting normal modes to operators that act on a Hilbert space of phonon states. The number operator for a specific mode is defined as n_{q s} = a_{q s}^† a_{q s}, which counts the number of phonons in that mode and obeys bosonic commutation relations. The total number of phonons in the system is then N = ∑{q s} n{q s}. These operators enable the description of occupation numbers, with eigenvalues giving the phonon occupancy. The vacuum state |0⟩ is the ground state with no phonon excitations, satisfying a_{q s} |0⟩ = 0 for all q, s, corresponding to the zero-point motion of the lattice. Coherent states can be constructed as displaced vacua, |α⟩ = exp(∑{q s} α{q s} a_{q s}^† - α_{q s}^* a_{q s}) |0⟩, which are eigenstates of the annihilation operators and represent classical-like phonon wavepackets with definite phase and amplitude. Multi-phonon states are built in the Fock space as tensor products of single-mode states, denoted |{n_{q s}}⟩ = ∏{q s} (a{q s}^†)^{n_{q s}} / √(n_{q s} !) |0⟩, where the energy of such a state is E = ∑{q s} n{q s} ℏ ω_{q s} + zero-point energy. This basis spans the full Hilbert space for non-interacting phonons, allowing arbitrary distributions of excitations. Compared to first quantization, which treats fixed numbers of distinguishable oscillators, second quantization excels in managing indistinguishable bosons with variable particle number, facilitating the inclusion of creation and annihilation processes in interactions without explicit symmetrization. This is essential for thermodynamic averages and response functions in solids. For anharmonic effects, which introduce interactions beyond the harmonic approximation, the potential energy terms (cubic, quartic, etc.) are expanded in powers of the displacement field ψ(r). In second quantization, these become perturbation Hamiltonians expressed in terms of a_{q s} and a_{q s}^†, such as three-phonon processes from cubic terms like ∫ d^3r ψ^3(r), enabling diagrammatic techniques for scattering and lifetime calculations.

Phonon Characteristics

Dispersion Relations

The phonon dispersion relations in characterize the dependence of vibrational mode frequencies \omega on the wavevector \mathbf{q} in reciprocal space, obtained as the square roots of the eigenvalues of the dynamical matrix, which encodes the interatomic force constants within the . These relations arise from the normal modes of lattice vibrations and reflect the periodic structure of the , with \omega(\mathbf{q}) being continuous within each branch but varying non-linearly due to the finite range of atomic interactions. Due to the lattice periodicity, the dispersion relations are invariant under translations by reciprocal lattice vectors \mathbf{G}, such that \omega(\mathbf{q} + \mathbf{G}) = \omega(\mathbf{q}), enabling a unique representation within the first . The first corresponds to the Wigner-Seitz cell in reciprocal space, constructed as the region closest to the origin bounded by perpendicular bisectors to neighboring points; this zone folding ensures that all distinct phonon modes are captured without redundancy. Critical points within the , where the gradient \nabla_{\mathbf{q}} \omega = 0, produce Van Hove singularities in the phonon g(\omega), manifesting as logarithmic or power-law divergences that significantly impact properties like and electronic interactions; these features were first theoretically described by Léon van Hove in the context of lattice vibrations. Phonon branches exhibit distinct behaviors depending on their type: acoustic branches display approximately linear dispersion \omega \approx v_s |\mathbf{q}| near the zone center \mathbf{q} = 0, where v_s is the , reflecting collective rigid-body-like translations of the . In contrast, optical branches in crystals with multiple atoms per feature a frequency gap \omega(0) > 0, resulting from relative oscillations between sublattices with differing masses or charges, leading to non-zero frequencies even at long wavelengths. Inelastic neutron scattering serves as the primary experimental technique to map these dispersion relations, probing the dynamic structure factor S(\mathbf{q}, \omega), which is proportional to \sum_{i,f} |\langle f | \sum_j e^{i \mathbf{q} \cdot \mathbf{r}_j} | i \rangle|^2 \delta(\omega - \omega_{fi}) and captures the intensity of phonon creation or annihilation transitions between initial |i\rangle and final |f\rangle states. This method resolves \mathbf{q} and \omega directly, often along high-symmetry paths in the Brillouin zone. For instance, in the face-centered cubic (FCC) lattice of aluminum, a monatomic metal, the three acoustic phonon branches (longitudinal and two transverse) are measured along directions like \Gamma-X, \Gamma-L, and \Gamma-K, showing linear rise from zero at \Gamma with longitudinal velocities around 6.4 km/s and transverse around 3.0 km/s, followed by flattening and avoided crossings near zone boundaries due to the cubic symmetry.

Acoustic and Optical Phonons

In crystals with a basis containing more than one atom per primitive , the phonon modes separate into acoustic and optical branches based on the relative motions of atoms within the . Acoustic phonons correspond to in-phase vibrations where all atoms in the move together, propagating as sound waves through the ; there are three acoustic branches in three dimensions—one longitudinal acoustic () mode and two degenerate transverse acoustic () modes—reflecting the three degrees of freedom per atom. At long wavelengths (small wavevector q), their is linear, \omega \propto |q|, with the proportionality constant being the in the material, which depends on the interatomic forces and atomic masses. Optical phonons, in contrast, arise from out-of-phase motions between atoms of different types or masses in the unit cell, leading to a relative that can couple to electromagnetic fields in ionic crystals. For a basis with p > 1 atoms, there are $3(p-1) optical branches, and these modes exhibit a finite at q = 0 due to the restoring forces from mass differences or electrostatic interactions between charged ions. A simple example is the one-dimensional diatomic lattice model, such as NaCl, where the unit cell has two atoms of different masses; this yields one acoustic branch (in-phase motion) and one optical branch (out-of-phase motion), with the optical mode remaining non-zero at the zone center. In polar crystals, the optical branches further split into longitudinal optical (LO) and transverse optical (TO) modes due to the introduced by long-range forces. The LO frequency is higher than the TO frequency because the longitudinal enhances the , increasing the restoring force; this LO-TO splitting is quantitatively described by the Lyddane-Sachs-Teller (LST) relation: \frac{\omega_{\mathrm{LO}}^2}{\omega_{\mathrm{TO}}^2} = \frac{\varepsilon(0)}{\varepsilon(\infty)}, where \varepsilon(0) and \varepsilon(\infty) are the static and high-frequency constants, respectively. These distinctions have important implications for experimental probes: optical phonons, with their dipole moments in ionic materials, are active in infrared absorption and , allowing direct optical access to their frequencies, whereas acoustic phonons, lacking such coupling, are primarily observed via inelastic .

Phonon Momentum and Interactions

Crystal Momentum

In , phonons are quasiparticles representing quantized vibrations, and their momentum is described by the crystal momentum \hbar \mathbf{q}, where \mathbf{q} is the phonon wavevector defined within the first and modulo a vector \mathbf{G}. This crystal momentum arises from the periodic potential, analogous to electron Bloch states, and governs the conservation laws in phonon interactions. Due to the of the , phonon eigenstates take a Bloch-like form: \exp(i \mathbf{q} \cdot \mathbf{R}) times a periodic , where \mathbf{R} denotes lattice sites, ensuring the wavefunction respects the crystal periodicity. In scattering processes, such as three-phonon interactions, of crystal dictates that the change in total wavevector \Delta \mathbf{q} = \mathbf{0} for processes, preserving overall within the . In contrast, umklapp processes allow \Delta \mathbf{q} = \mathbf{G} (with \mathbf{G} \neq \mathbf{0}), where the total is transferred to the , enabling momentum non-conservation relative to the extended zone scheme. processes redistribute among phonons without net to flow, while umklapp processes introduce irreversible that limits thermal conductivity, particularly at higher temperatures where they become prevalent. Experimentally, the momentum of phonons is probed via inelastic , where the transferred \hbar \mathbf{Q} from the to the alters the phonon wavevector by \Delta \mathbf{q} = \mathbf{Q}, allowing measurement of phonon and /annihilation events. The direction of energy propagation for a phonon is given by its \mathbf{v}_g = \nabla_{\mathbf{q}} \omega(\mathbf{q}), which aligns with the gradient of the \omega(\mathbf{q}) and reflects the 's role in determining properties.

Nonlinear Phonon Effects

In the harmonic approximation, lattice vibrations are modeled using a potential, resulting in non-interacting phonons with lifetimes and no . Real , however, include higher-order anharmonic terms, primarily cubic and quartic, which introduce nonlinearity and enable phonon interactions. The anharmonic contribution to the is expressed as H_{\text{anh}} = \sum \lambda_3 u^3 + \lambda_4 u^4, where u denotes atomic displacements and \lambda_3, \lambda_4 are the respective coupling coefficients. These anharmonicities give rise to phonon-phonon processes that limit phonon mean free paths and determine thermal properties. The dominant interactions at low orders are three-phonon processes driven by the cubic term, involving either the fusion of two phonons into one or the decay of one phonon into two, provided and crystal momentum are conserved (up to a vector, as detailed in the section on crystal momentum). The corresponding scattering rate \Gamma for such processes is proportional to \sum |V_3|^2 \delta(\omega_1 - \omega_2 - \omega_3) \delta(\mathbf{q}_1 - \mathbf{q}_2 - \mathbf{q}_3), where V_3 represents the three-phonon interaction vertex, \omega the frequencies, and \mathbf{q} the wave vectors. Four-phonon scattering, originating from the quartic term, provides higher-order corrections that become significant at higher temperatures or for long-wavelength modes. Anharmonicity also underlies thermal expansion, as volume changes shift phonon frequencies, quantified by the mode-specific Grüneisen parameter \gamma = -\frac{d \ln \omega}{d \ln V}, which gauges the anharmonicity's impact on vibrational modes. The resulting linewidth from spontaneous decay processes directly relates to the phonon lifetime via \tau = 1/\Gamma, reflecting the inverse of the scattering rate. Perturbative approaches employ to formulate the three-phonon vertex V_3 in terms of phonon creation (a^\dagger) and annihilation (a) operators, enabling calculations of interaction strengths from first-principles potentials.

Thermodynamic Aspects

Phonon Heat Capacity

In solids, the dominant contribution to the at constant volume, C_V, stems from the of phonons, which are quantized collective vibrations of the atomic lattice. Since phonons are bosons with zero , the average occupation number \langle n_{\mathbf{q}} \rangle for a labeled by wavevector \mathbf{q} and branch index (implicitly summed over) follows the Bose-Einstein distribution: \langle n_{\mathbf{q}} \rangle = \frac{1}{e^{\hbar \omega_{\mathbf{q}} / k_B T} - 1}, where \hbar is the reduced Planck's constant, \omega_{\mathbf{q}} is the phonon frequency, k_B is Boltzmann's constant, and T is the absolute temperature. This distribution arises from the canonical ensemble treatment of non-interacting harmonic oscillators representing the lattice modes. The total phonon internal energy U includes the zero-point energy and the thermally excited contribution, expressed as a sum over all normal modes: U = \sum_{\mathbf{q}} \hbar \omega_{\mathbf{q}} \left( \langle n_{\mathbf{q}} \rangle + \frac{1}{2} \right). The heat capacity is obtained by differentiating this energy with respect to temperature at constant volume: C_V = \left( \frac{\partial U}{\partial T} \right)_V. At sufficiently high temperatures, where k_B T \gg \hbar \omega_{\mathbf{q}} for typical phonon frequencies, the Bose-Einstein factor simplifies to \langle n_{\mathbf{q}} \rangle \approx k_B T / \hbar \omega_{\mathbf{q}}, yielding U \approx 3 N k_B T (neglecting the temperature-independent zero-point term). Thus, C_V \approx 3 N k_B, where N is the number of atoms; this is the classical Dulong-Petit limit, reflecting equipartition of energy with k_B T per vibrational degree of freedom (three per atom, each contributing kinetic and potential terms). This high-temperature saturation was first empirically observed for many elemental solids. To address the observed deviations at lower temperatures, early models approximated the phonon spectrum. The Einstein model treats the lattice as $3N independent harmonic oscillators, all with identical frequency \omega_E (chosen to fit experimental data, typically near the peak of the actual spectrum). The occupation simplifies to a single form, leading to the heat capacity: C_V = 3 N k_B \left( \frac{\theta_E}{T} \right)^2 \frac{e^{\theta_E / T}}{\left( e^{\theta_E / T} - 1 \right)^2}, where \theta_E = \hbar \omega_E / k_B is the Einstein temperature. This expression exhibits an exponential decay (C_V \propto e^{-\theta_E / T}) as T \to 0, correctly capturing the freezing out of high-frequency modes but overestimating the suppression at intermediate temperatures, as it neglects the spread in frequencies. The model marked a key application of quantum statistics to solids. The refines this by assuming a of modes with linear acoustic \omega_{\mathbf{q}} = v |\mathbf{q}| (where v is the , averaged over branches) up to a maximum Debye frequency \omega_D chosen such that the total number of modes is $3N. This introduces the Debye temperature \theta_D = \hbar \omega_D / k_B. At low temperatures (T \ll \theta_D), only long-wavelength modes are excited, and the heat capacity scales as C_V \propto T^3, arising from the phase-space volume available to low-frequency phonons. The full Debye expression involves integrals over the approximate density of states but recovers the Dulong-Petit limit as T \gg \theta_D. This continuum approximation successfully explains the T^3 law observed in insulators and metals (after subtracting electronic contributions). For precise computations beyond these approximations, the heat capacity requires evaluating U via integration over the exact phonon density of states g(\omega), which counts the number of modes per frequency interval and is derived from the full dispersion relations \omega(\mathbf{q}): U = \int_0^\infty g(\omega) \, \hbar \omega \left( \frac{1}{e^{\hbar \omega / k_B T} - 1} + \frac{1}{2} \right) d\omega, with C_V = \left( \frac{\partial U}{\partial T} \right)_V. In three dimensions, g(\omega) generally rises as \omega^2 at low frequencies due to the quadratic surface in \mathbf{q}-space, consistent with the Debye form, but deviates at higher \omega depending on the material's lattice dynamics. The phonon density of states g(\omega) is obtained from the dispersion relations covered in the Phonon Characteristics section.

Phonon Tunneling

Phonon tunneling describes the quantum mechanical penetration of vibration quasiparticles, known as phonons, through potential barriers in materials where classical is forbidden. This process is prominent in nanostructures, such as thin films, gaps, or periodic lattices, and at low temperatures where ballistic phonon prevails over anharmonic . In these confined geometries, phonons maintain over distances comparable to barrier widths, typically on the order of nanometers, enabling non-local that contrasts with bulk diffusive flow. The phenomenon arises from the wave-like of phonons, analogous to or tunneling, but governed by the crystal's acoustic or optical dispersion. For coherent phonons incident on a potential barrier, the transmission probability is often estimated using the semiclassical , expressed as T \approx \exp\left( -2 \int_{x_1}^{x_2} \sqrt{ \frac{2m (V(x) - E ) }{\hbar^2} } \, dx \right), where m is the effective , V(x) the barrier potential, E the phonon energy, \hbar the reduced , and the integral spans the turning points x_1 to x_2 of the forbidden region. This formula captures the exponential suppression of tunneling for thicker or higher barriers, with applications demonstrated in sonic analogs of black holes where phonons tunnel across horizons without backreaction effects. In nanostructures, the approximation highlights how low-energy acoustic phonons tunnel more readily than optical modes due to their smaller effective and . In double quantum wells and superlattices, phonon-assisted tunneling facilitates energy exchange between layers, where electrons or holes traverse barriers while emitting or absorbing phonons to conserve momentum and energy. In GaAs-AlGaAs double quantum wells, longitudinal optical phonon scattering dominates interwell transition rates, with tunneling probabilities peaking when the phonon energy matches the well separation. Similarly, in weakly coupled superlattices under , acoustic phonons assist sequential tunneling, leading to angle-dependent conductance peaks. These processes are crucial for understanding inelastic transport in heterostructures at cryogenic temperatures. Within phonon bandgaps of periodic structures like phononic , propagating modes are forbidden at specific frequencies, resulting in evanescent waves that decay exponentially and enable tunneling across finite slabs. These evanescent Bloch modes, characterized by complex wavevectors, dominate at interfaces, with decay lengths scaling inversely with bandgap width. In one-dimensional phononic , tunneling via evanescent fields transmits across gaps where direct is blocked, as seen in layered nanostructures with THz bandgaps. Experimental verification of phonon tunneling has been achieved through resonant tunneling spectroscopy in semiconductor heterostructures, where voltage-biased devices reveal phonon spectra via conductance resonances. In graphene-boron nitride heterostructures, inelastic tunneling features correspond to phonon energies from dispersion, with peaks at approximately 30-60 meV matching optical modes. These observations, conducted at low temperatures to suppress broadening, confirm tunneling-mediated phonon emission during transport. Additionally, vacuum phonon tunneling across nanometer gaps has been theoretically tied to experimental anomalies between metal surfaces, driven by evanescent coupling to interfacial phonons. A key distinction from electron tunneling lies in phonons' bosonic statistics, which permit multi-phonon bunching during coherent transport, enhancing probabilities through stimulated occupation unlike fermionic exclusion. This allows collective tunneling of multiple phonons in , as observed in multi-phonon inelastic processes in layered systems. Such bunching underpins coherent in phonon devices. Applications of phonon tunneling include phonon lasers, or SASERs (sound by of radiation), where tunneling in superlattices generates coherent THz phonons via -phonon in vertical cavities. Devices operating at 325 GHz demonstrate self-sustained oscillations through resonant phonon . Post-2010 advances in exploit tunneling asymmetry in confined structures, such as asymmetric nanowires or ribbons, where lateral phonon confinement creates diode-like flow with rectification ratios up to 1.4, arising from mismatched evanescent mode transmission under reversed biases. These developments enable nanoscale thermal diodes for .

Advanced Properties and Applications

Formal Operator Description

In the many-body formalism for phonons, the provides a key tool for describing the propagation and interactions of vibrations as bosonic quasiparticles. The phonon is defined in the time-ordered form as G(\mathbf{q}, \omega) = -i \langle T \psi(\mathbf{q}, t) \psi^\dagger(\mathbf{q}, 0) \rangle, where \psi(\mathbf{q}, t) is the phonon annihilation in the , and T denotes time ordering. This function encodes the correlation between phonon creation and annihilation, with its relating to the spectral properties of the . For response theory, the retarded G^R(\mathbf{q}, \omega) = -i \theta(t) \langle [\psi(\mathbf{q}, t), \psi^\dagger(\mathbf{q}, 0)] \rangle is particularly useful, as it determines linear response to external perturbations, such as those in or properties. Interactions beyond the harmonic approximation introduce self-energy corrections, captured by Dyson's equation: G(\mathbf{q}, \omega) = G_0(\mathbf{q}, \omega) + G_0(\mathbf{q}, \omega) \Sigma(\mathbf{q}, \omega) G(\mathbf{q}, \omega), where G_0(\mathbf{q}, \omega) = \frac{2\omega_{\mathbf{q}}}{\omega^2 - \omega_{\mathbf{q}}^2 + i\eta} is the bare phonon with phonon \omega_{\mathbf{q}}, and \Sigma(\mathbf{q}, \omega) is the arising from anharmonic phonon-phonon interactions. The \Sigma is computed perturbatively from Feynman diagrams representing three- and four-phonon processes, which account for thermal broadening and in real materials. Solving Dyson's equation self-consistently yields the interacting phonon , essential for understanding lifetimes and dispersion shifts due to . For electron-phonon interactions, the vertex function quantifies the coupling strength through matrix elements g_{\mathbf{k} \mathbf{q}} = \langle \mathbf{k} + \mathbf{q} | \frac{\delta V}{\delta u_{\mathbf{q}}} | \mathbf{k} \rangle \sqrt{\frac{\hbar}{2 M \omega_{\mathbf{q}}}}, where V is the ionic potential, u_{\mathbf{q}} the normal-mode displacement, M the ionic mass, and the states |\mathbf{k}\rangle are Bloch electrons. These elements enter diagrammatic expansions for both electronic and phononic self-energies, enabling the computation of scattering rates and transport coefficients. In vertex corrections, higher-order terms modify g_{\mathbf{k} \mathbf{q}} to include screening effects from the dielectric response. Coupled modes, such as magnon-phonon hybrids in magnetic crystals, require a to diagonalize the quadratic mixing bosonic operators. The transformation \begin{pmatrix} \alpha \\ \alpha^\dagger \end{pmatrix} = \begin{pmatrix} u & v \\ v^* & u^* \end{pmatrix} \begin{pmatrix} b \\ b^\dagger \end{pmatrix}, with |u|^2 - |v|^2 = 1, yields quasiparticle operators \alpha that remove the anomalous terms, revealing hybridized relations. This approach applies generally to any bilinear boson-boson coupling, stabilizing the spectrum against instabilities. Real-time dynamics under time-dependent perturbations, such as pulses, are handled in the , where operators evolve via i\hbar \frac{d}{dt} \psi(t) = [\psi(t), H(t)], with H(t) = H_0 + V(t) incorporating the perturbation. The time evolution of the then follows from solving the corresponding , facilitating simulations of nonequilibrium phonon populations. Ab initio computation of phonon operators relies on density functional perturbation theory (DFPT), which linearizes the Kohn-Sham equations to obtain the response of the to displacements, yielding dynamical matrices and thus phonon frequencies and eigenvectors. DFPT provides a nonempirical framework for the bare propagator G_0, serving as input for many-body like self-energies in subsequent diagrammatic treatments.

Phonons in Superconductivity

In conventional superconductivity, phonons play a central role by mediating an attractive interaction between electrons, enabling the formation of Cooper pairs. The Bardeen-Cooper-Schrieffer (BCS) theory, developed in 1957, posits that electrons near the Fermi surface experience an effective attractive potential V_{\text{eff}} = -g^2 / \omega_{\text{ph}} due to the exchange of virtual phonons, where g is the electron-phonon coupling strength and \omega_{\text{ph}} is the phonon frequency. This attraction occurs for electrons separated by energies less than the Debye energy \hbar \omega_D, leading to a superconducting energy gap \Delta = 1.76 k_B T_c at zero temperature in the weak-coupling limit, where T_c is the critical temperature. Eliashberg extended the BCS framework in 1960 to account for strong electron-phonon and retardation effects, incorporating the full frequency dependence of the interaction. In this theory, the retarded phonon propagator is given by D(\omega) = 2 \omega_q / (\omega^2 - \omega_q^2 + i \Gamma \omega), where \omega_q is the phonon frequency and \Gamma is a parameter, under the Migdal that neglects corrections. The electron-phonon \lambda quantifies the strength of this interaction and is defined as \lambda = 2 \int_0^{\infty} \alpha^2 F(\omega) / \omega \, d\omega, where \alpha^2 F(\omega) is the Eliashberg spectral function capturing the phonon weighted by the . An approximate formula for T_c in strong-coupling superconductors, proposed by McMillan in 1968, is T_c = \frac{\theta_D}{1.45} \exp\left[ -\frac{1.04 (1 + \lambda)}{\lambda - \mu^*} \right], where \theta_D is the Debye temperature and \mu^* is the Coulomb pseudopotential accounting for electron-electron repulsion. This expression highlights how stronger phonon-mediated attraction (higher \lambda) elevates T_c. The phonon mediation was experimentally confirmed by the isotope effect, first observed in 1950, where T_c \propto M^{-1/2} and M is the ionic mass, directly linking lattice vibrations to the pairing mechanism. In materials like Nb₃Sn, a conventional superconductor with T_c \approx 18 K, the electron-phonon coupling is strong with \lambda \sim 1.5, consistent with Eliashberg theory predictions for A15 compounds. However, this phonon-based framework fails to explain high-T_c cuprates, where pairing exhibits d-wave symmetry and is mediated by magnetic interactions rather than phonons, leading to T_c values far exceeding BCS limits.

Emerging Research Areas

Recent advances in phonon physics have explored topological properties, where phonon bands exhibit nontrivial Berry curvature, leading to protected modes in phononic crystals that enable dissipationless phonon analogous to topological insulators. These features were first demonstrated in through phononic crystals supporting one-way waves for both longitudinal and transverse polarizations, ensuring robustness against backscattering and defects. Further progress includes the identification of topological invariants in phonon spectra, with a comprehensive catalog of over 5,000 materials revealing widespread occurrence of topological phonons that could enhance thermal management and waveguiding in nanostructures. In two-dimensional materials like graphene, flexural phonons exhibit a distinctive quadratic dispersion relation, \omega \propto q^2, arising from the membrane's out-of-plane vibrations, which dominates low-frequency thermal transport. This dispersion results in an anomalously high density of states at low energies, contributing to divergent thermal conductivity in the absence of scattering, as confirmed by 2010s experiments on suspended graphene sheets showing ballistic phonon propagation over micrometer scales. Such anomalies have been observed in thermal conductance measurements, where flexural modes account for up to 50% of heat flow at room temperature, highlighting their role in engineering ultrahigh thermal conductivity in van der Waals heterostructures. Phonon hydrodynamics has emerged as a key paradigm for collective phonon flow in clean, nanoscale insulators, where mean free paths exceed sample dimensions, enabling —a wave-like propagation of temperature oscillations. This regime was directly observed in (SrTiO₃) in 2023 using picosecond laser heating and time-resolved thermoreflectance, revealing second sound velocities of approximately 100 m/s and propagation lengths up to several micrometers at cryogenic temperatures below 20 K. These findings underscore hydrodynamic effects in bulk dielectrics, with implications for nanoscale thermal rectification and reduced heat dissipation in quantum devices. Quantum phononics leverages coherent phonon manipulation in optical cavities to create hybrid , where phonons serve as information carriers with long times exceeding 1 ms. Advances in the include optomechanical coupling in superconducting circuits, enabling deterministic control of single phonons for quantum state transfer between distant qubits. Phonon qubits, encoded in acoustic resonators, have demonstrated entanglement with photons, paving the way for scalable quantum networks and operating at frequencies. Phonon-polaritons, hybrid modes arising from strong coupling between infrared photons and optical phonons in polar materials like hexagonal boron nitride, facilitate subwavelength confinement of light beyond the diffraction limit, with effective wavelengths as small as \lambda/100. These quasiparticles enable mid-infrared applications, such as ultra-compact waveguides and sensors, where polaritons in van der Waals crystals support resonant enhancements up to 10^4 in local fields for molecular detection. Recent demonstrations include tunable phonon-polariton cavities for dynamic beam steering in the 10-15 μm range, advancing nanophotonics for spectroscopy and thermal emitters. Machine learning has revolutionized phonon engineering through inverse design of phononic metamaterials, optimizing band structures for targeted wave manipulation without exhaustive simulations. Since 2023, generative neural networks have enabled on-demand , predicting complete phonon relations with over 95% accuracy and generating structures with bandgaps up to 50% of the . These AI-driven approaches, applied to metamaterials, facilitate custom designs for and acoustic , reducing design iterations from weeks to hours. As of 2025, research on chiral phonons has advanced, revealing that phonon chirality—circularly polarized vibrations—can be harnessed to control material properties such as heat flow, sound propagation, light-matter interactions, and in two-dimensional and topological materials. These developments build on topological phonon concepts, enabling novel encoding of and selective phonon manipulation for energy-efficient devices.

References

  1. [1]
    New Advances in Phonons: From Band Topology to Quasiparticle ...
    Phonons, the quantized vibrational modes of a crystal lattice, are ubiquitous quasiparticles in solid-state systems. They play a central role in a wide ...
  2. [2]
    [PDF] 4. Phonons - DAMTP
    This particle is a quantum of the lattice vibration. It is called the phonon. Note that the coupling between the atoms has lead to a quantitative change in the.
  3. [3]
    [PDF] Phonon Quasiparticle Studies of Anharmonic Properties of Solids
    A collective excitation of atoms in a crystalline material is defined as a phonon, which is a type of boson and obeys Bose-Einstein statistics. Phonons ,or ...
  4. [4]
    [PDF] Review Article Prediction of Spectral Phonon Mean Free Path and ...
    We give a review of the theoretical approaches for predicting spectral phonon mean free path and thermal conductivity of solids.
  5. [5]
    Electron-phonon interactions from first principles | Rev. Mod. Phys.
    Feb 16, 2017 · This article reviews the theory of electron-phonon interactions in solids from the point of view of ab initio calculations.
  6. [6]
    Crystal Physics - Toback Research Group - Texas A&M University
    Jan 19, 2024 · Phonons are a type of energy-carrying quasiparticle; in particular, they consist of quantized lattice oscillations that occur in several ways.
  7. [7]
    [PDF] Lecture 11: Phonons and Photons
    Our rst example is the theory of phonons, or vibrations of a lattice of atoms in a solid. This will let us compute the heat capacity of a solid and other useful ...
  8. [8]
    Phonon - an overview | ScienceDirect Topics
    Phonons describe the excitations and eigenstates of atomic motions in solids. Since atomic motions are easily visualized, phonons often serve as an ...
  9. [9]
    [PDF] Review o Phonon dispersion relations o Quantum nature of waves ...
    Phonons are bosonic particles so each mode (defined by its angular frequency 𝜔) can be occupied by more than one particle. The particle nature of lattice ...
  10. [10]
    [PDF] Phys 446: Solid State Physics / Optical Properties Lattice vibrations
    Similarly the energy levels of lattice vibrations are quantized. •The quantum of vibration is called a phonon. (in analogy with the photon - the quantum of the ...
  11. [11]
    [PDF] The dynamical theory of sound - hlevkin
    modes was subsequently given by Lord Rayleigh. The theory of vibrations normal to the plane is more intricate, since torsion is involved as well as flexure ...
  12. [12]
    Einstein Model of a Solid
    A useful step on the way to understanding the specific heats of solids was Einstein's proposal in 1907 that a solid could be considered to be a large number ...
  13. [13]
    [PDF] The Einstein specific heat model for finite systems - arXiv
    Essentially, the model proposed by Einstein in 1907 [1] to describe the thermal properties of a simple crystalline solid, treating the solid as an array of ...
  14. [14]
    Debye Model For Specific Heat - Engineering LibreTexts
    Sep 7, 2021 · Debye frequency and Debye Temperature ; Aluminum, 428K, Iron ; Cadmium, 209K, Lead ; Chromium, 630K, Manganese ; Copper, 343.5K · Nickel ; Gold, 165K ...
  15. [15]
    Periodic Boundary Condition - an overview | ScienceDirect Topics
    To eliminate such problems, a simple yet effective method proposed by Born and von Karman (1912) and generally referred to as the periodic boundary conditions ...
  16. [16]
    Brillouin Zones - Engineering LibreTexts
    Sep 7, 2021 · The concept of a Brillouin zone was first developed by Léon Brillouin (1889-1969), a French physicist. During his work on the propagation of ...<|separator|>
  17. [17]
    The Philosophy of Nature of the Natural Realism. The Operator ...
    ... Pascual Jordan and Eugene Wigner in 1928 [103] (see Figure 5). These two notions, using the further contribution of the Russian mathematician Vladimir Fock ...<|control11|><|separator|>
  18. [18]
    All-coupling theory for the Fr\"ohlich polaron | Phys. Rev. B
    Apr 18, 2016 · Such electron-phonon interactions were first described by Fröhlich who introduced a model where the electron emits and reabsorbs phonons [1]
  19. [19]
    [PDF] Chapter 4: Crystal Lattice Dynamics
    Jan 30, 2017 · Crystal lattice dynamics involves lattice vibrations, which contribute to thermal conductivity, scattering, and electron-phonon interactions. ...
  20. [20]
    [PDF] The vibrations of atoms inside crystals - lattice dynamics
    Appendices provide supplementary information and derivations for the Ewald method, statistical mechanics of lattice vibra- tions, Landau theory, scattering ...<|control11|><|separator|>
  21. [21]
    Lattice dynamics of Ge and Si using the Born-von Karman model
    Mar 15, 1979 · The Born-von Karman model of lattice dynamics of diamond structure has been extended to include up to 12th-neighbor interactions.
  22. [22]
    [PDF] Lecture 10 - 6.730 Physics for Solid State Applications
    Lattice Waves in 3-D Crystals. Displacements of basis atoms along three ... Elements of the Dynamical Matrix. 7. 3. 5. 2. 1. 0. 6. 4. 8 α1 α2. Example: 2-D ...
  23. [23]
    [PDF] From Lattice Waves to Phonons Classical Simple Harmonic Oscillator
    In this lecture you will learn: • Simple harmonic oscillator in quantum mechanics. • Classical and quantum descriptions of lattice wave modes. • Phonons – what ...
  24. [24]
  25. [25]
    [PDF] Part VII: Lattice vibrations – phonons 1 The simple harmonic oscillator
    Before going on, let me point out that it would have been slightly easier to define u = x−xeq as the displacement out of equilibrium, and use this as the ...
  26. [26]
    Dynamical Matrices and Interatomic-Force Constants from Wave ...
    Jul 8, 2004 · For a range of materials, phonon dispersion exhibits the close agreement with experiment that is now characteristic of first-principles schemes.
  27. [27]
    Brillouin Zone - an overview | ScienceDirect Topics
    The first Brillouin zone is the polyhedron associated with those planes that surround the origin; it is thus a Wigner–Seitz cell in the reciprocal lattice.
  28. [28]
    Acoustic Phonon - an overview | ScienceDirect Topics
    A phonon is a quantum of the lattice vibration, the collective motion of atoms constituting a crystal. There are two types of phonons: optical and acoustic. The ...
  29. [29]
    [PDF] 1 Phonon I: lattice waves - bingweb
    Feb 18, 2019 · H.P.Myers, Introductory Solid State Physics (Taylor & Francis,. 1990). A discrete vibrational state or mode has a well defined wave vector k and ...
  30. [30]
    [PDF] Solid State Theory I - TKM (KIT)
    • Crystal momentum. The vector ¯hk is not the momentum and the Bloch states are not eigenstates of the momentum operator. Indeed pψ n,k. = −i¯h∇ ψ n,k.
  31. [31]
    [PDF] Feng_book-chapter_four_phonon_2020.pdf
    The phonon momentum is ℏq. The processes with momentum conserved are normal processes. The others with momentum non-conserved are Umklapp processes, in which.
  32. [32]
    [PDF] Spectral Phonon Transport Properties of Silicon Based on Molecular ...
    which are subject to energy and momentum conservation constraints, can be classified as either normal processes, which conserve quasi-momentum, or umklapp ...
  33. [33]
    [PDF] LECTURE # 37
    Apr 20, 2009 · As the temperature rises, umklapp processes become less rare, and the conductivi- ty reaches a maximum when the mean free path due to phonon ...
  34. [34]
    [PDF] Lecture 1 Scattering and Diffraction - ORNL Neutron Sciences
    • Q is the momentum transfer. • r. 0 is the Thomson x-ray scattering length ... These frequencies map out the so-called phonon dispersion curves. phonon ...
  35. [35]
    [PDF] arXiv:2310.13592v1 [cond-mat.mes-hall] 20 Oct 2023
    Oct 20, 2023 · Throughout this paper, we will explore how the difference between crystal momentum and group velocity leads to novel subtleties and phenomena in ...
  36. [36]
    Phonon anharmonicity: a pertinent review of recent progress ... - arXiv
    Oct 21, 2021 · In this work, we review the recent development of the study on phonon anharmonic effect and summarize its origination, influence and mechanism, research ...
  37. [37]
    [2210.15537] Anharmonic phonon behavior via irreducible derivatives
    Oct 27, 2022 · Cubic phonon interactions are now regularly computed from first principles, and the quartic interactions have begun to receive more attention.
  38. [38]
    Three-phonon processes in solids | Phys. Rev. B
    Jan 15, 1979 · The problem of three-phonon processes in solids is investigated on the assumption that all the phonons contribute equally to the conduction ...Missing: seminal | Show results with:seminal
  39. [39]
    Grüneisen Parameters: origin, identity and quantum refrigeration
    Sep 12, 2019 · Abstract:In solid state physics, the Grüneisen parameter (GP), originally introduced in the study of the effect of changing the volume of a ...
  40. [40]
    Phonon Lifetime Investigation of Anharmonicity and Thermal ...
    Apr 8, 2013 · Inelastic neutron scattering measurements of individual phonon lifetimes and dispersion at 295 and 1200 K have been used to probe anharmonicity and thermal ...
  41. [41]
    C - Bose-Einstein distribution and the thermodynamic relations for ...
    C - Bose-Einstein distribution and the thermodynamic relations for phonons. Published online by Cambridge University Press: 17 February 2010. Martin T. Dove.
  42. [42]
    Vacuum Phonon Tunneling | Phys. Rev. Lett.
    Oct 11, 2010 · We show that phonon tunneling is driven by interfacial electric field and thermally vibrating image charges, and its rate is enhanced by surface electron- ...
  43. [43]
    [1108.2916] Phonon tunnels across a sonic horizon - arXiv
    Aug 15, 2011 · We consider phonon tunneling in sonic black hole by WKB approximation method without the backreaction, in which the relativistic momentum-energy ...
  44. [44]
    Phonon-assisted tunnelling in double quantum well structures
    We calculate the tunnelling rates for interwell transitions due to electron–longitudinal-optical phonon (LO-phonon) scattering in GaAs–AlGaAs DQW ...
  45. [45]
    Phonon-assisted tunneling in a superlattice in an applied magnetic ...
    Jul 31, 2009 · We have studied acoustic phonon-assisted tunneling in a weakly coupled GaAs/AlAs superlattice (SL) in a magnetic field.Missing: double | Show results with:double
  46. [46]
    Evanescent Bloch waves and the complex band structure of ...
    Sep 18, 2009 · Evanescent Bloch waves are involved in the diffraction of acoustic phonons at the interfaces of finite phononic crystal structures.
  47. [47]
    Phonon-Assisted Resonant Tunneling of Electrons in Graphene
    May 5, 2016 · The study observed resonant features in graphene-boron nitride transistors, caused by electron tunneling assisted by phonon emission, and ...
  48. [48]
    Two-photon blockade and photon-induced tunneling generated by ...
    Nov 27, 2019 · Multi-phonon blockade, which is a mechanical analog of multi-PB, was studied in Ref. [35] . Multi-PB in dissipation-free systems enables ...
  49. [49]
    Dynamics of a vertical cavity quantum cascade phonon laser structure
    Jul 25, 2013 · Here we fabricate a vertical cavity structure designed to operate as a saser oscillator device at a frequency of 325 GHz.
  50. [50]
    Many-Particle Physics - Book - SpringerLink
    The method of Green's functions has been used by many theorists to derive equations which, when solved, provide an accurate numerical description of many ...
  51. [51]
    Phonons and related crystal properties from density-functional ...
    Jul 6, 2001 · This article reviews the current status of lattice-dynamical calculations in crystals, using density-functional perturbation theory.
  52. [52]
    Many-body Green's function theory for electron-phonon interactions
    Jul 16, 2015 · We present a Kadanoff-Baym formalism to study time-dependent phenomena for systems of interacting electrons and phonons in the framework of many-body ...
  53. [53]
    Coupled Bogoliubov equations for electrons and phonons
    The Bogoliubov transform for phonons is non-Hermitian in the general case, and the corresponding time evolution is nonunitary. Several sufficient conditions ...Abstract · Article Text · INTRODUCTION · ALGEBRAIC FORM OF THE...
  54. [54]
    Theory of Superconductivity | Phys. Rev.
    A theory of superconductivity is presented, based on the fact that the interaction between electrons resulting from virtual exchange of phonons is attractive.
  55. [55]
    Transition Temperature of Strong-Coupled Superconductors
    The superconducting transition temperature is calculated using electron-phonon and electron-electron coupling constants within strong-coupling theory.Missing: Tc original
  56. [56]
    Phys. Rev. B 22, 1214 (1980) - Tunneling and the electron-phonon ...
    Aug 1, 1980 · The results satisfactorily describe the superconducting properties of Nb3Sn within the conventional framework of strong electron-phonon coupling ...Missing: lambda | Show results with:lambda
  57. [57]
    Berry phase and topological effects of phonons - Oxford Academic
    Aug 2, 2017 · In this Perspective, we will briefly introduce this emerging field and discuss the use of novel quantum degrees of freedom like the Berry phase and topology.Missing: post- | Show results with:post-
  58. [58]
    Topological Phononic Crystals with One-Way Elastic Edge Waves
    Sep 4, 2015 · We report a new type of phononic crystals with topologically nontrivial band gaps for both longitudinal and transverse polarizations, resulting in protected ...Missing: phonons | Show results with:phonons
  59. [59]
    Catalog of topological phonon materials - Science
    May 10, 2024 · Phonons play a crucial role in many properties of solid-state systems, and it is expected that topological phonons may lead to rich and ...
  60. [60]
    Flexural phonons and thermal transport in graphene | Phys. Rev. B
    Sep 15, 2010 · The ZA phonons in graphene have a quadratic dispersion over a wide range of the 2D Brillouin zone: ω ZA ( q ) = α ZA q 2 , where α ZA is a ...Abstract · Article Text · DENSITY OF STATES AND... · RESULTS AND DISCUSSION
  61. [61]
    Colloquium: Phononic thermal properties of two-dimensional materials
    Nov 13, 2018 · According to Eq. (1) , the anomalous thermal conductivity comes from the long-wavelength phonons or acoustic phonons near the Γ point ...
  62. [62]
    [PDF] arXiv:2212.03745v2 [physics.app-ph] 14 Jun 2023
    Jun 14, 2023 · This is the first 'direct' demonstration of second-sound propagation in. SrTiO3. By systematically changing the wavelength and fre- quency of ...
  63. [63]
    Ultra-confined mid-infrared resonant phonon polaritons in van der ...
    Jun 15, 2018 · We show that strongly subwavelength hexagonal boron nitride planar nanostructures can exhibit ultra-confined resonances and local field enhancement.
  64. [64]
    Surface phonon polaritons for infrared optoelectronics - AIP Publishing
    Jan 18, 2022 · We will assess the potential of surface phonon polariton-based nanophotonics for infrared (3–100 μ m) light sources, detectors, and modulators.
  65. [65]
    Deep learning for the design of phononic crystals and elastic ...
    Here, the recent progresses on deep learning for forward prediction, parameter design, and topology design of PnCs and EMs are reviewed.
  66. [66]
    Machine learning assisted intelligent design of meta structures
    In this review, we focus on the latest progress of ML in acoustic, elastic, and mechanical meta-structures from the aspects of band structures, wave ...
  67. [67]
    Inverse design of phononic meta-structured materials - ScienceDirect
    This review summaries the evolution of the inverse design of PMSMs from the fundamental methodology, diverse customized functionalities to representative ...