Fact-checked by Grok 2 weeks ago

Quasiparticle

In condensed matter physics, a quasiparticle is an emergent excitation in an interacting many-body system that behaves like a particle, allowing complex collective phenomena to be approximated as interactions among weakly coupled, well-defined entities. These excitations arise from the collective behavior of numerous fundamental particles, such as electrons or atoms, and are characterized by properties like effective mass, energy dispersion, and finite lifetime due to interactions. Quasiparticles simplify the description of intricate quantum systems by treating disturbances—such as lattice vibrations or electron correlations—as propagating entities with particle-like trajectories, often near the Fermi surface in metals where their decay rates are minimal at low temperatures. Prominent examples include phonons, which represent quantized lattice vibrations in solids and mediate electron-phonon interactions essential for thermal conductivity and ; polarons, electrons dressed by surrounding lattice distortions; and excitons, bound electron-hole pairs in semiconductors that influence . The concept originated in the 1930s with early work on collective excitations and was formalized in Lev Landau's in the 1950s, providing a framework to explain metallic properties beyond the independent-particle model by incorporating effects like mass enhancement. Quasiparticles extend to exotic cases in topological materials, such as Weyl fermions or skyrmions, enabling predictions of novel states of matter like quantum spin liquids and fractional quantum Hall effects. This approximation has proven vital for advancing technologies, including semiconductors, superconductors, and , by bridging microscopic interactions to macroscopic observables.

Fundamentals

Definition and Core Concept

In quantum many-body systems, numerous particles interact strongly, giving rise to collective correlations that cannot be adequately captured by non-interacting single-particle models. These systems, such as or atoms in quantum liquids, exhibit emergent behaviors where the collective response to perturbations simplifies the description of complex dynamics. Quasiparticles serve as effective entities in this context, representing excitations that mimic particles while encapsulating the influence of surrounding interactions, such as an in a solid coupled to vibrations. The core analogy for quasiparticles portrays them as "dressed" particles, wherein a bare particle acquires a surrounding cloud of excitations from interactions, thereby modifying its and response to external fields. This enables a quasiparticle to propagate coherently through the medium as if it were independent, despite the underlying many-body complexity. Introduced by Landau in the framework of , quasiparticles arise through adiabatic continuity, smoothly connecting the excitations of interacting and non-interacting systems without phase transitions. Unlike bare particles, quasiparticles exhibit renormalized properties including an effective that reflects interaction-induced , a potentially altered charge, and a finite lifetime determined by channels. The effective , for example, can differ significantly from the bare value due to screening and effects, while the lifetime ensures quasiparticles remain well-defined only over short timescales relative to their energy scales. These attributes underpin the utility of quasiparticles in approximating many-body , providing a bridge to intuitive physical interpretations of condensed phenomena.

Relation to Many-Body Quantum Mechanics

In quantum many-body systems, the dynamics are governed by a Hamiltonian that separates into a non-interacting part H_0 and an interaction term V, expressed generally as H = H_0 + V. For a system of fermions like electrons, H_0 = \sum_i \frac{\mathbf{p}_i^2}{2m} describes the kinetic energy of individual particles, while V encompasses pairwise or higher-order interactions, such as the Coulomb repulsion between electrons given by V = \frac{1}{2} \sum_{i \neq j} \frac{e^2}{|\mathbf{r}_i - \mathbf{r}_j|}. This form arises naturally in the second-quantized representation of condensed matter Hamiltonians, where interactions couple the degrees of freedom of multiple particles. Interactions in many-body systems, including electron-electron Coulomb forces, electron-phonon couplings, and spin-exchange terms, introduce strong correlations that entangle the wavefunction across particles, rendering exact solutions to the infeasible for large numbers of constituents beyond a few particles. These correlations manifest as collective behaviors that cannot be captured by treating particles independently, leading to phenomena like screening and effects. As a result, perturbative methods or mean-field approximations become essential to approximate the and excited states, though they often break down in strongly correlated regimes. Quasiparticles emerge as an effective description of low-energy excitations in these interacting systems, behaving like renormalized, non-interacting particles that propagate through the medium while accounting for the surrounding correlations. Near the in fermionic systems or close to the in bosonic ones, these excitations resemble the original particles but with modified relations and lifetimes due to with other . This concept allows the complex to be mapped onto a simpler gas of quasiparticles, preserving key physical properties like conservation laws. A cornerstone of this framework is Landau's , which posits that in three-dimensional interacting Fermi systems at low temperatures, quasiparticles represent the dominant low-energy excitations and obey Fermi-Dirac statistics, with a finite at the renormalized . These quasiparticles carry the same spin, charge, and momentum as the bare electrons but interact weakly among themselves, enabling thermodynamic and transport properties to be computed perturbatively around the non-interacting limit. This theory successfully explains the metallic behavior of electrons in solids despite strong interactions.

Theoretical Framework

Green's Function Formalism

In the Green's function formalism for many-body , the single-particle serves as the central object to describe the of quasiparticles, which emerge as effective excitations in interacting media. The retarded G^R(\mathbf{k}, \omega) is defined as the response to adding a particle with \mathbf{k} and energy \omega, while the advanced G^A(\mathbf{k}, \omega) corresponds to removing a particle; these functions encode the for such processes in the interacting system. Specifically, G^R(\mathbf{k}, \omega) = -i \int_0^\infty dt \, e^{i\omega t} \langle \psi | T[ c_{\mathbf{k}}(t) c_{\mathbf{k}}^\dagger(0) ] | \psi \rangle \theta(t), where T is the time-ordering operator, c_{\mathbf{k}}^\dagger creates a particle, and |\psi\rangle is the many-body , with the advanced counterpart obtained by G^A(\omega) = [G^R(\omega^*)]^*. This formulation captures how interactions modify the free-particle , linking directly to quasiparticle through the properties of G. The interacting Green's function G is related to the non-interacting one G_0 via the Dyson equation, G = G_0 + G_0 \Sigma G, where \Sigma(\mathbf{k}, \omega) is the self-energy that encapsulates all effects beyond the mean field. Inverting this yields G^{-1} = G_0^{-1} - \Sigma, highlighting how \Sigma dresses the bare to produce the quasiparticle spectrum. The self-energy \Sigma is generally complex, with its real part \mathrm{Re} \Sigma(\mathbf{k}, \omega) shifting the quasiparticle levels and renormalizing parameters like the effective , given by m^* = m \left(1 - \frac{\partial \mathrm{Re} \Sigma}{\partial \omega}\right)^{-1} evaluated at the quasiparticle energy, where m is the bare . The imaginary part \mathrm{Im} \Sigma(\mathbf{k}, \omega), on the other hand, accounts for decay processes, determining the quasiparticle lifetime \tau = -1 / (2 \mathrm{Im} \Sigma(\mathbf{k}, \omega)). Quasiparticles manifest as poles of the near the real axis, satisfying the pole condition \omega - \epsilon_{\mathbf{k}} - \mathrm{Re} \Sigma(\mathbf{k}, \omega) = 0, where \epsilon_{\mathbf{k}} is the bare . The strength of these poles, or the residue Z_{\mathbf{k}} = \left(1 - \frac{\partial \mathrm{Re} \Sigma(\mathbf{k}, \omega)}{\partial \omega}\right)^{-1} at the solution \omega = E_{\mathbf{k}}, quantifies the quasiparticle weight, with Z < 1 reflecting interaction-induced spectral broadening. This structure ensures that well-defined quasiparticles correspond to sharp peaks in the spectral function A(\mathbf{k}, \omega) = -\frac{1}{\pi} \mathrm{Im} G^R(\mathbf{k}, \omega), provided |\mathrm{Im} \Sigma| \ll | \mathrm{Re} \Sigma - \epsilon_{\mathbf{k}} |.

Quasiparticle Approximation and Dispersion

The quasiparticle approximation simplifies the complex dynamics of interacting many-body systems by treating excitations as weakly dressed particles that propagate with renormalized parameters, such as effective mass and lifetime. This approach is particularly applicable in weakly interacting regimes, like , where interactions lead to small corrections to the non-interacting description. The self-energy \Sigma(k, \omega), which encodes interaction effects within the Green's function formalism, plays a central role: its real part shifts the energy levels, while the imaginary part introduces damping. The validity of the quasiparticle approximation requires that the spectral function exhibits well-defined, Lorentzian-like peaks, which occurs when the imaginary part of the self-energy is much smaller than the energy distance to the quasiparticle pole: |\operatorname{Im} \Sigma(k, \omega)| \ll |\omega - \varepsilon_k - \operatorname{Re} \Sigma(k, \omega)|, evaluated near the solution \omega \approx E_k. This condition ensures minimal broadening and allows the Green's function to be approximated by a pole structure, G(k, \omega) \approx Z_k / (\omega - E_k + i \Gamma_k / 2), where Z_k is the quasiparticle residue or weight, given by Z_k = [1 - \partial \operatorname{Re} \Sigma(k, \omega) / \partial \omega]^{-1} at \omega = E_k. In , Z_k < 1 but finite for low-energy excitations near the Fermi surface, reflecting partial screening of interactions. Under this approximation, the effective quasiparticle dispersion relation is obtained by solving the Dyson equation on-shell: E_k \approx \varepsilon_k + \operatorname{Re} \Sigma(k, E_k), where \varepsilon_k is the bare . This leads to renormalized band structures, with the group velocity modified as v^*_k = \partial E_k / \partial k, often resulting in an enhanced effective mass m^* = \hbar k / v^*_k > m due to s. For example, in electron-phonon systems, Migdal's theorem justifies evaluating \Sigma at the bare energy for small coupling, preserving the form of the while incorporating polaronic shifts. The lifetime \tau_k = -1 / (2 \operatorname{Im} \Sigma(k, E_k)) determines the damping, with the \Gamma_k = -2 \operatorname{Im} \Sigma(k, E_k). In , \operatorname{Im} \Sigma is computed via , expressing the decay rate as a golden-rule transition probability to multi-particle states, \Gamma_k \propto \sum_f | \langle f | V | i \rangle |^2 \delta(E_i - E_f), where V is the . The approximation breaks down in strongly correlated systems where interactions are not perturbative, leading to Z_k \to 0 and incoherent spectral features without well-defined quasiparticles. A prominent example is the transition in the , where at half-filling and large on-site repulsion U, the quasiparticle weight vanishes, suppressing charge fluctuations and opening a despite nominal metallic filling. captures this by showing how the diverges at low frequencies, destroying Fermi liquid behavior.

Classifications and Distinctions

Quasiparticles versus Collective Excitations

Collective excitations represent coherent modes in many-body systems where numerous particles oscillate in synchrony, such as plasmons involving collective electron density fluctuations or spin waves arising from aligned spin precessions in magnetic materials. These excitations emerge from the cooperative dynamics of the system, often described as emergent phenomena with well-defined energy and momentum that differ from the properties of the underlying individual particles. In contrast, quasiparticles typically describe individual particle-like excitations that retain the quantum numbers of their constituent particles, such as charge or for dressed electrons or holes, allowing them to propagate as if they were weakly interacting entities within the many-body medium. The primary distinction lies in their nature: quasiparticles often embody localized, particle-like behavior with fermionic or bosonic tied to single-particle characteristics, whereas collective excitations inherently exhibit bosonic and rely on macroscopic across the system for their stability and propagation. Certain cases blur this boundary, as some collective excitations can be modeled as quasiparticles when they effectively behave like bosons with specific quantum numbers; for instance, magnons, which are of spin waves, are treated as spin-1 bosonic quasiparticles despite their collective origin in spin alignments. This overlap highlights how the quasiparticle approximation can encompass collective modes if they approximate free-particle relations. A notable example of this distinction appears in superconductors, where Cooper pairs—composite quasiparticles formed by paired electrons—carry the charge and fermionic heritage of their constituents but function as bosonic entities in the , while the Higgs mode represents a purely manifesting as oscillations of the superconducting order parameter without individual particle quantum numbers.

Elementary versus Composite Quasiparticles

Quasiparticles are broadly categorized into elementary and composite types based on their internal structure and formation mechanism within many-body systems. Elementary quasiparticles represent renormalized versions of bare single particles, where interactions with the surrounding medium effectively "dress" the original particle, modifying its effective mass and other properties without fundamentally altering its single-particle nature. For instance, in , conduction electrons in metals are described as dressed quasiparticles with a finite quasiparticle weight Z, where $0 < Z \leq 1, quantifying the overlap between the interacting quasiparticle state and the non-interacting bare particle state. This renormalization arises from perturbative corrections to the single-particle Green's function, ensuring that low-energy excitations behave as weakly interacting entities despite strong underlying correlations. In contrast, composite quasiparticles emerge as bound states involving multiple fundamental particles or excitations, forming stable entities with collective internal degrees of freedom. Examples include skyrmions in magnetic systems, which are topologically stable configurations of spins, and excitons as electron-hole pairs in semiconductors, though their detailed properties are discussed elsewhere. These composites differ from elementary quasiparticles by requiring an attractive interaction to form a bound state, often resulting in emergent properties not present in the constituent particles. Unlike collective excitations, which involve macroscopic coherent modes across the system, composite quasiparticles maintain localized, particle-like behavior due to their binding. The distinction between elementary and composite quasiparticles hinges on criteria such as binding energy, stability, and quantum numbers. For composites, a positive binding energy ensures stability against dissociation, while the total energy of the bound state lies below that of the separated constituents; weak binding, where the binding energy is much smaller than the rest mass energy, is particularly relevant for quasiparticles in condensed matter. Additionally, composites frequently exhibit fractional quantum numbers, such as charge or spin, reflecting their multi-particle origin. Stability is further governed by the system's topology or symmetry breaking, preventing decay into free particles. Elementary quasiparticles, by comparison, retain integer quantum numbers akin to their bare counterparts, with renormalization primarily affecting dispersion rather than composition. A prominent example of composite quasiparticles appears in the fractional quantum Hall effect (FQHE), where anyons emerge as fractionally charged excitations obeying exotic anyonic statistics. These anyons are composite fermionic quasiparticles, formed as topological bound states of electrons attached to an even number of magnetic flux quanta, enabling the observed fractional Hall conductance plateaus. This composite structure underlies the non-Abelian statistics crucial for potential topological quantum computing applications. Recent advances highlight topological quasiparticles like as composites in superconductors. In certain vortex configurations, these modes manifest as composite quasiparticles behaving as , arising from the pairing of electron and hole components in a topologically nontrivial superconductor. Their emergence as zero-energy bound states at defects underscores the role of composite formation in realizing fault-tolerant .

Key Examples

Phonons and Magnons in Solids

In crystalline solids, phonons emerge as quasiparticles that represent the quantized normal modes of collective lattice vibrations, arising from the harmonic approximation to the interatomic potential energy. These modes describe the oscillatory motion of atoms around their equilibrium positions, treated as a many-body system where the lattice is modeled as a set of coupled harmonic oscillators. The phonon dispersion relation \omega(\mathbf{q}), which relates the angular frequency \omega to the wavevector \mathbf{q} in the first Brillouin zone, exhibits distinct acoustic and optical branches. Acoustic branches correspond to in-phase vibrations of adjacent atoms, yielding a linear dispersion \omega(\mathbf{q}) \approx v_s |\mathbf{q}| at long wavelengths (small |\mathbf{q}|), where v_s is the speed of sound determined by the material's elastic constants; these modes propagate mechanical waves akin to sound. Optical branches, prevalent in multi-atom unit cells, involve out-of-phase motions and possess a finite frequency at \mathbf{q} = 0, typically in the infrared range, reflecting the restoring forces from short-range ionic or covalent bonds. The quantum mechanical description of phonons employs second quantization, expressing the lattice Hamiltonian in terms of bosonic creation and annihilation operators. For a system of normal modes, the phonon Hamiltonian takes the form \hat{H} = \sum_{\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 the sum runs over wavevectors \mathbf{q} and branch indices s (acoustic or optical), \omega_{\mathbf{q}, s} is the mode frequency, and \hat{a}_{\mathbf{q}, s}^\dagger, \hat{a}_{\mathbf{q}, s} satisfy bosonic commutation relations [\hat{a}_{\mathbf{q}, s}, \hat{a}_{\mathbf{q}', s'}^\dagger] = \delta_{\mathbf{q}, \mathbf{q}'} \delta_{s, s'}. This formulation underscores the bosonic statistics of phonons, enabling their treatment as indistinguishable particles with integer occupation numbers, and facilitates the computation of thermal and transport properties. Phonons obey , leading to phenomena like in certain low-dimensional systems under specific conditions. Phonons play a pivotal role in mediating electron-phonon coupling, where lattice vibrations scatter electrons and influence charge and heat transport; this interaction is central to the lattice contribution to thermal conductivity, as phonons carry heat via their propagation and scattering processes. Magnons, in contrast, are bosonic quasiparticles that embody quantized spin waves—collective excitations of the spin lattice in magnetically ordered solids, such as ferromagnets and antiferromagnets. In ferromagnetic materials, magnons describe transverse deviations from the uniform spin alignment, effectively reducing the total magnetization by \hbar per excitation; their dispersion at low wavevectors is quadratic, \omega(\mathbf{q}) \approx D |\mathbf{q}|^2, where D is the spin-wave stiffness constant reflecting the exchange interaction strength and spin magnitude. In antiferromagnets, magnons capture fluctuations around the staggered Néel order, typically exhibiting two degenerate linear branches with \omega(\mathbf{q}) \approx v_m |\mathbf{q}| near \mathbf{q} = 0, where v_m is the spin-wave velocity set by superexchange pathways; an energy gap may arise from anisotropy or external fields. Magnons, like phonons, follow bosonic statistics, allowing multiple excitations in the same mode. The theoretical framework for magnons relies on mapping the spin Hamiltonian to bosonic operators via the , which expands spin operators in powers of boson creation and annihilation operators for low-energy excitations: for a spin-S site, S^z_i = S - \hat{b}_i^\dagger \hat{b}_i and S^+_i \approx \sqrt{2S} \hat{b}_i, where \hat{b}_i^\dagger, \hat{b}_i are site-specific bosons, valid in the dilute magnon limit. This approximation linearizes the equations of motion, yielding the dispersion relations and enabling perturbative treatments of interactions. As composite quasiparticles, magnons arise from the collective alignment of many spins, distinguishing them from elementary excitations.

Excitons and Polarons in Semiconductors

In semiconductors, excitons form as bound states of an electron in the conduction band and a hole in the valence band, mediated by Coulomb attraction and stabilized by the material's dielectric screening. These quasiparticles are particularly prominent in direct-bandgap materials like , where they influence optical absorption and emission processes central to optoelectronic devices. Excitons in such systems are typically classified as Wannier-Mott type, characterized by large radii (tens to hundreds of lattice constants) and low binding energies (on the order of 1-100 meV), contrasting with the tightly bound Frenkel excitons in molecular solids. As composite quasiparticles, excitons exemplify the distinction from elementary excitations by combining two charge carriers into a neutral entity. The binding energy of a Wannier-Mott exciton follows a hydrogen-like model, given by E_b = \frac{\mu e^4}{2 \hbar^2 \varepsilon^2}, where \mu is the reduced mass of the electron-hole pair, e is the elementary charge, \hbar is the reduced Planck's constant, and \varepsilon is the static dielectric constant of the semiconductor. This formula arises from solving the Schrödinger equation for the relative motion under a screened Coulomb potential, yielding discrete energy levels analogous to the hydrogen atom but scaled by the effective masses and screening. In typical semiconductors like CdTe, E_b ranges from 10-50 meV, enabling thermal dissociation at elevated temperatures. The dispersion relation for excitons in semiconductors is parabolic, E(\mathbf{K}) = E_g - E_b + \frac{\hbar^2 K^2}{2M}, where \mathbf{K} is the center-of-mass momentum, E_g is the bandgap energy, and M = m_e^* + m_h^* is the total effective mass from the electron (m_e^*) and hole (m_h^*) contributions. This form reflects the translational invariance of the crystal lattice, allowing excitons to propagate as quasiparticles with bandwidths on the order of 10-100 meV. Excitons dissociate into free carriers when their kinetic energy exceeds the binding energy, which occurs for photon energies above the bandgap E_g, leading to free carrier generation in photoexcitation processes. Polarons emerge in polar semiconductors when charge carriers couple to longitudinal optical phonons, effectively "dressing" the electron or hole with a lattice distortion cloud that follows its motion. This electron-phonon interaction is captured by the Fröhlich Hamiltonian, H = \frac{p^2}{2m^*} + \sum_{\mathbf{q}} \hbar \omega \left( b_{\mathbf{q}}^\dagger b_{\mathbf{q}} + \frac{1}{2} \right) + \sum_{\mathbf{q}} V_q (b_{\mathbf{q}} + b_{-\mathbf{q}}^\dagger) e^{i \mathbf{q} \cdot \mathbf{r}}, where the first term is the carrier kinetic energy, the second describes free phonons, and the third mediates the coupling with strength V_q \propto 1/q for long-range Coulomb-like interactions. Polarons are categorized as large (weak coupling, radius much larger than lattice constant) or small (strong coupling, localized within a few sites), with the transition depending on the coupling constant \alpha = \frac{1}{2} \left( \frac{1}{\varepsilon_\infty} - \frac{1}{\varepsilon_0} \right) \sqrt{\frac{\hbar}{2 m^* \omega}}, where \varepsilon_\infty and \varepsilon_0 are high- and low-frequency dielectric constants, and \omega is the phonon frequency. The polaron radius in the weak-coupling (large polaron) regime is approximated as r_p \approx \sqrt{\frac{\hbar}{2 m^* \omega}}, providing a measure of the distortion extent; for \alpha < 1, r_p exceeds the lattice spacing, as in materials like GaAs where \alpha \approx 0.06. Polarons account for the observed reduction in charge carrier mobility in doped semiconductors, as the effective mass increases to m^* (1 + \alpha/6), scattering carriers and limiting drift velocities to 10-1000 cm²/V·s at room temperature. Conversely, excitons underpin photoluminescence in semiconductors, where radiative recombination of the bound electron-hole pair emits light at energies slightly below the bandgap, enabling efficient LEDs and lasers in materials like InGaN. Theoretical proposals suggest the possibility of room-temperature Bose-Einstein condensation of in two-dimensional , such as coupled to , where high binding energies (up to 500 meV) and reduced screening could stabilize dense excitonic gases that macroscopically occupy the ground state, opening pathways for coherent optoelectronic devices.

Physical Properties and Effects

Impact on Bulk Material Properties

Quasiparticles collectively underpin the macroscopic properties of materials, such as thermal conductivity, electrical transport, and optical response, by encapsulating the effects of many-body interactions into effective single-particle excitations. In this framework, the emergent behaviors of quasiparticles, including their dispersion relations and scattering processes, directly influence equilibrium thermodynamic quantities and linear response functions, enabling predictive models for bulk phenomena in solids, liquids, and other condensed matter systems. Thermal properties of materials are profoundly shaped by quasiparticle contributions, particularly from phonons and fermionic excitations. In insulating and semiconducting solids, phonons—quantized lattice vibrations—dominate the low-temperature specific heat according to the Debye model, which approximates the phonon density of states as quadratic in frequency and yields C_v \propto T^3 for temperatures much below the Debye temperature \Theta_D. This cubic dependence arises from the excitation of long-wavelength acoustic modes, aligning closely with experimental observations in non-metallic crystals and providing a cornerstone for understanding heat capacity in insulators. In metallic systems described by Fermi liquid theory, the electronic quasiparticles near the contribute a linear term to the specific heat, C = \gamma T, where the Sommerfeld coefficient \gamma = \frac{\pi^2}{3} k_B^2 N^*(0) reflects the enhanced density of states N^*(0) due to interactions, often manifesting as an effective mass enhancement m^* > m. Electrical conductivity in conductors emerges from the motion of charge-carrying quasiparticles, modulated by scattering events. The Drude model captures this through \sigma = \frac{n e^2 \tau}{m^*}, where n is the carrier density, e the charge, \tau the relaxation time determined by quasiparticle-phonon or quasiparticle-quasiparticle scattering, and m^* the effective mass incorporating band structure effects. This formulation explains the temperature dependence of resistivity in metals, with \tau decreasing at higher temperatures due to increased phonon scattering, and has been refined in quasiparticle theories to account for Fermi surface properties. Optical properties, including reflectivity and absorption, are governed by collective electronic excitations like plasmons. The dielectric function \varepsilon(\omega) incorporates plasmon contributions, leading to a screened response where the plasma frequency \omega_p = \sqrt{\frac{4\pi n e^2}{m}} marks the resonance of density oscillations, causing metallic reflection for \omega < \omega_p. In topological materials, such as Chern insulators, quasiparticles with nontrivial Berry curvature induce an intrinsic anomalous Hall effect, yielding a quantized transverse conductivity \sigma_{xy} = \frac{e^2}{h} C (with Chern number C) even without magnetization or external fields, as realized in thin films of magnetic topological insulators.

Interactions and Decay Mechanisms

Quasiparticles in condensed matter systems interact through various scattering processes that limit their coherence and transport properties. A primary interaction is electron-phonon , particularly in the Bloch-Grüneisen regime at low , where acoustic with wavelengths much longer than the lattice constant dominate. In this regime, the rate for electrons near the scales with as T^5 for three-dimensional metals, arising from the restrictions on phonon emission and that conserve both and . Electron-electron , often mediated by Umklapp processes, provides another key interaction channel, enabling relaxation essential for finite resistivity in metals. In Fermi liquids, Umklapp electron-electron contributes a [T^2](/page/T+2) dependence to the quasiparticle lifetime at low , distinguishing it from normal that conserves total . Quasiparticles also decay through specific channels that dissipate their energy into other excitations. For plasmons, represents a fundamental decay mechanism, where the collective electron density oscillation couples to and decays into single-particle electron-hole pairs across the Fermi sea, with the damping rate proportional to the available for such transitions. In semiconductors, excitons undergo Auger recombination as a non-radiative decay process, in which the recombination energy of an electron-hole pair excites a third to a higher state, leading to trion formation or free carrier heating; this process becomes dominant at high exciton densities. These decay channels highlight the of quasiparticles in interacting environments, where laws dictate the available final states. The finite lifetimes of quasiparticles are quantitatively described by Fermi's golden rule, which gives the transition rate from an initial state to a continuum of final states. The inverse lifetime is expressed as \tau^{-1} = \frac{2\pi}{\hbar} \sum_{k'} |M_{kk'}|^2 \delta(\epsilon_k - \epsilon_{k'}), where M_{kk'} is the matrix element of the interaction, and the delta function enforces energy conservation. For phonons in clean crystalline systems, lifetimes are primarily limited by anharmonic interactions, such as three-phonon scattering processes that split or combine phonons. At high temperatures, these anharmonicities yield a phonon lifetime scaling as \tau \propto T^{-1}, reflecting the increased phonon population and scattering opportunities, while higher-order processes can lead to stronger temperature dependences like T^{-2} or more in certain materials. In quantum devices, such as superconducting s, quasiparticle poisoning emerges as a critical decay-related issue, where nonequilibrium quasiparticles generated by absorption or break pairs, injecting excitations that relax qubit coherence times. This poisoning is exacerbated by resonant absorption at the Josephson junction, leading to bursts of quasiparticles that propagate and degrade performance across the device. The finite lifetime of quasiparticles corresponds to the imaginary part of the in the formalism, quantifying the broadening of spectral features.

Experimental Detection

Spectroscopic Techniques

Spectroscopic techniques play a crucial role in the direct observation of quasiparticles, particularly in bulk materials, by probing their and characteristics through or processes. These methods allow researchers to measure quasiparticle excitations, such as phonons and magnons, by detecting shifts in or energies corresponding to the creation or of these quasiparticles. Traditional approaches focus on ensemble-averaged responses in solids, providing insights into their relations and interactions without requiring surface sensitivity. Raman spectroscopy is a primary optical technique for detecting , relying on the of monochromatic light where incident photons exchange energy with lattice vibrations. In the Stokes process, the scattered photon loses energy equal to the phonon creation energy, resulting in a red-shifted line, while the anti-Stokes process involves phonon annihilation and a blue-shifted line, with the intensity ratio governed by the Boltzmann factor. Selection rules for Raman-active modes arise from the of the crystal lattice, requiring the tensor to change under the phonon's symmetry operations, thus only even-parity (gerade) modes in centrosymmetric crystals are typically observable. This technique has been instrumental in mapping phonon dispersions near the center, as seen in studies of and . Infrared absorption spectroscopy targets polar quasiparticles, particularly transverse optical (TO) phonons in ionic crystals, where the dipole moment induced by lattice vibrations couples to the of , leading to resonant absorption at frequencies matching the energy. For materials with ionic character, such as NaCl, this manifests as strong bands corresponding to TO modes, while longitudinal optical (LO) modes are often inactive due to no net dipole change. In polar materials, photons can couple with optical to form —hybrid quasiparticles exhibiting mixed electromagnetic and mechanical properties—observable as broadened or split features in the far- range. Neutron scattering provides a powerful momentum-resolved probe for both phonons and magnons, leveraging the 's magnetic and mass to interact with nuclear and magnetic in the sample. Inelastic measures the dynamic S(\mathbf{q}, \omega), where the differential cross-section is proportional to S(\mathbf{q}, \omega), encoding the space-time correlations of atomic displacements or spins, allowing full dispersion mapping across the . For phonons in metals like aluminum, this reveals acoustic and optical branches, while for magnons in antiferromagnets such as MnF₂, it probes spin-wave excitations with resolutions down to low energies. These techniques have resolution limits that influence their applicability: typically achieves ~1 meV energy resolution, suitable for zone-center phonons but limited for low-energy acoustic modes, whereas neutron scattering offers superior resolution (often <0.1 meV with advanced spectrometers) for momentum-dependent studies, though it requires large, bulk single-crystal samples due to the cross-section. Historically, quasiparticles like phonons were first inferred indirectly through specific heat measurements fitting the in the early , but direct spectroscopic confirmation came with the observation of from phonons in crystals in 1928.

Modern Probes like ARPES

Angle-resolved photoemission spectroscopy (ARPES) is a pivotal technique for probing quasiparticle properties in condensed matter systems, directly mapping the energy-momentum dispersion relation E(\mathbf{k}) of electrons near the surface. In ARPES, ultraviolet or soft X-ray photons illuminate the sample, ejecting photoelectrons whose kinetic energy E_{\text{kin}} and emission angle \theta encode the initial state's binding energy E_B and in-plane momentum \mathbf{k}_\parallel, via the relations E_B = h\nu - E_{\text{kin}} - \phi and \mathbf{k}_\parallel = \sqrt{2m E_{\text{kin}}} / \hbar \cdot (\sin\theta \cos\phi, \sin\theta \sin\phi), where h\nu is the photon energy, \phi is the azimuthal angle, and \phi is the work function. The spectral linewidth in ARPES spectra, often broadening to \sim \Gamma, reflects the quasiparticle lifetime \tau \approx \hbar / \Gamma, providing insights into scattering processes such as electron-phonon or electron-electron interactions. Modern ARPES setups leverage sources to achieve high energy and momentum resolution, often below 10 meV and 0.01 Å⁻¹, respectively, enabling precise quasiparticle band mapping. These facilities provide tunable energies and high , minimizing sample damage while allowing access to buried interfaces through increased depths at higher energies. However, the measured intensity is modulated by photoemission matrix elements, which depend on the , initial and final state wavefunctions, and orbital symmetries, sometimes suppressing certain bands and requiring careful polarization analysis for complete spectral reconstruction. ARPES has been instrumental in visualizing and quasiparticle effects in complex materials, particularly high-temperature superconductors. In cuprates like Bi₂Sr₂CaCu₂O₈₊δ, ARPES reveals a reconstructed with pseudogap features above the superconducting transition, alongside mass indicated by band flattening near the antinodal regions. These measurements quantify the \Sigma(\mathbf{k}, \omega), showing how interactions broaden and shift quasiparticle dispersions, essential for understanding pairing mechanisms. A hallmark application in cuprates is the confirmation of coherent quasiparticle bands, evidenced by dispersion kinks around 60-70 meV , attributed to strong electron-phonon coupling. These kinks manifest as abrupt changes in slope in the nodal band dispersion, disrupting the linear Dirac-like behavior and signaling phonon-mediated scattering, as observed across underdoped to overdoped regimes. Such features validate the quasiparticle picture while highlighting many-body effects beyond simple band theory. Recent advances in time-resolved ARPES (TR-ARPES), employing pump-probe schemes with high-harmonic generation or free-electron lasers, have extended these probes to quasiparticle dynamics on timescales. Post-2020 developments, including sub-10 resolution, allow tracking of nonequilibrium band and lifetime after photoexcitation, revealing ultrafast of electron-phonon interactions in transient states. In two-dimensional materials like , nano-ARPES variants with below 100 nm have uncovered spatially varying Dirac quasiparticle dispersions influenced by substrate interactions and , enabling studies of emergent topological states in moiré superlattices.

Historical Development

Early Theoretical Foundations

The foundational concepts of quasiparticles in originated in the early , as researchers sought to describe collective excitations in periodic structures like crystals. These efforts began with addressing the quantum mechanical behavior of individual particles in lattices, evolving toward recognizing emergent entities that simplify many-body interactions. By the mid-20th century, this led to a coherent framework for excitations with effective properties distinct from bare particles. A pivotal early contribution came from in 1928, who established the theoretical basis for electron quasiparticles in periodic potentials. Bloch showed that the for an electron in a crystal lattice—approximated as a strictly periodic potential—yields solutions in the form of plane waves modulated by periodic functions, known as Bloch waves. These wavefunctions describe delocalized electrons propagating through the lattice without scattering from individual ions, forming energy bands that underpin the quasiparticle picture of conduction electrons in metals and semiconductors. This work shifted the view of electrons from free particles to lattice-adapted quasiparticles, enabling the development of band theory. Parallel developments addressed lattice vibrations, with and providing the groundwork for phonons as quasiparticle normal modes. In their 1912 paper, extended through the 1920s, they modeled a as a of coupled oscillators under , describing the collective vibrational through discrete normal modes. These modes were later quantized as phonons, bosonic quasiparticles representing harmonic excitations of the entire rather than independent atomic motions, with frequencies determined by interatomic forces. This approach resolved inconsistencies in classical models and laid the foundation for quantum treatments of and elastic properties in solids, treating phonons as propagating as sound waves or optical vibrations. For instance, acoustic phonons correspond to long-wavelength lattice displacements, while optical phonons arise from relative motions of atoms in multi-atom unit cells. Building on these ideas, Building on Lev Landau's earlier introduction of the polaron concept in 1933, Herbert Fröhlich developed a model in 1950 to describe -lattice interactions in polar crystals. Fröhlich modeled an as coupled to longitudinal optical s via the long-range interaction, leading to a self-induced lattice that "dresses" the . The resulting quasiparticle exhibits a reduced effective and altered compared to a bare , as the cloud accompanies its motion. This perturbative treatment of the electron- coupling highlighted how interactions create composite excitations, influencing transport and optical properties in ionic materials like semiconductors. The culmination of these early foundations occurred with Lev Landau's in 1956–1957, which formalized quasiparticles in interacting many-body systems. Landau proposed that low-energy excitations in a degenerate , such as liquid helium-3 or conduction electrons, could be described as weakly interacting quasiparticles with renormalized masses, velocities, and finite lifetimes due to collision processes. Landau also introduced the term "quasiparticle" to describe these excitations. Unlike ideal fermions, these quasiparticles decay over time, with lifetimes inversely proportional to temperature squared near the , yet they retain Fermi-Dirac statistics at low temperatures. This phenomenological approach resolved the apparent paradox of strong interactions in dense systems by mapping them onto an effective non-interacting picture. Throughout this era, interactions were handled perturbatively, approximating quasiparticles as small corrections to free-particle states and focusing on scattering amplitudes rather than exact many-body solutions, which would emerge later.

Key Milestones and Advances

The concept of quasiparticles gained traction in the early 1930s through foundational work on collective excitations in solids. In 1930, introduced the idea of spin waves as low-energy excitations in ferromagnetic materials, describing how deviations from perfect spin alignment propagate as quantized units later termed magnons; this model explained the temperature dependence of in ferromagnets. The following year, Yakov Frenkel proposed excitons as bound electron-hole pairs that behave as neutral quasiparticles, capable of migrating through insulating crystals without net charge transfer, providing a framework for understanding optical absorption and in molecular solids. Building on these ideas, Frenkel coined the term "" in 1932 to denote the quantized modes of vibrations, analogous to photons for electromagnetic waves, which simplified the treatment of thermal and acoustic properties in crystals by treating vibrations as non-interacting bosonic quasiparticles. In 1933, advanced the concept, envisioning an electron dressed by a of distortions in polar materials, resulting in an effective particle with enhanced and reduced ; this self-trapping mechanism became essential for interpreting charge transport in ionic crystals and semiconductors. These developments marked a shift toward viewing complex many-body interactions as emergent, particle-like entities. A pivotal advance came in 1937 with Gregory Wannier's extension of the exciton model to larger-radius bound states in semiconductors, known as Wannier-Mott excitons, where the electron-hole separation spans multiple sites due to weaker binding; this complemented Frenkel's tightly bound s and proved crucial for interpreting excitonic effects in materials like Cu₂O. Landau's 1941 theory of superfluid helium II further solidified the quasiparticle paradigm by modeling the fluid's excitations as phonons at low energies and rotons at higher energies, enabling a two-fluid description that accounted for superfluidity's macroscopic quantum behavior without viscosity. The most general framework emerged in 1956 with Landau's Fermi liquid theory, which posited that low-temperature excitations in interacting Fermi systems—such as liquid ³He—could be described as weakly interacting quasiparticles with renormalized effective masses and lifetimes, preserving Fermi-Dirac statistics despite strong correlations; this phenomenological approach revolutionized the understanding of normal metals and degenerate gases. Subsequent refinements, including the 1958 work by Pines and Nozières, integrated diagrammatic to compute quasiparticle parameters, bridging Landau's ideas with quantum field methods. Later advances in the and beyond extended quasiparticles to novel contexts, such as Bogoliubov quasiparticles in superconductors (1957), which diagonalize the BCS Hamiltonian to reveal excitations, and the identification of fractional quasiparticles like anyons in two-dimensional systems (Laughlin, 1983), enabling descriptions of the . These milestones transformed quasiparticle theory from models into a cornerstone of , facilitating quantitative predictions for transport, , and phase transitions across diverse materials.

References

  1. [1]
    [PDF] What is a Quasi-Particle? * 1
    The concept of a quasi-particle excitation in an interacting many-body system will be discussed from both the physical and the mathematical points of view.
  2. [2]
    Fantastic beasts | Nature Physics
    Dec 1, 2016 · Elementary particles are the building blocks of matter, but there is also a zoo of quasiparticles that are crucial for understanding how this matter behaves.
  3. [3]
    [PDF] Chapter 5 Landau Theory of the Fermi Liquid
    The one-electron (particle) state becomes a quasiparticle which carries the same charge (−e) and spin (±1/2) of the bare electron. Non-interacting. Interacting.
  4. [4]
    None
    Below is a merged summary of quasiparticles from *Many-body quantum theory in condensed matter physics* by Henrik Bruus and Karsten Flensberg, consolidating all information from the provided segments into a comprehensive response. To retain maximum detail efficiently, I will use a combination of narrative text and a table for effective properties and key details, ensuring all information is preserved. The response is structured to avoid redundancy while capturing the full scope of the summaries.
  5. [5]
    [PDF] Quasiparticles in condensed matter systems
    Quasiparticles are a powerful concept of condensed matter quantum theory. In this review, the appearence and the properties of quasiparticles are presented in a ...
  6. [6]
    Many-Body Quantum Theory in Condensed Matter Physics
    This book is an introduction to the techniques of many-body quantum theory with a large number of applications to condensed matter physics.
  7. [7]
    [PDF] The Theory of a Fermi Liquid
    A theory of the Fermi licpid is constructed, based on the representation of the perturbation theory as a functional of the distribution function. The effective ...
  8. [8]
    [PDF] Quantum Theory of Many Particle Systems
    Mar 28, 2023 · We explicitly evaluate the expansion for a two-body interaction: V (ψ. ∗. (τ),ψ(τ)) = 1. 2. X αβγδ. (αβ |ˆv| γδ) ψ. ∗ α(τ)ψ. ∗ β(τ)ψδ(τ)ψγ(τ) ...
  9. [9]
  10. [10]
  11. [11]
  12. [12]
    [PDF] Dynamical mean-field theory of strongly correlated fermion systems ...
    We review in detail the recent progress in understanding the Hubbard model and the Mott metal-insulator transition within this approach, including some ...
  13. [13]
    [PDF] I. Collective Behavior, From Particles to Fields - MIT
    As in the case of phonons in solids or in superfluid helium, low energy excitations are typically collective modes, which involve the coordinated motions of ...
  14. [14]
    E 4 - UC Berkeley Ultracold Atomic Physics
    Apr 11, 2016 · Collective excitations are an emergent phenomena, modes of a many-body system with a well defined momentum and energy that can look unlike the constituent ...
  15. [15]
    Cooper pairs versus Higgs mode contribution | Phys. Rev. B
    May 25, 2016 · Here we show that this nonlinear optical process is dominated by light-induced excitation of Cooper pairs, while the collective amplitude (Higgs) fluctuations
  16. [16]
    Classification and characterization of nonequilibrium Higgs modes ...
    Jan 15, 2020 · Nonlinear optical effects and third-harmonic generation in superconductors: Cooper pairs versus Higgs mode contribution. Phys. Rev. B 93 ...
  17. [17]
    [PDF] Fermi-Liquid Theory - LPTMC
    Jan 16, 2025 · Besides Landau's original papers [1–3], there are excellent textbooks on Landau's. Fermi-liquid theory [4–7]. Sections 4.1, 4.2 and 4.3 rely ...
  18. [18]
    Structure of quasiparticles in a local Fermi liquid | Phys. Rev. B
    It clarifies the relationship between the single-particle eigenstates of the noninteracting effective quasiparticle Hamiltonian, and the associated many-body ...
  19. [19]
    [PDF] The Ontology of Compositeness Within Quantum Field Theory - arXiv
    Sep 8, 2024 · Compositeness is most useful when the object is weakly bound, that is the binding energy B is much less than the total rest mass M10. I claimed ...<|separator|>
  20. [20]
    Elementary and Composite Particles | Phys. Rev.
    The distinction between elementary and composite particle is related to a generalization of Levinson's theorem, and it is made plausible that this ...
  21. [21]
    Fractionally charged skyrmions in fractional quantum Hall effect - PMC
    Nov 26, 2015 · The fractional quantum Hall effect5 (FQHE) arises due to the formation of composite fermions, which are topological bound states of electrons ...
  22. [22]
    Anyons and the quantum Hall effect—A pedagogical review
    This paper is aimed at reviewing the physics of Anyons, particles whose statistics is neither fermionic not bosonic, and the way it is manifested in the ...
  23. [23]
    Field-driven spatiotemporal manipulation of Majorana zero modes in ...
    Dec 20, 2023 · This composite quasiparticle behaves as a non-Abelian anyon similar to a zero-energy Majorana state trapped by a vortex in superconductors.
  24. [24]
    Topological Quantum Materials for Realizing Majorana Quasiparticles
    A pivotal goal of this research has been to realize materials hosting Majorana quasiparticles, thereby making topological quantum computing a technological ...The Technological Driver... · Majorana Quasiparticles and... · Materials Exemplars
  25. [25]
    [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 ...
  26. [26]
    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.Article Text · Electron-phonon Interaction in… · Electron-phonon Effects in the...
  27. [27]
    Introduction to antiferromagnetic magnons | Journal of Applied Physics
    Oct 15, 2019 · Antiferromagnets are magnetic materials that have no net macroscopic magnetization and, therefore, are almost insensitive to external magnetic fields.INTRODUCTION · III. ANTIFERROMAGNETIC... · IV. ANTIFERROMAGNETIC...
  28. [28]
    Excitons and excitonic materials | MRS Bulletin
    Sep 1, 2024 · An exciton is a bound pair of negatively charged electron and positively charged hole (electron vacancy within a solid), both of which are held together by ...
  29. [29]
    [PDF] Polarons - arXiv
    It is customary to use the term “Fröhlich-polaron” or “large polaron” for the quasiparticle consisting of the electron (or hole) and the polarization due to the ...
  30. [30]
    Excitonic Photoluminescence in Semiconductor Quantum Wells
    Feb 13, 2004 · Excitonic Photoluminescence in Semiconductor Quantum Wells: Plasma versus Excitons. S. Chatterjee, C. Ell, S. Mosor, G. Khitrova, and H. M. ...
  31. [31]
  32. [32]
    Electron Interaction in Solids. Collective Approach to the Dielectric ...
    Collective Approach to the Dielectric Constant. Philippe Nozières and David Pines. Palmer Physical Laboratory, Princeton University, Princeton, New Jersey. PDF ...Missing: quasiparticles ω) plasma ω_p = √( 4π m)
  33. [33]
    Colloquium: Quantum anomalous Hall effect | Rev. Mod. Phys.
    Jan 23, 2023 · The quantum anomalous Hall (QAH) effect refers to QH effects that occur in the absence of external magnetic fields due to spontaneously broken time-reversal ...
  34. [34]
    Umklapp scattering as the origin of T -linear resistivity in the normal ...
    Dec 11, 2017 · We present a simple explanation for this behavior, in terms of the umklapp scattering of electrons. This fits within the general picture emerging from ...
  35. [35]
    Carriers, Quasi-particles, and Collective Excitations in Halide ...
    Jun 5, 2023 · ... decay due to Auger recombination occurs (Figure 12e,f). Furthermore, compared to CsPbBr3 NCs, longer HC cooling time due to the Auger ...
  36. [36]
    Lifetime of a quasiparticle in an electron liquid | Phys. Rev. B
    Feb 18, 2005 · We calculate the inelastic lifetime of an electron quasiparticle due to Coulomb interactions in an electron liquid at low (or zero) ...
  37. [37]
    Phonon anharmonicity, lifetimes, and thermal transport in from many ...
    Dec 8, 2016 · Here we address the anharmonic nature of phonons in CH 3 NH 3 PbI 3 from a theoretical perspective, with a particular focus on phonon ...Missing: clean | Show results with:clean
  38. [38]
    Quasiparticle Poisoning of Superconducting Qubits from Resonant ...
    Here we show that a dominant mechanism for quasiparticle poisoning is direct absorption of high-energy photons at the qubit junction.
  39. [39]
    Angle-resolved photoemission studies of quantum materials
    May 26, 2021 · E kin is the photoelectron kinetic energy and E B is the binding energy of the electron prior to emission. Because of the discrete in-plane ...
  40. [40]
    High-resolution angle-resolved photoemission spectroscopy and ...
    Dec 8, 2020 · This review outlines fundamental principles, instrumentation, and capabilities of angle-resolved photoemission spectroscopy (ARPES) and microscopy.
  41. [41]
    Understanding intensities of angle-resolved photoemission with ...
    Apr 15, 2009 · Here we report an extended study of the angle-resolved photoemission intensity of the Cu(111) Shockley surface state as measured in different ...Missing: setup | Show results with:setup
  42. [42]
    The relevance of ARPES to high-T c superconductivity in cuprates
    Jul 9, 2020 · Angle-resolved photoemission spectroscopy, visualizing the superconducting gap in k-space, plays a pivotal role in research on cuprates and other high-T c ...
  43. [43]
    Fermi-surface reconstruction and the origin of high-temperature ...
    Oct 18, 2010 · The cause of the reconstruction, and its implication for the origin of high-temperature superconductivity, is a subject of active debate.
  44. [44]
    Unmasking the Origin of Kinks in the Photoemission Spectra of ...
    Apr 8, 2021 · We report here ab initio calculations based on G ⁡ W perturbation theory and show that correlation-enhanced electron-phonon interaction in cuprates gives rise ...Abstract · Article Text
  45. [45]
    Dispersion kinks from electronic correlations in an unconventional ...
    Nov 17, 2024 · The kink initiates a high-energy section of steep dispersion with reduced intensity, which connects the coherent quasiparticle component of the ...
  46. [46]
    Time-resolved ARPES studies of quantum materials | Rev. Mod. Phys.
    Feb 27, 2024 · Time-resolved ARPES (TR-ARPES) can probe ultrafast electron dynamics and the out-of-equilibrium electronic structure, providing a wealth of information.
  47. [47]
    Recent technical advancements in ARPES: Unveiling quantum ...
    This topical review delves into four cutting-edge ARPES techniques that are currently at the forefront of development: time-resolved ARPES, nano ARPES, in-situ ...Missing: renormalization Tc
  48. [48]
    G 0 ⁡ Δ ⁢ W theory: Quasiparticle properties of two-dimensional ...
    Jan 2, 2024 · For example, the angle-resolved photoemission-spectroscopy (ARPES) measurements show that WS 2 band gap decreases about 140 meV on a graphite ...
  49. [49]
    Über die Quantenmechanik der Elektronen in Kristallgittern
    Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses durch ein zunächst streng dreifach periodisches Kraftfeld schematisieren.
  50. [50]
    [PDF] Freedom, Collectivism, and Quasiparticles: Social Metaphors in ...
    A short version of the history of quasiparticles was told by two physicists driven mainly by the curiosity to find out "who named the -on's" (both ...
  51. [51]
    Theory of the Superfluidity of Helium II | Phys. Rev.
    Phys. Rev. 60, 356 (1941) ... Theory of the Superfluidity of Helium II. L. Landau. Institute of Physical Problems, Academy of Sciences USSR, Moscow, USSR. PDF ...Missing: quasiparticles | Show results with:quasiparticles