Fact-checked by Grok 2 weeks ago

Drude model

The Drude model is a classical of electrical conduction in metals, proposed by Paul Drude in 1900 to explain the transport properties of electrons in materials, particularly how an applied induces a through the motion of free electrons. In this model, metals are envisioned as a of stationary positive ions surrounded by a gas of conduction electrons that behave like classical particles, accelerating under the influence of an but undergoing frequent collisions with the ions that randomize their velocities and establish a steady-state drift. These collisions are characterized by a relaxation time τ, typically on the order of 10⁻¹⁴ seconds at , during which electrons travel a before . The model's key achievement is deriving the electrical conductivity σ as σ = n e² τ / m, where n is the , e is the , and m is the , which quantitatively matches experimental values for metals (1–10 μΩ·cm resistivity) and underpins (J = σ E) in the linear response regime. It also successfully predicts the , , and the form of the Wiedemann-Franz law relating electrical and thermal conductivities (κ / σ T = (3/2) (k_B / e)², where κ is thermal conductivity and k_B is Boltzmann's constant), though the numerical prefactor differs from the experimental value. Drude's framework built on the recent discovery of the by J.J. Thomson in 1897 and drew analogies from , assuming no electron-electron interactions and treating scattering as abrupt events independent of the field. Despite its successes, the Drude model has notable limitations, as it fails to account for quantum mechanical effects, incorrectly attributing scattering solely to ion cores (whereas perfect lattices show no scattering in ), and predicting an erroneous temperature dependence for resistivity (proportional to √T rather than the observed linear T). These shortcomings were later addressed by refinements, such as Hendrik Lorentz's correction of a factor-of-two error in Drude's original and the quantum-based Sommerfeld model in , yet the Drude approach remains foundational for understanding classical transport and serves as a starting point for more advanced theories in .

Historical Development

Origins in Classical Physics

The foundations of the Drude model emerged from 19th-century advancements in kinetic theory, initially developed for gases but later adapted to conceptualize electrical conduction in solids. James Clerk Maxwell laid key groundwork in his 1860 work on the dynamical theory of gases, where he modeled gas particles as colliding spheres to derive the velocity distribution and transport properties, providing a framework for treating charge carriers similarly. extended this in 1872 with his transport equation, describing how particle distributions evolve under collisions and external forces, which proved essential for analyzing drift motion in conductors. Building on these ideas, Wilhelm Weber proposed in 1871 that electrical phenomena in matter arise from interactions between charged particles within atoms, envisioning positive charges surrounded by orbiting negative ones, an early atomic model influencing later views of conduction. Eduard Riecke advanced this in the by modeling metals as lattices of neutral atoms containing free s behaving like a gas, introducing concepts of under an to explain flow. Riecke's 1898 explicitly applied kinetic principles to estimate electron numbers and mobilities in metals, treating conduction as the collective motion of these particles scattered by lattice vibrations. The discovery of the by J.J. Thomson in 1897 provided the crucial empirical basis for such free-electron pictures, as his cathode-ray experiments identified negatively charged corpuscles far lighter than atoms, confirming the existence of mobile subatomic particles capable of carrying charge in metals. However, pre-Drude models like those of Weber and Riecke struggled with key observations, notably failing to explain the positive temperature dependence of resistivity, as their classical scattering assumptions predicted either temperature-independent or weakly varying resistance, contrary to experimental increases with thermal agitation. These classical precursors set the stage for Paul 's 1900 synthesis, which integrated kinetic theory with the confirmed to form a cohesive model of metallic conduction.

Key Formulations by Drude and Lorentz

In 1900, Paul published a seminal paper introducing a classical model for electrical conduction in metals, positing that conduction arises from a gas of free electrons undergoing random thermal motion akin to particles in an ideal gas. Drawing from classical kinetic theory, Drude envisioned these electrons as mobile charge carriers drifting through the metallic lattice under an applied electric field while colliding with ions. Drude's formulation included an initial estimate of electrical conductivity derived from the mean free path of electrons and their density, which he approximated using data from electrolysis experiments, leading to the prediction of approximately one conduction electron per atom in metals. This electron density assumption, grounded in Faraday's laws of electrolysis, marked a key conceptual advance, though it overestimated the number for polyvalent metals and was later refined in subsequent theories. In 1905, refined Drude's model through a series of papers on the motion of electrons in metallic bodies, introducing the relaxation time \tau as a parameter to describe the average time between collisions more realistically. Lorentz's approach, employing the Boltzmann transport equation under a relaxation-time approximation, corrected inconsistencies in Drude's treatment of scattering, such as the use of separate relaxation times for electrical and thermal processes. Lorentz also corrected a factor-of-two error in Drude's original by showing that the relaxation time is the same for electrical and thermal conductivity, improving the prediction for the Wiedemann-Franz law. These refinements established a unified framework that better aligned with experimental observations like the Wiedemann-Franz law and solidified the model's foundational equations for conductivity.

Core Assumptions

Electron Behavior in Metals

In the Drude model, metals are conceptualized as a regular lattice of positively charged ions that remain fixed in position, with the valence electrons detached from their parent atoms and forming a gas of free charge carriers capable of moving throughout the material. This "free electron gas" arises because the valence electrons are loosely bound and can wander freely, neutralizing the positive background charge of the ion lattice to maintain overall electrical neutrality. The model, introduced by Paul Drude in 1900, draws an analogy to the kinetic theory of gases, treating these electrons as classical particles in a container defined by the metal's boundaries. These free electrons are assumed to be in with the , exhibiting random thermal velocities due to their . According to the , each has an average kinetic energy of \frac{3}{2} kT, where k is Boltzmann's constant and T is the , leading to a root-mean-square speed of v_{\rms} = \sqrt{\frac{3kT}{m}}, with m the . At , this speed is on the order of $10^7 cm/s, far exceeding typical drift velocities under applied fields. A key simplification in the model is the neglect of interactions between themselves; instead, are treated as non-interacting particles that only experience from the fixed ions or impurities in the . This independent electron approximation allows the application of dilute gas to the dense electron gas in metals. The of these s, n, is estimated from the number of electrons per atom and the atomic density of the metal, typically yielding n \approx 10^{22} cm^{-3} for common metals like sodium or , where one or more electrons per atom contribute to the gas. For instance, in with one per atom, n \approx 8.5 \times 10^{22} cm^{-3}. This high density underscores the model's innovation in bridging dynamics with the phenomenon of conduction in solids, providing a foundational classical picture despite its simplifications.

Scattering and Relaxation Processes

In the Drude model, conduction electrons in metals are treated as a gas that undergoes frequent collisions primarily with ions, which are modeled as fixed centers. These collisions are assumed to be , preserving the electron's on average, but randomizing its direction, leading to a loss of directional drift and thus electrical . The average time between such collisions is denoted by the relaxation time τ, which characterizes the momentum relaxation process. The mean free path λ, representing the average distance an electron travels between collisions, is given by λ = v_rms τ, where v_rms = √(3k_B T / m) is the root-mean-square speed of the electrons, with k_B the , T the temperature, and m the . Typical values for τ at are on the order of 10^{-14} to 10^{-15} seconds, resulting in λ ≈ 10–100 for common metals like . The relaxation time τ exhibits a strong temperature dependence due to scattering by thermal vibrations of the lattice ions. In the classical Drude model, as temperature increases, the vibration amplitudes (∝ √T) and electron speeds (∝ √T) enhance the scattering rate, yielding τ ∝ T^{-1/2} for this contribution. In addition to scattering from lattice vibrations, impurities and defects introduce a temperature-independent scattering rate 1/τ_imp, which adds a fixed contribution to the total scattering rate via Matthiessen's rule: 1/τ = 1/τ_vib + 1/τ_imp. This impurity term accounts for the residual resistivity observed at low temperatures, where scattering from vibrations diminishes, leaving a finite resistivity even as T → 0. The model assumes a Markovian scattering process, wherein each collision completely randomizes the electron's velocity according to the equilibrium Maxwell-Boltzmann distribution, with no memory of its pre-collision state, justifying the use of a constant average τ independent of prior history.

Electrical Conductivity Derivation

Direct Current (DC) Response

In the Drude model, the (DC) response of a metal to a constant \mathbf{E} is analyzed through the classical equation of motion for a conduction , which balances the accelerating force from the field against a frictional drag due to collisions. The equation is m \frac{d\mathbf{v}}{dt} = -e \mathbf{E} - \frac{m \mathbf{v}}{\tau}, where m is the , e > 0 is the , \mathbf{v} is the electron velocity, and \tau is the average relaxation time between collisions. This form, introduced by Paul Drude in 1900, models collisions as randomizing the electron exponentially with timescale \tau. For steady-state DC conditions, the acceleration term vanishes (d\mathbf{v}/dt = 0), yielding the drift velocity \mathbf{v}_d = -\frac{e \tau}{m} \mathbf{E}. This represents the average velocity superimposed on the random motion of electrons, arising from the balance between field-induced and collision-induced deceleration. The resulting is then \mathbf{J} = -n e \mathbf{v}_d, where n is the , giving \mathbf{J} = \frac{n e^2 \tau}{m} \mathbf{E}. Thus, the DC electrical conductivity is \sigma = n e^2 \tau / m, which linearly relates current density to the applied field via Ohm's law in the form \mathbf{J} = \sigma \mathbf{E}. The corresponding resistivity is \rho = 1/\sigma = m / (n e^2 \tau). The temperature dependence of conductivity enters primarily through \tau(T); in the classical Drude model, phonon scattering leads to \tau \propto 1/\sqrt{T} and thus \sigma \propto 1/\sqrt{T}, though experiments show linear T at higher temperatures. This derivation of \sigma provides the electrical foundation for the Wiedemann-Franz law, where the shared \tau links electrical conductivity to thermal transport, predicting \kappa / (\sigma T) = constant (with the classical Lorentz number L = (3/2) (k_B / e)^2).

Alternating Current (AC) Response

The Drude model extends to alternating current (AC) by considering time-varying electric fields, where the oscillatory nature of the field introduces frequency dependence into the electron dynamics. In this framework, the electric field is assumed to take the form \mathbf{E}(t) = \mathbf{E}_0 e^{-i \omega t}, with the real part representing the physical field. The equation of motion for an electron under this field, incorporating the damping due to scattering, is given by m \frac{d \mathbf{v}}{dt} + \frac{m}{\tau} \mathbf{v} = -e \mathbf{E}_0 e^{-i \omega t}, where m is the electron mass, \mathbf{v} is the velocity, e is the electron charge, and \tau is the relaxation time. Assuming a steady-state solution of the form \mathbf{v}(t) = \mathbf{v}(\omega) e^{-i \omega t}, the velocity in frequency space becomes \mathbf{v}(\omega) = -\frac{e \mathbf{E}_0 / m}{-i \omega + 1/\tau}. This solution captures the balance between acceleration by the field, inertial effects from the frequency term, and frictional damping. The current density \mathbf{j}(\omega) = -n e \mathbf{v}(\omega), where n is the , leads to the complex \sigma(\omega) = \frac{n e^2 \tau / m}{1 - i \omega \tau} = \frac{\sigma_0}{1 - i \omega \tau}, with \sigma_0 = n e^2 \tau / m being the DC conductivity. The frequency-dependent conductivity relates to the dielectric function via in frequency space, yielding \varepsilon(\omega) = 1 + \frac{i \sigma(\omega)}{\varepsilon_0 \omega}, where \varepsilon_0 is the . Substituting \sigma(\omega) gives \varepsilon(\omega) = 1 + \frac{i \omega_p^2}{\omega (1/\tau - i \omega)}, revealing the plasma frequency \omega_p = \sqrt{n e^2 / \varepsilon_0 m}, which characterizes collective oscillations in metals. In the low-frequency limit (\omega \tau \ll 1), the conductivity approaches the value \sigma(\omega) \approx \sigma_0, and the response resembles steady-state conduction with minimal inertial effects. At high frequencies (\omega \tau \gg 1), the term -i \omega \tau dominates, making \sigma(\omega) \approx i n e^2 / (m \omega); here, electrons cannot follow the rapid field oscillations due to , leading to reduced and high reflectivity in metals as the dielectric function becomes negative for \omega < \omega_p. These features explain key of metals, such as the high reflectivity at and visible frequencies, where the model predicts a characteristic "Drude tail" in reflectivity spectra—a gradual decrease in reflectivity at higher frequencies approaching the plasma edge, beyond which the material becomes more transparent. This tail arises from the frequency dependence of \varepsilon(\omega) and is observed in spectra of simple metals like sodium and silver.

Thermal and Thermoelectric Properties

Thermal Conductivity Mechanism

In the Drude model, thermal conductivity arises from the transport of kinetic energy by free conduction electrons in response to a temperature gradient. Electrons in the hotter region possess higher average kinetic energies due to the local temperature and diffuse toward the cooler region between scattering events, establishing a net heat flux from hot to cold. This process mirrors the drift of electrons under an electric field for electrical conduction, but here the driving force is the spatial variation in thermal energy rather than charge separation. The model assumes that electrons behave as a classical gas, with collisions randomizing their velocities and the relaxation time τ governing both momentum and energy transfer. The heat current density \mathbf{J}_Q is derived from the flux of electron kinetic energy across a plane perpendicular to the temperature gradient \nabla T. Considering electrons crossing a unit area with mean free path λ and root-mean-square speed v_{\rms} = \sqrt{3 k_B T / m}, where k_B is Boltzmann's constant and m is the electron mass, the net energy transport yields Fourier's law \mathbf{J}_Q = -\kappa \nabla T. The thermal conductivity κ is then given by \kappa = \frac{1}{3} n v_{\rms} \lambda c_v, where n is the electron density, λ = v_{\rms} τ is the mean free path, and c_v = \frac{3}{2} k_B is the specific heat per electron following the Dulong-Petit law for a classical monatomic gas. This expression captures the electron contribution to heat conduction, neglecting lattice vibrations (phonons) as the dominant mechanism in metals at room temperature. Substituting the expressions for v_{\rms} and λ into the formula for κ, and assuming classical equipartition of energy, simplifies to \kappa = \frac{3}{2} \frac{n k_B^2 T \tau}{m}. However, to align with experimental observations in metals, where the classical specific heat overestimates the effective electron contribution at low temperatures, the model is often refined using Fermi-Dirac statistics in a semiclassical approximation, yielding \kappa = \frac{\pi^2}{3} \frac{n k_B^2 T \tau}{m}. This form treats the electrons as degenerate but retains the Drude scattering dynamics. The assumption of a single relaxation time τ for both energy and momentum transport is crucial here, linking thermal and electrical properties. A key prediction of the Drude framework is the Wiedemann-Franz law, which relates thermal and electrical conductivities through the ratio \frac{\kappa}{\sigma T} = L_0, where σ = \frac{n e^2 \tau}{m} is the electrical conductivity and e is the electron charge. Using the classical form, L_0 = \frac{3}{2} \left( \frac{k_B}{e} \right)^2 \approx 1.11 \times 10^{-8} W Ω K⁻², but the Fermi-adjusted expression gives L_0 = \frac{\pi^2}{3} \left( \frac{k_B}{e} \right)^2 \approx 2.45 \times 10^{-8} W Ω K⁻², known as the Lorenz number, which matches measurements for many metals. This law emerges directly from the shared τ and electron parameters, assuming isothermal electrical conditions and no phonon heat transport. Lorentz refined Drude's original derivation in 1905 by applying the Boltzmann transport equation, correcting a factor-of-two error in the mean free path and confirming the classical limit.

Thermopower and Seebeck Effect

The thermopower, also known as the S, quantifies the thermoelectric voltage generated across a material due to a and is defined as S = -\frac{[\Delta V](/page/Delta-v)}{\Delta T}, where \Delta V is the difference between hot and cold junctions separated by \Delta T. In the Drude model, this effect originates from the diffusive transport of charge carriers: electrons at the hotter end possess higher and thus greater average speeds, leading to a net flux toward the colder end. This diffusion creates a charge imbalance, with excess electrons accumulating at the cold junction, which in turn induces an electric field that opposes further net carrier flow in the steady state. Within the classical Drude framework, the Seebeck coefficient is derived by balancing the average electron drift velocity induced by the temperature gradient against that from the resulting electric field. The temperature-gradient contribution to the mean velocity is \mathbf{v}_Q = -\frac{\tau}{6} \frac{d \langle v^2 \rangle}{dT} \nabla T, where \tau is the relaxation time and \langle v^2 \rangle is the mean-square speed, while the field-induced velocity is \mathbf{v}_E = -\frac{e \tau}{m} \mathbf{E}. Setting the total velocity to zero for zero net current yields S = \frac{\mathbf{E}}{\nabla T} = -\frac{k_B}{2e}, assuming the classical specific heat per electron c_v = \frac{3}{2} k_B and deriving from the energy flux associated with carrier transport. This classical approximation can also be interpreted through entropy transport, where S \approx -\frac{k_B}{e} \frac{k_B T}{E_F}, reflecting the reduced effective entropy carried by carriers in a degenerate electron gas, though the Drude model originally treats electrons as a classical gas. A more detailed treatment in kinetic theory, which the Drude model simplifies, expresses the as S = -\frac{1}{eT} \frac{\int (\varepsilon - \mu) \sigma(\varepsilon) \left( -\frac{\partial f}{\partial \varepsilon} \right) d\varepsilon}{\int \sigma(\varepsilon) \left( -\frac{\partial f}{\partial \varepsilon} \right) d\varepsilon}, where \varepsilon is the , \mu is the , \sigma(\varepsilon) is the energy-dependent , and f is the . In the Drude approximation, assuming a Maxwellian distribution and energy-independent scattering time \tau, this integral reduces to the classical form S = -\frac{k_B}{2e} \approx -43 \, \mu\mathrm{V/K}, independent of temperature. For metals, incorporating partial degeneracy leads to a refined Drude-like expression S = -\frac{\pi^2 k_B^2 T}{3 e E_F}, where E_F is the , emphasizing the role of states near the . The Drude model predicts a negative Seebeck coefficient for electron conductors, consistent with the negative charge of carriers, with a magnitude on the order of k_B / e \approx 86 \, \mu\mathrm{V/K} scaled by factors near unity in the classical limit, or reduced to 10--100 \mu\mathrm{V/K} when accounting for degeneracy. However, the classical prediction overestimates the value by about two orders of magnitude at room temperature compared to typical metallic values of a few \mu\mathrm{V/K}, due to the neglect of Fermi-Dirac statistics and energy-dependent scattering. The Seebeck effect is thermodynamically linked to the Peltier effect through Kelvin relations, which state that the Peltier coefficient \Pi, representing heat transported per unit charge current, satisfies \Pi = S T, ensuring consistency in the coupled transport of charge and heat.

Model Validity and Limitations

Experimental Agreements and Discrepancies

The Drude model achieves notable success in predicting the order-of-magnitude value of electrical conductivity \sigma in metals, with experimental measurements for copper yielding \sigma \approx 6 \times 10^7 S/m at room temperature, aligning closely with the model's estimate using electron density n \approx 8.5 \times 10^{28} m^{-3} and relaxation time \tau \approx 2.5 \times 10^{-14} s. However, the classical Drude model predicts an incorrect \rho \propto \sqrt{T} temperature dependence for resistivity from phonon scattering, rather than the observed linear \rho \propto T at high temperatures. At low temperatures, the model fails to predict the observed \rho \propto T^5 dependence due to reduced phonon scattering, instead assuming classical behavior. Furthermore, the model explains the Hall effect through the coefficient R_H = -1/(n e), which matches experimental data for simple metals like alkali metals, confirming the negative charge of carriers and providing n values consistent with one conduction electron per atom; for instance, in sodium, theoretical and measured R_H \approx -2.5 \times 10^{-10} m³/C show strong agreement at low fields and room temperature. Despite these strengths, the Drude model encounters significant discrepancies in thermal properties, particularly specific heat. The model predicts an electronic contribution C_v = \frac{3}{2} n k_B, comparable to the term and violating the Dulong-Petit law, which experiments confirm applies only to ionic vibrations with total C_v \approx 3 N k_B per of atoms, while the observed electronic share is much smaller (\ll n k_B) at due to quantum degeneracy effects. The assumption of temperature-independent electron density n in the Drude model overlooks band structure influences, leading to overpredictions of carrier mobility at low temperatures, where experiments reveal or decreases due to reduced not captured by classical statistics. In , the model agrees well with the Drude tail in the infrared regime, reproducing high reflectivity via the Hagen-Rubens relation R \approx 1 - 2 \sqrt{\frac{m \omega}{2 \pi n e^2 \tau}} for metals like silver, where experimental spectra show near-unity below the plasma frequency \omega_p \approx 10^{15} rad/s. However, it fails in the , underestimating absorption because it neglects interband transitions between , which experiments attribute to quantum band structure effects peaking at energies around 3-5 eV in noble metals. Quantitatively, the n derived from roughly matches atomic valency (e.g., one per in monovalent metals), supporting the free-electron picture. Yet, the classical Drude model predicts a S \approx 0, which is much smaller than experimental values (several \muV/K in metals like ) and often the wrong sign, as non-zero S requires energy-dependent not captured classically.

Relation to Modern Theories

The Drude model served as a foundational classical framework that was significantly refined by in the late , particularly through Arnold Sommerfeld's incorporation of Fermi-Dirac statistics to describe the gas in metals. In this 1927 extension, Sommerfeld replaced the classical Maxwell-Boltzmann distribution with quantum statistics, which corrected the model's erroneous prediction of the electronic —classically constant at (3/2) n k_B but observed to vanish at low temperatures—yielding instead C_v \propto T for T \ll T_F, where T_F is the , while preserving the electrical conductivity \sigma \approx n e^2 \tau / m near room temperature due to the dominance of states near the . This quantum refinement established the Drude approach as the high-temperature limit of the model, where classical equipartition holds above T_F. In modern , the Drude model's assumption of electrons is extended by introducing the effective mass m^* to account for interactions with the crystal , allowing the formula \sigma = n e^2 \tau / m^* to describe structure effects in semiconductors and metals without fully resolving the periodic potential. Bloch's 1928 quantum mechanical treatment further advanced this by demonstrating that electrons in a periodic potential form Bloch waves, leading to energy rather than the Drude picture of unrestricted particles; this laid the groundwork for the , where weak periodic potentials perturb the states to produce gaps. Despite these quantum developments, the formula persists empirically in contemporary applications, such as modeling the frequency-dependent relaxation time \tau(\omega) for in plasmonics, where it describes the dielectric function \epsilon(\omega) = 1 - \omega_p^2 / (\omega(\omega + i/\tau)) for metals like in nanostructures. In disordered systems, the model remains a core for semiclassical transport calculations, fitting measured mobilities. For stronger electron correlations beyond the and approximations, the introduces on-site repulsion U to capture Mott insulation and other phenomena, yet retains the Drude-like mean-field core for weakly interacting regimes.

References

  1. [1]
    None
    ### Summary of the Drude Model
  2. [2]
    [PDF] Ohm's Law - UT Physics
    called the Drude–Lorentz model, or simply the Drude model, — a metal consists of motionless ions arranged in a crystalline ...
  3. [3]
    A history of the relation between fluctuation and dissipation
    Sep 22, 2023 · Riecke and Drude both relied on acrobatic kinetic-theoretical considerations in which all the electric particles are assumed to have the same ...
  4. [4]
    October 1897: The Discovery of the Electron
    Oct 1, 2000 · Thomson boiled down the findings of his 1897 experiments into three primary hypotheses: (1) Cathode rays are charged particles, which he called ...
  5. [5]
    [PDF] Physics 2415 Lecture 11: Microscopic Theory of Electric Current
    Experimentally (and theoretically) the electron mean free path is at least an order of magnitude more than Drude's model suggests, yet the mean free time ...
  6. [6]
    [PDF] The motion of electrons in metallic bodies. I - KNAW
    H.A. Lorentz, The motion of electrons in metallic bodies I, in: KNAW, Proceedings, 7, 1904-1905, Amsterdam, 1905, pp. 438-453. This PDF was made on 24 ...
  7. [7]
    [PDF] 2 Chapter 1 The Drude Theory of Metals - SIUC Physics WWW2
    Collisions in the Drude model, as in kinetic theory, are instantaneous events that abruptly alter the velocity of an electron. Drude attributed them to the ...Missing: explanation authoritative
  8. [8]
    [PDF] Handout 1 Drude Model for Metals - Cornell University
    1) Metals have a large density of “free electrons” that can move about freely from atom to atom (“sea of electrons”). 2) The electrons move according to.
  9. [9]
  10. [10]
    [PDF] τ τ ε - bingweb - Binghamton University
    Feb 3, 2020 · The theory on which Bohr based his study was the Lorentz-Drude model, according to which metals were depicted as gases of electrons moving ...
  11. [11]
    [PDF] Lecture1- Drude Model - IISc Physics
    It is a pre-quantum mechanical semi-classical model – still roughly applicable for simple alkaline metals. Applications of Drude model: DC electrical ...Missing: authoritative | Show results with:authoritative
  12. [12]
    Free Electron Gas and Drude Models – Joseph Henry Project
    The free electron gas model assumptions are used in looking at the movement of electrons: ion-electron and electron-electron interactions are neglected; ...
  13. [13]
    [PDF] Drude Theory of Metals
    The Drude Theory of Metals. J.J. Thomson. P. Drude. 1897. 1900 discovery of theory of metals the electron. Classical kinetic therry of electron gas.
  14. [14]
    The Drude Model - SciELO
    Oct 7, 2024 · In 1900, Paul Drude made a groundbreaking advancement with a model to describe the electrical conductivity of metals.Abstract · Text · References
  15. [15]
    [PDF] Prof.P. Ravindran, Drude-Lorenz Free Electron Theory
    average rms speed so at room temp. Page 14. P.Ravindran, PHY075- Condensed Matter Physics, Spring 2013 : Drude-Lorenz Free Electron Theory. Drude's classical ...Missing: citation | Show results with:citation
  16. [16]
    Drude model - Open Solid State Notes
    The Drude model is based on the kinetic theory of gases. In this exercise, we try to explain the temperature dependence of the resistivity using this theory.Missing: authoritative | Show results with:authoritative
  17. [17]
    [PDF] Drude, P., 1900, Annalen der Physik 1, 566
    Zur Elektronentheorie der Metalle; von P. Drude. I. Teil. Dass die Elektricitätsleitung der Metalle ihrem Wesen nach nicht allzu verschieden von der der ...Missing: paper | Show results with:paper
  18. [18]
    [PDF] Lorentz and Drude Models - EMPossible
    When describing metals, it is often more meaningful to put the equation in terms of the “mean collision rate” τ. This is also called the momentum scattering ...<|control11|><|separator|>
  19. [19]
    [PDF] the theory of electrons
    want to understand the way in which electric and magnetic properties depend on the temperature, the density, the chemical constitution or.
  20. [20]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    terms of a simple classical conductivity model, called the Drude model. This model is based on the classical equations of motion of an electron in an ...
  21. [21]
    [PDF] Lecture 2 Drude Theory of Thermal conductivity
    In Drude's model, thermal conductivity is related to electron thermalization at collisions, where electrons carry heat, and the thermalized heat current is ...
  22. [22]
    [PDF] The Drude Theory of Metals - Moodle@Units
    Dec 19, 2016 · The equation of motion reads dp(t) dt. = − p(t) τ. + f (t) f ... Q: thermopower. It cancels the effect of ∇T on v2 net flow of heat ...
  23. [23]
    Quantum Theory of Thermoelectric Power (Seebeck Coefficient)
    ... derivation of. Eq. (1.4) is given in Appendix. Setting cVequal to 3nkB/2, we obtain the classical formula for. thermopower: Sclassical =−kB. 2e=−0.43 ×10− ...
  24. [24]
    [PDF] Boltzmann Transport - Physics Courses
    (1.41). The peak at ω = 0 is known as the Drude peak. Here's an elementary derivation of this result. Let p(t) be the momentum of an electron, and solve the ...
  25. [25]
    [PDF] Solid State Physics - Kevin Zhou
    Next, we reflect on the successes and failures of these models. • Some quantities calculated in the Drude model, such as RH, are on the right order of magnitude.
  26. [26]
    [PDF] An Introduction to the Quantum Hall Effect - UBC Physics
    Note that Group 1 Metals, for which the. Drude Model applies well at room temperature, have a very good experimental agreement with theory [3]. Model ...
  27. [27]
    Zur Elektronentheorie der Metalle | The Science of Nature
    About this article. Cite this article. Sommerfeld, A. Zur Elektronentheorie der Metalle. Naturwissenschaften 15, 825–832 (1927). https://doi.org/10.1007 ...Missing: Arnold | Show results with:Arnold
  28. [28]
    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.
  29. [29]
    Surface plasmons in Drude metals - ScienceDirect.com
    We present here a detailed derivation of the dispersion relations for surface plasmons in classical Drude metals.
  30. [30]
    Modified Drude model for small gold nanoparticles surface plasmon ...
    Apr 16, 2020 · There are a lot of theoretical and experimental studies where the modified Drude model is used for showing size dependence of MNPs permittivity ...