Fact-checked by Grok 2 weeks ago

Free electron

In physics, a free electron is an electron that is not bound to an , , or , allowing it to move freely within a , , or . These electrons can originate from shells in solids, in gases or plasmas, or processes, and play key roles in electrical conduction, plasma dynamics, and various physical phenomena. In metals, free electrons—typically electrons detached from —form a gas-like with minimal interaction with the ionic , enabling efficient current flow in response to , as seen in conductors like or silver. The concept is central to the , the simplest framework for metallic properties, initially developed classically by Paul Drude and (1900–1905) using kinetic theory to explain electrical and thermal conductivity. The classical model treated electrons as particles in a , successfully predicting resistivity from collisions but failing to explain , as it significantly overestimated the electronic contribution. In 1927, provided a quantum refinement, incorporating wave-particle duality, the , and Fermi-Dirac statistics to describe electrons as a degenerate up to the (typically 2–10 eV in metals). Key assumptions neglect electron interactions and potential, with plane-wave wavefunctions and energy E = \frac{\hbar^2 k^2}{2m}. This model explains low-temperature electronic specific heat's linear temperature dependence and temperature-independent carrier density in metals, though it ignores band gaps and does not apply to insulators or semiconductors. It remains foundational in , aiding computations of , Fermi surfaces, and transport, and underpins advanced theories like the .

Definition and Fundamentals

Core Definition

A free electron is an electron that is not bound to an or , enabling it to move independently in response to external electric or . This unbound state distinguishes it from electrons in atomic orbitals, allowing participation in phenomena such as electrical conduction and particle . Free electrons occur in diverse physical contexts, including environments where they form directed beams in devices like electron microscopes and particle accelerators. In plasmas, they arise from processes, contributing to the material's responsiveness to electromagnetic fields. Within semiconductors, free electrons populate the conduction band, serving as charge carriers when excited across the band gap. In metals, valence electrons behave as conduction electrons, delocalized across the to facilitate high electrical conductivity. The free electron carries a negative elementary charge of -1.602 \times 10^{-19} C and has a rest mass of $9.109 \times 10^{-31} kg. As a fermion, it possesses an intrinsic spin quantum number of $1/2. For non-relativistic speeds, the kinetic energy E of a free electron is given by E = \frac{p^2}{2m}, where p is the electron's momentum and m is its mass.

Bound vs. Free Electrons

Bound electrons are those confined to specific atomic orbitals around individual atoms or ions, where they participate in chemical bonding and exhibit discrete energy levels. In isolated atoms or insulators, these electrons occupy quantized orbitals, such as the 1s orbital in , leading to well-defined energy spectra that determine atomic properties like emission lines. This confinement arises from the strong electrostatic attraction to the , restricting electron motion primarily to localized regions and contributing to the stability of molecules through orbital overlap in covalent or ionic bonds. In solids like insulators, bound electrons fill bands completely, preventing net charge transport due to the large energy gap to higher states. In contrast, free electrons are delocalized within a , such as the conduction electrons in metals, where they form continuous bands allowing high under applied fields. Unlike the quantized levels of bound electrons, free electrons experience nearly parabolic relations near band minima, resembling plane waves that propagate through the with minimal in ideal cases. This results in significantly higher for free electrons, enabling electrical conductivity, as they can respond collectively to external potentials without being tethered to specific atoms. In metals, valence electrons detach from atomic sites to occupy these extended states, forming a gas-like ensemble that underlies metallic properties. The transition from bound to free electrons occurs through ionization processes that supply sufficient energy to overcome the binding threshold, known as the . For isolated atoms like , this energy is 13.59844 , the minimum required to eject the electron from the to the . In solids, bound electrons can become free via thermal excitation, where heat provides energy to promote electrons across the band gap in semiconductors; the , in which photons above the liberate surface electrons; or field emission, where strong electric fields enable quantum tunneling from bound states. These processes are threshold-dependent, with the ionization energy setting the scale—for instance, around 5-10 for typical metals—beyond which electrons gain and contribute to conduction. In solid-state materials with free electrons, such as metals, the distribution of these electrons follows the , filling available states up to the , or , which represents the highest occupied energy at temperature. The lies within or at the top of the conduction band, determining which electrons can participate in transport; those near this energy have the longest mean free paths and highest velocities, dominating electrical and thermal properties. This concept highlights the distinction from bound systems, where no such collective Fermi sea exists due to the absence of overlapping bands.

Historical Context

Early Observations

The concept of free electrons emerged from pioneering experiments in the late that revealed the emission and behavior of negatively charged particles from matter. In 1897, J.J. Thomson conducted definitive experiments using cathode ray tubes, where he observed streams of particles emanating from the in a . By applying electric and magnetic fields to deflect these rays, Thomson measured their charge-to-mass ratio, finding it to be approximately $1.76 \times 10^{11} C/kg, a value thousands of times greater than that of ions, indicating lightweight, universal particles he termed "corpuscles" (later known as electrons). These particles were identical regardless of the cathode material, suggesting they were fundamental constituents of atoms capable of existing freely outside atomic structure. Building on such observations, the provided further evidence of free electron emission triggered by light. In 1887, , while experimenting with electromagnetic waves using a apparatus, noticed that light incident on a metal surface facilitated the discharge of sparks across a nearby gap, enhancing conductivity in a way that visible light could not. This puzzling phenomenon indicated that light of sufficient frequency could liberate electrons from metals, though Hertz offered no theoretical explanation. In 1905, provided the quantum interpretation, proposing that light consists of discrete energy packets (photons) that impart enough energy to eject electrons if the photon's frequency exceeds a metal-specific threshold, resolving the effect's dependence on frequency rather than intensity. For this seminal explanation, Einstein received the 1921 . Another key observation involved thermal emission of electrons, known as the Edison effect. In 1883, , while developing incandescent lamps, observed that a heated filament in a caused current to flow to a nearby cold electrode, attributing it to particles streaming from the hot wire without understanding the mechanism. This effect demonstrated that could free electrons from a metal surface. In 1901, Owen W. Richardson formulated a quantitative law describing this , expressing the current density D as D = A T^2 \exp(-W / kT), where A is a material constant, T is the temperature in , W is the (minimum energy to liberate an electron), and k is Boltzmann's constant. Richardson's work, which earned him the 1928 , established as a reliable process for generating free electrons.

Theoretical Advancements

In 1905, advanced the classical understanding of free electrons by formulating an electron theory of metals, modeling them as a gas of charged particles moving freely among fixed positive ions, akin to the . This approach, which refined Paul Drude's earlier ideas by incorporating Maxwell-Boltzmann statistics, treated s as classical point charges subject to random collisions with lattice ions, thereby providing a theoretical basis for metallic electrical conductivity and the . Lorentz's model predicted that electron velocities follow a thermal distribution, with mean free paths determining transport properties, though it failed to account for specific heat anomalies in metals. The transition to quantum paradigms began with Niels Bohr's atomic model, which introduced quantized for electrons in stable orbits around the , preventing classical radiative collapse and enabling discrete spectral lines. While primarily describing bound electrons, Bohr's quantization rule—L = nℏ, where n is an and ℏ = h/2π—provided initial quantum insights applicable to electrons transitioning to free states during or , influencing later models of electron emission from metals. A pivotal shift occurred in 1924 when hypothesized the wave-particle duality of matter, proposing that electrons, like photons, exhibit wave properties with λ = h/p, where h is Planck's constant and p is . This idea, extending wave-particle duality from light quanta to massive particles, suggested that free electrons could undergo and , a prediction experimentally verified in 1927 by and Lester Germer through from a , confirming de Broglie's relation for slow electrons. Building on de Broglie's hypothesis, developed wave mechanics in 1926, applying it to free particles via the time-dependent iℏ ∂ψ/∂t = - (ℏ²/2m) ∂²ψ/∂x², which yields solutions of the form \psi(x,t) = A \exp\left[i\left(kx - \omega t\right)\right], where k = 2π/λ is the wave number, ω = E/ℏ the , and A the normalization . This formulation described free electrons as propagating waves with definite p = ℏk and E = ℏω = p²/2m, unifying particle and wave descriptions without ad hoc quantization. Concurrently, in 1925, introduced the exclusion principle, asserting that no two identical fermions, such as electrons, can occupy the same simultaneously, characterized by a unique set of quantum numbers. For free electron systems in solids, this principle implies degeneracy at low temperatures, where electrons fill successive momentum states up to the , preventing collapse into the lowest energy state and enabling stable fermionic gases with Pauli blocking effects on scattering.

Physical Properties

Kinematics and Dynamics

The kinematics and dynamics of free electrons are described classically using Newtonian mechanics, treating the electron as a of charge -e and mass m_e. Under the influence of an external \mathbf{E}, the electron experiences a force \mathbf{F} = -e \mathbf{E}, resulting in an \mathbf{a} = - (e / m_e) \mathbf{E}. For a constant uniform field starting from rest, the velocity acquired after time t is \mathbf{v} = - (e \mathbf{E} / m_e) t, leading to a linear drift motion along the field direction. This underlies phenomena like electron drift in conductors, where the average velocity is proportional to the field strength. In the presence of a \mathbf{B}, the \mathbf{F} = -e (\mathbf{v} \times \mathbf{B}) dominates when \mathbf{E} = 0, causing the to undergo helical or if the has components parallel and to \mathbf{B}. For motion in a uniform \mathbf{B}, the is a circle with cyclotron radius r = m_e v_\perp / (e B), where v_\perp is the perpendicular speed, and the cyclotron frequency is \omega_c = e B / m_e. This frequency is independent of in the non-relativistic limit and governs oscillatory behavior in devices like or magnetrons. Scattering events interrupt this ideal motion, introducing a relaxation time \tau as the average interval between collisions with lattice ions, impurities, or phonons in solids, or atoms in gases. The mean free path \lambda, the average distance traveled between collisions, is given by \lambda = v \tau, where v is the electron speed (often the thermal or drift velocity). Electron mobility \mu, measuring the ease of motion under an electric field, follows from the Drude model as \mu = e \tau / m_e, linking scattering directly to transport efficiency; typical values in metals range from 10 to 100 cm²/V·s at room temperature. At high speeds approaching the c, relativistic corrections become necessary, modifying the effective mass via the \gamma = 1 / \sqrt{1 - v^2/c^2}. The is \mathbf{p} = \gamma m_e \mathbf{v}, and acceleration in fields adjusts accordingly, with the cyclotron frequency reduced to \omega_c = e B / (\gamma m_e); such effects are relevant for free electrons in particle accelerators or intense fields but negligible in typical condensed matter contexts. Collisions often involve drag forces that dissipate energy, particularly through inelastic scattering where electrons lose kinetic energy to lattice vibrations (phonons) in solids or excitation of atomic levels in gases. In solids, electron-phonon interactions dominate, with energy loss rates scaling with temperature and leading to a frictional drag that limits steady-state speeds; in dilute gases, collisions with neutrals cause similar deceleration via momentum transfer. These processes ensure that free electron motion reaches a balance between acceleration and dissipation, as captured in the relaxation time approximation.

Quantum Mechanical Description

In quantum mechanics, the behavior of a free electron, unbound by any potential, is described by the time-independent Schrödinger equation for a particle in one dimension: -\frac{\hbar^2}{2m} \frac{d^2\psi}{dx^2} = E \psi where \hbar is the reduced Planck's constant, m is the electron mass, E is the energy eigenvalue, and \psi(x) is the wave function. The general solutions to this equation are plane waves of the form \psi(x) = A e^{ikx} + B e^{-ikx}, where k = \sqrt{2mE}/\hbar is the wave number, representing propagating waves with definite momentum p = \hbar k. These solutions indicate that a free electron cannot be localized in space without a corresponding spread in momentum, as superpositions of plane waves form wave packets that disperse over time. The Heisenberg uncertainty principle further constrains the quantum description of free electrons, stating that the product of the uncertainties in and satisfies \Delta x \Delta p \geq \hbar/2. This relation arises from the non-commuting nature of the and operators in , implying that an electron cannot have both its and precisely determined simultaneously. For a free electron, attempting to localize it within a small region \Delta x requires a with a broad range of momenta \Delta p, leading to rapid spreading of the packet and preventing long-term confinement without external potentials. This probabilistic nature underscores the wave-like character of electrons, distinguishing quantum kinematics from classical trajectories. Free electrons also possess intrinsic spin angular momentum S = \hbar/2, which gives rise to a magnetic moment \mu = -g \mu_B \mathbf{S} / \hbar, where \mu_B = e \hbar / (2m) is the and g \approx 2 is the electron g-factor. The was proposed to explain in atomic spectra, with the g-factor of exactly 2 emerging naturally from the relativistic for the . In an external B, this moment interacts via the H_Z = -\mu \cdot B, splitting the spin states and producing the , where the energy levels shift by \pm g \mu_B B / 2. This splitting is observable in and confirms the 's role as a microscopic . For a three-dimensional ensemble of non-interacting electrons, the g(E), which gives the number of available states per unit interval, is derived from the phase space volume in : g(E) = \frac{V}{2\pi^2} \left( \frac{2m}{\hbar^2} \right)^{3/2} E^{1/2} where V is the system volume. This quadratic dependence arises from counting the number of plane-wave states within spherical shells in momentum space, accounting for degeneracy. The formula, central to statistical treatments of free electron systems, highlights how higher energies accommodate more states, influencing properties like thermal response. Quantum tunneling exemplifies the ability of free-like electrons to penetrate classical barriers. For an electron incident on a potential barrier higher than its E, the wave function decays exponentially inside the barrier but has a non-zero probability of , allowing the particle to emerge on the other side. This effect, first applied to processes, enables electrons to traverse thin insulating layers or potential steps that would be impenetrable classically, with probability T \approx e^{-2\kappa L} where \kappa = \sqrt{2m(V_0 - [E](/page/E!))}/\hbar and L is the barrier width.

Models in Condensed Matter

Classical Drude Model

The classical treats free electrons in metals as a gas of non-interacting classical particles undergoing random thermal motion, with velocities following a Maxwell-Boltzmann distribution, and subject to occasional collisions with fixed ion cores in the that randomize their but do not change their speed. These events occur on average every relaxation time \tau, assumed to be independent of electron and , allowing electrons to accelerate under an applied between collisions. The model neglects electron-electron interactions and quantum effects, viewing the electron gas as and the as a uniform background of positive charge. In the presence of an \mathbf{E}, the average \mathbf{v}_d of electrons reaches a where the acceleration from the field balances the deceleration from , given by m \frac{d\mathbf{v}_d}{dt} = -e \mathbf{E} - \frac{m \mathbf{v}_d}{\tau} = 0, yielding \mathbf{v}_d = -\frac{e \tau}{m} \mathbf{E}. The resulting \mathbf{j} = -n e \mathbf{v}_d leads to the electrical conductivity \sigma = \frac{n e^2 \tau}{m}, where n is the , e the magnitude, and m the ; this expression explains \mathbf{j} = \sigma \mathbf{E} and predicts that conductivity decreases with increasing temperature due to shorter \tau from enhanced vibrations. For the Hall effect, consider a current \mathbf{j} along the x-direction in a magnetic field \mathbf{B} along the z-direction, inducing a Lorentz force that deflects electrons toward one side of the sample, creating a transverse Hall field \mathbf{E}_y until the electric and magnetic forces balance. The steady-state Drude equation becomes $0 = - \frac{m \mathbf{v}}{\tau} - e (\mathbf{E} + \mathbf{v} \times \mathbf{B}), with components v_x = -\frac{e \tau}{m} E_x, v_y = -\frac{e \tau}{m} (E_y + v_x B_z). Since j_y = 0 in steady state, solving yields E_y = -\frac{j_x B_z}{n e}, so the Hall resistivity \rho_{xy} = E_y / j_x = - \frac{B_z}{n e} and Hall coefficient R_H = \frac{E_y}{j_x B_z} = -\frac{1}{n e}, which depends only on carrier density and sign, enabling experimental determination of n. The model extends to thermal transport by treating electrons as carriers of both charge and heat, with thermal conductivity \kappa arising from the diffusion of thermal energy via random motion between scatterings. Using classical specific heat per electron c_v = \frac{3}{2} k_B and mean square speed \langle v^2 \rangle = \frac{3 k_B T}{m}, the electron thermal conductivity is \kappa = \frac{1}{3} n c_v \langle v^2 \rangle \tau = \frac{3}{2} \frac{n k_B^2 T \tau}{m} . This implies a Lorenz number L = \frac{\kappa}{\sigma T} = \frac{3}{2} \left( \frac{k_B}{e} \right)^2 in the classical limit, but experimental observations follow the Wiedemann-Franz law \frac{\kappa}{\sigma} = \frac{\pi^2}{3} \left( \frac{k_B}{e} \right)^2 T, which the Drude model approximates but underestimates by a factor of approximately 2.2 due to its classical statistics. Despite its successes, the Drude model has key limitations: it fails at low temperatures where quantum effects like Pauli exclusion become prominent, leading to much longer mean free paths than classically predicted, and it incorrectly forecasts the electronic specific heat as the classical Dulong-Petit value C = \frac{3}{2} N k_B (independent of temperature), whereas experiments show a linear T dependence C = \gamma T from states near the Fermi level. These shortcomings are addressed in quantum refinements, such as the free electron gas model.

Quantum Free Electron Gas

The quantum free electron gas model extends the classical treatment of conduction electrons by incorporating quantum mechanical principles, particularly the and Fermi-Dirac statistics, to describe the behavior of electrons in metals as a degenerate at low temperatures. In this model, electrons occupy energy states up to the , leading to distinct thermodynamic and transport properties that align more closely with experimental observations in solids than classical predictions. The occupation of energy states in the quantum free electron gas is governed by the Fermi-Dirac distribution function, which gives the average number of electrons in a state of energy E as
f(E) = \frac{1}{\exp\left(\frac{E - \mu}{k_B T}\right) + 1},
where \mu is the chemical potential, k_B is Boltzmann's constant, and T is the temperature. At absolute zero temperature (T = 0), the distribution becomes a step function, with all states below the Fermi energy E_F fully occupied and those above empty, and \mu \approx E_F. The Fermi energy itself, representing the highest occupied energy level at T = 0, is derived from the electron density n for a three-dimensional degenerate electron gas as
E_F = \frac{\hbar^2}{2m} (3 \pi^2 n)^{2/3},
where \hbar is the reduced Planck's constant and m is the electron mass; this yields typical values of several electron volts for metals, indicating high degeneracy even at room temperature.
One key consequence of this quantum description is the electronic specific heat, which deviates markedly from the classical Dulong-Petit law. At low temperatures, the specific heat arises from excitations near the and takes the linear form
C_V = \frac{\pi^2}{2} n k_B \left( \frac{k_B T}{E_F} \right),
providing a T-linear contribution that dominates over contributions in metals at cryogenic temperatures and explains observed heat capacities without invoking classical equipartition.
The serves as the foundational approximation for more advanced band structure theories, such as the , where weak periodic potentials from the crystal lattice perturb the parabolic , opening energy gaps at boundaries while retaining much of the free-electron character for weakly bound electrons. Additionally, excitations in the electron gas manifest as plasma oscillations at the plasma frequency
\omega_p = \sqrt{\frac{n e^2}{\epsilon_0 m}},
where e is the charge and \epsilon_0 is the ; this frequency characterizes longitudinal waves and underlies phenomena like screening in metals.

Applications and Phenomena

In Metals and Conductors

In metals, free electrons serve as the primary charge carriers responsible for electrical conduction. These electrons, delocalized from their atomic orbitals, move through the under an applied , enabling high . The classical description, based on the , treats electrons as a gas of non-interacting particles subject to by lattice vibrations (phonons). The electrical resistivity \rho arises from this scattering and is given by the formula \rho = \frac{m}{n e^2 \tau}, where m is the electron mass, n is the free electron density, e is the electron charge, and \tau is the relaxation time between collisions. Resistivity increases with temperature because thermal agitation intensifies phonon scattering, reducing \tau. For example, in copper, the free electron density n \approx 8.5 \times 10^{28} m^{-3} (one per atom) contributes to its exceptionally high conductivity at room temperature, making it a standard material for electrical wiring. In superconductors, free electrons exhibit a remarkable transformation below a critical T_c, where they form Cooper pairs—bound pairs mediated by lattice phonons—leading to zero electrical resistance and perfect . This pairing allows electrons to bypass , enabling dissipationless current flow. The microscopic explanation, provided by Bardeen, , and Schrieffer in 1957 (BCS ), attributes the attraction to phonon-induced interactions overcoming repulsion. Semiconductors differ from metals in that free electrons are not inherently abundant; instead, doping introduces them into the conduction band. In n-type semiconductors, donor impurities (e.g., in ) donate electrons, creating mobile charge carriers with energies above . Unlike metals, where conduction relies on a high density of delocalized electrons with long mean free paths, semiconductors have lower carrier densities and mobilities, resulting in tunable but generally lower conductivity that increases with doping or temperature. At high frequencies and low temperatures, the anomalous skin effect modifies how free electrons interact with electromagnetic fields in metals. When the electron mean free path exceeds the classical skin depth, electrons traverse the surface layer ballistically rather than diffusively, leading to collective motion and a reduced effective for the field. This phenomenon, first analyzed by Reuter and Sondheimer, impacts applications like conductivity measurements in pure metals.

In Plasmas and Gases

In plasmas, free electrons constitute a key component of a quasi-neutral gas comprising charged particles—primarily electrons and ions—along with neutral species, where collective interactions dominate due to long-range forces. This state arises when a sufficient of atoms or molecules is ionized, allowing electrons to move freely while maintaining overall charge neutrality on scales larger than the , which quantifies the distance over which electric fields are screened by mobile charges. The is defined as \lambda_D = \sqrt{\frac{\epsilon_0 k_B T_e}{n_e e^2}}, where \epsilon_0 is the vacuum permittivity, k_B is the Boltzmann constant, T_e is the electron temperature, n_e is the electron density, and e is the elementary charge; typical values range from micrometers in dense laboratory plasmas to astronomical units in sparse astrophysical ones. Ionization processes in gases generate these free electrons through mechanisms such as high-energy collisions between particles, strong that initiate Townsend avalanches—where initial electrons accelerate and ionize neutrals, exponentially multiplying charge carriers—or by radiation, often culminating in gas discharge phenomena like glow or arc discharges. In the Townsend avalanche, the first ionization coefficient \alpha describes the number of ionizations per unit length, scaling with the strength E and gas p, typically following \alpha / p = A \exp(-B p / E) for , enabling controlled discharges in devices like neon lamps. These processes are prevalent in low-pressure environments, where mean free paths allow electrons to gain sufficient energy between collisions. In conditions, free s are harnessed as collimated beams by accelerating them via to energies from kiloelectronvolts (keV) in tubes (CRTs), where they strike phosphors to produce images, to megaelectronvolts (MeV) in particle accelerators for scientific probing of matter. These beams propagate ballistically without scattering, guided by magnetic fields to maintain focus, and find applications in medical radiotherapy, materials sterilization, and high-energy physics experiments. For instance, linear accelerators use radiofrequency cavities to boost electron velocities near the , achieving beam currents up to amperes. Recombination counterbalances ionization, with free electrons recombining with ions through radiative processes, where an electron is captured into a bound state while emitting a photon, or three-body collisions in denser plasmas. The radiative recombination rate coefficient scales approximately as \alpha_{RR} \propto Z^2 T_e^{-1/2}, decreasing with temperature but dominant in low-density, high-temperature regimes like fusion plasmas; for hydrogen-like ions, cross sections peak at low electron energies around 1 . These rates are critical for modeling plasma evolution, as they determine equilibrium ionization states. Astrophysically, free electrons abound in stellar interiors, where full at temperatures exceeding 10 million leads to opacity \kappa_{es} = 0.2 (1 + X) cm²/g (with X the mass fraction), impeding radiative transport and shaping . In the , free electrons—number densities around 0.01–1 cm⁻³—arise from and stellar winds, contributing to opacity via free-free absorption and causing that broadens radio source images through density fluctuations. These electrons also facilitate wave propagation and heating in H II regions around hot stars.

References

  1. [1]
    The Free Electron - The Atom and Elements - NDE-Ed.org
    Free electrons are valence electrons that break away from an atom when they gain sufficient energy and can be shared among neighboring atoms.
  2. [2]
    Free Electron Model of Metals – University Physics Volume 3
    The simplest model of a metal is the free electron model. This model views electrons as a gas. We first consider the simple one-dimensional case.
  3. [3]
    [PDF] Free Electron Fermi Gas (Kittel Ch. 6) - SMU Physics
    Starting Point for Understanding. Electrons in Solids. • Nature of a metal: Electrons can become. “free of the nuclei” and move between nuclei.
  4. [4]
    electron mass - CODATA Value
    The NIST Reference on Constants, Units and Uncertainty. Fundamental Physical ... electron mass $m_{\rm e}$. Numerical value, 9.109 383 7139 x 10-31 kg.
  5. [5]
    Electrons, Ions and Plasma - NASA
    A near-vacuum is necessary for any such procedure to be effective, because in ordinary air free electrons collide with molecules, lose their energy and are ...Missing: semiconductors | Show results with:semiconductors
  6. [6]
    DOE Explains...Electrons - Department of Energy
    Electrons can gain enough energy to leave atomic nuclei behind. These free electrons mix with ions to form a plasma. Electrons can separate from atoms ...
  7. [7]
    Electrical Conduction in Semiconductors - MKS Instruments
    In a metal, the valence band and the conduction band overlap, so electrons are free to move into many available and vacant energy levels. When an electrical ...
  8. [8]
    elementary charge - CODATA Value
    elementary charge $e$. Numerical value, 1.602 176 634 x 10-19 C. Standard uncertainty, (exact). Relative standard uncertainty, (exact).CODATA Value
  9. [9]
    Electron spin - HyperPhysics Concepts
    An electron spin s = 1/2 is an intrinsic property of electrons. Electrons have intrinsic angular momentum characterized by quantum number 1/2.Missing: source | Show results with:source
  10. [10]
    [PDF] The Schrödinger Equation in One Dimension
    Hence, K = p2/2m, and E represents the total mechanical energy (i.e., the sum of the kinetic and potential energies, not the relativistic mass-energy).<|control11|><|separator|>
  11. [11]
    Atomic Data for Hydrogen (H ) - Physical Measurement Laboratory
    These wavelengths are from the corresponding weighted-average energies. H I Ground State 1s 2S1/2. Ionization energy 109678.7717 cm-1 (13.598433 eV) Ref. MK00a.
  12. [12]
    [PDF] Atomic Structure & Chemical Bonding - Projects at Harvard
    The electron configuration of an atom describes how electrons in that atom are arranged into electron shells and orbitals. For example, carbon, which has six ...
  13. [13]
    [PDF] 11. Solids - UF Physics
    The basic idea is that one or more electrons of each atom in the solid are detached from their atomic location and free to move throughout the crystal as a free ...
  14. [14]
    [PDF] Energy Bands in Crystals
    Energy bands are nearly continuous energy levels for electrons in crystals, separated by band gaps, and are crucial for electron conductivity.
  15. [15]
    [PDF] Bound-Bound and Bound-Free Transitions Initial questions
    A bound-free transition is not sharply defined in energy, since in principle the ionized electron can have anywhere from zero energy (if it was barely ionized) ...
  16. [16]
    [PDF] Recombination and Ionization Measurements at the Heidelberg ...
    In direct ionization a free electron collides with an ion transferring energy to, and freeing, a bound electron. The most common indirect process is excitation ...
  17. [17]
    [PDF] Section 7: Free electron model
    A free electron model is the simplest way to represent the electronic structure of metals. Although the free electron model is a great oversimplification of ...
  18. [18]
    [PDF] Lecture 13: Metals
    A step beyond the free electron gas is the nearly free electron gas. Our treatment of the nearly- free gas was somewhat heuristic. The required calculations ...
  19. [19]
    The Discovery of the Electron (JJ Thomson) - Purdue University
    When Thomson's data are converted to SI units, the charge-to-mass ratio of the particles in the cathode-ray beam is about 108 coulomb per gram. Thomson found ...
  20. [20]
    Photoelectric Effect - Galileo
    Hertz used a piece of copper wire 1 mm thick bent into a circle of diameter 7.5 cm, with a small brass sphere on one end, and the other end of the wire was ...Hertz Finds Maxwell's Waves... · Hallwachs' Simpler Approach
  21. [21]
    Albert Einstein – Facts - NobelPrize.org
    In one of several epoch-making studies beginning in 1905, Albert Einstein explained that light consists of quanta—packets with fixed energies corresponding to ...
  22. [22]
    Einstein and The Photoelectric Effect - American Physical Society
    Jan 1, 2005 · In 1887, German physicist Heinrich Hertz noticed that shining a beam of ultraviolet light onto a metal plate could cause it to shoot sparks.
  23. [23]
    Thermionic Emission: A Comprehensive Guide - Electrical4U
    Jun 18, 2023 · This phenomenon, also called thermal electron emission or the Edison effect, was first observed by Thomas Edison in 1883. Thermionic emission ...
  24. [24]
    Nobel Prize in Physics 1928
    ### Summary of Owen Richardson's Work and Nobel Prize
  25. [25]
    Diffraction of Electrons by a Crystal of Nickel | Phys. Rev.
    Feb 3, 2025 · Davisson and Germer showed that electrons scatter from a crystal the way x rays do, proving that particles of matter can act like waves. See ...
  26. [26]
    An Undulatory Theory of the Mechanics of Atoms and Molecules
    The paper gives an account of the author's work on a new form of quantum theory ... The wave-equation and its application to the hydrogen atom. §5. The ...
  27. [27]
    [PDF] Classical Electromagnetism - Richard Fitzpatrick
    In conclusion, electric and magnetic fields look simple in the near field region (they are just. Coulomb fields, etc.), and also in the far field region ...
  28. [28]
    [PDF] SOLID STATE PHYSICS PART I Transport Properties of Solids - MIT
    ... electrons are nearly free and the weak binding approxima- tion describes these electrons quite well. The Fermi surface is nearly spherical and the band gaps ...
  29. [29]
    [PDF] Quantisierung als Eigenwertproblem
    Schrodinger hangt in eminentem MaBe von dem Zustand ab, in welchem das Atom sich gerade befindet , wahrend die erzwungenen. Schwingungen einer Membran ...
  30. [30]
    [PDF] 1.3 THE PHYSICAL CONTENT OF QUANTUM KINEMATICS AND ...
    The Franck-Hertz collision experiments allow one to base the measurement of the energy of the atom on the measurement of the energy of electrons in rectilinear ...
  31. [31]
    [PDF] The Quantum Theory of the Electron - UCSD Math
    Oct 26, 2013 · By P. A. M. DIRAC, St. John's College, Cambridge. (Communicated by R. H. Fowler, F.R.S.-Received January 2, 1928.) The new quantum mechanics, ...Missing: g- | Show results with:g-
  32. [32]
    [PDF] Zur Elektronentheorie der Metalle - Gilles Montambaux
    SOMMERFELD: Zur Elektronentheorie der Metalle. 377 weil andernfalls alle Elektronen das Metall ver- lassen und dieses seinen metallischen Charakter.
  33. [33]
    [PDF] Gamow's alpha-decay theory revisited - arXiv
    In this case quantum tunnelling plays the role of nuclear disintegration. Fig. 2 Unequal fragments in disintegration. It is reasonable to think that a naturally ...
  34. [34]
    [PDF] Physics of Semiconductors and Nanostructures - Debdeep Jena
    Apr 16, 2017 · Icing on the cake: The Weidemann-Franz Law. One of the major successes of the Drude electron model of electrical and thermal conductivity of ...
  35. [35]
    [PDF] Handout 1 Drude Model for Metals - Cornell University
    ... Drude Model for Metals. In this lecture you will learn: • Metals, insulators, and semiconductors. • Drude model for electrons in metals. • Linear response ...
  36. [36]
    Drude model - Open Solid State Notes
    The goal is to derive the Hall voltage and coefficient. As before, we consider a 2D sample in the xy plane. We push a current through the sample by ...
  37. [37]
    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.
  38. [38]
    Oscillations in Ionized Gases - PNAS
    Soc., 117, 660 (1928). OSCILLA TI&NS IN IONIZED GASES. By IRVING LANGMUIR. RESEARCHLABORATORY, GsERALu EIXCTRIc Co., SCH NsCTADY, N.Missing: original | Show results with:original
  39. [39]
    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.
  40. [40]
    Semiconductors and Doping – University Physics Volume 3
    In an n-type semiconductor, majority carriers are free electrons contributed by impurity atoms, and minority carriers are free electrons produced by thermal ...
  41. [41]
    The Anomalous Skin Effect - NASA/ADS - Astrophysics Data System
    The anomalous skin effect arises in good conductors at low temperatures and high frequencies when the electronic mean free path becomes comparable with or ...
  42. [42]
    [PDF] Heuristic derivations of basic plasma parameters - 2025 Intro Course
    Jun 27, 2016 · Analo- gously, for an ion of charge Ze and mass mi, the ion plasma frequency can be defined as: ωpi = p(Z2e2ni)/(mi 0).Missing: original | Show results with:original
  43. [43]
    [PDF] Chapter 1 Introduction
    A plasma is a gas in which an important fraction of the atoms is ionized, so that the electrons and ions are separately free. When does this ionization occur?
  44. [44]
    Townsend Avalanche - an overview | ScienceDirect Topics
    A Townsend avalanche is when an electron collides with gas particles, creating more electrons and ions, leading to a stable discharge.
  45. [45]
    Detector (gas amplification) | University of
    Apr 10, 2024 · The Townsend avalanche is an ionisation process for gases where free electrons are accelerated by an electric field, collide with gas molecules.
  46. [46]
    How Particle Accelerators Work | Department of Energy
    Jun 18, 2014 · Electromagnets steer and focus the beam of particles while it travels through the vacuum tube. Electric fields spaced around the accelerator ...
  47. [47]
  48. [48]
    Radiative Recombination Plasma Rate Coefficients for Multiply ...
    Nov 3, 2004 · The radiative recombination (RR) of an ion refers to the capture of a free electron under the emission of one or several photons [1,2,3,4].
  49. [49]
    [PDF] INTRODUCTION TO RECOMBINATION IN PLASMAS - DTIC
    Radiative recombination of an electron and a proton, for example, is just the inverse of the photoionization of the hydrogen atom. If we know the cross section ...
  50. [50]
    [PDF] Lecture 7 – Part 1 Sources of Stellar Opacity
    At temperatures higher than a few tens of million K, stellar interiors are virtually completely ionized, and electron scattering is the dominant opacity source.Missing: interstellar medium
  51. [51]
    [PDF] The Interstellar Medium - Homepages of UvA/FNWI staff
    by free electrons caused a high opacity in the medium keeping the electrons and photons in thermal equilibrium. After decoupling, the loss of free electrons ...