Fact-checked by Grok 2 weeks ago

Fermi level

The Fermi level, also known as the , is a key concept in that denotes the energy at which the probability of occupation by a fermionic particle, such as an , is exactly one-half according to the . At temperature, it represents the highest energy level occupied by electrons in a system, forming the surface of the so-called "Fermi sea" where all states below this energy are fully occupied and those above are empty, due to the . This level arises from the quantum statistical mechanics described by , which govern the behavior of indistinguishable fermions. In metals, the Fermi level typically lies within the conduction band, resulting in a high at this energy and enabling significant electrical conductivity even at low temperatures, as electrons near the Fermi level can readily respond to . For instance, in , the is approximately 7 , corresponding to a Fermi temperature of about 82,000 K, far exceeding , which underscores the degenerate nature of the gas in metals. The Fermi level also influences thermal properties, such as the low specific heat of metals at low temperatures, since only electrons within ~ of the Fermi level (where is Boltzmann's constant and is ) can participate in thermal excitations. In semiconductors, the position of the Fermi level relative to the band edges is crucial for determining concentrations and behavior. In intrinsic semiconductors, it resides near the midpoint of the , where and concentrations are equal and thermally generated, as given by n_i = (N_c N_v)^{1/2} exp(-E_g / 2kT), with E_g being the . Doping shifts this level: in n-type materials, donor impurities raise the Fermi level closer to the conduction edge, increasing concentration via E_F ≈ E_i + kT ln(N_D / n_i), while in p-type materials, acceptors lower it toward the , boosting concentration as E_F ≈ E_i - kT ln(N_A / n_i). These shifts enable the tailoring of electrical properties in like transistors and diodes. The Fermi level's role extends to interfaces and non-equilibrium conditions, where quasi-Fermi levels describe separate and distributions under applied or illumination, essential for understanding photovoltaic and optoelectronic phenomena. Overall, the Fermi level provides a unifying framework for interpreting the electronic structure and transport properties across diverse materials, from metals to insulators.

Fundamentals

Definition and origin

The concept of the Fermi level emerged from the early development of quantum statistics for identical particles obeying the . In 1926, formulated the of an of such particles in his seminal paper "Zur Quantelung des idealen einatomigen Gases," published in Zeitschrift für Physik, where he derived the distribution now known as Fermi-Dirac statistics for systems like electrons. Independently, Paul A. M. Dirac arrived at an equivalent formulation in his work "On the Theory of Quantum Mechanics" that same year, applying it to the quantization of the . The Fermi level, denoted E_F, represents the highest occupied in a fermionic at temperature (T = [0](/page/0) K), where all states below E_F are fully occupied and those above are empty, in accordance with the . In thermodynamic terms, it corresponds to the , or \mu, which governs the average occupancy of quantum states in the grand ; specifically, \mu = E_F at T = [0](/page/0). At finite temperatures, the Fermi level serves as an approximation for \mu(T), the at which the probability of state occupation is 50% according to the Fermi-Dirac distribution. For a three-dimensional free electron gas model, the Fermi energy can be derived by filling the available states in momentum space up to the Fermi wavevector k_F = (3\pi^2 n)^{1/3}, where n is the electron density. This yields the expression E_F = \frac{\hbar^2}{2m} (3\pi^2 n)^{2/3}, with m the electron mass and \hbar the reduced Planck's constant; this formula establishes the scale of electron energies in metals, typically several electron volts./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/09%3A_Condensed_Matter_Physics/9.05%3A_Free_Electron_Model_of_Metals) While the terms are sometimes used interchangeably, the Fermi energy E_F strictly denotes the fixed energy value at T = 0, whereas the Fermi level more generally refers to the temperature-dependent \mu(T) in contexts like solid-state systems, where thermal excitations cause slight shifts but \mu remains near E_F for degenerate electron gases.

Fermi-Dirac statistics

Fermi-Dirac statistics provide the quantum mechanical framework for describing the thermal distribution of identical , which are particles with half-integer spin that obey the , preventing more than one particle from occupying the same . This statistical approach was independently developed by and in 1926, building on and indistinguishability of particles. Unlike the classical Maxwell-Boltzmann statistics applicable to distinguishable particles at low densities, Fermi-Dirac statistics account for the antisymmetric wavefunction of fermions under particle exchange, leading to a fundamentally different occupation probability for energy states. The core of Fermi-Dirac statistics is the f(E), which gives the average occupation number of a with energy E: f(E) = \frac{1}{1 + \exp\left(\frac{E - \mu}{k_B T}\right)}, where \mu is the , k_B is the , and T is the absolute . This formula emerges from the in , where the partition function for fermions is constructed by summing over antisymmetric states, ensuring compliance with the Pauli principle. In contrast to the exponential decay of the Maxwell-Boltzmann distribution f(E) \approx \exp\left(-\frac{E - \mu}{k_B T}\right) for classical dilute gases, the Fermi-Dirac form saturates at 1 for E < \mu and approaches 0 for E \gg \mu, reflecting the exclusion effect even at high temperatures. At absolute zero temperature (T = 0), the distribution simplifies to a step function: f(E) = \theta(\mu - E), where \theta is the Heaviside step function, fully occupying all states below the chemical potential \mu (identified as the E_F at T = 0) and leaving higher states empty. For finite but low temperatures (k_B T \ll E_F), the sharp step smears over an energy width of order k_B T, and the Sommerfeld expansion approximates integrals involving f(E) by expanding around the T = 0 limit. This expansion, developed by in 1928, yields corrections to the chemical potential as \mu(T) \approx E_F \left[1 - \frac{\pi^2}{12} \left(\frac{k_B T}{E_F}\right)^2 \right], enabling precise calculations of thermodynamic properties like energy and pressure for degenerate systems. A key application of Fermi-Dirac statistics is in identifying degenerate Fermi gases, where quantum effects dominate classical behavior when the thermal energy is much smaller than the Fermi energy, i.e., E_F \gg k_B T. This degeneracy condition arises naturally from the filling of states up to E_F, leading to phenomena such as the pressure of a zero-temperature electron gas, which persists even without interactions due to the Pauli principle. In equilibrium, the chemical potential \mu serves as the Fermi level at low temperatures, parameterizing the distribution.

Application in solids

Band structure context

In crystalline solids, the electronic wavefunctions of electrons are described by the Bloch theorem, which states that the solutions to the Schrödinger equation in a periodic potential take the form of plane waves modulated by a periodic function with the same periodicity as the lattice. This theorem, proposed by , leads to the formation of energy bands where electron energies are quantized into allowed ranges separated by forbidden gaps. In insulators and semiconductors, the periodic potential results in a filled valence band below a bandgap E_g and an empty conduction band above it, with E_g determining the material's insulating or semiconducting nature. The position of the Fermi level E_F within this band structure dictates the occupation of electronic states according to the Pauli exclusion principle, which prohibits more than one fermion from occupying the same quantum state. In metals, E_F lies within the conduction band, allowing electrons near E_F to contribute to conduction. In intrinsic semiconductors, E_F is located approximately at the mid-gap position, ensuring equal numbers of electrons in the conduction band and holes in the valence band at thermal equilibrium. This filling up to E_F at absolute zero temperature minimizes the total energy of the system under the constraints of quantum statistics. The electron density n in the solid is obtained by integrating the density of states g(E), derived from the band structure, with the occupation probability given by the Fermi-Dirac distribution f(E): n = \int_0^\infty g(E) f(E) \, dE, where g(E) accounts for the number of available states per unit energy interval, varying with the band dispersion relation. The Fermi-Dirac occupation f(E) briefly referenced here provides the probabilistic filling of states. This integral encapsulates how the band structure influences the total electron population. The Fermi level plays a crucial role in electrical conductivity by determining the availability of states for thermal excitation. In metals, modeled as a free electron gas, the carrier concentration is approximately n \approx \frac{1}{3\pi^2} \left( \frac{2m E_F}{\hbar^2} \right)^{3/2}, where m is the electron mass and \hbar is the reduced Planck's constant; this relation highlights how E_F sets the scale for the number of conduction electrons, enabling metallic transport. In semiconductors, the position of E_F relative to the band edges governs the excitation of carriers across E_g, influencing the material's response to external stimuli.

Doping and position in materials

In intrinsic semiconductors, the Fermi level E_F is positioned approximately at the midpoint of the band gap, between the valence band maximum E_v and the conduction band minimum E_c, reflecting equal concentrations of electrons and holes. Doping introduces impurities to create extrinsic semiconductors, shifting E_F and altering carrier concentrations to enable controlled electrical properties. In n-type semiconductors, donor impurities (e.g., phosphorus in silicon) add extra electrons, increasing the electron concentration n \approx N_d, where N_d is the donor concentration. This shifts E_F toward E_c, making electron excitation to the conduction band more probable. The position is approximated by E_F \approx E_c - kT \ln \left( \frac{N_c}{N_d} \right), where kT is the thermal energy (≈0.026 eV at 300 K), and N_c is the effective density of states in the conduction band (≈2.8 × 10¹⁹ cm⁻³ for silicon). For example, in silicon doped with phosphorus at N_d = 10^{17} cm⁻³, E_F lies ≈0.146 eV below E_c. At higher doping levels (e.g., N_d > 10^{18} cm⁻³), E_F can enter the conduction band, leading to degenerate semiconductors with metallic-like behavior. In p-type semiconductors, acceptor impurities (e.g., in ) create holes by accepting electrons from the , increasing the hole concentration p \approx N_a, where N_a is the acceptor concentration. This shifts E_F toward E_v. The position is approximated by E_F \approx E_v + kT \ln \left( \frac{N_v}{N_a} \right), where N_v is the effective in the (≈1.04 × 10¹⁹ cm⁻³ for ). For p = 10^{14} cm⁻³ in , E_F lies ≈0.31 eV above E_v. High acceptor doping can similarly cause degeneracy within the . These shifts determine the semiconductor type (n or p) and are essential for fabricating devices like p-n junctions in diodes, where the differing E_F positions in adjacent regions create a built-in potential for .

Equilibrium properties

Chemical potential relation

In , the Fermi level, often denoted as \zeta or E_F, is identical to the \bar{\mu} for in a system, which remains constant throughout. The \mu is defined as the change in the of the system when an additional is added (or removed) while keeping the T, V, and total number of particles N constant, i.e., \mu = \left( \frac{\partial G}{\partial N} \right)_{T,V}. This equivalence arises because, at , E_F marks the highest occupied energy state, and at finite temperatures, it determines the occupation probability via the Fermi-Dirac distribution, aligning precisely with the thermodynamic role of \bar{\mu} in controlling particle exchange. In spatially non-uniform systems, such as heterostructures or interfaces, the local \mu(\mathbf{r}) varies with position due to differences in composition or density, while the electrostatic potential \phi(\mathbf{r}) causes . The \bar{\mu} = \mu(\mathbf{r}) - e \phi(\mathbf{r}) (for electrons with charge q = -e) remains constant across the system in equilibrium, where e is the . This constancy of the (the Fermi level E_F) ensures no net current flows, as electrons redistribute to align \bar{\mu} throughout. Thus, the local is \mu(\mathbf{r}) = E_F + e \phi(\mathbf{r}), reflecting how band bending via \phi(\mathbf{r}) shifts the effective energy landscape relative to the constant Fermi level. In closed systems where different phases or materials are in contact, the electrochemical potential \bar{\mu} is conserved and equalizes across interfaces, driving charge transfer until equilibrium is reached. This process minimizes the free energy by aligning the Fermi levels relative to a common reference, such as the vacuum level, preventing further diffusion or drift currents. For instance, in a metal-semiconductor junction, initial differences in \bar{\mu} lead to depletion or accumulation regions until the potentials balance. The connection to the work function \Phi in metals further illustrates this thermodynamic link: \Phi = -\mu / e, where \mu is the of s referenced to the level just outside the surface. This expression quantifies the minimum energy required to remove an electron from the Fermi level to , directly tying \mu to surface properties and enabling predictions of potentials between materials.

Temperature dependence

In metals, the μ(T), which corresponds to the at finite , shows only a weak dependence on due to the high degeneracy of the gas. The Sommerfeld yields the μ(T) ≈ E_F [1 - (π²/12)(kT/E_F)²], where E_F is the zero- , k is the , and T is the . This indicates a slight decrease in μ with rising T, typically less than 1% at for metals with E_F in the range of 5–10 . In intrinsic semiconductors, the Fermi level μ_i(T) lies near the center of the band gap, with its position given by μ_i(T) = (E_c + E_v)/2 + (kT/2) ln(N_v / N_c), where E_c and E_v are the conduction and valence band edges, respectively, and N_v and N_c are the effective densities of states in the . The logarithmic term introduces a linear temperature shift, which depends on the ratio of density-of-states effective masses (m_{dh}^* / m_{de}^) since N_v ∝ (m_{dh}^)^{3/2} and N_c ∝ (m_{de}^)^{3/2}; for example, in where m_{de}^ > m_{dh}^*, resulting in N_c > N_v and ln(N_v / N_c) < 0, this shift moves μ_i slightly toward the band. Furthermore, the band edges shift with owing to lattice expansion and electron-phonon coupling, generally narrowing the band gap E_g = E_c - E_v by about 10^{-4} eV/K, which causes μ_i to move toward the conduction band as T increases to maintain charge neutrality. For doped semiconductors, the Fermi level's temperature dependence reflects distinct regimes influenced by doping. At low temperatures (below ~100 K), carrier freeze-out pins μ near the dopant levels—close to the donor energy E_d (∼0.05 eV below E_c) for n-type materials—since thermal energy kT is insufficient to ionize most impurities, resulting in low free carrier density. In the intermediate extrinsic regime (∼100–400 K), full ionization occurs, and μ stabilizes at a doping-dependent position, such as μ ≈ E_c + kT ln(N_c / N_D) for n-type with donor concentration N_D, lying several kT below E_c for moderate doping. At high temperatures (above ∼400 K), intrinsic excitation dominates, and μ approaches the intrinsic value μ_i(T) as the intrinsic carrier density n_i exceeds the dopant density. Doping strongly influences the Fermi level by shifting it from the intrinsic position toward the majority carrier band, with the extent depending on impurity type and concentration. The degeneracy parameter η = (E - μ)/kT characterizes these behaviors; in non-degenerate semiconductors, η ≫ 1 (e.g., μ ≪ E_c for electrons), enabling classical Maxwell-Boltzmann statistics, whereas η ≈ 0 signals degeneracy near the band edges. Typical plots of μ(T) illustrate near-constancy in metals, a slow drift toward mid-gap in intrinsic semiconductors, and stepwise transitions in doped cases from dopant-pinned to intrinsic positions.

Measurement techniques

Voltage-based methods

Voltage-based methods for determining the rely on measuring equilibrium potentials or voltage responses that arise from the alignment of electrochemical potentials in materials, particularly in contact with metals, electrolytes, or under magnetic fields. These techniques exploit the fact that the , as the chemical potential of electrons at absolute zero, equilibrates across interfaces, manifesting as measurable voltages without net current flow. In practice, such methods provide indirect access to the position relative to reference scales like the or . The Kelvin probe technique measures the contact potential difference (CPD) between a vibrating reference electrode and the sample surface, which directly relates to the difference in work functions and thus of the two materials. The CPD, denoted as \Delta \Phi, is given by \Delta \Phi = \frac{\mu_1 - \mu_2}{e}, where \mu_1 and \mu_2 are the chemical potentials ( at low temperature) of the sample and probe, respectively, and e is the elementary charge. This voltage is induced by the capacitive coupling from the probe's vibration, nullifying the electric field between the surfaces. The method, originally developed by in 1898 and refined into scanning (KPFM) for nanoscale resolution, allows mapping of local variations, such as those due to surface states or doping gradients, with sensitivities down to millielectronvolts. In semiconductors, Mott-Schottky analysis extracts the flat-band potential from capacitance-voltage measurements at a semiconductor-electrolyte interface, linking it to the bulk . By plotting the square reciprocal of the space-charge capacitance $1/C^2 versus applied voltage V, the x-intercept yields the flat-band potential V_{fb}, which corresponds to the position of the bulk relative to the reference electrode potential through alignment with the electrolyte's redox level. This linear relation assumes a depleted space-charge region and negligible surface states, with the slope providing the doping density N_D or N_A. The technique, rooted in the 1939 theory by N.F. and further developed by , is widely applied to characterize band edges in photoelectrochemical materials like , where V_{fb} shifts with pH due to Fermi level alignment with the redox couple. The Hall effect enables inference of the Fermi level in metals and degenerate semiconductors by measuring the transverse Hall voltage V_H under a magnetic field, which yields the carrier density n = \frac{IB}{V_H e t}, where I is current, B is magnetic field, t is thickness, and e is charge. For free-electron-like systems, the Fermi energy is then calculated as E_F = \frac{\hbar^2}{2m} (3\pi^2 n)^{2/3}, relating the equilibrium carrier density directly to the occupied states up to E_F. In non-degenerate semiconductors, temperature-dependent Hall data combined with mobility allow positioning E_F relative to band edges via . This equilibrium method, discovered by in 1879, is essential for quantifying degeneracy and has been used to determine E_F \approx 7 eV in metals like copper from room-temperature measurements. Electrochemical cells determine the Fermi level through the open-circuit voltage V_{oc} between a working electrode (e.g., semiconductor or metal) and a reference redox couple, where V_{oc} = \frac{\mu_{redox} - \mu_{electrode}}{e}, equilibrating the Fermi level of the electrode with the electrochemical potential of the solution. In a setup like a with a ferrocene/ferrocenium couple, the measured V_{oc} positions \mu_{electrode} on the absolute potential scale, calibrated against the , which corresponds to -4.44 eV versus vacuum. This approach, formalized in , reveals Fermi level shifts with doping or illumination and is crucial for assessing stability in batteries and solar cells.

Spectroscopic approaches

Spectroscopic approaches provide direct probes of the by measuring the energy distribution of electrons or photons associated with occupied and unoccupied electronic states. These techniques leverage the or tunneling processes to map the (DOS) relative to the Fermi energy E_F, offering insights into without relying on electrical transport. (UPS), (IPES or BIS), (ARPES), and (STS) are among the primary methods, each targeting specific aspects of the electronic structure around E_F. Ultraviolet photoelectron spectroscopy (UPS) measures the kinetic energies of photoelectrons emitted from occupied electronic states up to the when a sample is irradiated with ultraviolet photons, typically from a helium discharge lamp (h\nu \approx 21.2 eV for He I). The spectrum reveals the valence band structure, with the Fermi edge marking the highest occupied state at T = 0 K, and the binding energy referenced to E_F. The secondary electron cutoff, the low-energy edge of the spectrum, allows determination of the work function \Phi, defined as the energy difference between the vacuum level and E_F: \Phi = h\nu - E_{\text{cutoff}} where E_{\text{cutoff}} is the minimum kinetic energy of secondary electrons. This technique is surface-sensitive (top 1-2 nm) and widely used for metals and semiconductors to establish E_F positioning relative to surface potentials. Seminal developments in UPS, including its application to valence states, trace back to early implementations in the 1970s. Inverse photoemission spectroscopy (IPES, also known as bremsstrahlung isochromat spectroscopy or BIS) complements UPS by probing unoccupied electronic states above E_F. Low-energy electrons (typically 5-50 eV) are incident on the sample, and those that decay into empty states emit photons of fixed energy, detected to map the conduction band DOS starting from E_F. The energy of the emitted photon corresponds to the difference between the incident electron energy and the binding energy of the unoccupied state, providing electron affinity and band gap information. IPES is particularly valuable for semiconductors, where it reveals the conduction band minimum relative to E_F, with energy resolution around 0.3-1 eV. A comprehensive review of IPES principles and applications highlights its role in studying unoccupied states since the 1980s. Angle-resolved photoemission spectroscopy (ARPES) extends UPS by measuring both the energy and momentum of photoelectrons, enabling mapping of band dispersions E(k) across the Brillouin zone with E_F set as the zero-energy reference. The binding energy is given by E_b = h\nu - E_k, where E_k is the kinetic energy, and the in-plane momentum k_\parallel is derived from the emission angle \theta via k_\parallel = \sqrt{2m E_k}/\hbar \cdot \sin\theta. Sharp features at E = 0 (Fermi surface crossings) confirm the E_F position, often calibrated using a metallic reference sample. ARPES is essential for visualizing how bands cross E_F in metals or lie within the gap in insulators, with resolutions down to 1-10 meV in energy and 0.01 Å^{-1} in momentum. High-resolution ARPES has been pivotal in studies of topological materials and superconductors. Scanning tunneling spectroscopy (STS), performed using a scanning tunneling microscope (STM), probes the local density of states (LDOS) near E_F by measuring the differential tunneling conductance dI/dV as a function of bias voltage V. At zero bias (V = 0), the tip and sample Fermi levels align, and dI/dV \propto \rho(E_F), the LDOS at E_F; peaks or features in dI/dV(V) indicate energy positions relative to this reference. This technique offers atomic-scale spatial resolution (sub-nm) and energy resolution (~1-10 meV), ideal for heterogeneous materials like or quantum dots where E_F varies locally. STS has revealed E_F shifts due to doping or strain in 2D materials.

Non-equilibrium dynamics

Transport and injection

In biased p-n junctions, carrier injection under forward bias leads to non-equilibrium conditions where minority carriers are injected across the junction, altering local carrier concentrations and thereby shifting the local chemical potential μ. Electrons injected into the p-region increase the local electron density n, raising μ relative to its equilibrium value, while holes injected into the n-region similarly elevate the local hole density p, lowering μ in that region to maintain charge neutrality away from the junction. This process is modeled within the , where current densities for electrons and holes are given by \mathbf{J}_n = q \mu_n n \mathbf{E} + q D_n \nabla n and \mathbf{J}_p = q \mu_p p \mathbf{E} - q D_p \nabla p, with spatial variations in n and p driving the deviations in μ. The Boltzmann transport equation provides a more fundamental description of these non-equilibrium dynamics, governing the evolution of the distribution function f(\mathbf{r}, \mathbf{k}, t) for carriers in phase space: \frac{\partial f}{\partial t} + \mathbf{v} \cdot \nabla_{\mathbf{r}} f - \frac{q}{\hbar} \mathbf{E} \cdot \nabla_{\mathbf{k}} f = \left( \frac{\partial f}{\partial t} \right)_{\text{coll}}, where the collision term accounts for scattering, often approximated via the relaxation time τ. Under bias, f deviates from the equilibrium , resulting in spatially varying effective Fermi levels that reflect local imbalances in carrier populations and energies. In steady-state conditions, this spatial variation of the effective Fermi level is solved iteratively with to capture drift, diffusion, and electrostatic effects. In light-emitting diodes (LEDs) and semiconductor lasers, forward bias injects electrons and holes into the active region, splitting the effective by an amount approximately equal to qV, where V is the applied voltage and q is the elementary charge. This splitting, denoted as ΔE_F = E_{F,n} - E_{F,p}, enables population inversion when ΔE_F exceeds the bandgap energy E_g, as the occupation probability for states in the conduction band near the band edge surpasses that in the valence band, favoring stimulated emission over absorption. For instance, in direct-bandgap materials like GaAs, this condition is achieved at biases where qV > E_g, leading to net optical gain essential for lasing. Thermoelectric effects further illustrate the role of the Fermi level in transport, particularly through the Seebeck coefficient S, which relates temperature gradients to voltage via the Mott formula derived from the Boltzmann equation in the low-temperature limit: S = -\frac{\pi^2 k_B^2 T}{3e} \left. \frac{d \ln \sigma(E)}{dE} \right|_{E=E_F}, where σ(E) is the energy-dependent conductivity, k_B is Boltzmann's constant, T is temperature, and e is the electron charge. This expression links the position of E_F relative to band features—such as the density of states or mobility edges—to the sign and magnitude of S, with p-type materials exhibiting positive S when E_F lies below the valence band maximum and n-type showing negative S above the conduction band minimum, driving voltage gradients in response to heat flow.

Quasi-Fermi levels

In non-equilibrium conditions characterized by high injection and low recombination rates, the distribution of s and s in semiconductors deviates from a single Fermi-Dirac function but can be approximated using separate quasi-Fermi levels for each type. The occupation probability is then given by f(E) \approx f_{\text{FD}}(E - E_{Fn}), where f_{\text{FD}} is the standard Fermi-Dirac distribution and E_{Fn} is the quasi-Fermi level. Similarly, the occupation probability is f_h(E) \approx 1 - f_{\text{FD}}(E - E_{Fp}), with E_{Fp} as the quasi-Fermi level. This concept was introduced to describe distributions in p-n junctions under bias, enabling the extension of to steady-state non-equilibrium scenarios. The separation between the quasi-Fermi levels, \Delta E_F = E_{Fn} - E_{Fp}, equals the applied voltage times the in , \Delta E_F = q V_{\text{applied}}, assuming negligible voltage drops across neutral regions. This relation arises from the constancy of the product of and quasi-Fermi potentials across the device in forward bias. The approximation holds under conditions of low temperature, where the Fermi-Dirac tails are sharp, and short , which maintain distinct populations; it breaks down in regimes of strong that prevent local thermalization of carrier distributions within each . Quasi-Fermi levels are central to analyzing device performance, particularly in solar cells, where the power conversion \eta scales with the normalized splitting (E_{Fn} - E_{Fp}) / E_g, with E_g the bandgap; maximum occurs when this splitting approaches E_g under . In recombination processes, rates such as Shockley-Read-Hall are expressed using quasi-Fermi levels, yielding R = (np - n_i^2) / \tau, where n and p derive from E_{Fn} and E_{Fp}, n_i is the intrinsic carrier concentration, and \tau the effective lifetime, facilitating compact modeling of non-radiative losses.

Technical considerations

Nomenclature variations

The term "Fermi level" originated in the context of Enrico Fermi's 1926 work on the quantization of the ideal monatomic gas, where it described the highest occupied energy state for fermions at temperature under Fermi-Dirac statistics. This concept was initially applied to gases in . Its adoption in occurred in the late 1920s and 1930s, particularly following Bloch's 1928 analysis of waves in periodic crystal lattices, which introduced band structures and extended the Fermi-Dirac framework to electrons in solids, leading to the notion of a filled Fermi sea within energy bands. A persistent source of inconsistency in the literature arises from the interchangeable or context-dependent use of "" and "Fermi level." The , often denoted E_F, strictly refers to the energy of the highest occupied state at T = 0 K, representing a fixed value for a given system. In contrast, the Fermi level corresponds to the \mu (sometimes denoted \zeta or E_F) at finite temperatures, where it shifts slightly due to thermal excitation while remaining approximately constant for typical solid-state systems. This distinction is not always maintained; many texts and papers use E_F for both, especially in approximations where \mu \approx E_F for metals at , leading to terminological overlap across theoretical and experimental works. In device physics, the is frequently termed the "Fermi potential," denoted \psi_F or \phi_F, to emphasize its role in relating concentrations to applied voltages in semiconductors. It represents the electrostatic potential shift of the Fermi level relative to the intrinsic level, which is zero in intrinsic semiconductors. For p-type materials, \phi_F \approx \frac{[kT](/page/KT)}{[q](/page/Q)} \ln\left(\frac{N_A}{n_i}\right); for n-type, \phi_F \approx -\frac{[kT](/page/KT)}{[q](/page/Q)} \ln\left(\frac{N_D}{n_i}\right), where N_A and N_D are acceptor and donor densities, and n_i is the intrinsic concentration. This contrasts with thermodynamic contexts, where it is simply the \mu, highlighting a disciplinary for applications involving electrostatic potentials. In surface chemistry, "Fermi energy" is commonly used to denote the energy reference for at interfaces, such as in adsorption processes, but care must be taken to distinguish it from the level to avoid conflation with measurements.

Energy referencing conventions

In energy band diagrams for solids, a standard convention is to set the Fermi level E_F to zero, which simplifies the visualization of the relative positions of the conduction band minimum E_c and valence band maximum E_v with respect to the occupied states. This relative referencing highlights the band gap and doping-induced shifts without needing an absolute energy scale, as commonly practiced in analysis. For absolute energy positioning, the level E_\mathrm{vac} serves as a , defined such that the \Phi = E_\mathrm{vac} - E_F, where E_\mathrm{vac} is typically set to 0 and E_F is negative relative to it—the minimum energy required to remove an from E_F to just outside the surface. This convention links the internal electronic structure to external potentials, enabling comparisons across materials via parameters like \chi = E_\mathrm{vac} - E_c. Internal referencing relative to band edges, such as E_c - E_F, is also prevalent, as it directly relates to concentrations; for instance, in n-type semiconductors, E_c - E_F quantifies the distance from the Fermi level to the conduction band, influencing without invoking external levels. However, the vacuum level reference introduces challenges due to its sensitivity to surface conditions. , involving atomic rearrangements or formation, can shift the local E_\mathrm{vac} relative to the bulk, altering the apparent by several electron volts and causing misalignment errors in band diagrams. In heterostructures, this non-invariance exacerbates issues, as interface s and charge transfer disrupt simple vacuum-level alignment, leading to inaccuracies in predicting barrier heights or offsets between dissimilar materials like GaAs/AlGaAs. To address these limitations in multi-material systems, internal referencing schemes are recommended, such as aligning bands using the \mu (equivalent to E_F at zero temperature) or average valence band maximum positions derived from core-level . Band-aligned scales, which match internal energy levels across interfaces without relying on surface-sensitive vacuum references, provide more reliable offsets for device modeling in complex heterojunctions.

Nanoscale and discrete effects

In nanoscale systems, the continuum approximation of the Fermi level breaks down due to quantum confinement, leading to energy levels that dominate the electronic structure. In quantum dots, these levels result in a stepwise variation of the Fermi level as electrons are added one by one, particularly in the regime where the charging energy U = e^2 / (2C), with C being the dot , greatly exceeds the k_B T. This blockade suppresses electron tunneling until the applied bias or gate voltage overcomes U, manifesting as sharp conductance peaks corresponding to alignment of individual levels with the Fermi energy of the leads. U = \frac{e^2}{2C} \gg k_B T Such behavior is prominent in semiconductor quantum dots fabricated from materials like GaAs, where confinement in all three dimensions quantizes the spectrum, and the addition of each electron shifts the chemical potential (Fermi level) by approximately U, reflecting the discrete nature of charge quantization. In two-dimensional systems, such as graphene, the linear dispersion of Dirac fermions further modifies the Fermi level position, which depends on the carrier density n and can be continuously tuned via electrostatic gating. The relation is given by E_F = \hbar v_F \sqrt{\pi n}, where v_F is the Fermi velocity, approximately $10^6 m/s. This density-dependent tuning allows the Fermi level to cross the Dirac point at zero density, transitioning from electron to hole doping without a bandgap, enabling unique transport properties in graphene nanoribbons or constrictions where edge effects introduce additional quantization. Finite-size effects introduce broadening to these discrete levels, characterized by \Gamma \approx \hbar / \tau, where \tau is the lifetime of the states due to coupling to leads or phonons, causing the effective Fermi level to smear over multiple discrete states when \Gamma exceeds the level spacing \delta. In small quantum dots, this broadening, often on the order of meV from tunneling rates, blurs sharp steps in the density of states, but the discrete regime persists when k_B T and \Gamma remain below \delta. For instance, in single-electron transistors based on quantum dots, the gate voltage V_g electrostatically tunes the position of the dot's discrete levels relative to the Fermi level of the source and drain, revealing Coulomb blockade oscillations; this quantized tuning is essential for devices below approximately 10 nm, where the continuous Fermi sea model fails and level spacing \delta \gtrsim 10 meV dominates over thermal smearing.

References

  1. [1]
    Fermi level and Fermi function - HyperPhysics
    The Fermi level is the surface of that sea at absolute zero where no electrons will have enough energy to rise above the surface. The concept of the Fermi ...
  2. [2]
    [PDF] Lecture 13: Metals
    Sometimes the Fermi energy is also called the Fermi level. At any temperature T >0, some fermions will have energies higher than "F, and therefore some states ...
  3. [3]
    [PDF] Contents - Physics Courses
    Metals are characterized by a finite electronic density of states f(εF) at the Fermi level at zero tempera- ture. This entails a number of salient features, ...<|control11|><|separator|>
  4. [4]
    [PDF] Semiconductors
    Fermi Distribution, Fermi Energy, and Fermi Level​​ It is the chemical potential (µ) of the electrons in the semiconductor. probability of occupation is ½.
  5. [5]
    [PDF] The Fermi function and the Fermi level • Effective density of states
    is the effective, or quasi-, Fermi level for electrons. Similarly, there is a quasi-Fermi level for holes, E fp. , and the holes are.
  6. [6]
    [PDF] Conduction of Electricity in Solids - UF Physics Department
    Note that all states with energies up to the Fermi energy EF " 7 eV are occupied, and all those with energies above the Fermi energy are vacant. (b) The same ...
  7. [7]
    On the Quantization of the Monoatomic Ideal Gas - cond-mat - arXiv
    Dec 14, 1999 · An english translation of the original work of Enrico Fermi on the quantization of the monoatomic ideal gas, is given in this paper.Missing: 1926 | Show results with:1926
  8. [8]
    The Fermi–Dirac Distribution - DoITPoMS
    At absolute zero the value of the chemical potential, μ , is defined as the Fermi energy. At room temperature the chemical potential for metals is virtually ...
  9. [9]
    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.
  10. [10]
    [PDF] Energy bands - Rutgers Physics
    Equation (14) is known as Bloch theorem, which plays an important role in electronic band structure theory. Now we discuss a number of important conclusions ...
  11. [11]
    [PDF] Electrons and Holes in Semiconductors
    The electron and hole concentrations are closely linked to the Fermi level. The materials introduced in this chapter will be used repeatedly as each new device ...
  12. [12]
    [PDF] Electron Density, Band Structure, and Density of States
    In order to count this number of allowed states, we can formally introduce the so-called “density of states” (DOS) by the following integral over the Brillouin ...
  13. [13]
    [PDF] Free Electron Fermi Gas (Kittel Ch. 6) - SMU Physics
    - topmost filled energy level (Fermi level). Page 17. Electron orbitals. • D(E)= (V/2π2) E1/2 (2m / h2)3/2. • Now we need to figure out how many electrons are ...
  14. [14]
    Extending the Concept of Defect Chemistry from Semiconductor ...
    Feb 27, 2014 · The chemical potential of the electron μ e (or Fermi energy, E F ) represents the electron reservoir and accounts for the possibility of ...
  15. [15]
    [PDF] 6.730 Physics for Solid State Applications - Lecture 25 - MIT
    Apr 7, 2004 · Lecture 25: Chemical Potential and Equilibrium. • Microstates and ... The electrochemical potential, a.k.a, the fermi level is. The ...
  16. [16]
    [PDF] Title goes here - Umberto Ravaioli
    • So, the Fermi level is the same as the electro- chemical potential in equilibrium. 8. Page 9. Two materials in contact at equilibrium. • The two materials ...
  17. [17]
    One-dimensional plasmons confined in bilayer graphene - junctions
    ... electrostatic potential and kinetic energy of electrons (at the Fermi level) is constant. ... The relation between the local chemical potential and the ...<|control11|><|separator|>
  18. [18]
    6.13 Fermi-Dirac Distribution - FAMU-FSU College of Engineering
    In particular, a Fermi-level electron has an energy equal to the chemical potential. The term electrochemical potential needs some additional comment. The ...
  19. [19]
    Band Alignment and the Built-in Potential of Solids | Phys. Rev. Lett.
    Nov 9, 2018 · The built-in potential is of central importance to the understanding of many interfacial phenomena because it determines the band alignment at the interface.
  20. [20]
    Work Function: Fundamentals, Measurement, Calculation ...
    ... Fermi level is the chemical potential of electrons in the material. Thus, if the two materials have different work functions, electrons will flow from the ...
  21. [21]
    6.16 Chemical Potential and Diffusion
    The chemical potential, or Fermi level, that appears in the Fermi-Dirac ... But if the work function of the second metal is different from the first ...
  22. [22]
    Sommerfeld Expansion - Richard Fitzpatrick
    is the Fermi function. This expansion is known as the Sommerfeld expansion, after its inventor, Arnold Sommerfeld. is known as the Fermi temperature.Missing: level | Show results with:level
  23. [23]
    Fermi energy of an intrinsic semiconductor - TU Graz
    The Fermi energy is in the middle of the band gap (E c + E v )/2 plus a small correction that depends linearly on the temperature.
  24. [24]
    [PDF] Fermi Levels of Intrinsic Semiconductors with Effective Mass in ...
    Using this we solve for the concentration of each and develop an expression for the Fermi level. We then explore what happens to the Fermi level at different.
  25. [25]
    [PDF] Temperature Dependence of Semiconductor Conductivity
    In this intrinsic region, carrier concentration increases with temperature as shown in Equation 3 because ni has become the dominant term of. Equation 4. 0. 2.
  26. [26]
    [PDF] Lecture 16 - MIT OpenCourseWare
    Prentice Hall, 1988. 5. Page 6. Degenerate and non-degenerate semiconductors. In non-degenerate SC the chemical potential satisfies: E c. -μ >> k. B. T μ - E v.
  27. [27]
    Kelvin probe force microscopy and its application - ScienceDirect
    Kelvin probe force microscopy (KPFM) is a tool that enables nanometer-scale imaging of the surface potential on a broad range of materials.
  28. [28]
    Flat-Band Potential of a Semiconductor: Using the Mott–Schottky ...
    Determination of the flat-band potential of a semiconductor using electrochemical impedance spectroscopy (EIS) and the Mott–Schottky equation.Missing: original | Show results with:original
  29. [29]
    [PDF] HALL COEFFICIENT AND RESISTIVITY OF SEMICONDUCTORS ...
    ... Fermi level, and f is the Fermi-. Dirac distribution function: 1 f(E) ... Long, Donald; and Myers, John: Hall Effect and Impurity Levels in Phosphorus-.
  30. [30]
    I. Bridging Open-Circuit Voltage of Electrochemical Cells and ... - NIH
    Jul 25, 2022 · Bridging Open-Circuit Voltage of Electrochemical Cells and ... (Fermi level) and a point outside the solution.” This definition has ...
  31. [31]
    [PDF] 1 Boltzmann equation - Quantum Transport Theory Group
    Note that in conditions of equilibrium (zero bias) only the solution of Poisson's equation is necessary, since the equilibrium. Fermi level is constant and ...
  32. [32]
    [PDF] The Junction Characteristics and Current Conduction ... - DTIC
    (a) Schematic of a semiconductor p-n junction, and (b) an energy-band diagram of a forward-biased p-n junction showing. Fermi-level splitting. 29. Page 49 ...
  33. [33]
    [PDF] ECE 340 Lecture 29 : LEDs and Lasers
    To obtain population inversion, frequently the separation between quasi-Fermi levels is greater than the band gap in the inversion area.Missing: splitting ΔE_F > E_g
  34. [34]
    [PDF] T h e P h y sic s a n d Pe r fo r m a n c e of In GaA s Qu a n t u m D ...
    In a simple p-n junction laser, population inversion is achieved by forward biasing a heavily doped p-n junction. In equilibrium, the Fermi level, Ef, lies ...
  35. [35]
    [PDF] Thermoelectric Transport in Nanoscale Materials - Projects at Harvard
    The Mott formula is a function of the conductivity, whose expression contains the scattering time, τ. If τ is a function of the energy, τ(E) ∝ Ep, then ...
  36. [36]
    Transport Function Analysis of Thermoelectric Properties
    Seebeck Formulas. We like to think of the Seebeck coefficient as primarily a measure of the Fermi Level, relative to the band edge. This can be seen in ...
  37. [37]
    [PDF] shockley-pn-junction.pdf
    Contact potentials across pn junctions, carry- ing no current, may develop when hole or electron injection occurs. The principles and theory of a pa-p ...
  38. [38]
    Quasi-Fermi level splitting and sub-bandgap absorptivity from ...
    Nov 5, 2014 · van Roosbroeck and Shockley (vRS) famously demonstrated the relationship between the volumetric rate of internal spontaneous radiative ...
  39. [39]
    [PDF] Detailed Balance Limit of Efficiency of pn Junction Solar Cells
    Comparison of the "semiempirical limit" of efficiency of solar cells with the "detailed balance limit," derived in this paper. ... SHOCKLEY AND H. J. QUEISSER.
  40. [40]
    [PDF] Statistics of the Recombinations of Holes and Electrons
    f(E) is the fraction of states of energy E that are occupied by electrons. The probability that an electron be emittedfrom the trap into the band of energies in ...
  41. [41]
    Indistinguishable elements in the origins of quantum statistics. The ...
    Jan 5, 2022 · In this paper, we deal with the historical origins of Fermi–Dirac statistics, focusing on the contribution by Enrico Fermi of 1926.
  42. [42]
  43. [43]
    [PDF] Elementary Semiconductor Device Physics
    Energy band diagram of a forward-biased p-n junction with quasi-Fermi levels. At zero bias (V = 0), carrier drift and diffusion are balanced (p. 175), and ...<|separator|>
  44. [44]
    [PDF] Fermi level, work function and vacuum level
    Oct 7, 2015 · Electronic levels and energies of a solid, such as Fermi level, vacuum level, work function, ionization energy or electron affinity, ...
  45. [45]
    Fermi level, work function and vacuum level - RSC Publishing
    The goal of this brief focus article is to provide definitions of key concepts and review simple mechanisms that affect these fundamental quantities.Missing: reference | Show results with:reference
  46. [46]
    Opportunities for energy level tuning at inorganic/organic ...
    Dec 27, 2021 · In practical terms and for solids, the IE and EA are given with respect to the vacuum level ... If the surface reconstruction involves ...
  47. [47]
    Understanding band alignments in semiconductor heterostructures ...
    Apr 12, 2017 · Common choices of universal alignment levels are the vacuum level following Anderson's electron-affinity rule [30] , characteristic marker ...Missing: issues | Show results with:issues
  48. [48]
    [PDF] Nanoscale device modeling: the Green's function method
    A discrete level is coupled to a reservoir with a Fermi level µ. The broadening is the same regardless of whether the reservoir states are empty or filled.
  49. [49]
    [PDF] 1 Single-Electron Devices - MPI-FKF
    For the metal single- electron transistor, the transport characteristics are periodic in V€S : With each gate voltage change AV€S = e / C € , the same ...