Fact-checked by Grok 2 weeks ago

Degenerate semiconductor

A degenerate semiconductor is a material doped to such a high concentration—typically on the order of 10¹⁹ to 10²⁰ atoms per cm³—that the penetrates into the conduction band for n-type materials or the valence band for p-type materials, leading to degenerate or statistics and behavior more akin to a metal than a conventional . In this regime, the carrier concentration exceeds the effective near the band edge, invalidating the Maxwell-Boltzmann approximation and requiring the full Fermi-Dirac distribution for accurate modeling of carrier densities and transport properties. Unlike non-degenerate semiconductors, where the Fermi level resides in the bandgap and carrier concentrations follow classical statistics (with the chemical potential at least 3kT away from band edges), degenerate cases exhibit enhanced conductivity due to the filling states up to the , resulting in minimal temperature dependence of carrier density at low temperatures and a near-elimination of the freeze-out regime. High doping also induces effects such as bandgap narrowing (e.g., up to 200 meV in heavily doped ) from electron-electron interactions and many-body effects, alongside potential band-tail states that further modify optical and electrical responses. These properties arise primarily in materials like , , or when impurity atoms (donors or acceptors) form extended bands rather than discrete levels in the bandgap. Degenerate semiconductors play a critical role in advanced device applications where high carrier densities are essential for efficient performance. They are commonly employed in ohmic contacts for low-resistance electrical interfacing in integrated circuits, laser diodes leveraging in the degenerate regime, and Zener diodes for voltage regulation via tunneling effects enhanced by narrow effective bandgaps. Additionally, they enable high-injection regimes in heterojunction bipolar transistors (HBTs) and optoelectronic components, though excessive doping can degrade performance through mechanisms like reduced gain from degenerate statistics. Emerging uses include and ionic-gated devices for modulating or Hall effects in oxide semiconductors.

Fundamentals

Definition and characteristics

A degenerate semiconductor is defined as a material in which the doping concentration is sufficiently high—typically exceeding $10^{19} cm^{-3} for (Si) and ~$10^{18} cm^{-3} for (GaAs), depending on the effective near the band edge—that the penetrates into the conduction band for n-type doping or the valence band for p-type doping. This high impurity level causes the donor or acceptor states to form an impurity band that merges with the respective host band, ensuring nearly complete ionization of dopants and resulting in carrier concentrations comparable to those in metals. Consequently, the material exhibits metallic-like electrical conductivity, with transport properties dominated by intraband rather than across the bandgap. The primary characteristics of degenerate semiconductors stem from the elevated carrier density, which creates a degenerate or gas governed by quantum mechanical effects. Under these conditions, the prevents multiple fermions from occupying the same , leading to a filled Fermi sea up to the and enabling conduction even at low temperatures without thermal excitation. Representative examples include heavily n-type doped with concentrations above $10^{19} cm^{-3}, which displays ohmic behavior akin to metals, and (InAs), which can achieve degeneracy at lower doping levels (around $10^{17} cm^{-3}) owing to its narrow bandgap of approximately 0.35 eV and low effective . This quantum nature manifests in phenomena such as the Burstein-Moss bandgap shift, where the onset of optical absorption blue-shifts due to blocked low-energy transitions in the filled band states. The concept of degenerate semiconductors emerged in the mid-20th century, particularly during the , as advanced with the development of technology and deeper understanding of band theory. It built upon Sommerfeld's early 1920s free electron model for metals, which introduced the idea of a degenerate to explain metallic properties through quantum statistics, later adapted to heavily doped semiconductors.

Comparison with non-degenerate semiconductors

Non-degenerate semiconductors are characterized by relatively low doping levels, where the lies within the bandgap, typically at least 3kT away from the conduction edge (E_C) for n-type or the edge (E_V) for p-type materials, allowing the use of the classical Boltzmann approximation to describe carrier statistics. In this regime, the carrier distribution follows Maxwell-Boltzmann tails, with electron concentration n approximated as n = N_C \exp\left(-\frac{E_C - E_F}{kT}\right), where N_C is the effective in the conduction , leading to thermally activated behavior and exponential dependence on . This contrasts sharply with degenerate semiconductors, where high doping positions the within the conduction or (E_F > E_C or E_F < E_V), requiring the full Fermi-Dirac statistics and resulting in a near step-like occupation function that fills states up to the Fermi energy, akin to metallic behavior. The degeneracy parameter η = (E_F - E_C)/kT (for n-type) exceeds approximately 0 in this case, causing saturation in carrier mobility due to dominant ionized impurity scattering that renders mobility largely -independent at low temperatures, and introducing non-linear responses in transport properties such as enhanced tunneling and reduced temperature sensitivity. Degeneracy arises when the doping concentration N_D (or N_A for p-type) greatly exceeds the intrinsic carrier concentration n_i, such that n_i << N_D, ensuring nearly all dopants are ionized and contribute carriers without significant thermal activation. For example, in silicon at room temperature (where n_i ≈ 1.5 × 10^{10} cm^{-3}), the Boltzmann approximation breaks down for N_D ≳ 10^{18} cm^{-3} (η ≳ -3, errors exceeding 10%), while degeneracy sets in around N_D ≈ 10^{19} cm^{-3} (η ≈ 0). At these levels, the inter-dopant spacing becomes comparable to the Debye length, promoting band-like merging of impurity levels rather than isolated states. In modeling, non-degenerate cases simplify calculations using exponential carrier distributions, suitable for standard device simulations under moderate doping and temperature. Degenerate regimes, however, demand numerical evaluation of for carrier concentrations, such as n = N_C F_{1/2}(η), reflecting the abrupt filling of states and enabling accurate prediction of metallic-like conductivity without freeze-out effects at cryogenic temperatures. This distinction is crucial for applications like ohmic contacts and tunnel junctions, where degeneracy ensures high carrier densities and robust performance.

Physics of degeneracy

Fermi level and carrier concentration

In degenerate n-type semiconductors, the Fermi level E_F shifts above the conduction band edge E_C, while in degenerate p-type semiconductors, it shifts below the valence band edge E_V. This positioning arises from high doping concentrations that fill states up to energies beyond the band edges, following rather than classical . The electron carrier concentration n in the conduction band for degenerate conditions is given by n = N_c F_{1/2}(\eta), where N_c = 2 \left( \frac{2\pi m_e^* kT}{h^2} \right)^{3/2} is the effective density of states in the conduction band, \eta = (E_F - E_C)/kT is the reduced Fermi energy, and F_{1/2}(\eta) is the Fermi-Dirac integral of order 1/2, defined as F_{1/2}(\eta) = \frac{2}{\sqrt{\pi}} \int_0^\infty \frac{\sqrt{x}}{1 + e^{x - \eta}} \, dx. An equivalent form expresses n directly as n = \frac{1}{2\sqrt{2}} \left( \frac{2\pi m_e^* kT}{h^2} \right)^{3/2} F_{1/2}(\eta). A similar expression holds for hole concentration p in the valence band, with N_v and \eta = (E_V - E_F)/kT. These relations account for the Pauli exclusion principle, which limits occupancy to one electron per state at the Fermi level. For strong degeneracy where \eta \gg 1, the Fermi-Dirac integral approximates to F_j(\eta) \approx \frac{\eta^{j+1}}{j+1}, yielding for j = 1/2, F_{1/2}(\eta) \approx \frac{2}{3} \eta^{3/2}. Thus, n \approx N_c \frac{2}{3} \left[ \frac{E_F - E_C}{kT} \right]^{3/2}, implying E_F - E_C \approx \frac{1}{2} (3\pi^2 n)^{2/3} \frac{\hbar^2}{m_e^*} from the free-electron Fermi energy at zero temperature. This approximation simplifies calculations for highly doped materials, where the carrier density approaches the dopant concentration N_D (for n-type) with minimal thermal excitation. In gallium arsenide (GaAs), an n-type doping of $10^{19} cm^{-3} results in \eta \approx 10 at 300 K, given N_c \approx 4.7 \times 10^{17} cm^{-3} and m_e^* = 0.067 m_0, placing E_F approximately 0.26 eV above E_C and confirming degenerate behavior since \eta > 2. Such levels are common in device contacts and enable metallic-like conduction. The degeneracy persists at low temperatures, as the thermal energy kT becomes negligible compared to the Fermi energy spacing. The degeneracy temperature T_D = E_F / k marks the onset of degeneracy, typically \approx 3000 K for doping around $10^{19} cm^{-3} in GaAs, ensuring quantum effects dominate even near room temperature in heavily doped samples.

Band filling and degeneracy parameter

In degenerate n-type semiconductors, the Fermi level E_F lies within the conduction band, above the conduction band edge E_C, resulting in partial filling of the available states up to E_F. This configuration forms a degenerate electron gas, where the electrons occupy states according to the Pauli exclusion principle, filling all quantum states up to E_F at absolute zero temperature. The total electron occupancy is determined by integrating the density of states g(E) with the Fermi-Dirac distribution function from E_C to the Fermi level. For p-type degenerate semiconductors, the valence band is partially empty below E_F, creating a degenerate hole gas in an analogous manner. The degeneracy parameter \eta, which quantifies the degree of degeneracy, is defined for electrons as \eta = (E_F - E_C)/kT, where k is Boltzmann's constant and T is the temperature; a similar definition applies for holes as \eta_v = (E_V - E_F)/kT. Degeneracy becomes significant when \eta > 2, marking the transition from non-degenerate Boltzmann statistics to Fermi-Dirac degeneracy effects. In the strongly degenerate limit, the Fermi energy for a three-dimensional parabolic conduction band can be approximated using the free-electron model as E_F = \frac{\hbar^2}{2m^*} (3\pi^2 n)^{2/3}, where \hbar is the reduced Planck's constant, m^* is the electron effective mass, and n is the electron concentration. Quantum effects in degenerate semiconductors arise from the Pauli exclusion principle, which enforces full occupancy of states below E_F and prevents further excitation of electrons to already filled levels, a phenomenon known as Pauli blocking. This blocking suppresses low-energy transitions and alters the material's response to perturbations compared to non-degenerate cases. Narrow-bandgap materials like InSb exemplify these effects, where the small effective mass and bandgap enable degeneracy at relatively modest doping levels, leading to metallic-like behavior.

Material properties

Electrical conductivity

In degenerate semiconductors, the electrical conductivity follows the given by \sigma = n e \mu, where n is the carrier concentration, e is the , and \mu is the mobility; the high n (often exceeding $10^{19} cm^{-3}) arising from heavy doping leads to metallic-like conductivities. At these high carrier densities, ionized scattering becomes the dominant mechanism limiting \mu, causing it to decrease and eventually saturate with further doping, unlike the more gradual decrease in non-degenerate cases at lower doping levels, typically reaching values around 10–50 cm²/V·s in heavily doped materials. While the temperature dependence of due to ionized can be classically described by models such as the Brooks-Herring formula, yielding \mu \propto T^{3/2} from increased screening and reduced cross-section at higher temperatures, in the degenerate this overestimates values, and overall \mu often exhibits weaker dependence or at low temperatures where degeneracy effects fix velocities near the Fermi speed. The in degenerate semiconductors retains the classical form for the Hall coefficient R_H = 1/(n e) for single-band parabolic transport, but effects—such as non-sphericity or multi-valley structures in materials like —introduce corrections, modifying the effective Hall factor and leading to deviations from simple free-electron predictions. At high frequencies, the anomalous emerges when the exceeds the classical skin depth, resulting in non-local current distribution and altered ac conductivity, observable in degenerate semiconductors due to their high carrier densities and conductivities. Experimental measurements in degenerate n-type silicon with doping levels above $10^{19} cm^{-3} yield conductivities exceeding $10^3 S/cm, approaching metallic values; for instance, heavily phosphorus-doped silicon nanowires exhibit \sigma \approx 1660 S/cm at room temperature. In heavily doped polycrystalline silicon, such as boron-doped films used in thermoelectric applications, conductivities on the order of $10^3 S/cm are reported, highlighting the impact of grain boundaries on transport while maintaining high overall values due to degeneracy.

Optical absorption

In degenerate semiconductors, particularly n-type materials, the optical absorption edge is influenced by two competing effects: the Burstein-Moss blue-shift and bandgap narrowing. The Burstein-Moss effect arises because high carrier concentrations fill the lower energy states in the conduction band, blocking vertical transitions from the valence band to those occupied states due to the ; thus, absorption onset requires photons with energy exceeding the original . For parabolic conduction bands, the shift in absorption onset energy due to Burstein-Moss is approximately \Delta E = \frac{\hbar^2}{2 m_e^*} (3 \pi^2 n)^{2/3}, where m_e^* is the electron effective mass and n is the carrier concentration, resulting in an effective increase of the optical band gap by \Delta E. This shift scales with n^{2/3} and is prominent in materials like indium-doped zinc oxide at carrier densities above $10^{19} cm^{-3}. However, heavy doping also induces bandgap narrowing \Delta E_g (up to ~100-200 meV) from electron-electron and electron-impurity interactions, which reduces the fundamental and partially counteracts the Burstein-Moss shift; the net optical band gap is thus E_g + \Delta E - \Delta E_g, with the dominant effect depending on material and doping level. High free carrier densities also give rise to a frequency \omega_p = \sqrt{\frac{n e^2}{\epsilon_0 \epsilon_r m^*}}, where \epsilon_r is the . Below \omega_p, typically in the , the real part of the function becomes negative, leading to high reflectivity akin to metallic behavior. Free carrier absorption dominates in this regime, with the absorption coefficient \alpha following \alpha \propto \lambda^2 due to momentum-conserving phonon-assisted intraband transitions. Degeneracy further quenches by Pauli blocking, which inhibits radiative electron-hole recombination as the final states in the conduction or bands are already occupied by . In degenerate GaAs, this manifests as reduced efficiency at high doping levels (e.g., above $10^{18} cm^{-3}), limiting emission in structures where carrier injection creates degenerate conditions.

Thermal properties

In degenerate semiconductors, the electronic specific heat at low temperatures deviates significantly from classical expectations due to Fermi-Dirac statistics. For degenerate fermions, the specific heat C is linear in temperature T, given by C = \frac{\pi^2}{3} k_B^2 T g(E_F), where k_B is the and g(E_F) is the at the E_F. This arises because only electrons within approximately k_B T of E_F contribute to the , leading to a much smaller value than the classical \frac{3}{2} Nk_B per carrier. In contrast, non-degenerate semiconductors exhibit an electronic specific heat that follows an exponential activation law due to the thermally activated carrier concentration across the bandgap. The thermal conductivity \kappa in degenerate semiconductors is primarily dominated by the electronic contribution, as the high carrier density enhances electron-mediated heat transport. This is described by the Wiedemann-Franz law, which holds well in the degenerate regime: \kappa = \frac{\pi^2}{3} \frac{k_B^2 T}{e^2} \sigma, where \sigma is the electrical and e is the charge. The contribution to \kappa, which typically dominates in non-degenerate semiconductors, is relatively suppressed in heavily doped degenerate cases due to the increased electronic term and scattering effects. Thermoelectric effects in degenerate semiconductors feature a Seebeck coefficient S that is notably high and inversely proportional to E_F, approximated by the Mott formula S \approx \frac{\pi^2 k_B^2 T}{3 e E_F} for energy-independent scattering, reflecting the asymmetry in carrier transport near the . This energy-dependent scattering enhances S compared to non-degenerate cases. For example, in degenerate variants of \mathrm{Bi_2Te_3}, such as those tuned via adjustment through doping, S values around 150–200 \muV/K are achieved at , enabling high thermoelectric performance.

Fabrication and doping

High doping techniques

Ion implantation is a widely used method for achieving high dopant concentrations in semiconductors, enabling degenerate doping levels by accelerating ions into the to create precise profiles. This technique is particularly effective for selective area doping, as demonstrated in (SiC) and (GaN), where it introduces impurities beyond solubility limits. Following implantation, high-temperature annealing is essential to repair radiation-induced damage and electrically activate the s, often reaching activation efficiencies near 100% in materials like 4H-SiC for n-type doping up to degenerate concentrations of approximately 10^{20} cm^{-3}. Molecular beam epitaxy (MBE) provides exceptional control over incorporation during layer growth, ideal for fabricating degenerate semiconductors with abrupt interfaces and minimal defects. In III-V compounds like cubic , MBE enables germanium doping up to 3.7 \times 10^{20} cm^{-3}, resulting in degenerate n-type behavior evidenced by spectral broadening in spectra. Similarly, heavy doping in via MBE achieves concentrations exceeding 10^{20} cm^{-3} while suppressing surface segregation through optimized growth conditions. (CVD), including variants like CVD, supports uniform high-dose doping across large areas, as seen in n-type layers with or precursors yielding concentrations over 10^{20} cm^{-3} for epitaxial growth. Common dopant selections depend on the host material: in group IV semiconductors like , group V elements such as or serve as n-type donors, while group III elements like act as p-type acceptors, with exhibiting a solid limit of approximately 3 \times 10^{20} cm^{-3} at 1100^\circ C. In III-V semiconductors like GaAs, n-type doping typically employs group IV () or group VI () impurities to donate electrons, and p-type uses group II ( or ) acceptors to provide holes, enabling degenerate levels while respecting constraints. Delta-doping, often implemented via interrupted growth in or CVD, confines dopants to a sub-monolayer plane, inducing two-dimensional degeneracy by forming a high-density (2DEG) in structures like Si:P layers with sheet densities up to 10^{14} cm^{-2}. Recent advancements include laser annealing techniques, which rapidly heat the surface to activate dopants with negligible diffusion, preserving shallow junctions essential for scaling. In SiGe alloys for 2020s technologies, nanosecond UV laser annealing of gallium-implanted layers achieves high activation rates over 80% at concentrations near 10^{20} cm^{-3}, mitigating strain relaxation and enabling sub-5 nm node performance.

Challenges in degenerate doping

Achieving degenerate doping levels in semiconductors often exceeds the solubility limits of dopants, leading to and clustering that render portions of the dopants electrically inactive. For instance, in , the solid of is approximately 10^{20} cm^{-3} at typical processing temperatures around 800–900°C, beyond which excess phosphorus forms precipitates such as phases or interstitial clusters, introducing recombination centers that degrade device performance. These precipitates not only reduce the effective carrier concentration but also create structural inhomogeneities, complicating uniform doping across the material. High dopant concentrations also promote the formation of lattice defects, including vacancies and dislocations, which arise from strain induced by the incorporation of foreign atoms into the crystal lattice. In wide-bandgap semiconductors, native defects exhibit amphoteric behavior, acting as compensating centers that donate or accept carriers depending on the position of the Fermi level relative to a stabilization energy, thereby limiting the net carrier density even at high nominal doping levels. For example, in n-type doping, self-compensation occurs through the formation of acceptor-like vacancies or interstitials, reducing the effective electron concentration and mobility, while dislocations further scatter carriers and exacerbate non-uniformity. This defect proliferation is particularly pronounced in materials like or ZnO, where doping beyond 10^{19}–10^{20} cm^{-3} triggers significant compensation, often halving the expected carrier density. Maintaining degenerate doping profiles is challenged by thermal instability, as elevated temperatures during processing cause that broadens and degrades the intended concentration gradients. Electrical efficiency typically falls below 100% due to inactive dopants trapped in or complexes, with only a fraction contributing to free carriers even after annealing. To mitigate these issues, rapid thermal annealing () is employed, providing short, high-temperature pulses (e.g., 1000–1100°C for seconds) to enhance while minimizing and growth. However, must be carefully optimized, as over-annealing can still lead to or defect annealing that inadvertently increases compensation.

Applications

In semiconductor devices

Degenerate semiconductors are integral to tunnel diodes, devices that exploit band-to-band tunneling in heavily doped p-n junctions to achieve negative differential resistance. In these structures, the Fermi level in the p-type region extends into the valence band, while in the n-type region it enters the conduction band, creating overlapping band edges that allow direct quantum tunneling of electrons from the valence band of the p-side to the conduction band of the n-side under forward bias. This tunneling current initially increases with voltage until the bands misalign, causing the current to decrease and producing the characteristic negative resistance region. The seminal Esaki diode, developed using degenerate germanium junctions with doping concentrations exceeding 10^{19} cm^{-3}, demonstrated this effect with a peak current density of approximately 1000 A/cm² and a peak-to-valley ratio up to 100, enabling applications in high-frequency oscillators and amplifiers. Heavily doped degenerate regions are also critical for ohmic contacts in semiconductor devices, where they reduce effects at metal-semiconductor interfaces by promoting tunneling over . In n-type contacts, for instance, degenerate doping narrows the depletion width to less than 10 nm, allowing electrons to tunnel through the barrier with minimal resistance, resulting in linear current-voltage characteristics and specific contact resistivities as low as 10^{-7} Ω·cm² for materials like n-GaAs. This tunneling-dominated transport ensures non-rectifying behavior, essential for efficient current injection in diodes, transistors, and integrated circuits, where undoped or lightly doped interfaces would otherwise introduce significant voltage drops. In high-speed transistors such as MOSFETs, degenerate doping of and extensions significantly lowers parasitic series resistance, improving overall device performance in scaled technologies. For FinFETs at sub-5 nm nodes, n-type /drain regions doped to 10^{20} cm^{-3} or higher reduce external resistance by up to 50% compared to non-degenerate cases, enabling higher drive currents (over 1 mA/μm) and faster switching speeds exceeding 100 GHz while mitigating short-channel effects. This approach, often implemented via raised /drain structures or in-situ doping during epitaxial growth, is vital for maintaining electrostatic integrity and boosting on-state conductance in advanced nodes like 3 nm FinFETs used in logic processors.

Advanced uses in nanotechnology

Modulation doping in and quantum wells, including degenerate regimes at higher concentrations, enables precise control over the , facilitating advanced optoelectronic devices such as quantum dot lasers with enhanced performance characteristics. In InAs/GaAs quantum dot structures, p-type modulation doping at concentrations up to ~10^{18} cm^{-3} in the active region or spacer layers can induce degeneracy, shifting the quasi-Fermi level to favor in excited states, improving temperature stability and reducing threshold current density in lasers operating at wavelengths. This approach has demonstrated CW lasing up to 105–167°C in p-doped designs, by mitigating carrier escape from the dots through filled states. In , degenerate ferromagnetic semiconductors like (Ga,Mn)As serve as efficient spin injectors into non-magnetic semiconductors, leveraging high hole densities to mediate carrier-induced . With Mn concentrations of 5-8 at.%, (Ga,Mn)As exhibits p-type degeneracy where the enters the valence band, enabling spin-polarized current injection into GaAs with polarizations up to 50% at low temperatures via structures. Efforts to enhance the , typically around 110-200 K in optimized samples, include co-doping and nanostructuring, which strengthen exchange interactions and have achieved values approaching in thin films, supporting applications in spin-field-effect transistors and magnetoresistive devices. For two-dimensional materials, degenerate doping via electrostatic gating or chemical methods in and MoS_2 unlocks plasmonic functionalities rooted in their unique electronic structures. In gated , Fermi levels shifted beyond 0.5 eV above the Dirac point induce degenerate carrier densities exceeding 10^{13} cm^{-2}, supporting tunable Dirac plasmons with wavelengths in the mid-infrared for applications in modulators and detectors, as demonstrated in plasmonic waveguides. Similarly, n-type degenerate doping in MoS_2, achieved through vacuum annealing to carrier densities around 10^{13} cm^{-2}, reveals two-dimensional plasmonic polarons—hybrid quasiparticles of plasmons and phonons—enabling enhanced light-matter interactions for nanoscale sensing and in 2020s research.

References

  1. [1]
    [PDF] Degenerate and non-degenerate Semiconductors - Gyan Sanchay
    A degenerate semiconductor is a semiconductor with such a high level of doping that the material starts to act more like a metal than as a semiconductor. What ...
  2. [2]
    [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.
  3. [3]
    None
    ### Summary of Degenerate Semiconductors
  4. [4]
    [PDF] doped GaAs - Purdue e-Pubs
    For doping concentrations greater than 1 x 102' cm - 3, effects due to degenerate Fermi statistics oppose the band-gap shrinkage effects; consequently, a ...<|control11|><|separator|>
  5. [5]
    [PDF] Electrons and Holes in Semiconductors
    Different surface orientations have different properties such as the rate of oxidation and the electronic quality of the oxide/semiconductor interface.
  6. [6]
    Selective Doping in Silicon Carbide Power Devices - PMC - NIH
    In general, for a p-type-doped 4H-SiC semiconductor with a doping ... The data are from [56,61,83,84,86] and refer to high doping levels (>1018 cm−3).
  7. [7]
    Summary of the Basic Free Electron Transport Characteristics in ...
    ... Doped Silicon at Any Doping Level. The estimation of the free RM electron ... On the other hand, at doping densities smaller than 1018 cm−3 all ...<|separator|>
  8. [8]
    Electron depletion at InAs free surfaces: Doping-induced ...
    May 23, 2006 · For almost all n -type zinc-blende III-V semiconductor free surfaces a depletion layer is observed. InAs is an exception, since n -type InAs ...
  9. [9]
    New Insights on the Burstein-Moss Shift and Band Gap Narrowing in ...
    Oct 30, 2015 · ... degenerate semiconductor. A comparison between the band gap energy shift obtained from experimental data and theoretical prediction of the ...
  10. [10]
    Arnold Sommerfeld - Biography - University of St Andrews
    Sommerfeld's approach was to regard electrons in a metal as a degenerate electron gas. He was able to explain features which were unexplained by the earlier ...
  11. [11]
    [PDF] Physics of Semiconductor Devices
    ... Degenerate Semiconductors. As shown in Fig. 8, for degenerate levels where n- or p-concentrations are near or beyond the effective density of states (N, or ...
  12. [12]
    [PDF] NUMERICAL CALCULATIONS OF IMPURITY SCATTERING ...
    It is accepted to assume with error 8 %, that ξ = –2 is the degeneracy boundary for charge carrier gas between degeneracy and no degeneracy state. At ξ < –2 the.
  13. [13]
    Electronic properties of crystalline Ge1-xSbxTey thin films
    Sep 6, 2012 · ... mobility saturation ∝ T0, i.e., ionized impurity scattering in a degenerate semiconductor.25. On the other hand, the interpretation of the ...
  14. [14]
    [PDF] Semiconductor Physics and Devices - OptiMa-UFAM
    ... Doping and Low. Injection 203. 6.3.3 Applications of the Ambipolar Transport ... Fermi–Dirac Probability Function 91. 3.5.3 The Distribution Function and ...
  15. [15]
    None
    Below is a merged summary of the segments from *Advanced Semiconductor Fundamentals* by Robert F. Pierret, consolidating all information into a concise yet comprehensive response. To handle the dense and detailed nature of the data, I will use tables in CSV format where appropriate to organize key concepts, numerical examples, and GaAs-specific details efficiently. The narrative will cover all topics (Degenerate Doping, Fermi Level Position, Carrier Density Calculation, Degeneracy Parameter, Approximations, Numerical Examples, and GaAs Specifics) while referencing the tables for clarity.
  16. [16]
    Approximation for the Fermi–Dirac integral with applications to ...
    To justify its applicability the formula has been used to calculate the effective charge carrier concentration of degenerate heavily doped material.
  17. [17]
    [PDF] Monte Carlo Simulation of Electron Transport in Degenerate ... - arXiv
    Many of today's interesting microelectronic devices are working in high doping concentrations up to 10. 20 cm. -3 . Degenerate semiconductors are important for.
  18. [18]
    [PDF] Lecture 3 - Carrier Statistics in Equilibrium (cont.) February 9, 2007 ...
    Feb 9, 2007 · How can one compute ni? • Where is the Fermi level in a given semiconductor? How does its location depend on doping level? Cite as ...
  19. [19]
    [PDF] 13 Classical and quantum statistics - userhome.brooklyn...
    The range of validity of the approximation is limited to EF – EC ≤ 2 kT, i. e. to weak degeneracy. ... where η = E/kT and ηF = –(EC – EF)/kT are reduced energies.
  20. [20]
    Realizing p-type InSb with enhanced thermoelectric performance via ...
    Jan 23, 2025 · The decrease in σ with increasing temperature indicates that In(1−x)CdxSb (x = 0.0025–0.10) samples are degenerate semiconductors [Fig. 4(a)], ...
  21. [21]
    [PDF] iv ELECTRONIC AND THERMAL BEHAVIOR OF GERMANIUM ...
    𝜎 = n e2𝜏. m∗. (2.13). One often sees the electrical conductivity is expressed as σ = n e µ, where µ is the “electronic mobility” because it ...
  22. [22]
  23. [23]
    Ionized Impurity Scattering of Charge Carriers in Crystalline ...
    It is shown that the Brooks-Herring formula for μBH ∝ T3/2/N overestimates the value of mobility. Comparison of mobility calculation in degenerate crystals with ...
  24. [24]
    [PDF] Electrical transport and the Hall effect in metals and semiconductor
    On the basis of a simple free-electron Fermi gas model, the Hall coefficient is equal to: RH = -1/ne (in SI units), where e is the electron charge and n is the ...
  25. [25]
    [PDF] thermomagnetic effects in an electron gas of semiconductors heated ...
    Anomalous and normal skin effects are investigated in the case of strong fields. Degenerate semiconductors and semimetals are considered separately. It is indi-.
  26. [26]
    Enhanced thermoelectric figure of merit in highly-doped silicon ...
    Dec 15, 2023 · ... the high doping concentration. Consequently, two P-doped CSiNWs with DNW of 223 and 173 nm demonstrate σ values of 1660 and 1515 S·cm–1.
  27. [27]
    (PDF) High Figures of Merit in Degenerate Semiconductors. Energy ...
    Oct 6, 2015 · Heavily boron-doped polycrystalline silicon has been reported to be characterized by somewhat unexpectedly high power factor.
  28. [28]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    the infrared region of the spectrum. This formula also applies to degenerate semiconductors below the plasma frequency. 5.4 Direct Interband Transitions. To ...
  29. [29]
    Electronic excitations stabilized by a degenerate electron gas in ...
    Jul 26, 2018 · Several observations based on Fig. 1 contradict the standard model for the behaviour of highly doped semiconductors. First, the high doping ...
  30. [30]
    Photoluminescence in heavily doped GaAs. I. Temperature and hole ...
    Jul 15, 1980 · The luminescence across the gap gives evidence for impurity states associated with the spin-orbit-split band. At 2.1 K this luminescence ...Missing: quenching degenerate Pauli blocking
  31. [31]
    [PDF] A.8. Degenerate Fermion Systems - UBC Physics
    However the specific heat at low T is still linear in T; the phonons give a contribution ~ T3, as we would expect from a massless bosonic gas, and the ...
  32. [32]
    [PDF] Thermoelectric figure of merit of degenerate and nondegenerate ...
    Jul 1, 2015 · into sharp focus two of the most important characteristics of degenerate semiconductors: (1) The higher the temperature, the better is the ...
  33. [33]
    Ion Implantation Doping in Silicon Carbide and Gallium Nitride ...
    Therefore, ion implantation represents a method of choice to achieve selective doping in SiC devices, and electrical activation is achieved by post- ...
  34. [34]
    Active dopant profiling and Ohmic contacts behavior in degenerate n ...
    Jul 6, 2020 · This Letter reports on the active dopant profiling and Ohmic contact behavior in degenerate P-implanted silicon carbide (4H-SiC) layers.
  35. [35]
    [PDF] Chapter 9: Ion Implantation
    Ion implantation is a low-temperature technique for introducing impurities into semiconductors, where atoms are accelerated and directed at a target.
  36. [36]
    Germanium doping of cubic GaN grown by molecular beam epitaxy
    Mar 6, 2019 · With further increased doping, the peaks merge and spectral broadening occurs, indicating degenerate doping. The 3.7 × 1020 cm−3 doped ...
  37. [37]
    Heavy arsenic doping of silicon by molecular beam epitaxy
    Active Control of Charge Density Waves at Degenerate Semiconductor Interfaces. 2017, Scientific Reports. Vapor phase doping with N-type dopant into silicon by ...
  38. [38]
    High concentration n-type doping in Si layers epitaxially grown by ...
    Classical semiconductor. High concentration n-type doping in Si layers epitaxially grown by ultra-high vacuum chemical vapor deposition with cracking heater.
  39. [39]
    [PDF] Chapter 8: Diffusion
    These three elements are highly soluble in silicon with solubilities exceeding 5 x 1020 atoms / cm3 in the diffusion temperature range (between 800oC and 1200 ...
  40. [40]
    [PDF] Doping control at the atomic scale in III-V semiconductors
    Mar 17, 2023 · N-type doping introduces electrons into the material, shifting the Fermi level up in energy, closer to the conduction band. P- type doping adds ...
  41. [41]
    Electronic transport in two-dimensional Si:P δ -doped layers
    Mar 12, 2013 · We investigate theoretically two-dimensional (2D) electronic transport in Si:P 𝛿 -doped layers limited by charged-dopant scattering.
  42. [42]
    Delta-Doping of Semiconductors - ResearchGate
    [7,9] In such delta-doped semiconductors, 2D electron gases (2DEGs) are formed at the interfaces which are localized within a "V" shape potential well ...
  43. [43]
    Stress relaxation and dopant activation in nsec laser annealed SiGe
    The study investigates damage and strain evolution in SiGe, its impact on dopant distribution, and the efficiency of dopant activation during nanosecond laser ...
  44. [44]
    Segregation and activation of Ga in high Ge content SiGe by UV ...
    Jun 3, 2019 · We used an UV excimer nanosecond laser annealing and studied the segregation and activation of dopants in a Ga-implanted SiGe 50% epilayer.
  45. [45]
  46. [46]
    (PDF) Heavily doped Si: P emitters of crystalline Si solar cells
    Aug 9, 2025 · The high concentration of phosphorus that exceeds solubility limits forms phosphorous precipitates that function as defects, causing the excess ...
  47. [47]
    [PDF] Intrinsic limitations to the doping of wide-gap semiconductors
    It can be relatively easily doped n-type but p-type doping is very difficult to accomplish and only recently doping with reactive nitrogen was successful in ...
  48. [48]
    Suppression of compensating native defect formation during ...
    Jun 21, 2016 · Two donor doping concentrations were considered: undoped and moderately doped (Nd = 1017 cm−3). Additionally, three photon fluxes were applied: ...Missing: solubility precipitation
  49. [49]
  50. [50]
    Rapid thermal annealing of high-melting-point films on ... - NASA ADS
    Rapid thermal annealing which involves fast heating and cooling rates, is used to activate dopants in thin-film structures yet minimize the dopant diffusion ...
  51. [51]
    [PDF] Metal-Semiconductor Contacts - TU Graz
    Ohmic contacts. Thermionic emission. Tunnel contacts. Page 3. Photoelectric ... Degenerate doping at a tunnel contact. Page 23. Contacts. Page 24. Transport ...<|control11|><|separator|>
  52. [52]
    [PDF] Advanced Source/Drain and Contact Design for Nanoscale CMOS
    May 20, 2010 · This work investigates key design issues surrounding DSS. MOSFETs from both a modeling and experimental perspective, including the effect of.
  53. [53]
    [PDF] Quantum Dot Lasers for Silicon Photonics - eScholarship
    Recently, replacing the QW active region with InAs quantum dots (QDs) in GaAs-based ... By introducing just 5×1017 cm-3 doping to a 10 nm GaAs layer in the spacer ...
  54. [54]
    [PDF] GaAs-Based Quantum Dot Lasers
    The highest reported uncooled operating temperatures for InGaAs/GaAs QD lasers have been quite impressive, demonstrating lasing operation beyond 195!C from the.
  55. [55]
    Spin-dependent phenomena and device concepts explored in (Ga ...
    Jul 11, 2014 · This is a review of the basic material properties that make (Ga,Mn)As a favorable test-bed system for spintronics research.
  56. [56]
    Very high spin polarization in GaAs by injection from a (Ga,Mn)As ...
    May 15, 2025 · Electron spin polarization was analyzed by oblique Hanle effect and vanishes at 120 K, the Curie temperature of the (Ga, Mn)As injector. The ...
  57. [57]
    Plasmonics of 2D Nanomaterials: Properties and Applications - Li
    Feb 16, 2017 · The strong reconstruction of the massless Dirac Fermions (MDFs) chiral spectrum near the Dirac point of doped graphene supports that the self- ...Missing: MoS2 | Show results with:MoS2
  58. [58]
    Two-dimensional plasmonic polarons in n -doped monolayer M o S 2
    May 27, 2021 · Degenerate n doping was achieved by repeated in situ vacuum annealing cycles (with final annealing for 12 h at 850 K), which is known to ...Missing: fermion | Show results with:fermion