Fact-checked by Grok 2 weeks ago

Direct and indirect band gaps

In semiconductors, the band gap is the energy difference between the top of the valence band and the bottom of the conduction band, determining the material's electrical and optical properties. Direct and indirect band gaps classify this energy gap based on the alignment of the band extrema in momentum space (k-space). In a direct band gap semiconductor, the maximum of the valence band and the minimum of the conduction band occur at the same crystal momentum (k-vector), allowing electrons to transition between bands with negligible momentum change, typically via photon absorption or emission. This enables highly efficient radiative recombination, where an electron-hole pair recombines to emit a photon with energy approximately equal to the band gap. Examples include gallium arsenide (GaAs, band gap ~1.42 eV at room temperature) and gallium nitride (GaN, ~3.4 eV). In contrast, an indirect band gap semiconductor has its valence band maximum and conduction band minimum at different k-vectors, requiring a change in both and for band-to-band transitions. Such transitions cannot occur solely with a , as photons carry negligible (p = E/c, where c ≈ 3 × 10^8 m/s and E ≈ 10^{-19} J for typical band gaps), necessitating assistance from a ( vibration) to conserve . This three-particle process (, ) makes radiative transitions much less probable, often resulting in non-radiative recombination via heat. Common examples are (Si, band gap ~1.12 ) and (Ge, ~0.67 ). The distinction profoundly impacts applications: direct band gap materials excel in , powering light-emitting diodes (LEDs), diodes, and efficient photodetectors due to short lengths (~1 μm) and strong . Indirect band gap materials, while dominant in electronics (e.g., in integrated circuits), require thicker layers (hundreds of μm) for sufficient light in cells and are unsuitable for light without enhancements. Band gap engineering, such as through alloying or , can sometimes shift materials from indirect to direct types, broadening their utility in .

Basic Concepts

Electronic Band Structure

In crystalline solids, the electronic structure is governed by the periodic arrangement of atoms, which creates a periodic potential for the electrons. According to , the wavefunctions of electrons in such a potential can be expressed as plane waves modulated by a with the same periodicity as the . Mathematically, the Bloch wavefunction takes the form \psi_{\mathbf{k}}(\mathbf{r}) = e^{i \mathbf{k} \cdot \mathbf{r}} u_{\mathbf{k}}(\mathbf{r}), where \mathbf{k} is the crystal wavevector, and u_{\mathbf{k}}(\mathbf{r}) is periodic with the lattice periodicity. This form arises from the translational symmetry of the crystal, allowing solutions to the Schrödinger equation to be labeled by \mathbf{k} within the first Brillouin zone. The energy eigenvalues associated with these Bloch states form continuous bands as a function of \mathbf{k}, due to the delocalization of electrons across the lattice. In solids, these bands originate from the overlap and hybridization of atomic orbitals from neighboring atoms. For instance, in a simple one-dimensional chain, s-orbitals from adjacent atoms split into bonding and antibonding combinations, broadening into a (filled with electrons) below the and a conduction band (empty) above it. This band formation is a quantum mechanical consequence of the tight-binding model, where the extent of overlap determines the . The band gap is defined as the energy difference between the highest occupied state at the top of the valence band and the lowest unoccupied state at the bottom of the conduction band, typically on the order of a few volts in semiconductors and insulators. This gap arises from the forbidden region where no electron states exist, separating the filled valence band from the empty conduction band at temperature. The foundational understanding of electronic band structure emerged in the late 1920s and 1930s, with Felix Bloch's 1928 thesis introducing the theorem for electrons in periodic potentials, followed by extensions from and others that incorporated zone boundaries and nearly models. A typical band structure plot displays the E(\mathbf{k}) along high-symmetry directions in the , often showing parabolic-like minima or maxima near zone center or edges, with the valence band curving upward to its maximum and the conduction band starting from its minimum, separated by the band gap. Such plots, computed via methods like , illustrate how electron effective mass varies with band curvature, \frac{1}{m^*} = \frac{1}{\hbar^2} \frac{d^2 E}{d k^2}.

Definition of Direct and Indirect Band Gaps

In physics, the refers to the energy difference between the top of the valence band and the bottom of the conduction band. A direct band gap occurs when the maximum of the valence band and the minimum of the conduction band align at the same wavevector k in the , typically at the Γ point (k = 0). This alignment allows for vertical transitions in the energy-momentum (E-k) diagram, where an electron can move from the valence to the conduction band without a change in crystal . In contrast, an indirect band gap arises when the valence band maximum and conduction band minimum occur at different k-points, necessitating a change in for interband transitions. Schematic representations in E-k diagrams illustrate this distinction: for direct band gaps, the band edges form a vertical step, enabling straightforward optical transitions; for indirect band gaps, the edges are offset, resulting in slanted transitions that require additional momentum compensation, often involving phonons. Representative examples include gallium arsenide (GaAs), a direct band gap semiconductor with an energy gap of approximately 1.42 eV at the Γ point, and silicon (Si), an indirect band gap material with a minimum gap of about 1.12 eV between the Γ and X points in its diamond lattice structure. The type of band gap is influenced by factors such as material composition and lattice structure; for instance, polar compounds like those in the zincblende structure (e.g., GaAs) tend toward direct gaps due to stronger ionic bonding effects, whereas more covalent diamond-structured materials (e.g., ) favor indirect gaps. Alloying or can further tune these properties by shifting band extrema relative to each other in .

Momentum Conservation in Transitions

Role of Crystal Momentum

In solid-state physics, crystal momentum, denoted as \hbar \mathbf{k}, represents the quasi-momentum of electrons within the periodic potential of a crystal lattice, confined to the first Brillouin zone as per Bloch's theorem. This concept arises from the wave-like nature of electrons in a crystal, where the wavefunction takes the form \psi_{n\mathbf{k}}(\mathbf{r}) = e^{i\mathbf{k}\cdot\mathbf{r}} u_{n\mathbf{k}}(\mathbf{r}), with \mathbf{k} labeling states in the reciprocal lattice. The conservation of crystal momentum serves as a fundamental selection rule for electronic transitions, particularly those induced by photons in optical processes across the band gap. In direct band gap materials, where the conduction band minimum and valence band maximum occur at the same \mathbf{k}-point (typically the zone center \Gamma), vertical transitions in \mathbf{k}-space are allowed. Photons carry negligible momentum (\hbar \mathbf{k}_\gamma \approx 0) compared to electrons, enabling momentum-conserving transitions without additional carriers of momentum. Consequently, the transition probability is proportional to a delta function enforcing conservation: \delta(\mathbf{k}_c - \mathbf{k}_v - \mathbf{k}_\gamma), where \mathbf{k}_c and \mathbf{k}_v are the wavevectors in the conduction and valence bands, respectively, and \mathbf{k}_\gamma \approx 0. In indirect band gap semiconductors, the band extrema occur at different \mathbf{k}-points, violating direct momentum conservation for photon-induced transitions. This prohibition arises because the dipole matrix element for electric-dipole transitions, \langle c | \hat{\mathbf{e}} \cdot \mathbf{p} | v \rangle, is nonzero only when \Delta \mathbf{k} = 0, as the perturbation from the photon's acts locally within the unit cell. As a result, pure photon-mediated transitions across indirect gaps are inefficient, often requiring assistance to bridge the momentum mismatch.

Phonon Involvement

Phonons, the quantized collective excitations of vibrations in a , carry \mathbf{q} and are essential for enabling optical transitions in indirect semiconductors, where the valence band maximum and conduction band minimum occur at non-equivalent wavevectors \mathbf{k} in the . Unlike direct transitions, where conservation is approximately satisfied by the negligible of the , indirect transitions require participation to supply the required difference \Delta \mathbf{k}. This involvement occurs through electron- interactions, modeled via , allowing the overall process to conserve both and . In indirect light absorption, an is excited from the to the conduction via the simultaneous of a and either the or of a , bridging the \Delta \mathbf{k} mismatch; for example, facilitates upward transitions in space when the initial state requires additional \mathbf{q}. Conversely, in radiative recombination, an electron-hole pair annihilates, emitting a and involving or to ensure matching between the recombining carriers. These processes are symmetric in their mechanics but differ in the direction of energy flow. The energy balance in these transitions is adjusted by the phonon energy, yielding effective thresholds of E_g \pm \hbar \omega_{\text{phonon}}, where E_g is the indirect band gap and \hbar \omega_{\text{phonon}} typically ranges from a few meV for acoustic phonons to 10–60 meV for optical phonons, introducing weak sidebands in the absorption or emission spectra. Phonons are categorized as acoustic (with linear dispersion and low energy, predominant in intravalley scattering within the same conduction or valence band valley) or optical (with higher energy and nearly flat dispersion, key for intervalley scattering between distinct valleys, as seen in silicon's \Gamma-to-X transitions). The probability of phonon-assisted transitions depends strongly on through the Bose-Einstein occupation factor for s, given by n(\omega) = \frac{1}{\exp(\hbar \omega / k_B T) - 1}, where processes involving phonon absorption scale with n(\omega) (reflecting the available phonon population) and those with emission scale with n(\omega) + 1 (accounting for ); thus, higher temperatures increase the phonon density, enhancing overall transition rates in indirect semiconductors.

Optical Processes

Light Absorption

Light absorption in semiconductors with direct and indirect band gaps exhibits distinct characteristics due to the momentum conservation requirements in optical transitions. In direct band gap materials, such as (GaAs), vertical transitions at the same crystal momentum k allow for strong absorption via electric dipole interactions, resulting in a sharp onset of the absorption coefficient α(ω) near the band gap energy E_g. Conversely, indirect band gap materials like (Si) require phonon assistance to conserve momentum, leading to weaker absorption and a more gradual edge often featuring an Urbach tail, an exponential increase in absorption below E_g attributed to disorder or thermal effects. For direct band gaps, the absorption process involves allowed electric dipole transitions between the valence band maximum and conduction band minimum at the same k-point, typically at the Brillouin zone center (Γ point). The absorption coefficient near the band edge follows a square-root dependence, α(ω) ∝ (ħω - E_g)^{1/2}, arising from the joint density of states (JDOS) for three-dimensional parabolic bands. To derive this, consider the transition rate from time-dependent perturbation theory, where the absorption is proportional to the matrix element |<c| \mathbf{e} \cdot \mathbf{p} |v>|^2 (with \mathbf{p} the momentum operator and \mathbf{e} the light polarization) times the JDOS. For parabolic bands, E_c(k) = E_g + ħ^2 k^2 / (2 m_e^) and E_v(k) = - ħ^2 k^2 / (2 m_h^), the energy conservation ħω = E_c(k) - E_v(k) yields a spherical surface in k-space. The JDOS ρ(ω) is then obtained by integrating over this surface: \rho(\omega) = \frac{1}{4\pi^2} \left( \frac{2\mu}{\hbar^2} \right)^{3/2} (\hbar \omega - E_g)^{1/2}, where μ = (m_e^* m_h^) / (m_e^ + m_h^*) is the reduced effective mass. The full absorption coefficient includes a constant prefactor involving the matrix element and is thus α(ω) = (C / ω) (ħω - E_g)^{1/2} for ħω > E_g, with C encapsulating material parameters. In indirect band gap semiconductors, direct transitions are forbidden by momentum conservation, so absorption is phonon-assisted, involving the absorption or emission of a phonon with energy ħω_ph to bridge the k-mismatch between band extrema. This results in a weaker coefficient with quadratic dependence, α(ω) ∝ (ħω - E_g ± ħω_ph)^2, where the ± accounts for phonon absorption or emission, respectively. The derivation parallels the direct case but incorporates the electron-phonon interaction vertex, leading to a JDOS modified by phonon dispersion; for low temperatures, the dominant term involves phonon emission, and the squared dependence emerges from the 3D parabolic band integration with momentum offset Δk between extrema. This phonon involvement makes indirect absorption typically orders of magnitude smaller than direct, with characteristic energies ħω_ph ~ 10-60 meV in common materials. Experimentally, the type of band gap is determined by measuring the absorption coefficient via transmittance , where thin samples allow extraction of α(ω) = (1/d) ln(I_0 / I), with d the thickness and I, I_0 the transmitted and incident intensities. Spectroscopic provides a model-independent approach by fitting the complex function ε(ω) = ε_1 + i ε_2, where Im[ε(ω)] relates to α(ω) = (ω / c) Im[√ε(ω)], revealing sharp critical points for gaps or broader features for indirect ones; for example, in Ge_{1-x}Sn_x alloys, distinguishes the crossover from indirect to gaps as Sn content increases.

Radiative Recombination

Radiative recombination is the process in which an in the conduction annihilates with a in the valence , releasing in the form of a whose corresponds to the . This direct band-to-band transition is the primary mechanism for light emission in semiconductors. In direct band gap materials, such as (GaAs), radiative recombination occurs with high efficiency due to favorable momentum conservation, allowing vertical transitions in without additional assistance. High-quality GaAs exhibits internal quantum efficiencies exceeding 90% under appropriate conditions, with minority radiative lifetimes on the order of nanoseconds (τ_rad ≈ 1–10 ns). In contrast, indirect band gap materials like silicon (Si) have low radiative efficiency, typically less than 0.1%, because momentum conservation requires participation to bridge the k-space mismatch between the band extrema, making the process improbable and dominated by non-radiative pathways. The rate of bimolecular radiative recombination is described by the equation R_{\text{rad}} = B n p, where n and p are the and concentrations, respectively, and B is the radiative recombination coefficient. For band gap semiconductors, B ranges from $10^{-9} to $10^{-11} cm³/s, which is orders of magnitude larger than in indirect materials ($10^{-13} to $10^{-15} cm³/s), reflecting the enhanced transition probability. The quantum yield \phi, representing the fraction of carrier recombinations that result in photon emission, is given by \phi = \frac{\tau_{\text{non-rad}}}{\tau_{\text{rad}} + \tau_{\text{non-rad}}}, where \tau_{\text{non-rad}} is the non-radiative lifetime. In indirect band gap semiconductors, low \phi due to short \tau_{\text{non-rad}} necessitates defect engineering strategies, such as reducing point and extended defects, to suppress non-radiative recombination and enhance radiative efficiency. Temperature influences radiative recombination through the band gap energy E_g(T), which decreases with increasing according to the Varshni relation: E_g(T) = E_g(0) - \frac{\alpha T^2}{T + \beta}, where \alpha and \beta are material-specific constants. This variation shifts the emission wavelength to longer values () and can modulate recombination rates, with indirect materials showing particularly strong temperature dependence due to involvement.

Applications and Implications

Optoelectronic Devices

Optoelectronic devices, such as light-emitting diodes (LEDs) and photodetectors, rely heavily on the nature of the to achieve efficient light emission and detection. Direct materials facilitate vertical electronic transitions without momentum mismatch, enabling high radiative recombination rates essential for and sensitive absorption. In contrast, indirect materials require assistance, leading to lower efficiency in optical processes but offering advantages in certain spectral ranges. This distinction profoundly influences device design, with direct semiconductors dominating applications requiring strong light-matter interactions. In LEDs and lasers, direct band gap materials like (GaN) and (InP) are preferred for their efficient , as electrons and holes recombine radiatively with minimal non-radiative losses. For instance, GaN-based blue LEDs achieve high brightness through direct transitions, powering displays and lighting. Conversely, indirect band gap (Si) exhibits poor performance in such devices due to inefficient light emission, with radiative efficiency often requiring structures to enhance carrier confinement and indirect-to-direct-like behavior. The historical shift toward III-V direct band gap semiconductors in the 1960s, pioneered by developments in GaAsP and GaP epitaxial growth, revolutionized by enabling practical visible LEDs beyond inefficient early Si-based attempts. Photodetectors also benefit from direct s for fast and sensitive responses, as the strong coefficient allows thin active layers and high in detecting photons near the band edge. Materials like InGaAs, with direct s, provide near-unity for near- detection. Indirect band gap materials, such as , offer broader spectral coverage into the but suffer from lower due to weaker , necessitating thicker layers that can introduce response delays. This trade-off guides material selection, with direct gaps favored for high-speed applications like fiber-optic receivers. Engineering approaches, such as strain or alloying, address limitations of indirect materials by inducing direct band gap transitions. For example, tensile strain in () reduces the direct relative to the indirect one, enabling light emission in otherwise inefficient group IV semiconductors. Such techniques, including biaxial strain via epitaxial growth on Si substrates, have demonstrated direct behavior in strained , paving the way for integrated photonic devices. Performance metrics highlight the impact: direct LEDs often exceed 80% internal (IQE), reflecting efficient radiative recombination, while indirect Si-based emitters typically achieve less than 1% IQE without enhancements. This disparity underscores the radiative recombination efficiency's role in device viability.

Photovoltaic Cells

In photovoltaic cells, the type of band gap significantly influences light absorption efficiency and overall device design. Direct band gap materials exhibit strong absorption coefficients, typically on the order of 10^4 to 10^5 cm⁻¹ near the band edge, allowing for efficient capture of photons in ultra-thin absorber layers. For instance, copper indium gallium selenide (CIGS) solar cells, which utilize a direct band gap of approximately 1.1–1.2 eV, achieve sufficient absorption in thicknesses of 1–2 μm, enabling lightweight, flexible thin-film devices with reduced material usage. In contrast, indirect band gap semiconductors like silicon (Si), with a band gap of 1.12 eV, have much weaker absorption coefficients (around 10^2–10^3 cm⁻¹), necessitating thicker absorbers—typically 100–200 μm—to capture a comparable fraction of the solar spectrum and minimize transmission losses. This difference arises from the momentum mismatch in indirect transitions, requiring phonon assistance and resulting in lower probability for optical absorption. The Shockley-Queisser provides a theoretical for single-junction solar cell , capping it at approximately 33% for a of 1.1 eV under AM1.5 illumination, assuming between and radiative recombination. Direct band gap materials more readily approach this due to their efficient above-band-gap and minimal non-radiative losses in optimized structures, as the strong facilitates high . Indirect band gap materials, however, face practical challenges in reaching this , including parasitic free-carrier in thicker layers and increased non-radiative recombination volumes, which introduce additional losses beyond the radiative . For cells, these factors contribute to real-world efficiencies of 20–25%, well below the theoretical maximum despite mature processing. To surpass single-junction limitations, multi-junction solar cells stack layers with progressively narrower direct s to split the solar and minimize thermalization losses. High-efficiency designs often employ direct III-V semiconductors, such as (GaAs, 1.42 ) for the middle junction and gallium indium phosphide (GaInP, ~1.8 ) for the top, enabling current matching and external quantum efficiencies exceeding 80% across visible wavelengths. These lattice-matched structures, typically grown via metalorganic vapor-phase epitaxy, have achieved efficiencies over 30% under one-sun illumination and up to 47% under concentration, far outperforming single-junction indirect cells by better utilizing the full AM1.5 . Recent advances since 2010 have leveraged direct hybrid perovskites, such as methylammonium lead (MAPbI₃) with a of ~1.55 eV, to enhance photovoltaic performance through solution-processable thin films (300–500 nm thick) that exhibit direct transitions with high coefficients (>10^5 cm⁻¹). These materials have driven single-junction efficiencies beyond 25%, with certified records reaching 26.7% as of 2025, owing to low binding energies and defect tolerance that promote long carrier diffusion lengths. When integrated into tandems with bottom cells, perovskite/ hybrids exploit the direct for top-cell capture, reaching overall efficiencies up to 34.9% as of 2025. Despite these advantages of direct band gap materials, indirect band gap continues to dominate the photovoltaic market, holding approximately 95% share as of 2025 due to its abundance, low cost (from established ), and reliable scalability. This prevalence persists even though 's weaker demands thicker, more material-intensive cells, underscoring a where economic and infrastructural factors outweigh optical limitations in commercial deployment.

References

  1. [1]
    Band Gap – dielectrics, semiconductors, metals, energy, electronic ...
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same crystal momentum ( k vector). In an ...Semiconductors · Methods for Measuring Band... · Effects of Doping on the Band...
  2. [2]
    Direct and Indirect Band Gap Semiconductors - DoITPoMS
    In a direct band gap semiconductor, the top of the valence band and the bottom of the conduction band occur at the same value of momentum.
  3. [3]
    22.5: Direct and Indirect Band Gap Semiconductors
    ### Summary of Direct and Indirect Band Gap Semiconductors
  4. [4]
    Basics of Semiconductors - BYU Physics & Astronomy
    In semiconductors it is the difference in energy between the lowest energy state in the conduction band and the highest energy state in the valence band. It ...Missing: explanation | Show results with:explanation
  5. [5]
    [PDF] About the Quantum Mechanics of the Electrons in Crystal Lattices
    Nov 29, 2018 · Abstract: This article aims to review Felix Bloch theorem of electron motion in a crystal lattice through his seminal paper that has also acted ...
  6. [6]
    5.1: Introduction to Bloch's Theorem - Chemistry LibreTexts
    Apr 4, 2024 · Bloch's theorem identifies the important features of basis functions for the group of lattice translation operations and creates a foundation for solving Schrö ...
  7. [7]
    2.2: Bands of Orbitals in Solids - Chemistry LibreTexts
    Apr 8, 2021 · The bands of orbitals arising in any solid lattice provide the orbitals that are available to be occupied by the number of electrons in the crystal.
  8. [8]
    Introduction to Energy Bands - DoITPoMS
    When two valence electron atomic orbitals in a simple molecule such as hydrogen combine to form a chemical bond, two possible molecular orbitals result. One ...
  9. [9]
    Band Gap | PVEducation
    The band gap (EG) is the gap in energy between the bound state and the free state, between the valence band and conduction band. Therefore, the band gap is the ...
  10. [10]
    Topological Quantum Chemistry, the band theory of solids is now ...
    Jul 20, 2017 · Extended and refined by Bloch and others during the 1930s, Bloch's theory, known as the band theory of solids, accounts very well for the ...
  11. [11]
    [PDF] The development of the quantum-mechanical electron theory of metals
    a metal that took the ions into account, Bloch's paper laid the foundations of band theory. As Bloch showed, the electron energy-momentum relation was no longer.
  12. [12]
    [PDF] Band Structures and the Meaning of the Wave Vector k - UCSB MRL
    Band structures are a representation of the allowed electronic energy levels of solid materials and are used to better inform their electrical properties.
  13. [13]
    [PDF] Semiconductors and Insulators - Physics Courses
    6.1.4 Direct versus indirect gaps. Fig. 6.5 shows the cases of direct and indirect gap semiconductors. In a direct gap material, the conduction band minimum ...
  14. [14]
    [PDF] 29 Band structures - David Miller
    The center is the Γ (gamma) point. X. L k x k y k z. Page 35. Γ. Band structures ... Note that GaAs is a direct gap semiconductor unlike Si which is ...
  15. [15]
    Electronic Properties of Materials - Physics of Semiconductors
    Nov 7, 2017 · AlGaAs: (Known as Al gaas) GaAs is a direct bandgap semiconductor, and AlAs is an indirect bandgap semiconductor. These two can be combined to ...
  16. [16]
    Silicon Solar Cells - Stanford University
    Nov 13, 2015 · Silicon, on the other hand, has an indirect band gap of 1.1 eV. Silicon is not the ideal solar cell, but it provides several advantages: silicon ...
  17. [17]
    [PDF] Lecture 3 Introduction to Semiconductors and Energy Bandgaps
    Later we will see that covalent bonds tend toward “indirect bandgap” (defined later) materials whereas polar bonds (ionic and partially ionic) tend toward “ ...
  18. [18]
    Bandgap engineering of two-dimensional semiconductor materials
    Aug 24, 2020 · We provide a review of the basic physical principles of these various techniques on the engineering of quasi-particle and optical bandgaps.
  19. [19]
    [PDF] Fundamentals of Semiconductors - Sites do IFGW
    ... Yu. Manuel Cardona. Fundamentals of Semiconductors. Physics and Materials Properties. Fourth Edition. 123. Page 5. Professor Dr. Peter Y. Yu. University of ...
  20. [20]
    [PDF] Optical Transitions in Semiconductors - Cornell University
    Note that the momentum conservation principle is stated in terms of the crystal momentum of the electrons. This principle will be derived later. E. ECE 407 ...
  21. [21]
    [PDF] 32 Optical absorption in semiconductors - David Miller
    correspond to conservation of crystal momentum ). In this case, the sum ... the direct optical transitions are essentially “vertical” on the energy ...
  22. [22]
  23. [23]
    [PDF] Phonon-assisted optical processes
    → Atomic vibrations provide the additional momentum to enable phonon-assisted optical transitions across the indirect band gap of silicon, and enable the ...
  24. [24]
  25. [25]
    Electron-phonon interaction effects on the direct gap transitions of ...
    The Bose-Einstein distribution predicts that the average number of phonons increases with increase in temperature and thereby increasing the e-p interactions.
  26. [26]
    [PDF] SOLID STATE PHYSICS PART II Optical Properties of Solids - MIT
    5.5 Conservation of Crystal Momentum in Direct Optical Transitions . . . . . . ... • Yu and Cardona, Fundamentals of Semiconductors, pp. 251-258. • Madelung ...
  27. [27]
    URBACH'S TAIL IN INDIRECT BAND-GAP SEMICONDUCTORS
    It is established that the optical absorption edge in these materials obeys Urbach's Rule. It is also found that the structural disorder in monocrystalline ...
  28. [28]
    [PDF] Lecture 6 Absorption and Photo-generation
    Light with photon energy, hν < Eg is not easily absorbed. A convenient expression for the energy of light is E=1.24/λ where λ is the wavelength of the light ...
  29. [29]
    [PDF] RECOMBINATION PROCESSES IN SEMICONDUCTORS
    ... HALL, L. H.: 'Indirect Transitions from the Valence to the Conduction Bands', Photoconductivity. Conference (John Wiley, 1956), p. 146. (19) MACFARLANE ...
  30. [30]
    Fundamental band gap and direct-indirect crossover in Ge 1
    Nov 11, 2013 · The lowest direct band gap was measured using spectroscopic ellipsometry ... Direct measurements of the indirect gap for samples lattice ...
  31. [31]
    Types of Recombination - PVEducation
    Radiative recombination is the recombination mechanism that dominates in direct bandgap semiconductors. ... (GaAs etc) and radiative recombination dominates.<|separator|>
  32. [32]
    Minority carrier diffusion length, lifetime and mobility in p-type GaAs ...
    Sep 19, 2017 · With respect to the higher doped samples, the radiative lifetime is estimated to be 6 ± 3 ns for these samples. This is just slightly longer ...INTRODUCTION · Cathodoluminescence setup... · III. RESULTS AND DISCUSSION
  33. [33]
    Silicon LED strides ahead - Physics World
    Aug 22, 2001 · Silicon LEDs currently have efficiencies in the range 0.01 to 0.1%, but the device developed by Green and co-workers exceeds a value of 1 ...
  34. [34]
    Radiative Recombination - an overview | ScienceDirect Topics
    Values range from 10− 9 to 10− 11 cm3/s for direct bandgap materials and from 10− 13 to 10− 15 cm3/s for indirect bandgap materials.
  35. [35]
    Point and extended defects engineering as a key to advancing the ...
    The luminescence efficiency is determined by the ratio between the rates of radiative and nonradiative recombination. ... indirect semiconductors, for example, to ...
  36. [36]
    Temperature dependence of the energy gap in semiconductors
    A relation for the variation of the energy gap (Eg) with temperature (T) in semiconductors is proposed. Eg ≐ E0 - αT2/( where α and β are constants.
  37. [37]
    III-V Semiconductor - an overview | ScienceDirect Topics
    The binary nitride semiconductors of group-III, such as GaN, AlN, and InN, show direct bandgap. At room temperature, the bandgap of GaN, AlN, and InN are 3.4 ...
  38. [38]
    III-V Semiconductors - Beneq
    ... LEDs and lasers. III-V semiconductors, like GaN, GaAs, and InP, have a “direct” bandgap. Direct bandgaps require only one pathway for radiative ...
  39. [39]
    Comparing GaAs, InP, and GaN in RF and Photonics Applications
    Jul 8, 2025 · InP's direct bandgap enables the production of photonic devices operating at longer wavelengths, a necessity for long-distance communication.
  40. [40]
    High-efficiency silicon light emitting diodes - ScienceDirect.com
    The reason usually given for silicon's poor performance is that silicon is an indirect bandgap material and hence optical processes are weak, involving at least ...Missing: inefficiency | Show results with:inefficiency
  41. [41]
  42. [42]
    Why are direct band gap materials preferred over indirect ... - Quora
    Jul 22, 2018 · A direct gap semiconductor can generate photons with less heat produced because it doesn't generate a phonon with each electron-hole photo-recombination.
  43. [43]
    What devices work better using indirect bandgap semiconductors?
    Jul 5, 2017 · In an indirect bandgap semiconductor, a photon must couple to a phonon to be absorbed. This makes the absorption coefficient much lower (and ...
  44. [44]
    Direct-bandgap light-emitting germanium in tensilely strained ...
    Theoretical studies have predicted that tensile strain in Ge lowers the direct energy bandgap relative to the indirect one.
  45. [45]
    Direct bandgap emission from strain-doped germanium - Nature
    Jan 19, 2024 · We propose incorporating Li atoms into the Ge to cause lattice expansion to achieve the desired tensile strain for a transition from an indirect to a direct ...
  46. [46]
    Near-Unity Internal Quantum Efficiency of Luminescent Silicon ...
    Jun 17, 2015 · (4-6) Despite the indirect band gap of bulk silicon, strong visible luminescence is observed when reduced to nanometer size scale ...
  47. [47]
    CIGS Solar Cells Overview - PVEducation.Org
    The CIGS material has a direct bandgap and high absorption coefficient. Efficient sunlight absorption can be achieved in CIGS layers as thin as 1 µm, 100 times ...
  48. [48]
    Silicon Solar Cell Parameters | PVEducation
    In particular, silicon's band gap is slightly too low for an optimum solar cell and since silicon is an indirect material, it has a low absorption co-efficient.Missing: absorber | Show results with:absorber
  49. [49]
    Detailed Balance Limit of Efficiency of p‐n Junction Solar Cells
    The maximum efficiency is found to be 30% for an energy gap of 1.1 ev and fc = 1. Actual junctions do not obey the predicted current‐voltage relationship, and ...
  50. [50]
    Band Gap Engineering of Multi-Junction Solar Cells - Nature
    May 11, 2017 · Our results demonstrate that appropriate bandgap engineering may lead to significantly higher conversion efficiency at illumination levels above ~1000 suns.
  51. [51]
    Photovoltaics Reaching for the Shockley–Queisser Limit
    Sep 1, 2020 · The much lower effective bandgap (1.4 vs 1.6 eV) in the new organic materials also creates potential for higher achievable efficiency ...
  52. [52]
    [PDF] Direct or Indirect? Unraveling the bandgap nature of metal halide ...
    This thesis explores the bandgap nature of metal halide perovskites, specifically whether they are direct or indirect.
  53. [53]
    How crystalline silicon will dominate global energy by 2050 - PV Tech
    May 6, 2025 · Crystalline silicon technology has become the industry standard, accounting for roughly 95% of the global PV market.