Fact-checked by Grok 2 weeks ago

Band offset

Band offset refers to the discontinuity in the at the between two different materials, manifesting as differences in the energy levels of the conduction band minimum (ΔE_c) and valence band maximum (ΔE_v). These offsets determine the relative alignment of the bands across the and are fundamental to the behavior of charge carriers at such interfaces. In heterostructures, band offsets lead to three primary types of band alignments: Type I (straddled), where both the conduction and bands of one material lie within of the other, confining carriers to the narrower- region; Type II (staggered), where the bands overlap partially, promoting spatial separation of electrons and holes; and Type III (broken-), where the band maximum of one material exceeds the conduction band minimum of the other, enabling interband tunneling. The band offset (VBO) and conduction band offset (CBO) are related by the difference in band s of the two materials (ΔE_g = ΔE_c + ΔE_v), with typical values ranging from 0.1 to 1 eV depending on the material pair. Band offsets play a critical role in the design and performance of optoelectronic and electronic devices, such as light-emitting diodes (LEDs), solar cells, and high-electron-mobility transistors (HEMTs), by influencing carrier confinement, injection barriers, and recombination processes. For instance, a large conduction band offset (>1 eV) is essential for gate dielectrics in metal-oxide-semiconductor field-effect transistors (MOSFETs) to minimize leakage currents, while Type II alignments enhance charge separation in photovoltaic heterojunctions. Experimental determination of offsets often involves techniques like (XPS) or internal photoemission, while theoretical predictions rely on (DFT) calculations. Early models for predicting band offsets include the Anderson rule (1960s), which estimates ΔE_c as the difference in electron affinities of the two semiconductors, assuming vacuum level alignment, though it often deviates from experiments due to interface dipoles and charge transfer effects. More advanced approaches, such as the charge neutrality level model, account for interface states and differences to align a reference across the junction, improving accuracy for lattice-matched systems. These models, refined through first-principles computations, remain essential for engineering novel heterostructures in modern .

Basic Concepts

Definition

In semiconductor physics, band offset refers to the discontinuity in the energy levels of the valence band maximum (ΔE_v) and conduction band minimum (ΔE_c) at the interface between two different , forming a . These discontinuities arise due to differences in the electronic structures of the materials and play a crucial role in determining confinement and transport properties in heterostructure devices. A is an interface between two dissimilar , in contrast to a homojunction, which occurs within a single material where the band structure is continuous except for doping-induced changes. The valence band offset (VBO), denoted as ΔE_v, represents the difference in valence band maximum positions, while the conduction band offset (CBO), denoted as ΔE_c, represents the difference in conduction band minimum positions across the interface. These offsets are related to the bandgaps of the two , E_{g1} and E_{g2}, through the equation \Delta E_c + \Delta E_v = E_{g1} - E_{g2} (assuming the wider-bandgap material is labeled as 1, with E_{g1} > E_{g2}, and offsets taken as positive magnitudes), which follows from the alignment of band edges relative to a common reference such as the vacuum level. The valence band offset can be expressed as \Delta E_v = E_{v2} - E_{v1}, where E_v denotes the energy of the valence band edge relative to the vacuum level for each semiconductor. This formulation assumes no interface dipole effects and aligns the materials based on their intrinsic properties. The concept of band offset was first introduced by R. L. Anderson in 1960 as a fundamental parameter for understanding carrier behavior in heterojunctions, particularly for enabling confinement in devices like transistors and optoelectronics.

Band Diagrams at Interfaces

Band diagrams at heterojunction interfaces are constructed by plotting the conduction band edge E_c and valence band edge E_v as functions of position along a spatial coordinate to the . In isolated semiconductors, these bands are flat, but upon forming the heterojunction, discontinuities appear at the : the valence band offset \Delta E_v = E_{v2} - E_{v1} and conduction band offset \Delta E_c = E_{c2} - E_{c1}, where subscripts 1 and 2 denote the two materials. These offsets manifest as abrupt steps in the band edges, with the direction and magnitude depending on the relative positions of the bands in the two semiconductors. The construction often begins by aligning the vacuum levels or using differences to position the bands before contact. The relationship between the offsets and the bandgaps of the materials is given by \Delta E_g = \Delta E_v + \Delta E_c, where \Delta E_g = E_{g1} - E_{g2} represents the difference in bandgaps (E_g = E_c - E_v) between the two semiconductors, assuming a type I alignment where both offsets contribute additively to the gap difference. This holds under the assumption of conserved charge neutrality far from the interface and is fundamental to predicting carrier confinement in heterostructures. At equilibrium, the E_F aligns continuously across the , driving and until a built-in potential balances the offsets. This charge redistribution creates regions (depletion or accumulation layers) on either side of the , inducing electrostatic where the bands slope linearly or parabolically over distances of tens to hundreds of nanometers, depending on doping levels. The bending reflects the internal \mathcal{E} = -\nabla \phi, with potential \phi governed by in the region. For instance, in an n-type GaAs/AlGaAs , s transfer from the narrower-gap GaAs to the wider-gap AlGaAs, forming a in GaAs and upward there. Ideal band diagrams employ the flat-band approximation, depicting uniform bands far from the with sharp offsets and no bending, which simplifies analysis but neglects charge transfer. In reality, this approximation is limited by interface effects such as —localized energy levels within the bandgap arising from dangling bonds or defects—that can pin the and introduce interface dipoles, altering the effective offsets by 0.1–0.5 eV. These states lead to additional band bending even in undoped structures and are particularly pronounced in lattice-mismatched interfaces, as seen in Ge/GaAs heterojunctions where experimental diagrams show smoothed transitions over layers rather than ideal steps.

Types of Band Alignments

Type I (Straddling Gap)

In Type I band alignment, also referred to as straddling gap alignment, the conduction and bands of the narrower-bandgap are fully nested within the bandgap of the wider-bandgap at the interface. This arrangement ensures that both the valence band offset \Delta E_v and the conduction band offset \Delta E_c have the same sign, typically positive when defining offsets from the narrower-gap material to the wider-gap barrier, leading to effective potential wells for both charge carriers. The nested band structure confines both electrons and holes to the narrower-bandgap region, promoting strong spatial overlap of their wavefunctions and minimizing leakage into the barrier material. This carrier confinement is particularly advantageous for quantum well designs, where it enhances radiative recombination efficiency and supports the development of low-threshold optoelectronic devices. The close proximity of electron and hole wavefunctions in Type I alignments facilitates exciton formation, with the binding energy approximated by the hydrogenic model as E_b = \frac{13.6 \, \mathrm{eV} \cdot (m_r / m_0)}{\epsilon_r^2}, where m_r is the exciton reduced mass, m_0 is the free electron mass, and \epsilon_r is the relative dielectric constant of the confining material; in quantum-confined structures, this energy can be enhanced due to reduced dimensionality. A classic example is the GaAs/AlGaAs , where the narrower-bandgap GaAs layer (bandgap ~1.42 eV) is surrounded by wider-bandgap AlGaAs barriers (bandgap tunable up to ~2.16 eV for AlAs), enabling robust confinement for applications in high-performance lasers and lasers.

Type II (Staggered Gap)

In type II band alignment, also known as staggered gap alignment, the conduction band minimum of one material lies above the conduction band minimum of the adjacent material, while the valence band maximum of the first lies below that of the second, resulting in partial overlap of the bandgaps at their interface. This staggered arrangement does not fully confine both charge carriers within a single material but instead promotes their distribution across the interface, with the conduction band offset (ΔE_c) and valence band offset (ΔE_v) having opposite signs. The spatial separation of electrons and holes in type II alignment drives electrons toward one material and holes toward the other, facilitating efficient charge separation. This separation reduces the probability of radiative recombination between electrons and holes, as they are localized in different spatial regions, thereby suppressing non-productive carrier losses. However, the indirect nature of the electron-hole overlap across the leads to weaker optical transitions, characterized by reduced compared to direct-gap configurations. The effective bandgap in a type II heterostructure, denoted as E_g^{II}, is given by the expression
E_g^{II} = E_{g1} + E_{g2} - \Delta E_v - \Delta E_c,
where E_{g1} and E_{g2} are the bandgaps of the two materials, and \Delta E_v and \Delta E_c are the respective . This effective bandgap determines the minimum energy required for interband excitations across the staggered interface and is typically smaller than the individual material bandgaps due to the offset contributions.
A representative example of type II alignment is found in InAs/GaSb heterostructures, where the staggered configuration enables absorption in the mid-infrared regime. These structures are particularly advantageous for long-wavelength photodetectors, as the carrier separation enhances while the reduced minimizes dark current contributions from unwanted radiative processes.

Type III (Broken Gap)

In type III band alignment, also known as the broken gap configuration, the conduction band edge of one lies energetically below the valence band edge of the adjacent at the , resulting in no overlap between the band gaps of the two materials. This misalignment creates a scenario where the valence band maximum of one material exceeds the conduction band minimum of the other, enabling direct spatial overlap of and states across the and facilitating interband transport processes. Such alignments are distinct from overlapping gap types due to the potential barrier that electrons must through to transition between bands. The carrier dynamics in type III heterostructures are dominated by interband tunneling mechanisms, particularly Zener tunneling, where electrons can quantum mechanically tunnel from the valence band of one material directly into the conduction band of the other under an applied . This process leads to unique electrical characteristics, including negative differential resistance (NDR), where the current decreases with increasing voltage beyond a peak value due to the depletion of available tunneling states. The efficiency of this tunneling is highly sensitive to the interface quality and applied bias, making these structures promising for high-speed switching devices. A representative example of type III alignment is found in HgTe/CdTe superlattices, where HgTe acts as a zero-gap paired with the CdTe, resulting in a broken gap that allows tunable band crossing. These superlattices, pioneered through growth in the , have been extensively studied for their semimetallic properties and applied in detection owing to the ability to engineer narrow effective gaps via layer thickness control. The historical development highlighted the potential of such systems for optoelectronic applications, building on early theoretical predictions of band discontinuities in semimetal-semiconductor interfaces. The probability of interband tunneling in these broken gap systems can be estimated using the Wentzel-Kramers-Brillouin (WKB) approximation: P \approx \exp\left( -2 \int_{x_1}^{x_2} \kappa(x) \, dx \right), where \kappa(x) = \sqrt{2m^* \left( V(x) - E \right) / \hbar^2}, m^* is the effective carrier mass, V(x) is the local potential profile across the junction, E is the carrier energy, and the integral spans the classically forbidden region from turning points x_1 to x_2. This semiclassical expression underscores the exponential sensitivity of tunneling to barrier thickness and height, guiding the design of low-power tunneling devices.

Theoretical Determination

Electron Affinity Rule

The electron affinity rule, proposed by R. L. Anderson in 1960, serves as a foundational theoretical model for estimating band offsets at abrupt semiconductor heterojunctions. This rule posits that the vacuum levels of the two isolated semiconductors align upon contact, allowing the conduction band offset \Delta E_c to be directly determined by the difference in their electron affinities \chi: \Delta E_c = \chi_1 - \chi_2. The valence band offset \Delta E_v is then derived from the band gap difference \Delta E_g = E_{g1} - E_{g2}: \Delta E_v = \Delta E_g - \Delta E_c. The model relies on key assumptions, including the absence of charge transfer across the and negligible effects from interface dipoles or rearrangements, which simplifies the band alignment to a direct superposition of bulk energy levels referenced to the vacuum. It is particularly suited to ideal, abrupt junctions without significant chemical bonding or strain at the boundary. Despite its simplicity, the electron affinity rule exhibits limitations in real heterojunctions, often overestimating offsets by 0.2-0.5 due to unaccounted interface dipoles and effects. For instance, in the GaAs/ system, the rule predicts a valence band offset of 0.49 , while experimental measurements yield approximately 0.15 . Its predictive accuracy is around 70% for -matched III-V semiconductors, where errors are typically smaller, but it performs less reliably for systems involving different chemical bonding or mismatch. The rule has been refined in later models to incorporate these effects, improving agreement with experimental data.

Interface Dipole Models

Interface dipole models address limitations in simpler theories like the electron affinity rule by incorporating charge redistribution and asymmetry at the , which generates an electric layer and a corresponding potential step ΔV. This arises primarily from bond charge asymmetry or electron tunneling between the across the , leading to a net charge separation. The potential associated with this is given by \Delta V = \frac{e}{\varepsilon A} \int \rho(z) \, z \, dz, where e is the elementary charge, \varepsilon is the permittivity, A is the interface area, \rho(z) is the charge density profile along the interface normal direction z, and the integral captures the dipole moment from the spatial distribution of charge. This correction is essential for realistic predictions, as it accounts for microscopic effects not captured in macroscopic alignments. Key models within this framework include the common anion rule, which predicts band offsets for heterojunctions sharing the same anion, attributing small discontinuities to the dominance of anion p-states in the band maximum. Proposed by Wei and Zunger, the rule uses core-level shifts to estimate offsets, with the band discontinuity ΔE_v approximated as the difference between apparent and true chemical shifts in core levels, reflecting cation differences. However, the rule's accuracy is limited in systems where cation d-orbitals hybridize strongly with anion states, as seen in examples like CdTe/ZnTe, where calculated ΔE_v ≈ 0.13 eV deviates from the ideal zero offset due to such contributions. Another approach involves self-consistent solutions to the Poisson-Schrödinger equations, which iteratively compute the electrostatic potential and wavefunctions to determine the charge-induced . These solutions model the interface as a quantum system, solving ∇·(ε ∇V) = -ρ/ε_0 coupled with the for electron densities, yielding the equilibrium that stabilizes the band alignment. The overall valence band offset is refined by adding the interface dipole potential to the Anderson rule prediction: \Delta E_v = \Delta E_v^{\text{Anderson}} + \Delta V_{\text{interface}}, where ΔE_v^{Anderson} derives from differences in electron affinities and band gaps, and ΔV_{interface} provides the microscopic correction. This formulation, as in Tersoff's effective model, stems from charge transfer across the , treating the as an effective shift from valence band penetration into the adjacent barrier. Such models enhance accuracy in lattice-mismatched systems like Si/, where strain and intermixing induce dipoles that adjust the nominal 0.74 eV valence band offset by up to 0.2 eV, depending on interfacial composition. (DFT) calculations further refine these predictions by directly computing ρ(z) and ΔV for specific interfaces, achieving errors below 0.1 eV for Si/ through and slab models that capture charge asymmetry. Recent advances include machine learning frameworks, such as graph neural networks integrated with DFT data, enabling rapid and accurate band offset predictions for novel heterostructures as of 2024.

Experimental Methods

Photoemission Techniques

Photoemission techniques, particularly X-ray photoemission spectroscopy (XPS) and ultraviolet photoemission spectroscopy (UPS), provide direct measurements of band offsets at semiconductor surfaces and interfaces by probing the electronic structure near the Fermi level and core levels. These methods rely on the photoelectric effect, where incident photons eject electrons from the material, allowing determination of binding energies relative to the Fermi energy. XPS uses higher-energy X-rays (typically Al Kα at 1486.6 eV) to access core-level electrons, while UPS employs lower-energy ultraviolet photons (e.g., He I at 21.2 eV) for valence band states, offering complementary insights into band alignment. In , the valence band offset (ΔE_v) is determined by measuring the energy difference between core-level binding energies and the valence band maximum (VBM) for each material, both separately and at the . The involves first obtaining reference spectra for materials A and B to establish the core-level to VBM separations, (E_{CL}^A - E_{VBM}^A) and (E_{CL}^B - E_{VBM}^B); at the , the interfacial core-level shift (ΔE_{CL}) is measured, yielding ΔE_v = (E_{CL}^A - E_{VBM}^A) - (E_{CL}^B - E_{VBM}^B) + ΔE_{CL}. The conduction band offset (ΔE_c) is then inferred using the known bandgaps of the materials, E_g^A and E_g^B, via ΔE_c = ΔE_v + E_g^A - E_g^B, assuming no interface states significantly alter the gap. complements this by directly mapping the valence band and onset, providing the VBM position with high resolution for clean surfaces, often used in conjunction with for confirmation. These techniques achieve an energy resolution of approximately 0.03–0.1 for VBM determination, enabling precise offset measurements, though surface sensitivity (typically 5–10 nm probing depth) necessitates (UHV, <10^{-10} Torr) conditions to avoid . Challenges include at surfaces, which can shift apparent offsets, and the need for growth (e.g., via ) to access clean interfaces without exposure to air. For instance, measurements on GaAs/AlAs heterojunctions have yielded a ΔE_v of about 0.4 , establishing a type I alignment consistent with theoretical expectations. Similarly, has been applied to organic-inorganic interfaces, confirming offsets with comparable accuracy. Advancements since the early 2000s include angle-resolved XPS (ARXPS) and soft angle-resolved photoemission spectroscopy (SX-ARPES), which enhance depth profiling and access buried interfaces by varying electron emission angles or using higher-energy photons for greater (up to 20–50 ). These techniques mitigate surface sensitivity limitations, allowing band offset determination in multilayer structures without destructive , as demonstrated in studies of superconductor/ interfaces where offsets are resolved with ~0.1 eV precision.

Internal Photoemission

Internal photoemission (IPE) is an optical-electrical technique used to measure conduction band offsets (ΔE_c) in by detecting the threshold energy for photocarrier excitation across the interface barrier. The method involves fabricating a device structure (e.g., a Schottky-like heterojunction or p-n ) and illuminating it with monochromatic light while measuring the resulting under reverse bias. The (hν) at which the onset occurs corresponds to the minimum energy required for electrons (or holes) to surmount the band offset, yielding ΔE_c ≈ hν_threshold after correcting for applied bias and image force lowering. For valence band offsets, hole excitation can be probed in appropriate configurations. IPE is particularly advantageous for buried interfaces and operational devices, as it does not require UHV or analysis. The technique achieves high energy resolution of ~0.01–0.05 by analyzing the photocurrent yield as a function of , often fitted to Fowler's for thermionic-like emission over the barrier. Challenges include distinguishing states or defect-related from true band offset transitions, and sensitivity to doping levels and quality, which can broaden the threshold. It is commonly performed at low temperatures to reduce thermal excitation noise. For example, IPE measurements on GaAs/AlGaAs heterojunctions have confirmed conduction band offsets scaling with Al composition, with values around 0.2–0.6 for x=0.2–0.4, aligning with results. Advancements include multiphoton IPE for lower barriers and integration with for non-destructive probing.

Electrical Measurements

Electrical measurements of band offsets in semiconductor heterojunctions primarily involve capacitance-voltage (C-V) profiling and current-voltage (I-V) characteristics to infer barrier heights from macroscopic device behaviors, providing empirical validation of alignment types such as type I or II. These techniques are particularly useful for isotype heterojunctions, where direct probing of band discontinuities occurs through carrier depletion and transport properties. Unlike spectroscopic methods, electrical approaches capture the effective offsets influenced by doping, interfaces, and defects in operational devices. Capacitance-voltage (C-V) profiling determines conduction or band offsets by analyzing the depletion as a function of applied , which reveals the built-in potential and doping profiles across the . In isotype n-n heterojunctions, the apparent carrier concentration profile from C-V data exhibits a step or peak at the due to the band discontinuity, allowing extraction of \Delta E_c from the intercept of 1/C²-V plots or numerical fitting of the full C-V curve. The , pioneered by Kroemer, is independent of grading and applicable to nonabrupt s, with the band offset derived from the difference in flat-band voltages or charge neutrality conditions. For example, in strained Si/SiGe quantum wells, C-V fitting yields band offsets accurate to within 10-20 meV by solving Schrödinger-Poisson equations self-consistently with measured profiles. However, charges and effects can introduce errors, requiring for precise results. Current-voltage (I-V) measurements assess band offsets through thermionic emission over the barrier, where the saturation current reflects the effective barrier height for carrier transport. In forward bias, the diode equation J = J_s [\exp(eV / n k T) - 1], with J_s = A^* T^2 \exp(-\Phi_b / k T), governs the current density J, and temperature-dependent I-V data enable extraction of the barrier height \Phi_b. For heterojunction Schottky diodes, this relates the conduction band offset to the measured barrier via \Phi_b = \Delta E_c - e V_{bi}, where V_{bi} is the built-in potential determined from doping levels. Analysis via Richardson plots, plotting \ln(J / T^2) versus $1 / T, yields \Phi_b from the slope and effective Richardson constant A^* from the intercept, assuming dominant thermionic emission. Modified Richardson plots account for barrier inhomogeneities using Gaussian distributions, improving accuracy in defective interfaces. In practice, Schottky diode structures on heterojunctions like n-GaN/p-Si or n-ZnO/p-Si demonstrate these techniques, with I-V yielding \Phi_b \approx 0.95 eV consistent with known \Delta E_c, and Richardson constants near theoretical values (e.g., 32 A cm⁻² K⁻² for Si). C-V on InGaN/GaN interfaces measures polarization-induced offsets around 0.2-0.3 eV, while I-V on MoS₂/Si confirms offsets via temperature activation. These methods achieve accuracies of ~0.05 eV in ideal cases but are sensitive to defects, tunneling, and series resistance, often requiring complementary temperature sweeps for reliability.

Applications and Implications

In Optoelectronic Devices

In optoelectronic devices, band offsets play a crucial role in carrier confinement, particularly in Type I alignments where both electrons and holes are confined within the same spatial region of a . This configuration is essential for quantum well lasers and light-emitting diodes (LEDs), enabling efficient radiative recombination by localizing carriers and reducing leakage. For instance, in InGaN/ quantum wells used in blue LEDs, the conduction band offset provides effective electron confinement while maintaining overlap with hole wavefunctions for high internal . Type II staggered band offsets contribute to efficiency enhancements in photovoltaic devices like solar cells by promoting charge separation and minimizing non-radiative recombination losses. In such alignments, electrons and holes are confined in adjacent materials, reducing overlap and thus suppressing recombination pathways that degrade performance. For example, in GaSb/GaAs intermediate band solar cells, the Type II alignment spatially separates carriers, lowering thermal emission rates to around 2.85 × 10^7 s^{-1} at 300 K and enabling efficiencies up to 13% under AM1.5G illumination by curbing leakage currents. Recent advancements in / cells leverage optimized offsets to achieve efficiencies exceeding 34% as of 2025, surpassing single-junction limits through complementary absorption spectra. Favorable conduction offsets (CBO) in the range of 0–0.3 eV at the / facilitate efficient charge while minimizing recombination, as demonstrated in monolithic tandems reaching 34.85% power conversion efficiency. These offsets ensure spike-like barriers that block minority carriers, boosting and fill factor in developments through the . However, challenges arise from interface traps, which can alter effective band offsets through charge accumulation and , exacerbating efficiency droop in LEDs. In InGaN/GaN structures, polarization-induced interface charges create net negative charges, modifying the electrostatic field and , which promotes carrier leakage and non-radiative losses at high injection currents. This trap-mediated alteration reduces the apparent confinement potential, contributing significantly to droop observed in blue LEDs, where can drop by up to 50% at elevated densities.

In Field-Effect Transistors

In high-electron-mobility transistors (HEMTs) based on /AlGaN heterostructures, a large conduction band offset ΔE_c between the wider-bandgap AlGaN barrier and the narrower-bandgap channel drives the formation of a high-density (2DEG) at their interface, enabling superior and gate control for high-frequency and applications. The sheet carrier density n_s of this 2DEG can be approximated by the relation n_s = ε ΔE_c / (e d), where ε is the dielectric permittivity of the barrier, e is the , and d is the AlGaN barrier thickness; this model highlights how larger offsets and thinner barriers increase density while maintaining confinement. Such configurations achieve 2DEG densities exceeding 10^{13} cm^{-2}, significantly enhancing and reducing on-resistance compared to conventional silicon-based transistors. Type II band offsets in Si/Ge strained-channel heterostructures further improve performance in p-channel field-effect transistors (pFETs) by confining holes to the lower effective mass region of the SiGe layer, which boosts hole mobility by factors of 2-4 relative to unstrained channels through reduced intervalley and band warping effects. This mobility enhancement arises because the valence band offset localizes carriers away from interface defects, with the overall mobility governed by the Drude relation μ = q τ / m^, where q is the carrier charge, τ is the momentum relaxation time, and m^ is the effective mass; offsets prolong τ by minimizing and . In practice, compressive in Si_{1-x}Ge_x layers with x ≈ 0.2-0.4 yields peak enhancements, supporting scaled integration with improved drive currents and lower power dissipation. Post-2015 advancements in two-dimensional () material heterostructures, such as MoS_2 channels encapsulated by hexagonal boron nitride (hBN), leverage van der Waals band offsets to realize low-power FETs with mobilities over 100 cm²/V·s and minimal . The type I alignment at the MoS_2/hBN interface, with a band offset of approximately 1.2 , screens charge traps and preserves intrinsic transport, enabling sub-60 /dec subthreshold swings ideal for energy-efficient logic and sensing devices beyond limits. These structures demonstrate on/off ratios exceeding 10^6 while consuming sub-femtojoule switching energies, positioning them for flexible and ultralow-power electronics.

References

  1. [1]
    Computing with DFT Band Offsets at Semiconductor Interfaces - PMC
    Jun 16, 2021 · The determination of the band alignment (also called band offset) at semiconductor interfaces is a subject of high interest as it allows ...
  2. [2]
    Band offsets, Schottky barrier heights, and their effects on electronic ...
    Aug 27, 2013 · They can be thought of as the dangling bond states of the broken surface bonds of the semiconductor, dispersed across its “Penn gap.”5 ΦS is ...<|control11|><|separator|>
  3. [3]
    None
    ### Summary of Main Models for Band Offsets in Semiconductor Heterostructures
  4. [4]
    Band Offset - an overview | ScienceDirect Topics
    Band offset is defined as the energy difference between the conduction band edges of two materials at their interface, which determines the barrier for ...
  5. [5]
    Theoretical study of band offsets at semiconductor interfaces
    May 15, 1987 · We present a first-principles approach to deriving the relative energies of valence and conduction bands at semiconductor interfaces.<|control11|><|separator|>
  6. [6]
    [PDF] XPS and IPE Determination of Band Offsets of Germanium Based ...
    [22] These offsets are defined as the discontinuity of the band edge at the interface, the VBO being the discontinuity of the valance band maximum (VBM) and the ...
  7. [7]
    [PDF] Band Offsets in Semiconductor - Heterojunctions
    The band offsets AE, and AE, are abrupt discontinuities in the band edges at the heterojunction interface.
  8. [8]
    Heterojunction - an overview | ScienceDirect Topics
    A heterojunction is defined as a junction formed between two different semiconductors, typically one n-type and one p-type, which have distinct properties ...
  9. [9]
    [PDF] Chapter 2 Semiconductor Heterostructures - Cornell University
    Semiconductor heterostructures are devices with two or more different semiconductors, including pn, nn, pp heterojunctions, and quantum wells, wires, and dots.
  10. [10]
  11. [11]
    [PDF] Chapter 2 Semiconductor Heterostructures Band Diagrams in Real ...
    Electron flow from semiconductor (1) to semiconductor (2) and hole flow from semiconductor (2) to semiconductor (1) continues.
  12. [12]
    [PDF] Band discontinuities at heterojunctions between crystalline and ...
    With this convention, the band-offset values should add up to the difference in band gap, i.e.,. ΔEv. ΔEc. ΔEg . B. Hydrogenated a-Si. Once again, we first ...Missing: equation | Show results with:equation
  13. [13]
    [PDF] Herbert Kroemer - Nobel Lecture
    ly was the electron affinity rule (Anderson, 1960), according to which the con- duction band offset should be equal to the difference in electron affinity at.
  14. [14]
    None
    ### Summary of Type I Band Alignment from the Document
  15. [15]
    Understanding band alignments in semiconductor heterostructures ...
    Apr 12, 2017 · While the zb-GaN/ band edges consistently show a type-I alignment, the relative position of fundamental band edges changes to a type-II ...
  16. [16]
    Exciton binding energy in quantum wells | Phys. Rev. B
    Aug 15, 1982 · Variational calculations are presented of the ground exciton state in quantum wells. For the GaAs-GaAlAs system, the results obtained from a trial wave functionMissing: AlGaAs Type alignment
  17. [17]
    Exciton Binding Energy - an overview | ScienceDirect Topics
    2.4 The exciton binding energy ... where μ is the reduced effective mass, which equals to 1 / ( 1 m e + 1 m h ) , me and mh are the effective mass of electron and ...
  18. [18]
    [PDF] Band Offsets in Heterojunctions Formed by Oxides with Cubic ...
    Parameters determining the valence band offset ΔEv in the energy band diagram of studied heterojunctions. (all energies are in eV). Heterojunction. Ev2. Ev1. ΔV.
  19. [19]
    Band Gap Alignment - an overview | ScienceDirect Topics
    The type II band alignment (staggered gap) offers the best charge separation because the relative positions of the valence and conduction bands allow the ...
  20. [20]
    Electric-field tunable Type-I to Type-II band alignment transition in ...
    May 14, 2024 · This staggered alignment causes the photoexcited electrons and holes to relax to different materials. This facilitates exciton dissociation and ...
  21. [21]
    Heterojunction photocatalysts: where are they headed?
    Apr 2, 2025 · While this spatial separation reduces recombination rates, it also diminishes the redox power, potentially limiting the applicability of ...
  22. [22]
    Evidence of Type-II Band Alignment in III-nitride Semiconductors
    Oct 6, 2014 · ... reduces the oscillator strength and consequently the radiative recombination. That is also a typical character of type-II MQWs. We further ...
  23. [23]
    [PDF] Heterostructure Fundamentals - nanoHUB
    Feb 25, 2013 · The purpose of these notes is to explain the basic concepts needed to draw energy band diagrams for heterostructure devices. Before we turn to ...Missing: limitations | Show results with:limitations
  24. [24]
    [PDF] Engineered type-II heterostructures for high efficiency solar cells
    The advantages in electronic and optical characteristics of the type-II heterostructures ... 021 superlattice with effective bandgap energy of 1eV. Figure 2.4 ...
  25. [25]
    A Comparative Study on Optical Characteristics of InGaAsP QW ...
    In this paper, we have configured InGaAsP QW (quantum well) heterostructures of type-I and type-II band alignments and simulated their optical ...
  26. [26]
    [PDF] A Comparative Study on Optical Characteristics of InGaAsP ... - Neliti
    energy gap between ... InGaAsP STIN QW heterostructure with type-II band alignment is shown. ... InGaAsP semiconductor compound on InP substrate is shown.
  27. [27]
    Broken-Gap Type-III Band Alignment in WTe2/HfS2 van der Waals ...
    Aug 26, 2019 · We show first-principles evidence that WTe 2 /HfS 2 vdWH possesses the long-sought type-III band alignment with a broken gap, providing a promising platform ...Introduction · Results and Discussion · Conclusions · Supporting Information
  28. [28]
    Robust broken-gap van der Waals heterostructure | Phys. Rev. B
    Dec 5, 2022 · On the other hand, in the less-studied broken-gap (or type III) band alignment [7–9] , the bottom of the conduction band of one semiconductor ...
  29. [29]
    A resilient type-III broken gap Ga2O3/SiC van der Waals ... - Nature
    Jun 3, 2024 · A unique Ga 2 O 3 /SiC vdW bilayer heterostructure with inherent type-III broken gap band alignment has been revealed through first-principles calculation.Missing: offset | Show results with:offset
  30. [30]
    Negative Differential Resistance with Ultralow Peak-to-Valley ...
    Sep 13, 2024 · Broken-gap (type-III) two-dimensional (2D) van der Waals heterostructures (vdWHs) offer an ideal platform for interband tunneling devices due to ...<|control11|><|separator|>
  31. [31]
    Topological transition and edge states in HgTe quantum wells from ...
    May 28, 2014 · The resulting band structure is called to be an inverted one since the Γ 8 v ( 4 ) - Γ 6 c ( 2 ) s p gap E g is negative with values E g = − 0.6 ...
  32. [32]
    Layer intermixing in HgTe‐CdTe superlattices - AIP Publishing
    Jun 9, 1986 · We find appreciable intermixing of the HgTe and CdTe layers at temperatures as low as 110 °C. Such results have serious implications for the use ...<|control11|><|separator|>
  33. [33]
    [PDF] Esaki diodes in van der Waals heterojunctions with broken-gap ...
    allowed to tunnel. TWKB is the WKB tunneling probability term. We convert the sum over crystal momentum ∑ (…) k into the integral LxLyLz/(2𝜋)3 ...
  34. [34]
    [PDF] Interfacial atomic structure and band offsets at semiconductor ...
    The relationship between interfacial atomic structure and band offsets at semiconductor heterojunctions is explored through first-principles local density ...
  35. [35]
    An effective dipole model for predicting band offsets ... - AIP Publishing
    Dec 1, 1986 · A simple model to predict the band offsets in semiconductor heterojunctions is presented. The model is based on the formation of an ...
  36. [36]
    Calculation of the valence band offsets of common‐anion ...
    Jul 1, 1987 · Calculation of the valence band offsets of common‐anion semiconductor heterojunctions from core levels: The role of cation d orbitals Available.
  37. [37]
    Self-consistent solution of the Schrödinger-Poisson equations for ...
    Feb 10, 2010 · A self-consistent one-dimensional solution of the Poisson and Schrödinger equations is developed to describe the electronic properties of the hole gas confined ...Missing: offset dipole
  38. [38]
    Precise Determination of the Valence-Band Edge in X-Ray ...
    Jun 16, 1980 · A highly precise method for locating the valence-band edge in x-ray photoemission spectra is reported. The application to measuring semiconductor interface ...Missing: offsets 1989
  39. [39]
    Measurement of GaAs/InP and InAs/InP heterojunction band offsets ...
    May 8, 1989 · J. R. Waldrop, R. W. Grant, E. A. Kraut; Measurement of GaAs/InP and InAs/InP heterojunction band offsets by x‐ray photoemission spectroscopy.
  40. [40]
    Band offsets of ultrathin high- oxide films with Si | Phys. Rev. B
    In this paper, we present the results of a combined ultraviolet photoemission spectroscopy (UPS) (to probe the occupied electronic state) and inverse ...
  41. [41]
    Momentum-resolved electronic structure and band offsets in an ...
    Dec 22, 2021 · We use soft x-ray angle-resolved photoelectron spectroscopy to directly measure the momentum-resolved electronic band structures on both sides of the Schottky ...
  42. [42]
    Electron transport in unipolar InGaN/GaN multiple quantum well ...
    May 8, 2015 · It was evident that the largest effect of the higher indium composition is the conduction band barrier being ∼0.2 eV higher for the 19% sample ...
  43. [43]
    Impact of band alignment at interfaces in perovskite-based solar cell ...
    Oct 30, 2023 · This review provides detailed information on the significance of optimization of conduction and valance band offsets in the perovskite solar cells.
  44. [44]
    Efficiency Models for GaN-Based Light-Emitting Diodes - MDPI
    Nov 17, 2020 · InGaN grown on GaN results in a negative interface charge (cf. Figure 7) while growing AlGaN on GaN gives a positive interface charge.2. Carrier Transport Models · 3. Quantum Well Carrier... · 4. Light Extraction Models<|control11|><|separator|>
  45. [45]
    [PDF] Simulation of zincblende AlGaN/GaN high electron ... - ICORLAB
    Jun 12, 2017 · However, as xAl decreases, the conduction band offset (AEC) also decreases adversely impacting the 2DEG sheet density (ns). This in turn ...
  46. [46]
    Strained Si, SiGe, and Ge channels for high-mobility metal-oxide ...
    Dec 9, 2004 · This article reviews the history and current progress in high-mobility strained Si, SiGe, and Ge channel metal-oxide-semiconductor field-effect transistors ( ...
  47. [47]
    Charge Scattering and Mobility in Atomically Thin Semiconductors
    Mar 18, 2014 · The mobility is calculated in the relaxation-time approximation (RTA) of the Boltzmann transport equation. The results shed light on the ...
  48. [48]
    [PDF] ARTICLE - Center for Integrated Quantum Materials
    Jun 17, 2015 · We demonstrate fabrication of hBN-encapsulated. MoS2 FETs contacted by graphene electrodes. These. vdW heterostructure devices fulfill ...
  49. [49]
    Band alignment of two-dimensional h-BN/MoS2 van der Waals ...
    Sep 5, 2020 · We have measured the valence band offset (VBO) of h-BN/MoS 2 heterojunction by X-ray photoelectron spectroscopy (XPS) experimentally.