Fact-checked by Grok 2 weeks ago

Heterojunction

A heterojunction is an formed between two different materials, where the abrupt change in leads to discontinuities in the energy band edges, such as the conduction and valence bands. These band offsets, determined by factors like and bandgap differences, enable unique electronic and optical properties not possible in homojunctions, which involve the same material with varying doping. Common examples include GaAs/AlGaAs and / interfaces, where the materials' distinct bandgaps—such as 1.43 eV for GaAs and 1.80 eV for Al_{0.3}Ga_{0.7}As—allow for tailored carrier confinement and transport. The theoretical foundations of heterojunctions emerged in the mid-20th century, with early models proposed by researchers like in 1957, emphasizing band alignment rules. Practical fabrication advanced in the 1970s through epitaxial growth techniques, including (MBE) and metalorganic chemical vapor deposition (MOCVD), enabling high-quality interfaces free of defects. Landmark demonstrations, such as quantum wells in GaAs/AlGaAs structures by et al. in 1974, highlighted their potential for quantum confinement effects. Band offsets at these interfaces are governed by models like the Anderson rule, though experimental measurements using techniques such as have refined predictions for specific material pairs. Heterojunctions are classified by their band alignment into three types: Type I, where both conduction and valence bands of one material straddle those of the other (e.g., GaAs/AlAs); Type II, featuring staggered overlaps that promote charge separation (e.g., in ); and Type III, involving broken gaps that can form potential barriers. These configurations influence carrier dynamics, with Type II structures particularly effective for reducing recombination losses. Transport across the junction can involve tunneling, , or , depending on the offset magnitudes and doping. In applications, heterojunctions underpin modern and , including light-emitting diodes (LEDs), lasers, and high-efficiency cells. For instance, multijunction cells leveraging GaAs-based heterostructures achieve efficiencies exceeding 32%, far surpassing single-junction limits. They also enable lasers and heterojunction bipolar transistors (HBTs) for high-speed communications. Recent studies using scanning ultrafast electron have visualized hot carrier dynamics in Si/Ge heterojunctions, revealing trapping effects that inform designs for and .

Fundamentals

Definition and Basic Principles

A heterojunction refers to the between two dissimilar solid-state materials, most commonly semiconductors with different band gaps, which gives rise to distinctive electronic and optoelectronic properties at the boundary due to discontinuities in energy levels. This contrasts with a homojunction, where the materials on either side share the same composition, and the unique characteristics stem from the mismatch in electronic structure, such as varying lattice constants or chemical compositions. To understand heterojunction formation, key prerequisite concepts from physics are essential. In a , the valence represents the range of states that are fully occupied at , while the conduction consists of unoccupied states above the forbidden gap, or , where can move freely as charge carriers when excited. The denotes the at which the probability of occupancy is 50%, serving as a reference for the . Doping modifies this structure: n-type doping introduces donor impurities with extra valence , elevating the closer to the conduction and increasing concentration; conversely, p-type doping adds acceptor impurities that capture , creating holes in the valence and lowering the . At the heterojunction interface, particularly for an n-type and p-type semiconductor pair, the differing Fermi levels drive charge carrier diffusion across the boundary upon contact, leading to the formation of a space-charge or depletion region depleted of mobile carriers. This charge separation induces band bending in both materials, establishing a built-in electric field that opposes further diffusion and creates potential barriers for electrons and holes, confining carriers and enabling device functionality like rectification. The equilibrium band diagram for a generic abrupt heterojunction illustrates this: the conduction and valence bands of the two semiconductors align with offsets at the interface due to differences in electron affinity (χ) and band gap (E_g), featuring upward or downward bending in the depletion regions on the n-side and p-side, respectively, to equalize Fermi levels across the structure. The magnitude of this built-in potential, V_bi, quantifies the electrostatic barrier and is fundamentally given by V_{bi} = \frac{[kT](/page/KT)}{[q](/page/Q)} \ln \left( \frac{N_A N_D}{n_i^2} \right) for homojunctions, where is Boltzmann's constant, is , is the , N_A and N_D are acceptor and donor concentrations, and n_i is the intrinsic carrier concentration. In heterojunctions, this expression is adapted to incorporate differing electron affinities and band gaps, typically as q V_{bi} = (\chi_p - \chi_n) + E_{g,p} - ([\delta_p](/page/Fermi_level) + [\delta_n](/page/Fermi_level)), where χ_p and χ_n are the electron affinities of the p- and n-type materials, E_{g,p} is the band gap of the p-type material, and δ_p and δ_n are the distances of the from the valence band maximum in the p-material (δ_p = E_{f,p} - E_{v,p}) and from the conduction band minimum in the n-material (δ_n = E_{c,n} - E_{f,n}), respectively; these can be approximated for non-degenerate doping as δ_p ≈ ln(N_{v,p} / N_A) and δ_n ≈ ln(N_{c,n} / N_D), with N_v,p and N_{c,n} the effective densities of states in the valence and conduction bands, and N_A, N_D the doping concentrations. This accounts for the asymmetric band alignment and carrier confinement.

Historical Development

The theoretical foundations of heterojunctions were laid in the early 1960s with the introduction of models for band alignment at semiconductor interfaces. In 1960, R. L. Anderson proposed the electron affinity rule, which predicts the conduction band offset between two semiconductors as the difference in their electron affinities, providing an initial framework for understanding heterojunction band offsets despite later refinements showing its limitations. This rule marked a shift from homojunctions, where materials are identical, to heterojunctions enabling bandgap engineering for improved device performance. The concept of heterostructures for practical devices emerged independently in 1963 from and . Kroemer proposed hetero-junction injection lasers using double heterostructures to confine carriers and enhance efficiency, envisioning applications in . Alferov, along with Rudolf Kazarinov, filed a Soviet patent for double-heterostructure lasers that similarly exploited carrier confinement in GaAs-based systems. These proposals laid the groundwork for bandgap engineering, transforming heterojunctions from theoretical constructs to viable device architectures and earning Kroemer and Alferov the 2000 for heterostructures. Advancements in fabrication techniques during the and were crucial for realizing heterojunctions. The development of epitaxial growth methods, such as liquid-phase epitaxy, enabled precise layering of different semiconductors in the . A pivotal milestone was the invention of (MBE) in 1968 by Alfred Y. Cho and John R. Arthur Jr. at Bell Laboratories, which allowed atomic-level control over heterostructure growth in , facilitating high-quality interfaces essential for device functionality. Key experimental demonstrations followed in the early 1970s. In 1970, Alferov's group at the A. F. Ioffe Physico-Technical Institute achieved continuous-wave operation of GaAs/GaAlAs double-heterostructure lasers at with low thresholds, a breakthrough that enabled practical lasers for and . This success highlighted the role of heterojunctions in . Concurrently, the first practical heterojunction bipolar transistors (HBTs) emerged in the 1970s, building on Kroemer's earlier theoretical work from on wide-bandgap emitters; these devices, often using AlGaAs/GaAs, offered superior speed and gain over homojunction counterparts. By the 1980s, heterojunction research expanded significantly into and high-speed electronics, driven by improved like and metalorganic . This era saw widespread adoption of heterostructure lasers and transistors in commercial applications, solidifying their impact on modern semiconductor technology.

Classification

Types Based on Band

Heterojunctions are classified into three primary types based on the of their conduction and bands at the , which governs the behavior of charge carriers such as confinement, separation, and recombination. This classification arises from the relative positions of the band edges of the two semiconductors, influencing device performance in and energy applications. The band offsets, denoted as \Delta E_c for the conduction band and \Delta E_v for the valence band, are key parameters in determining the type of alignment. A foundational model for estimating these offsets is the electron affinity rule, introduced by Anderson in 1960, which posits that \Delta E_c = \chi_1 - \chi_2, where \chi_1 and \chi_2 are the electron affinities of the two semiconductors, and \Delta E_v = E_{g1} - E_{g2} - \Delta E_c, with E_g being the bandgap energies. For heterojunctions sharing a common anion (e.g., both containing arsenic), the valence band offset \Delta E_v tends to be small, primarily due to the dominance of cation electronegativity differences, while \Delta E_c accommodates most of the bandgap difference.

Type I: Straddling Gap

In Type I heterojunctions, the bandgap of the narrower-gap is fully contained within the bandgap of the wider-gap material, leading to both electrons and holes being confined in the narrower-gap region. The conduction band minimum of the narrower-gap material lies above that of the wider-gap material by \Delta E_c > 0, and the valence band maximum lies below by \Delta E_v > 0, such that the entire bandgap E_{g,\text{narrow}} fits between the band edges of the wider-gap (E_{g,\text{wide}}). A classic example is the GaAs/AlGaAs heterojunction, where GaAs has the narrower gap (E_g \approx 1.42 eV) compared to AlGaAs (E_g > 1.42 eV), with typical offsets of \Delta E_c \approx 0.2 eV and \Delta E_v \approx 0.15 eV for low Al content. This alignment promotes efficient carrier confinement without spatial separation, making Type I heterojunctions ideal for quantum confinement effects in structures like quantum wells, where discrete energy levels enhance radiative recombination for applications in lasers and light-emitting diodes. The band illustrates a nested : \begin{array}{c} \text{Wider-gap material (e.g., AlGaAs):} \\ E_c^{\text{wide}} \quad \quad \quad \quad \Delta E_c \\ \downarrow \quad E_{g,\text{wide}} \quad \downarrow \\ E_v^{\text{wide}} \quad \quad \quad \quad \Delta E_v \\ \text{Narrower-gap material (e.g., GaAs):} \\ E_c^{\text{narrow}} \quad E_{g,\text{narrow}} \quad E_v^{\text{narrow}} \end{array}

Type II: Staggered Gap

Type II heterojunctions feature a staggered arrangement where the conduction band minimum of one is higher than that of the other, and the valence band maximum is lower, resulting in electrons and holes being spatially separated across the . Here, \Delta E_c > 0 and \Delta E_v < 0 (or vice versa), with partial overlap of the bandgaps enabling charge transfer. An exemplary system is the strained Si/Ge heterojunction, where the band offsets drive electrons into the Si and holes into the Ge. This spatial separation reduces recombination rates by localizing opposite charges on different sides, facilitating efficient charge separation crucial for photovoltaic devices and photocatalysis. The band diagram shows the staggered offsets: \begin{array}{c} \text{Material 1 (e.g., Ge):} \\ E_c^1 \quad \quad \Delta E_c > 0 \\ \downarrow \quad E_{g1} \\ E_v^1 \quad \Delta E_v < 0 \\ \text{Material 2 (e.g., Si):} \\ E_c^2 \quad E_{g2} \quad E_v^2 \end{array}

Type III: Broken Gap

In Type III heterojunctions, the bandgaps do not overlap, with the valence band maximum of one semiconductor lying above the conduction band minimum of the other, creating a broken-gap alignment. This results in \Delta E_v > \Delta E_c, often with negative effective overlap. The InAs/GaSb system exemplifies this, with InAs's conduction band below GaSb's valence band by about 0.15 , enabling direct interband tunneling without thermal activation. Such configurations are suited for interband tunneling devices like Esaki diodes and tunnel field-effect transistors, where the broken gap promotes band-to-band tunneling for low-power switching. The band diagram depicts the non-overlapping gaps: \begin{array}{c} \text{Material 1 (e.g., GaSb):} \\ E_c^1 \quad \quad \text{(below } E_v^2\text{)} \\ \downarrow \quad E_{g1} \\ E_v^1 \quad \Delta E_v > \Delta E_c \\ \text{Material 2 (e.g., InAs):} \\ E_c^2 \quad E_{g2} \quad E_v^2 \end{array}

Types Based on Structure and Dimensions

Heterojunctions are classified based on their structural configuration and dimensional scale, which directly impact the quality of the , distribution, and overall performance. Structurally, they can be abrupt or graded. Abrupt heterojunctions feature a sharp where the transition between the two semiconductors occurs over a very short distance, typically on the order of a single atomic layer, enabling precise control over electronic properties but potentially introducing high defect densities if mismatch is present. Graded heterojunctions, in contrast, involve a gradual compositional change across the , which smooths the band profile and reduces accumulation, thereby improving carrier transport and minimizing dislocations. Another structural distinction is between isotype and anisotype heterojunctions, based on the doping types of the constituent semiconductors. Isotype heterojunctions form when both materials share the same conductivity type—either both n-type or both p-type—resulting in a homojunction-like behavior but with band offsets that can enhance confinement without a built-in field from doping differences. Anisotype heterojunctions occur when the semiconductors have opposite doping types (n-p or p-n), creating a pronounced depletion region and strong built-in electric field similar to a conventional p-n junction, which is advantageous for rectification and separation of charge carriers. For instance, an anisotype GaAs/n-AlGaAs heterojunction demonstrates efficient electron injection due to the valence band offset. In terms of dimensions, heterojunctions span from three-dimensional (3D) bulk structures to lower-dimensional configurations, each offering unique advantages in interface area and quantum effects. Bulk heterojunctions are 3D, disordered networks where donor and acceptor materials are intermixed at the nanoscale, as seen in organic photovoltaics where phase separation enhances dissociation without requiring epitaxial growth. Two-dimensional (2D) heterostructures, often assembled via van der Waals stacking, enable atomically clean interfaces without lattice matching constraints, exemplified by /hexagonal (h-BN) stacks that exhibit high carrier mobilities due to weak interlayer coupling. One-dimensional (1D) and zero-dimensional (0D) heterojunctions, such as core-shell nanowires (e.g., GaAs/AlGaAs) or assemblies, provide radial or point-like junctions that amplify surface-to-volume ratios for enhanced light-matter interactions. A critical aspect of heterojunction structure is matching, which determines and . -matched systems, like GaAs/AlGaAs with a mismatch of only about 0.16%, allow for pseudomorphic growth where the epilayer conforms to the substrate up to a critical thickness, maintaining a defect-free coherent that preserves electronic integrity. In contrast, lattice-mismatched heterojunctions, such as InP/GaAs with a 3.7% mismatch, introduce that can lead to misfit dislocations if exceeding the pseudomorphic limit, though controlled relaxation techniques enable functional devices. This engineering is pivotal for tailoring properties like band offsets while mitigating defects.

Physical Properties

Energy Band Alignment

The energy band alignment at a heterojunction interface refers to the relative positioning of the conduction and valence band edges of the two constituent semiconductors, characterized by the conduction band offset \Delta E_c and valence band offset \Delta E_v. These offsets determine the potential barriers for and transport across the and are crucial for device performance. The sum of the offsets relates to the bandgap difference and any level discontinuity: \Delta E_c + \Delta E_v = \Delta E_g + \Delta V, where \Delta E_g = E_{g2} - E_{g1} is the bandgap difference between the two materials and \Delta V is the interface-induced shift in levels (often small or zero under the common assumption). One of the earliest theoretical models for predicting band offsets is Anderson's electron affinity rule, proposed in 1960, which assumes alignment of the vacuum levels at the interface. According to this model, the conduction band offset is \Delta E_c = \chi_2 - \chi_1, where \chi_1 and \chi_2 are the electron affinities of the two semiconductors, and the valence band offset is \Delta E_v = E_{g1} - E_{g2} + \Delta E_c, with E_{g1} and E_{g2} being the respective bandgaps. This simple empirical approach works reasonably well for many lattice-matched systems but fails to account for interface-specific effects. A refinement came from Tersoff's charge neutrality model in 1984, which incorporates an interface dipole arising from quantum-mechanical charge transfer to maintain local charge neutrality. In this model, the dipole modifies the naive electron affinity difference, leading to \Delta E_c = \chi_2 - \chi_1 + \delta, where \delta represents the dipole potential shift calculated from the difference in the charge neutrality levels of the bulk materials. For specific material systems like III-V semiconductors, empirical rules provide practical corrections to these models. The 60:40 rule, widely adopted for GaAs/AlGaAs heterojunctions, states that the bandgap discontinuity partitions approximately 60% to the conduction band offset and 40% to the valence band offset, as confirmed by experimental measurements and ab initio calculations. Another guideline is the common anion rule, introduced in 1975, which observes that in heterojunctions sharing the same anion (e.g., As in GaAs/AlAs), the valence band offset is small (~0-0.2 eV), placing most of the discontinuity in the conduction band due to the anion-derived nature of valence band states. Several factors influence the actual band alignment beyond bulk properties. Interface states, such as defect-induced midgap levels, can pin the and alter offsets through charge redistribution, as emphasized in dipole models. Strain from lattice mismatch modifies the band edges via deformation potentials; for instance, compressive strain in the narrower-gap material typically increases \Delta E_v by shifting the valence band upward. Ordering effects in alloy layers, like cation sublattice ordering, introduce local bandgap variations that perturb the interface and offsets by up to 0.1-0.3 eV. To engineer band offsets for desired alignments, techniques like bandgap grading involve gradually varying the composition across the interface to create a smooth potential profile, reducing abrupt barriers and minimizing carrier scattering; this is commonly implemented in graded-index separate-confinement heterostructures. Interface passivation, using of thin oxide layers or monolayer surfactants, suppresses dangling bonds and interface states, thereby stabilizing and tuning offsets by 0.1-0.5 while improving alignment predictability.

Effective Mass Mismatch and Carrier Dynamics

In heterojunctions, the effective mass of charge carriers often differs between the adjacent materials owing to variations in the of their bands. The effective mass m^* quantifies this through the relation m^* = [\hbar](/page/H-bar)^2 \left( \frac{d^2 [E](/page/Energy)}{d[k](/page/K)^2} \right)^{-1}, derived from the second derivative of the E with respect to the wavevector k near the band extrema. This mismatch introduces discontinuities in carrier at the interface, since is given by v = \frac{[\hbar](/page/H-bar) [k](/page/K)}{m^*}, thereby influencing the overall characteristics across the junction. To properly describe carrier wavefunctions in such structures, the BenDaniel-Duke boundary conditions are applied at the interface. These require continuity of the envelope function, \psi_1 = \psi_2, and continuity of the , expressed as \frac{1}{m_1^*} \frac{d\psi_1}{dx} = \frac{1}{m_2^*} \frac{d\psi_2}{dx}, where subscripts 1 and 2 denote the two sides of the junction. These conditions arise from the need to conserve in the effective mass approximation, accounting for the abrupt change in m^*. The effective mass mismatch significantly alters carrier quantization in quantum wells by modifying wavefunction into the barriers, leading to shifts in subband energies and increased effective masses for confined excitons in narrow wells. It also promotes interface scattering, where s experience relaxation due to the discontinuity, impacting overall . In high-electron-mobility transistors (HEMTs), such as those based on AlGaAs/GaAs or InGaAs channels, the reduced effective mass in the two-dimensional electron gas layer—combined with effective mass differences across the interface—facilitates high velocities and mobilities exceeding 10,000 cm²/V·s at . Carrier dynamics at heterojunctions are dominated by thermionic emission, where carriers gain sufficient thermal energy to surmount band offsets, and tunneling, which allows penetration through thin barriers with transmission probabilities enhanced or reduced by the mass mismatch. The latter effect modifies tunneling rates via the WKB approximation, where the decay constant depends inversely on \sqrt{m^*}. Additionally, mobility enhancements stem from the spatial separation of carriers from ionized impurities, a process amplified by effective mass variations that confine electrons to low-mass regions, reducing scattering and enabling terahertz-frequency performance in HEMTs.

Fabrication

Synthesis Techniques

Heterojunctions are fabricated through a variety of techniques that enable the precise control of interfaces between dissimilar semiconductors, ensuring minimal defects and optimal band alignment. These methods are broadly categorized into epitaxial growth processes, which provide atomic-level precision, and alternative approaches suitable for thin films or organic materials. The choice of technique depends on factors such as material compatibility, size, and required interface quality, with epitaxial methods dominating for high-performance devices due to their ability to manage mismatch effectively. Epitaxial growth techniques, such as (), offer unparalleled atomic-layer precision for heterojunction formation. In , elemental sources are evaporated in an environment (typically 10^{-10} ) and directed toward a heated , allowing layer-by-layer deposition at temperatures around 500-600°C. This method excels in producing abrupt interfaces with low defect densities, making it ideal for III-V heterojunctions, though it is limited to small sizes due to the need for high-vacuum conditions. Metal-organic chemical vapor deposition (MOCVD), also known as organometallic vapor phase (OMVPE), is widely used for scalable production of heterojunctions on larger wafers. It involves the of metal-organic precursors, such as trimethylgallium (TMGa) for gallium-based compounds, in a or carrier at elevated temperatures of 700-800°C. MOCVD enables uniform growth over substrates up to 8 inches in diameter and is particularly suited for compound semiconductors like GaAs/AlGaAs, though it requires careful control of precursor flows to avoid unintentional doping. Liquid phase (LPE) provides a cost-effective alternative for heterojunction synthesis, particularly for thicker layers. In this technique, a saturated melt of the material is brought into contact with a , allowing epitaxial growth via and reprecipitation at temperatures typically below 1000°C. LPE is advantageous for its simplicity and low equipment cost, often used in early developments of III-V heterostructures, but it offers less interface sharpness compared to vacuum-based methods. For thin-film heterojunctions, physical vapor deposition methods like and pulsed laser deposition (PLD) are employed. involves bombarding a target material with ions to eject atoms that deposit onto a , enabling room-temperature growth of polycrystalline or amorphous interfaces. PLD uses a high-power to ablate a target, creating a plume that deposits material conformally, suitable for heterojunctions at substrate temperatures up to 800°C. These techniques are versatile for non-epitaxial applications but can introduce more defects due to energetic particle bombardment. Solution-based methods, such as spin-coating, are prevalent for and heterojunctions. In spin-coating, a precursor is dispensed onto a spinning , forming uniform thin films through and , often followed by annealing to enhance crystallinity. This approach is low-cost and scalable for , though it challenges control over abruptness in multilayer stacks. A key challenge in heterojunction synthesis is managing lattice mismatch between constituent materials, which can lead to and dislocations if the critical thickness (typically 10-100 nm) is exceeded. Techniques like relaxation through buffer layers or graded compositions are employed to minimize defects such as threading dislocations and antiphase domains, ensuring high carrier at the . Recent trends in synthesis include (ALD) for conformal, pinhole-free heterojunction coatings. ALD proceeds via sequential, self-limiting surface reactions of precursors, enabling precise thickness control down to the scale at moderate temperatures (100-300°C), ideal for complex geometries in nanoscale devices. Hybrid approaches combining epitaxial growth with 2D material transfer, such as van der Waals stacking, are also emerging to create lattice-mismatch-tolerant interfaces for advanced heterostructures.

Common Material Pairs

One of the most widely studied heterojunction pairs in III-V semiconductors is GaAs/AlGaAs, which exhibits excellent lattice matching with a of approximately 5.653 for GaAs and a nearly identical value of 5.661 for AlAs, enabling pseudomorphic growth without significant strain. This pair forms a Type I band alignment, where the bandgap of GaAs is 1.42 eV and that of AlAs is 2.16 eV, providing effective carrier confinement for applications such as lasers. The bandgap contrast arises from the higher aluminum content in AlGaAs, which increases the conduction to about 0.3-0.4 eV and valence to 0.2-0.3 eV, depending on the aluminum fraction. Another prominent III-V pair is InGaAs/InP, often employed in strained configurations for high-speed , where the lattice mismatch can reach up to 2-3% for non-standard In compositions (e.g., In0.7Ga0.3As), inducing biaxial that modifies carrier effective masses and enhances mobility. In lattice-matched variants (In0.53Ga0.47As on InP), the is 5.869 for both, with a Type I alignment featuring a narrow bandgap of 0.75 eV for InGaAs and 1.34 eV for InP, resulting in a conduction of approximately 0.3 eV. in mismatched structures alters the band structure, splitting the valence band and increasing hole mobility by up to 50%, though it is limited by critical thicknesses on the order of 10-20 nm to avoid formation. In II-VI semiconductors, CdSe/ZnS forms a popular core-shell heterojunction in quantum dots, characterized by a Type I band alignment that confines both electrons and holes within the CdSe core (bandgap ~1.74 ) due to the wider bandgap of ZnS (~3.6 ), with valence band offset of ~1.0 and conduction band offset of ~0.8 . Despite a lattice mismatch of about 11% (CdSe at 6.05 versus ZnS at 5.41 ), the thin shell (typically 1-5 monolayers) accommodates coherently, improving and stability without introducing defects. Similarly, ZnO/GaN heterojunctions leverage their wide bandgaps of 3.37 for ZnO and 3.4 for , forming a Type II alignment with a valence band offset of ~0.8 , suitable for UV optoelectronics; their lattice constants (ZnO 3.25 , GaN 3.19 ) yield a small mismatch of ~2%, allowing epitaxial growth with minimal . Beyond compound semiconductors, Si/Ge heterojunctions utilize for mobility enhancement, where compressive strain in the Ge layer (lattice constant 5.658 Å versus 5.431 Å for ) on relaxed SiGe virtual substrates increases by 2-4 times and hole by up to 20 times through band warping and valley repopulation. The bandgap of is 1.12 eV and is 0.66 eV, creating a Type II with offsets of ~0.2 eV in the conduction band; however, effects limit critical thicknesses to 5-50 nm depending on Ge content (e.g., ~10 nm for 30% Ge), beyond which misfit dislocations relax the and degrade performance. In , pentacene/C60 heterojunctions form a donor-acceptor with pentacene ( ~5.0 eV, LUMO ~3.2 eV) as the p-type material and C60 (LUMO ~4.5 eV) as the n-type, enabling efficient dissociation via a Type II-like with an of ~0.6 eV at the ; this pair is vacuum-deposited for thin-film devices, with no inherent mismatch but relying on molecular ordering for charge transport.

Applications

Optoelectronic Devices

Heterojunctions play a pivotal role in optoelectronic devices by enabling precise control over carrier injection, confinement, and recombination, which enhances efficiency and performance compared to homojunction counterparts. In light-emitting devices such as lasers and LEDs, the heterostructure design facilitates both electrical and optical confinement, reducing losses and allowing operation at lower currents. This was demonstrated in the development of GaAs/AlGaAs double heterojunction lasers in the early , where the wider bandgap AlGaAs cladding layers surround a narrower bandgap GaAs , confining electrons and holes to the active layer while guiding light via differences. These structures achieved continuous-wave room-temperature operation, marking a breakthrough that enabled practical lasers for applications in and . In quantum well-based heterostructures, further improvements arise from the quantization of states in thin active layers, leading to lower current densities. The current density J_{th} scales inversely with the well width L_w, as the required density for increases with decreasing L_w due to the two-dimensional , but the overall injection efficiency improves from better confinement. J_{th} \propto \frac{1}{L_w} This relationship allows for densities as low as 100-500 A/cm² in optimized GaAs/AlGaAs quantum well lasers, compared to several kA/cm² in bulk double heterostructures, enabling higher quantum efficiencies exceeding 50% and faster modulation speeds up to tens of GHz. For LEDs, similar heterojunction configurations in materials like InGaN/ have boosted internal quantum efficiencies to over 80% through enhanced localization and reduced non-radiative recombination. Photodetectors benefit from heterojunctions by tailoring band alignments to extend spectral response and improve , particularly in the . Type II band alignments, where and wavefunctions are spatially separated across the , minimize recombination losses and enhance collection efficiency. An example is the InGaAs/GaAs heterojunction, which leverages strain-induced offsets for NIR detection up to 1.7 μm, achieving responsivities greater than 0.8 A/W at low bias voltages due to improved separation of photogenerated carriers. This design outperforms single-material detectors by providing higher detectivity, often exceeding 10^{10} Jones, and faster response times below 10 , critical for high-speed systems developed in the 1980s-1990s. In transistor-based optoelectronic integrated circuits, heterojunction bipolar transistors (HBTs) and high-electron-mobility transistors (HEMTs) incorporate band offsets for superior performance. HBTs, such as those using AlGaAs/GaAs or InGaP/GaAs, achieve high current gain \beta through reduced base-emitter recombination, approximated by \beta = \frac{D_n n_i^2 L_p}{D_p N_D W_n} where D_n and D_p are diffusion coefficients, n_i is the intrinsic concentration, L_p is the diffusion length, N_D is the base doping, and W_n is the base width; the heterojunction bandgap discontinuity amplifies \beta to values over 100, enabling power outputs up to watts at frequencies beyond 100 GHz. HEMTs exploit modulation doping in GaAs/AlGaAs structures to form a (2DEG) at the interface, yielding electron mobilities above 10^6 cm²/V·s and transconductances over 1000 mS/mm, which support modulation speeds exceeding 200 GHz in optoelectronic switches and amplifiers from the onward. Overall, these heterojunction devices have driven key metrics like above 70%, modulation bandwidths to 100+ GHz, and integration densities that underpin modern fiber-optic networks.

Photovoltaic and Energy Conversion Devices

Heterojunctions play a pivotal role in photovoltaic devices by facilitating efficient charge separation and reducing recombination losses, thereby enhancing overall . In silicon heterojunction (SHJ) solar cells, thin layers of hydrogenated amorphous silicon (a-Si:H) are deposited on (c-Si) to form passivating contacts that minimize surface recombination while enabling selective carrier transport. This configuration has achieved certified efficiencies up to 27.8% as of 2025. Tandem solar cells leverage heterojunctions for spectrum splitting, pairing wide-bandgap materials like with narrower-bandgap to capture a broader range of the solar spectrum. For instance, monolithic perovskite/ tandems with a perovskite absorber tuned to a bandgap of approximately 1.68 eV have reached certified efficiencies of 34.9% as of 2025, where the heterojunction interface ensures efficient current matching and voltage addition. Bandgap in these devices optimizes , with the top absorbing higher-energy photons and transmitting lower-energy ones to the bottom . In organic photovoltaics, bulk heterojunctions (BHJs)—a type of interpenetrating network structure—enable efficient dissociation in blends like poly(3-hexylthiophene) (P3HT) donor and [6,6]-phenyl-C61-butyric acid methyl (PCBM) acceptor. Optimal yields donor-acceptor domains on the order of 10 nm, matching the typical length in conjugated polymers and thereby maximizing charge generation efficiency. Classic P3HT:PCBM BHJs have demonstrated power conversion efficiencies around 5%, highlighting the role of nanoscale heterojunction morphology in overcoming the limited charge transport in organics. Beyond solar cells, heterojunctions enhance performance in other energy conversion technologies. In thermoelectrics, structures of Bi2Te3/Sb2Te3 exploit energy filtering at interfaces to increase the , with core-shell nanostructures showing improved through reduced thermal conductivity and enhanced . Similarly, in photoelectrochemical (PEC) water splitting, type-II heterojunctions such as TiO2/BiVO4 promote spatial separation of photogenerated electrons and holes, improving stability and photocurrent density for . A key metric in these devices is the open-circuit voltage (V_{oc}), given by V_{oc} = \frac{kT}{q} \ln \left( \frac{J_{sc}}{J_0} + 1 \right), where k is Boltzmann's constant, T is temperature, q is the elementary charge, J_{sc} is the short-circuit current density, and J_0 is the saturation current density. Heterojunction barriers reduce J_0 by impeding minority carrier recombination, thus elevating V_{oc} and overall efficiency in both photovoltaic and PEC systems.

Advanced Developments

Nanoscale and Quantum Heterojunctions

Nanoscale heterojunctions exploit quantum confinement effects when the dimensions of the heterostructure approach or fall below the , typically on the order of 10-100 , leading to discrete energy levels and modified carrier dynamics distinct from behaviors. In these systems, the spatial restriction of charge carriers in one or more dimensions enhances optical and electronic properties, such as increased energies and tunable emission wavelengths, enabling advanced device functionalities. Quantum wells and superlattices represent foundational periodic heterostructures at the nanoscale, consisting of alternating thin layers of semiconductors with different bandgaps, such as GaAs wells embedded in AlGaAs barriers. The first experimental observation of carrier confinement in such GaAs/AlGaAs quantum wells was reported in 1974, demonstrating quantized subband energies through shifts. In these structures, the confinement energy for electrons or holes scales inversely with the square of the well width L, as E \propto 1/L^2, arising from the particle-in-a-box model adapted to the effective mass approximation with boundary conditions at the interfaces. Superlattices, proposed theoretically in 1970, extend this periodicity over multiple periods, enabling miniband formation and negative differential resistance due to resonant tunneling. A key application of these nanoscale periodic heterojunctions is in quantum cascade lasers, first demonstrated in 1994 using GaAs/AlGaAs active regions, where intersubband transitions across multiple quantum wells allow tunable mid-infrared emission without reliance on interband processes. Quantum dots and nanowires further exemplify nanoscale heterojunctions in zero- and one-dimensional forms, respectively, where confinement in multiple dimensions amplifies quantum effects. Core-shell quantum dots, such as CdSe cores overcoated with ZnS shells, exhibit enhanced quantum yields up to 50% due to passivation of surface traps by the wide-gap shell, with the effective band offsets influenced by shell thickness through altered carrier wavefunction penetration. In these type-I heterojunctions, the shell thickness tunes the confinement potential, shifting emission peaks by 10-50 meV as the shell grows from 1 to 5 monolayers. For nanowires, axial 1D heterojunctions, like GaAs/ segments, facilitate directional transport along the wire axis, with abrupt interfaces enabling p-n junctions or superlattice-like behaviors for photonic and electronic applications. These structures support efficient axial electron flow, with mobilities exceeding 10,000 cm²/V·s in high-quality samples. Key quantum effects in these nanoscale heterojunctions include quantized energy levels, where carriers occupy discrete subbands rather than continuous bands, leading to shell-like (DOS) in quantum wells and delta-function-like DOS in quantum dots. This quantization manifests in , observed in single-electron transistors based on quantum dots, where charging energy E_c = e^2 / 2C (with C as ) prevents current flow below a , enabling precise control of single-electron transport at in small dots. The modified DOS enhances oscillator strengths for optical transitions and suppresses in certain regimes. Despite these advantages, nanoscale heterojunctions face challenges such as interface roughness , which limits in GaAs/AlGaAs quantum wells to below 10^6 cm²/V·s in narrow wells due to fluctuations in well width causing energy broadening up to 1-5 meV. Size uniformity is another critical issue, particularly in colloidal quantum dots and nanowires, where polydispersity greater than 5% broadens emission linewidths and reduces ensemble , necessitating precise growth controls like for wells or size-selective precipitation for dots.

Emerging Materials and Recent Advances

Recent advancements in two-dimensional (2D) materials have centered on van der Waals heterojunctions, enabling precise band structure engineering through weak interlayer interactions. For instance, MoS₂/ heterostructures fabricated via direct exhibit enhanced charge transfer and optoelectronic performance, supporting applications in flexible . Studies on twisted MoSe₂/WSe₂ heterojunctions have demonstrated twist-angle-dependent ultrafast transient dynamics beyond the Mott transition. dichalcogenides (TMDs) like MoSi₂N₄/MoS₂ have shown type-II band alignment in van der Waals stacks, promoting efficient carrier separation for , with recent dry-transfer methods enabling scalable integration on patterned substrates. Hybrid perovskite heterojunctions, particularly those combining methylammonium lead iodide (MAPbI₃) with , have advanced through interface engineering to improve and . In MAPbI₃/ photodetectors, antisolvent dripping optimizes film morphology, yielding responsivities over 10⁴ A/W while mitigating degradation under ambient conditions. enhancements via doping, such as Co ions in MAPbI₃ layers, extend operational lifetimes by 50% in heterojunction cells, addressing and migration issues. heterojunction (SHJ) cells reached a 27% in 2025, incorporating nanocrystalline passivation for fill factors above 86% and cell-to-module ratios of 98.6%, driven by optimized transparent conductive oxides. Other innovations include ternary heterojunctions for , such as g-C₃N₄/AgCl/FeOCl systems synthesized via and , which exhibit enhanced visible-light absorption and charge separation for pollutant degradation rates 3–5 times higher than binaries. Nd-doped CuO/ZnO heterojunctions, prepared through wet-chemical methods, demonstrate superior antibacterial activity against pathogens like E. coli, achieving 99.9% inhibition via generation, alongside UV blocking capabilities. Porous heterojunctions, exemplified by UiO-66/TDCOF composites, enable strategic energy-level modulation from type-I to type-II alignment, boosting gas sensing sensitivities by over 200% through customized pore defects and interfaces. Emerging trends encompass AI-optimized interfaces and bio-inspired designs. Machine learning frameworks, integrating crystal graph convolutional neural networks, accelerate heterojunction discovery by predicting dual-active sites in transition metal chalcogenides, reducing design iterations by 70%. Bio-heterojunctions, such as ultrasound-activated herbal variants, facilitate self-catalytic therapy by inducing bacterial cuproptosis-like death, promoting implant-associated wound healing with 95% sterilization efficiency. Ultra-high linearity in Ga₂O₃-based cascade heterojunctions, leveraging hole-trapping mechanisms, achieves responsivities exceeding 10⁴ A/W in deep-ultraviolet optoelectronic synapses, with third-order intercept points over 20 dBm for high-frequency applications.

References

  1. [1]
    [PDF] Chapter 2 Semiconductor Heterostructures - Cornell University
    2.2 A pn Heterojunction Diode. Consider a junction of a p-doped semiconductor (semiconductor 1) with an n-doped semiconductor (semiconductor 2).
  2. [2]
    Electronic Properties of Materials - Physics of Semiconductors
    Nov 7, 2017 · Many other devices are made at the semiconductor heterojunctions, that is at an interface of differently doped different crystalline materials.
  3. [3]
    [PDF] Heterostructure and Quantum Well Physics William R. Frensley May ...
    Mar 25, 1994 · An ideal heterojunction consists of a semiconductor crystal (in the sense of a regular network of chemically bonded atoms) in which there exists ...
  4. [4]
    A review of semiconductor heterojunctions
    The article serves as an introduction to a comprehensive list of references on semiconductor heterojunctions. Several methods of producing such structures ...
  5. [5]
    [PDF] UC Santa Barbara - eScholarship
    Oct 1, 2024 · Semiconductor heterojunctions have gained significant attention for efficient opto- electronic devices owing to their unique interfaces and ...<|control11|><|separator|>
  6. [6]
    Heterojunction Band Alignment
    The central feature of a heterojunction is that the bandgaps of the participating semiconductors are usually different.
  7. [7]
    [PDF] Heterojunctions
    Aug 29, 1997 · In this case, the conduction band of one semiconductor lies below the valence band of the other. Transport is complicated by the fact that the ...Missing: physics | Show results with:physics
  8. [8]
    [PDF] Electrons and Holes in Semiconductors
    One may say that semiconductors differ from insulators in that semiconductors can be made N type or P type with low resistivities through impurity doping. This ...
  9. [9]
    [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.
  10. [10]
    Zhores I. Alferov – Facts - NobelPrize.org
    In 1963, at the same time as but independently of Herbert Kroemer, Zhores Alferov built a heterostructure that acted as a laser. Semiconductor lasers have ...Missing: proposal | Show results with:proposal
  11. [11]
    [PDF] Zhores I. Alferov - Nobel Lecture
    The first important step was done in our laboratory in 1970: in the paper [30] we reported that various lattice-matched heterojunctions based on quaternary. III ...
  12. [12]
    Milestones:Molecular Beam Epitaxy, 1968–1970
    Oct 24, 2025 · Between 1970 and 1974 Al Cho published numerous papers related to MBE growth of GaAs and related materials, their doping, multilayered stacks, ...
  13. [13]
    [PDF] Band Offsets in Semiconductor - Heterojunctions
    The band offsets AE, and AE, are abrupt discontinuities in the band edges at the heterojunction interface.
  14. [14]
    Two-Dimensional Semiconductor Heterojunctions for ... - Frontiers
    For example, heterojunctions with type-III band alignment can actuate Esaki diodes with significant negative differential resistance, heterojunctions with p–n ...
  15. [15]
  16. [16]
    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.
  17. [17]
    Band edge alignment of pseudomorphic on GaAs | Phys. Rev. B
    In this paper, we measure two sets of carefully designed quantum well (QW) samples with different barrier configurations to precisely determine the conduction ...
  18. [18]
    Quantum confinement effects in Si/Ge heterostructures with spatially ...
    Jun 3, 2010 · Here we report on the study of confinement properties of two-terminal devices, based on the quantum tunneling through Ge QDs of definite spatial ...Missing: implications | Show results with:implications
  19. [19]
    Charge Separation in Type-II Semiconductor Heterodimers
    Type-II semiconductor heterodimers with a staggered alignment of band edges at the heterointerface can be synthesized by seeded growth or ion exchange.
  20. [20]
    Integration of broken-gap heterojunction InAs/GaSb Esaki tunnel ...
    Nov 16, 2015 · This study entails a comparison of the broken-gap InAs/GaSb heterojunction system on two different substrates, including Si and native GaSb ...
  21. [21]
    Interband quantum tunneling at the band-edges in III-V ...
    This thesis explores interband tunneling in semiconductor heterojunctions where a density-of-states switching mechanism can be used to sharply modulate the ...
  22. [22]
    Heterojunction Device - an overview | ScienceDirect Topics
    Following the patent on heterostructures granted to Shockley in 1951, Kroemer wrote a paper on the theory and advantages of wide-gap emitters for bipolar ...Missing: origin | Show results with:origin
  23. [23]
    2D Heterostructures for Ubiquitous Electronics and Optoelectronics
    Feb 8, 2022 · In general, the formation of heterostructures with materials of different dimensionalities (2D/0D, 2D/1D, 2D/1.5D, 2D/2D, and 2D/3D) as ...
  24. [24]
    Van der Waals Heterostructures by Design: From 1D and 2D to 3D
    Feb 3, 2021 · We provide a comprehensive review of VDW heterostructures assembled from low-dimensional materials, including 1D/1D, 0D/2D, 1D/2D, 2D/2D, 2D/3D, and 3D/3D ...
  25. [25]
    Mixed-Dimensional 1D/2D van der Waals Heterojunction Diodes ...
    Jan 11, 2022 · We fabricate and study this atomically thin one-dimensional/two-dimensional (1D/2D) van der Waals heterojunction and employ it as the gate of a 1D ...
  26. [26]
    [PDF] Comparing AlGaAs-GaAs Heterojunction Materials with CdS-InP ...
    Feb 21, 2013 · AlxGa1-xAs fulfills the need for an adjustable band-gap counterpart for GaAs with an exceptionally good lattice match (only 0.16% mismatch ...
  27. [27]
    [PDF] Nearly ideal InP/In0.53Ga0.47As heterojunction regrowth on ...
    InP and AlGaAs are of course lattice-matched heterostructures. The Si-H surface is measured under acid. “All defects” refers to the worst possible case of ...
  28. [28]
    Elasticity theory of pseudomorphic heterostructures grown on ...
    Sep 25, 2006 · A theoretical model for lattice mismatched pseudomorphically grown heterostructure, which is based on continuum elasticity theory is described.Missing: heterojunction | Show results with:heterojunction
  29. [29]
    Determination of band offsets at the interfaces of NiO, SiO2, Al2O3 ...
    Jun 17, 2024 · The Al2O3/AlN interfaces are type-II (staggered) heterojunctions with a conduction band offset of −0.47 eV and a valence band offset of 0.6 eV.
  30. [30]
    Theory of semiconductor heterojunctions: The role of quantum dipoles
    Oct 15, 1984 · At any semiconductor heterojunction there is an interface dipole associated with quantum-mechanical tunneling, which depends on the band lineup between the two ...
  31. [31]
    Bandgap and band offsets determination of semiconductor ...
    Sep 14, 2009 · The direct-gap Γ band offset ratio r ≡ Δ E C / Δ E V is found to be 60.4:39.6 (±2%), which agrees with the 60:40 rule.8. FIG. 3. Al composition ...
  32. [32]
    Strain Dependent Electronic Structure and Band Offset Tuning ... - NIH
    Feb 14, 2017 · Our results reveal significant strain tuning of conduction band offsets using epitaxial buffer layers, with strain-induced offset differences as ...Missing: ordering | Show results with:ordering
  33. [33]
    Engineering the Band Alignment in QD Heterojunction Films via ...
    Nov 19, 2019 · Our work demonstrates the possibility to tune the band offset of QD heterostructures via control of the chemical species passivating the QD ...
  34. [34]
    Interface engineering and defect passivation for enhanced hole ...
    Mar 5, 2024 · This study's innovative approach to investigating the band alignment, band offsets, charge transport, and recombination losses in the ...
  35. [35]
    Effective-mass approximation in semiconductor heterostructures
    Dec 15, 1988 · Here, I analyze the EMM in a simple case of one-dimensional heterostructures. Being able to obtain the exact solutions (using the transfer- ...Missing: paper | Show results with:paper
  36. [36]
    Space-Charge Effects on Electron Tunneling | Phys. Rev.
    Rev. Letters 16, 1108 (1966); D. J. Bendaniel and C. B. Duke, General Electric Research and Development Center Report 66-C-331, 1966 (unpublished); F. W. J. ...Missing: original | Show results with:original
  37. [37]
    Effect of carrier confinement on effective mass of excitons ... - Nature
    Jul 7, 2017 · Significant increase of effective mass is observed for the confined exciton in narrow QWs. The foremost reason behind such an observation is due ...
  38. [38]
    [PDF] Quantum coupling and electrothermal effects on electron transport in ...
    May 8, 2019 · in GaN HEMTs. The effective electron mass mismatch at the interface can cause the occurrence of the current collapse when the electron ...
  39. [39]
    Carrier transport across heterojunction interfaces - ScienceDirect.com
    Comparison with the classical thermionic emission model is made to show the significance of tunneling and effect of quantum mechanical reflection.
  40. [40]
    Tunneling-assisted transport of carriers through heterojunctions
    Oct 1, 2017 · The formulation of carrier transport through heterojunctions by tunneling and thermionic emission is derived from first principles. ... physics ...
  41. [41]
    InGaAs based heterojunction phototransistors: Viable solution for ...
    Apr 22, 2019 · A recently studied EI detector—HPT with type-II band alignment—exhibited a Fano factor of 0.5 at a 1 V bias voltage.15 Also, in Ref. 16 for an ...
  42. [42]
    Monolithic integration of visible GaAs and near-infrared InGaAs for ...
    Dec 9, 2019 · To perform multicolor detection in integrated structures, GaAs PDs were transferred onto InGaAs PDs by using a Y2O3 bonding layer to ...Missing: heterojunction | Show results with:heterojunction
  43. [43]
    [PDF] 5.7. Heterojunction Bipolar Transistors
    Where ∆Eg is the difference between the bandgap energy in the emitter and the bandgap energy in the base. The current gain depends exponentially on this ...
  44. [44]
    Silicon heterojunction solar cells achieving 26.6% efficiency on ...
    Apr 17, 2024 · SHJ solar cell was developed to reach 26.6% efficiency, breaking the record for p-type silicon solar cells. The cell structure is illustrated in ...
  45. [45]
    Monolithic perovskite/silicon tandem solar cell with >29% efficiency ...
    Dec 11, 2020 · We report a monolithic perovskite/silicon tandem with a certified power conversion efficiency of 29.15%. The perovskite absorber, with a bandgap of 1.68 ...Missing: tuning | Show results with:tuning
  46. [46]
    Spin-enhanced organic bulk heterojunction photovoltaic solar cells
    Sep 4, 2012 · This domain size is ideally suited to the commonly accepted 10 nm exciton diffusion length in the P3HT domains. The relative intensity of ...
  47. [47]
    Recent Progress on Semiconductor Heterojunction‐Based ...
    Mar 13, 2022 · The generation of PEC water splitting is a three-step procedure: 1) Light absorption, in which the semiconductor is excited by light irradiation ...Introduction · Definition of a Heterojunction · Effective Strategies to Promote...
  48. [48]
    (CdSe)ZnS Core−Shell Quantum Dots: Synthesis and ...
    This paper describes the synthesis and characterization of a series of room-temperature high quantum yield (30%−50%) core−shell (CdSe)ZnS nanocrystallites with ...Introduction · II. Experimental Section · III. Results and Analysis · IV. DiscussionMissing: seminal | Show results with:seminal
  49. [49]
  50. [50]
    Interface roughness scattering in GaAs/AlAs quantum wells
    We study experimentally and theoretically the influence of interface roughness on the mobility of two‐dimensional electrons in modulation‐doped AlAs/GaAs ...
  51. [51]
    Twist-Angle-Dependent Ultrafast Transient Dynamics of MoSe 2 ...
    Sep 5, 2025 · Two-dimensional van der Waals heterostructures (HS) exhibit twist-angle (θ)-dependent interlayer charge transfer, driven by moiré potential ...
  52. [52]
    Type-II MoSi 2 N 4 /MoS 2 van der Waals Heterostructure with ...
    Apr 12, 2023 · Here, we review recent advances in the use of graphene and other 2D materials in catalytic applications, focusing in particular on the ...
  53. [53]
    Dry Transfer of van der Waals Junctions of Two-Dimensional ...
    Oct 31, 2024 · In this paper, we present a method for transferring exfoliated 2D crystal flakes from SiO 2 substrates onto patterned substrates using a poly(vinyl chloride) ( ...
  54. [54]
    Effect of antisolvent dripping time on the photodetection ...
    Sep 15, 2025 · In this work, we developed a silicon/CH3NH3PbI3 (MAPbI3) heterojunction-based photodetector under ambient conditions using the antisolvent ...
  55. [55]
    Improved performance and stability in CH 3 NH 3 PbI 3 /Si ...
    Jul 18, 2023 · On this basis, MAPbI3/Si heterojunction PDs with improved performance and stability have been demonstrated. The Co ion doped MAPbI3 thin films ...
  56. [56]
    27%-efficiency silicon heterojunction cell with 98.6% cell-to-module ...
    Oct 24, 2025 · The new record is 26.81%, with a fill factor (FF) above 86%, on the basis of the implementation of a P-side nanocrystalline silicon (ncSi) ...Missing: pre- | Show results with:pre-
  57. [57]
    Solar-driven photocatalysis using a new ternary g-C3N4/AgCl ...
    Sep 27, 2025 · In this context, a new ternary heterojunction g-C3N4/AgCl/FeOCl (AFxCN) was developed using a combined calcination and coprecipitation method to ...
  58. [58]
    Nd-doped CuO/ZnO and ZnO/CuO heterojunctions for simultaneous ...
    Sep 30, 2025 · Investigation of photocatalytic, antibacterial and antioxidant properties of environmentally green synthesized zinc oxide and yttrium doped zinc ...
  59. [59]
    Strategic energy-level modulation in porous heterojunctions - Nature
    The precise modulation of energy-level brings energy-level alignment between UiO-66 and TDCOF, enabling the customization of porous ...
  60. [60]
  61. [61]
    Ultrasound activated herbal bio-heterojunctions for self-catalytic ...
    Sep 19, 2025 · Ultrasound activated herbal bio-heterojunctions for self-catalytic regulation and bacterial cuproptosis-like death in the treatment of implant ...
  62. [62]
    Ultra-highly linear Ga2O3-based cascade heterojunctions ... - Nature
    Sep 30, 2025 · This hole-trapping-induced gain endows the cascade heterojunctions optoelectronic synapses with an ultrahigh DUV responsivity of over 104 A/W, ...