Fact-checked by Grok 2 weeks ago

Epitaxy

Epitaxy is a method of crystal growth in which a thin crystalline layer, or overlayer, is deposited onto a crystalline substrate such that the overlayer's crystal orientation is determined by and aligned with that of the substrate, resulting in an epitaxial film with precise structural registry. This process, derived from the Greek words epi (upon) and taxis (in an ordered manner), enables the formation of high-quality single-crystal films essential for advanced materials. First described in 1928 by French mineralogist L. Royer, epitaxy has evolved into a cornerstone technique in materials science, particularly for semiconductors. There are two primary types of epitaxy: homoepitaxy, where the deposited material is the same as the substrate (e.g., on ), allowing for purer layers with controlled doping; and heteroepitaxy, involving different materials with compatible crystal structures (e.g., on aluminum gallium arsenide), which facilitates the creation of heterostructures for complex devices. Key growth techniques include vapor phase epitaxy (VPE), which uses at high temperatures around 1200°C for ; liquid phase epitaxy (LPE), involving growth from a liquid solution; and (MBE), a vacuum-based method developed in the late that enables atomic-layer precision at rates of 0.01–0.3 μm/min under conditions (10⁻⁸ to 10⁻¹⁰ ). Pioneered in 1951 by Gordon Teal and Howard Christensen at , epitaxial deposition marked a significant advancement in fabrication by enabling thinner, higher-purity active regions on substrates. In modern applications, epitaxy is indispensable for manufacturing, producing epitaxial layers typically 0.5 to 20 microns thick that enhance device performance in integrated circuits, such as improving doping control, reducing defects, and minimizing issues like in VLSI technologies. It supports optoelectronic devices including LEDs, lasers, and quantum wells through heteroepitaxial structures like GaAs/AlGaAs superlattices, and extends to for multilayer films in displays, , and magneto-optical systems. The technique's ability to grow materials below their melting points has revolutionized the production of high-quality crystals unattainable by other methods, driving innovations in and photonics.

Fundamentals

Definition and Principles

Epitaxy refers to the oriented overgrowth of a crystalline layer on a crystalline , where the atoms in the deposited material align with the 's in a specific crystallographic orientation. This process, derived from the Greek words "epi" (upon) and "" (arrangement), results in the epitaxial layer extending the substrate's , facilitating the creation of interfaces with minimal defects such as dislocations or grain boundaries. The fundamental principles of epitaxy revolve around achieving matching between the and overlayer to reduce interfacial , alongside the minimization of and the influence of thermodynamic driving forces. matching ensures that the periodic atomic arrangement of the growing film corresponds closely to that of the , promoting coherent interfaces where atoms maintain positional registry across the boundary. minimization dictates that adatoms preferentially occupy sites that lower the overall of the system, often favoring two-dimensional layer-by-layer growth over three-dimensional clustering. Thermodynamic driving forces, primarily arising from of the vapor or solution phase, provide the gradient necessary for atoms or molecules to incorporate into the with epitaxial alignment. In contrast to non-epitaxial growth, which produces polycrystalline films or amorphous deposits lacking long-range order and atomic registry with the , epitaxial growth enforces a template-directed assembly that preserves crystallinity throughout the overlayer. This registry is essential for maintaining low defect densities, as deviations lead to energetically unfavorable misalignments. When lattice parameters differ, misfit \epsilon accumulates in the overlayer, contributing elastic strain energy density expressed as E = \frac{\mu (1 + \nu)}{1 - \nu} \epsilon^2, where \mu is the shear modulus and \nu is Poisson's ratio of the material; this energy influences the stability and quality of the epitaxial interface.

Historical Development

The concept of epitaxy originated from observations of natural oriented overgrowth in minerals during the early 20th century, with systematic studies emerging in the 1920s. In 1928, French mineralogist Louis Royer coined the term "epitaxy" (from Greek, meaning "arranged upon") to describe the epitaxial growth of ionic crystals, such as sodium chloride on calcite, primarily from aqueous solutions onto substrates like mica, where the overgrowth crystals aligned crystallographically with the substrate. Royer's work established key conditions for oriented overgrowth, including lattice matching between substrate and deposit, laying the groundwork for later artificial applications. Artificial epitaxy advanced significantly in the mid-20th century amid the rise of semiconductor research. In 1951, Gordon Teal and coworkers at Bell Laboratories demonstrated the first controlled epitaxial deposition of germanium layers using a horizontal pulling technique, enabling high-purity single-crystal films essential for early transistors. By 1960, Henry Theurer's team at Bell Labs achieved the first vapor-phase epitaxial growth of silicon via chemical vapor deposition (CVD) from silicon tetrachloride and hydrogen, producing thin, doped layers that improved transistor performance by reducing base width and enhancing carrier mobility. These developments in the 1950s and 1960s shifted epitaxy from natural phenomena to engineered processes, supporting the transistor revolution and integrated circuit fabrication. The 1970s and 1980s marked a surge in sophisticated epitaxial techniques, driven by demands for precise heterostructures in . In 1968–1970, Alfred Y. Cho at Bell Laboratories pioneered (), a ultra-high-vacuum method evaporating elemental sources to deposit atomically precise layers, first demonstrated on (GaAs) for high-quality interfaces. enabled abrupt heterojunctions, revolutionizing design. Concurrently, refinements in vapor-phase methods, such as metalorganic (MOCVD), emerged in the late 1960s, with Harold Manasevit reporting GaAs growth on in 1968, facilitating scalable of III-V compounds. These innovations culminated in the 2000 awarded to Zhores I. Alferov and for pioneering semiconductor heterostructures grown via epitaxy, which enabled efficient lasers and high-speed electronics. Post-2000 advancements have refined epitaxy for nanoscale precision and novel materials. Tuomo Suntola's atomic layer epitaxy (ALE), invented in 1974 for films, saw significant post-2000 enhancements in self-limiting surface reactions, enabling angstrom-level control in III-V and oxide heterostructures for quantum devices. Integration with proliferated, particularly in van der Waals epitaxy for 2D materials, where epitaxial on substrates was demonstrated in 2004, offering large-area, high-mobility films. By the 2020s, epitaxial growth of 2D heterostructures, such as transition metal dichalcogenides on buffers, advanced via MOCVD and , supporting and ; notable 2025 progress includes direct epitaxial synthesis of single-crystal MoS2 for scalable . These trends underscore epitaxy's evolution toward atomic-scale engineering for .

Types

Homoepitaxy

Homoepitaxy refers to the epitaxial deposition of a crystalline layer onto a substrate made from the same material, such as on , which ensures identical parameters and enables coherent growth with negligible at the . This perfect matching promotes dislocation-free and allows for the formation of smooth, uniform films with high crystalline perfection, often surpassing the substrate's quality in terms of purity and structural integrity. Consequently, homoepitaxial layers exhibit minimal defects, facilitating precise control over thickness and orientation for advanced applications. The primary advantages of homoepitaxy include achieving high material purity by growing cleaner layers on potentially impure substrates, resulting in low densities that enhance device performance through uniform electrical and . For instance, homoepitaxial of on silicon wafers is widely employed in fabrication, where it provides ultra-pure epitaxial layers essential for high-density, high-performance semiconductors. In ideal conditions, the growth rate v follows v = \Omega J, where \Omega represents the atomic volume and J the impinging flux, underscoring the direct proportionality to deposition flux without complicating factors like mismatch-induced . Despite these benefits, homoepitaxy faces challenges such as autodoping, where impurities from the evaporate and inadvertently incorporate into the growing layer, compromising purity in doped systems. Additionally, on vicinal surfaces—substrates slightly misoriented from low-index planes—step-flow can lead to morphological instabilities like step bunching, complicating the achievement of atomically flat layers. These issues necessitate careful surface preparation and optimized parameters to maintain the desired high-quality epitaxy.

Heteroepitaxy

Heteroepitaxy refers to the epitaxial growth of a crystalline layer of one material onto a substrate of a different material, where the lattice parameters of the epilayer and substrate are typically mismatched. This process introduces strain at the interface due to the lattice mismatch, which can be accommodated elastically or plastically depending on the epilayer thickness and mismatch magnitude. Subtypes include pseudomorphic growth, where the epilayer remains coherently strained to match the substrate lattice without defects, and relaxed growth, where misfit dislocations form to relieve the strain, leading to partial or full lattice matching but introducing defects. The transition from pseudomorphic to relaxed regimes occurs at a critical thickness h_c, beyond which the elastic exceeds the required to introduce dislocations. A widely used model for this critical thickness, developed by Matthews and Blakeslee, is given by h_c = \frac{b}{4\pi \epsilon} \left( \frac{1 - \nu \cos^2 \theta}{1 + \nu} \right) \ln \left( \frac{h_c}{b} \right), where b is the , \epsilon is the misfit strain, \nu is , and \theta is the angle between the dislocation line and (typically ≈60° for common dislocations). This equilibrium model predicts the onset of plastic relaxation but often overestimates h_c compared to kinetic growth conditions, where dislocations nucleate earlier due to surface steps or impurities. To manage lattice mismatch and minimize defects, techniques such as buffer layers and structures are employed. Buffer layers, often compositionally graded, provide a gradual transition in lattice parameter between and epilayer, bending misfit dislocations sideways to reduce threading s that propagate into the active layer. Strained-layer (SLS), consisting of alternating thin layers of materials with complementary strains, distribute the mismatch over multiple interfaces, suppressing propagation and enabling higher-quality growth for larger total thicknesses. A representative example is the heteroepitaxy of (GaAs) on (Si) substrates, driven by the need for optoelectronic devices integrated with electronics, despite a ~4% lattice mismatch. In this system, growth beyond the critical thickness (~10-20 nm) leads to misfit dislocations at the interface, many of which convert to threading dislocations that thread through the epilayer, degrading carrier mobility and luminescence efficiency in optoelectronic applications. Advanced schemes, such as step-graded InGaAs layers, can reduce threading dislocation densities to below 10^6 cm^{-2}, improving device performance. Key challenges in heteroepitaxy include the formation of antiphase domains (APDs), particularly in polar-on-nonpolar systems like III-V on , where the lack of inversion causes regions of reversed atomic bonding, leading to recombination centers that reduce device efficiency. Additionally, mismatch between epilayer and induces biaxial during cooling from growth temperatures, often resulting in wafer bowing or cracking; for GaAs on , this mismatch (GaAs TEC ≈2.2 times larger than Si's) generates tensile in the GaAs layer exceeding 200 MPa, necessitating low-temperature buffers or patterned substrates to mitigate.

Growth Mechanisms

Atomic-Level Processes

In epitaxial growth, surface adsorption of precursor atoms or molecules initiates the deposition process, where involves weak van der Waals interactions with low energies (typically <0.1 eV), allowing reversible attachment far from the surface, while chemisorption forms strong chemical bonds with higher activation energies (0.5–2 eV), leading to irreversible precursor dissociation and stable adatom formation closer to the substrate lattice sites. These activation energies determine the rate of precursor attachment, with chemisorption dominating in vacuum-based techniques to ensure ordered layer-by-layer growth. Adatom diffusion on the substrate surface plays a crucial role in nucleation, enabling mobile atoms to migrate across terraces and attach to step edges or form stable clusters, with mobility governed by the D = D_0 \exp(-E_d / kT), where D is the diffusion coefficient, D_0 is the pre-exponential factor (often ~10^{-4} cm²/s), E_d is the diffusion barrier (0.5–1.5 eV for semiconductors), k is , and T is temperature. Low adatom mobility at lower temperatures promotes nucleation of new islands on terraces, whereas high mobility favors attachment to existing steps, reducing defect density and enabling smoother epitaxial layers. Incorporation kinetics at growth fronts involve adatoms descending from upper terraces to lower ones, often hindered by the , an additional energy obstacle (~0.2–0.5 eV) at step edges that traps adatoms on upper levels, leading to multilayer mound formation if not overcome. This barrier influences the balance between step-flow and island nucleation modes, with effective ES values determined experimentally via growth rate measurements on vicinal surfaces. Substrate temperature profoundly affects these atomic processes, as higher temperatures (e.g., 500–800°C for ) increase adatom diffusion rates and lower ES barrier impacts, accelerating incorporation and minimizing defects like vacancies or dislocations, while excessively low temperatures (<400°C) slow kinetics, promoting amorphous or polycrystalline deposition with higher defect densities. For instance, in , optimal temperatures around 750°C yield the lowest threading dislocation densities (~10^8 cm^{-2}), balancing mobility and thermodynamic stability. In-situ techniques such as reflection high-energy electron diffraction (RHEED) enable real-time observation of atomic steps during growth, revealing intensity oscillations corresponding to monolayer completion and step propagation, with streaky patterns indicating smooth, two-dimensional epitaxy. These observations provide direct insights into adatom dynamics and surface reconstruction.

Growth Modes

In epitaxial growth, the morphology of the deposited film is determined by the interplay of surface and interface energies, leading to three primary growth modes: , , and . These modes describe how adatoms assemble on the substrate, influencing the resulting film's structure and properties, such as smoothness or the formation of nanostructures. The selection of a mode depends on thermodynamic favorability, where the change in surface energy \Omega = \gamma_f + \gamma_i - \gamma_s dictates wetting behavior, with \gamma_f, \gamma_i, and \gamma_s representing the overlayer-vacuum surface energy, overlayer-substrate interface energy, and substrate-vacuum surface energy, respectively. The Frank-van der Merwe mode, also known as layer-by-layer growth, occurs when the overlayer wets the substrate completely, resulting in smooth, two-dimensional (2D) film expansion. This mode is favored when overlayer-substrate adhesion exceeds overlayer cohesion, i.e., when \Omega < 0, allowing each monolayer to complete before the next begins. It is common in homoepitaxy or low-misfit heteroepitaxy systems where strong bonding promotes flat interfaces. This growth was theoretically described in early models of vapor deposition. In contrast, the Volmer-Weber mode involves three-dimensional (3D) island formation, where adatoms preferentially nucleate into clusters rather than spreading across the substrate. This arises when overlayer cohesion is stronger than overlayer-substrate adhesion, leading to \Omega > 0 and poor , often observed in metal deposits on insulating substrates like on . The islands grow laterally and vertically, eventually coalescing into a continuous but rough film. This mode was first identified in nucleation studies of supersaturated vapors. The Stranski-Krastanov mode combines elements of the other two, starting with initial layer-by-layer growth that transitions to islands after a critical thickness due to accumulated lattice mismatch . Initially, \Omega < 0 enables wetting layers (typically 1-10 monolayers), but strain energy buildup makes further growth unstable, prompting island formation to relieve stress; this is exemplified in semiconductor quantum dot systems like germanium on silicon with ~4% misfit. The transition thickness increases with decreasing supersaturation. This mixed mode was proposed to explain oriented crystal precipitation. The wetting angle criterion provides a thermodynamic basis for mode selection, derived from Young's equation for equilibrium at the three-phase contact line. For complete wetting (Frank-van der Merwe), the contact angle \theta = 0^\circ, satisfying \gamma_s = \gamma_f + \gamma_i; partial wetting (\theta > 0^\circ) leads to Volmer-Weber or Stranski-Krastanov modes. Factors influencing the mode include substrate preparation, which affects \gamma_i through surface cleanliness or reconstruction, and deposition rate, where higher rates can kinetically favor by limiting adatom . These considerations, unified in phenomenological , guide control of epitaxial morphologies.

Techniques

Vapor-Phase Epitaxy

Vapor-phase epitaxy encompasses techniques where precursor materials are transported in the gas phase to a heated , enabling the deposition of epitaxial layers through chemical or physical processes. The transport mechanisms primarily involve either diffusion-limited or reaction-limited . In diffusion-limited , the rate is controlled by the of precursors through a near the , which predominates at higher temperatures and pressures, leading to uniform deposition over large areas but potential depletion effects. In contrast, reaction-limited occurs at lower temperatures where surface dominate, allowing finer control over incorporation but risking incomplete reactions. Key methods in vapor-phase epitaxy include (CVD), metalorganic chemical vapor deposition (MOCVD), and vapor phase epitaxy (HVPE). CVD typically uses inorganic precursors like chlorides or s, reacting on the substrate to form the epitaxial layer, suitable for and compound semiconductors. MOCVD employs metalorganic precursors, such as trimethylgallium for gallium-based III-V materials, enabling precise control in structures like AlGaAs through adjustable flow rates. HVPE, a variant of CVD, utilizes precursors for high growth rates, often exceeding 100 μm/h for . Growth parameters critically influence the quality and uniformity of epitaxial layers. Precursor flow rates, typically on the order of 10-100 μmol/min for III-V semiconductors, determine the growth rate, which can range from 0.1 to 10 μm/h depending on the . Substrate temperatures for III-V materials like GaAs or are commonly 500-1000°C; for instance, MOCVD growth of occurs at 1000-1100°C to ensure high crystallinity. Pressure effects vary: atmospheric or low-pressure (10-100 ) conditions in CVD and MOCVD promote uniformity across wafers up to 200 mm in diameter. These techniques offer advantages such as for large-area deposition and precise over in or alloys, essential for optoelectronic devices. For example, MOCVD has been pivotal in producing GaN-based light-emitting diodes (LEDs), enabling commercial and LEDs with efficiencies over 100 lm/W through layered heterostructures grown on substrates. HVPE complements this by providing thick, low-defect GaN templates at rates up to 200 μm/h, serving as pseudo-substrates for subsequent MOCVD overgrowth. Doping can be achieved via vapor precursors, such as trimethylindium for n-type incorporation, though detailed mechanisms are covered elsewhere.

Molecular Beam Epitaxy

Molecular beam epitaxy () is an -based technique for epitaxial growth, where elemental sources are evaporated from cells to form directed molecular beams that deposit onto the , providing monolayer-level thickness control without chemical reactions. operates in an ultra-high vacuum environment (around 10^{-11} ), minimizing impurities for abrupt interfaces. Substrate temperatures for III-V materials in MBE, such as GaAs, are typically around 550-600°C to balance adatom mobility and incorporation. Growth rates are low, on the order of 0.1 to 1 μm/h, enabling precise control for complex heterostructures. MBE's conditions promote high-purity layers essential for advanced devices.

Liquid-Phase Epitaxy

Liquid-phase epitaxy (LPE) involves the epitaxial of crystalline layers from a molten onto a single-crystal , where the solute diffuses from the saturated melt to the substrate surface under near-equilibrium conditions. The process typically employs a or vertical furnace setup, with the brought into contact with the melt, allowing controlled to drive layer deposition without requiring systems. Common techniques include the sliding boat method, where a boat with compartments for melt and is used, and the is slid under the for and termination, and the tipping method, where is tilted to wet the with the melt. LPE is particularly suited for III-V compound semiconductors such as InP, where growth occurs from group III-rich melts like or In solvents, enabling the formation of high-quality layers for optoelectronic devices. For InP, typical growth rates range from 1 to 10 μm/h, depending on and temperature, allowing for the deposition of relatively thick films. is critical, maintained at 400–800°C to ensure minimal (often 5–10°C below equilibrium) and prevent spontaneous or substrate dissolution, with ramp cooling or constant-temperature approaches used to regulate growth. The advantages of LPE include low defect densities, often on the order of 10^4 cm⁻² or lower, due to the near-equilibrium growth conditions that promote high crystalline quality akin to bulk materials. It is also cost-effective, requiring simple equipment and offering high precursor utilization efficiency, making it ideal for producing bulk-like layers in compound semiconductors. However, limitations arise from challenges in achieving uniform thickness over large areas, as the diffusion-limited growth and melt-substrate contact can lead to variations in layer morphology and planarity. In heteroepitaxial LPE, lattice mismatch can introduce , though this is managed through careful control.

Solid-Phase Epitaxy

Solid-phase epitaxy (SPE) involves the epitaxial regrowth of a crystalline layer from a solid precursor, such as an amorphous or damaged layer, directly onto a crystalline through short-range atomic and rearrangements at the solid-solid . This typically requires elevated temperatures to enable the attachment of atoms from the amorphous phase to the underlying crystal lattice, without involving or vapor transport. In , SPE commonly occurs in the temperature range of 500–700°C, where the advances continuously, restoring single-crystal order. The primary mechanism is the propagation of the crystalline-amorphous via thermally activated bond-breaking and reformation events, often at ledge sites on low-index planes like {111}, allowing atoms to incorporate into the with minimal long-range . This contrasts with random in the bulk amorphous material, as the template dictates the epitaxial orientation. simulations confirm that the growth proceeds layer-by-layer, with orientation-dependent rates due to varying atomic configurations at the . Key methods include SPE from amorphized layers, where creates an amorphous region that is subsequently annealed to regrow the crystal epitaxially. This technique, pioneered in studies, achieves high-quality regrowth with defect densities as low as 10^7 cm^{-2} after optimization. Another approach is epitaxy, applied to or molecular materials, where amorphous films of molecules like oligothiophenes are crystallized epitaxially on suitable substrates through solid-state annealing, enabling oriented growth for device applications. A major application of SPE is the recrystallization of ion-implanted layers in processing, where implantation damage is repaired and dopants are activated while preserving the substrate's crystallinity, often at temperatures below 600°C to avoid diffusion broadening. In silicon-on-insulator (SOI) , SPE facilitates the formation of thin, defect-free layers over masks, supporting advanced device isolation. Additionally, SPE is utilized in thin-film transistor production to convert into polycrystalline channels with large grain sizes, improving carrier mobility for display and sensor technologies. The growth kinetics of SPE are governed by an Arrhenius relation for the interface velocity v, given by
v = v_0 \exp\left(-\frac{E_a}{kT}\right),
where v_0 is the pre-exponential factor (approximately $4.6 \times 10^6 m/s for Si(001)), E_a is the activation energy (2.7 eV for silicon), k is Boltzmann's constant, and T is the absolute temperature. This yields rates from ~0.1 Å/s at 500°C to several nm/s at 700°C, with the process exhibiting a strong exponential temperature dependence rather than linearity, though log-rate plots appear linear over narrow ranges. These kinetics relate to underlying atomic processes of attachment-limited growth at the interface.

Doping in Epitaxy

Incorporation Mechanisms

Dopants are incorporated into epitaxial layers to modify electrical properties, primarily through intentional addition of impurities during or after growth. In-situ doping involves introducing precursors directly into the growth environment, such as (SiH₄) for n-type doping in III-V semiconductors like GaAs, where it decomposes to provide atoms that substitute on sites, acting as donors. This method ensures uniform distribution and avoids post-growth processing damage. Alternatively, ex-situ doping uses to introduce dopants into the substrate or pre-grown layers, followed by epitaxial overgrowth to encapsulate and activate the impurities, as demonstrated in Ge-on-Si structures where enhances beyond limits. During epitaxial growth, dopant segregation and diffusion govern their distribution between the growth surface and the crystal lattice. The surface segregation coefficient, defined as k = C_s / C_l, where C_s and C_l are the dopant concentrations in the solid lattice and liquid (or vapor/liquid interface) phases, respectively, quantifies this partitioning; values of k < 1 indicate preference for the liquid phase, leading to pile-up at the interface. In molecular beam epitaxy (MBE) of SiGe, for instance, boron segregation decreases with increasing Ge content, altering diffusion profiles and requiring temperature control to achieve desired uniformity. Compensation effects occur when unintended impurities or self-compensation neutralize dopant activity, while solubility limits cap the maximum incorporable concentration before precipitation or deactivation. In silicon epitaxy, phosphorus solubility reaches approximately $5 \times 10^{20} cm⁻³, beyond which electrical activation saturates due to vacancy-dopant complexes like P-V pairs, reducing free carrier density. These limits are material-specific; in InGaAs, silicon doping achieves metastable concentrations up to $5 \times 10^{19} cm⁻³ via growth techniques, but compensation from native defects lowers activation efficiency. Dopants influence epitaxial growth kinetics by altering surface mobility and aiding defect passivation. For example, indium in GeSn MBE acts as a surfactant, accumulating on the surface to form mobile Sn-In droplets that enhance adatom diffusion but limit incorporation to $2.8 \times 10^{18} cm⁻³ and promote Sn segregation, potentially nucleating defects at low temperatures. In Si epitaxy, heavy antimony doping modifies surface reconstruction, increasing adatom mobility and passivating vacancies to improve layer quality, though excessive levels introduce compensation. Advanced control techniques like delta-doping enable atomically sharp dopant profiles in heterostructures by momentarily interrupting growth to deposit a sub-monolayer of dopant, minimizing diffusion. In Si/Ge systems grown by MBE, this involves depositing two Ge monolayers at 275°C followed by Si capping, creating confined quantum wells with precise positioning verified by resonant X-ray scattering, ideal for high-mobility transistors.

Effects on Material Properties

Doping during epitaxial growth enables precise tuning of carrier concentrations in semiconductors, fundamentally altering their electrical properties to suit specific device requirements. In gallium arsenide (GaAs), silicon (Si) acts as a donor dopant, facilitating n-type conductivity by providing free electrons, with carrier concentrations controllable up to approximately 10^{18} cm^{-3} in molecular beam epitaxy (MBE)-grown layers. Conversely, zinc (Zn) serves as an acceptor, enabling p-type conductivity in epitaxial GaAs, where doping levels around 10^{18} cm^{-3} convert the material to p-type, as evidenced by capacitance-voltage measurements in Zn-doped epilayers. These electrical modifications are critical for forming p-n junctions in devices like heterojunction bipolar transistors, where n-type Si-doped GaAs emitters achieve high current gains exceeding 200. Optically, heavy doping induces the Burstein-Moss effect, where the apparent bandgap widens due to the filling of conduction band states by excess carriers, blocking low-energy optical transitions. In Si-doped n-type GaAs epilayers, this shift increases the interband transition energy, observable in absorption spectra of layers with donor concentrations above 10^{18} cm^{-3}. Similar effects occur in Te-doped InGaP epilayers grown by liquid-phase epitaxy, where the Burstein-Moss shift raises the absorption edge, impacting the performance of optoelectronic devices like light-emitting diodes. Structurally, dopants with atomic sizes differing from host atoms modify the lattice constant and introduce strain in epitaxial layers. For instance, carbon (C) doping in , where C atoms are smaller than Ga or As, causes lattice contraction, reducing the lattice parameter by up to 0.1% at concentrations of 10^{19} cm^{-3} in metalorganic molecular beam epitaxy (MOMBE)-grown films. This size mismatch can also alter epitaxial strain profiles, as seen in films where dopant substitution leads to measurable lattice parameter changes, influencing overall film quality and defect formation. Doping interacts with defects to enhance material purity through gettering, where impurities are trapped away from active regions. In epitaxial silicon, carbon doping promotes gettering of metallic contaminants like iron and nickel, forming sinks that reduce their concentration in the channel area and thereby improve electron mobility by minimizing scattering. This effect is pronounced in carbon-cluster-implanted epitaxial Si wafers, where gettering efficiency surpasses that of bulk silicon, leading to higher device yields in CMOS image sensors. Quantitative assessments via Hall effect measurements reveal that increasing dopant concentrations degrade transport properties due to enhanced ionized impurity scattering. In n-type epitaxial GaAs, electron Hall mobility decreases from over 8000 cm²/V·s at low doping (∼10^{16} cm^{-3}) to below 2000 cm²/V·s at 10^{18} cm^{-3}. Similarly, minority carrier lifetimes shorten at high doping levels owing to increased recombination at dopant-related centers; in n-type GaAs epilayers, lifetimes drop from nanoseconds at 10^{16} cm^{-3} to picoseconds above 10^{19} cm^{-3}. These reductions limit device efficiency but can be optimized for applications requiring specific carrier dynamics, such as high-speed transistors.

Natural Epitaxy

Isomorphic Minerals

Isomorphic minerals, characterized by identical crystal structures and space groups but differing chemical compositions, enable natural epitaxial growth through compatible lattice parameters that support coherent interfaces. In the spinel group, for instance, all minerals share the cubic space group Fd3m, facilitating epitaxial overgrowths such as magnetite (Fe₃O₄) onto chromite (FeCr₂O₄) substrates, where the structural similarity minimizes mismatch strains. The thermodynamic favorability of these epitaxial relationships stems from low interfacial energy due to structural continuity, which lowers the nucleation barrier and promotes oriented crystal attachment over random growth. Small differences in lattice parameters between the isomorphic phases further reduce strain energy at the interface, making epitaxy energetically preferred in systems where compositional variations occur without altering the overall symmetry. A representative example is the epitaxial overgrowth of on in metasomatic environments. These formations arise in low-temperature metasomatic or metamorphic settings with minimal strain, where slow diffusion and fluid-mediated processes allow for precise compositional substitution without polymorphic disruption. Observation of such epitaxial relations in isomorphic minerals relies on techniques like electron backscatter diffraction (EBSD), which maps local crystallographic orientations to verify parallel lattice alignments and quantify misorientation angles, often revealing dispersion patterns consistent with epitaxial control over growth direction. EBSD analysis, typically conducted in a scanning electron microscope, distinguishes coherent overgrowths by identifying low-angle grain boundaries (e.g., <5°) that indicate minimal dislocation accumulation at the interface.

Polymorphic Minerals

In polymorphic minerals, epitaxial relationships arise when different crystal phases of the same chemical composition form oriented overgrowths due to kinetic factors during natural crystallization processes. This oriented attachment occurs when the lattice parameters of the overgrowing phase align closely with those of the substrate phase, minimizing interfacial energy and promoting coherent growth. A prominent example is the epitaxial overgrowth of (a metastable tetragonal polymorph) on (the stable tetragonal polymorph), both of , observed in natural mineral specimens from localities such as , where rutile crystals form parallel, needle-like attachments on anatase bipyramids. The driving forces behind these epitaxial relationships primarily involve substrate templating, where the underlying crystal structure dictates the orientation of the overlying phase, thereby stabilizing metastable polymorphs that would otherwise transform to the stable form. In the case of TiO₂, the lattice mismatch between anatase and rutile can be as low as ~4% in specific epitaxial orientations, such as rutile(110)//anatase(100), allowing epitaxial stabilization of anatase even under conditions favoring rutile. This templating reduces the nucleation barrier for the metastable phase by providing a low-energy interface, as demonstrated in computational models of polymorph growth on homologous substrates. Similarly, in silica (SiO₂) systems, epitaxial lattice matching facilitates the oriented deposition of high-temperature polymorphs like tridymite or cristobalite during vapor-phase crystallization in volcanic environments. A key example of such epitaxial growth in silica polymorphs is the oriented overgrowth of (the low-temperature stable phase) on (a high-temperature metastable phase) in volcanic rocks, such as those from rhyolitic lavas. Here, lattice matching occurs with minimal mismatch in d-spacings (e.g., tridymite's pseudo-hexagonal layers aligning with quartz's trigonal structure at angles near 0°–5° misorientation), enabling coherent interfaces during cooling and transition from vapor-phase deposition. Impurities like (declining from ~4 wt% in cores to <0.2 wt% in rims) further aid this process by modulating lattice expansion, allowing epitaxial progression without disrupting the oriented attachment. Stability in these systems is maintained by transformation barriers that prevent immediate phase conversion despite thermodynamic favorability. For instance, in TiO₂, the energy barrier for anatase-to-rutile transition (~60 meV/atom) is heightened by epitaxial strain (up to 10 meV/atom), which locks the metastable structure via coherent interfaces with coincident site lattice areas ≤400 Ų. In silica, kinetic barriers from slow cooling in volcanic settings preserve tridymite or cristobalite overgrowths on quartz substrates, with impurities expanding unit cells (e.g., α-cristobalite d(101) ~4.1 Å) to inhibit inversion to quartz until temperatures drop below ~200–300°C. These barriers ensure the persistence of oriented polymorphs in geological records. Analytical methods for identifying these epitaxial overgrowths in polymorphic minerals commonly rely on X-ray diffraction (XRD) to confirm phase identity and orientation in the overgrowths, revealing peak shifts indicative of strain (e.g., broadened reflections at 2θ ~25°–35° for TiO₂ polymorphs or ~20°–30° for SiO₂). Transmission electron microscopy (TEM) complements XRD by visualizing interfacial coherence and d-spacings at the nanoscale, while electron backscatter diffraction (EBSD) maps misorientation angles (<5°) across overgrowth boundaries. These techniques have been pivotal in documenting epitaxial features in natural samples, such as the core-rim transitions in volcanic cristobalite-tridymite-quartz assemblages.

Specific Geological Examples

One prominent example of natural epitaxy involves the oriented growth of (TiO₂) on (Fe₂O₃), where rutile crystals form parallel intergrowths within the hematite host due to the matching of {110} planes, resulting in low-energy interfaces that dictate 12 possible exsolution directions. These intergrowths, appearing as platy rutile pillars interwoven in a hematite matrix, are commonly observed in hydrothermal quartz veins, such as those in the Swiss Alps, formed through topotaxial reactions during fluid-mediated mineral replacement. The epitaxial relationship minimizes lattice mismatch, promoting coherent growth at the interface. Another key case is the epitaxial overgrowth of hematite (α-Fe₂O₃) on magnetite (Fe₃O₄) in banded iron formations (BIFs), exhibiting cubic-to-hexagonal orientation relationships that facilitate the transformation and layering during oxidation processes. In the presence of pre-existing hematite, magnetite oxidizes epitaxially to form additional hematite layers, preserving structural continuity across the phase boundary in these Precambrian sedimentary sequences. This occurs in low-temperature diagenetic or early metamorphic environments typical of BIFs, such as those in South Africa and Australia. Additional examples include oriented overgrowths of pyrite (FeS₂) on galena (PbS) in hydrothermal ore deposits, where electron backscatter diffraction (EBSD) reveals low misorientation angles (typically <5°) at the interface, indicating epitaxial alignment driven by close lattice matching between the cubic structures. Such occurrences are documented in Mississippi Valley-type deposits, like those in the Tri-State district (USA), where pyrite cubes grow coherently on galena faces, reflecting sequential precipitation from evolving ore fluids. These natural epitaxial features serve as geological indicators of formation conditions, with the rutile-hematite system, for instance, implying hydrothermal temperatures of 200–400°C and pressures below 3.5 kbar, as inferred from fluid inclusion and geothermometry data in associated quartz veins. Similarly, hematite-magnetite epitaxy in BIFs points to oxidative environments at near-surface temperatures (<200°C) during sedimentation or early burial. Pyrite-galena overgrowths suggest episodic fluid pulses at 150–300°C, aiding reconstruction of ore deposit paragenesis. In modern materials science, these natural mineral epitaxies inspire biomimetic approaches to synthetic crystal growth, such as oriented hydroxyapatite deposition for bone repair, where low-misorientation interfaces mimic geological overgrowths to achieve coherent, defect-minimized layers under ambient conditions.

Applications

Semiconductor Devices

Epitaxy plays a pivotal role in the fabrication of semiconductor devices by enabling the precise control of material composition, thickness, and interfaces, which is essential for forming high-performance junctions and enhancing electronic transport properties. In electronic devices such as transistors, epitaxial growth techniques like and allow for the deposition of heterostructures that optimize carrier mobility and reduce parasitic capacitances, leading to improved speed and efficiency in logic and switching applications. In bipolar junction transistors (BJTs), epitaxial base layers are crucial for achieving high-speed switching performance, particularly in silicon-germanium (SiGe) heterojunction bipolar transistors (HBTs). The graded Ge composition in the epitaxial SiGe base creates a built-in electric field that accelerates minority carriers, reducing base transit time and enabling cutoff frequencies (f_T) exceeding 300 GHz through epitaxial strain engineering. For instance, selective epitaxial growth of SiGe layers in BiCMOS processes integrates these HBTs with CMOS logic, supporting millimeter-wave applications with f_T/f_max values of 300/330 GHz. Field-effect transistors (FETs) benefit from channel epitaxy to form high-mobility conduction paths, as seen in FinFETs and high-electron-mobility transistors (HEMTs). In FinFETs, epitaxial growth of strained silicon or III-V channels on silicon substrates enhances drive current and short-channel control, while in GaN/AlGaN HEMTs, the epitaxial heterostructure induces a two-dimensional electron gas (2DEG) at the interface, enabling high-frequency operation in power amplifiers. These structures, grown via metalorganic CVD (), achieve electron mobilities over 2000 cm²/V·s, supporting applications in RF and high-power electronics. Defect management is critical in epitaxial layers to minimize scattering and leakage, with epitaxial lateral overgrowth (ELO) effectively reducing threading dislocations in mismatched heterostructures like GaN on Si. By selectively growing over patterned masks, ELO bends and terminates dislocations in the coalesced regions, lowering densities from 10^9 to 10^5 cm⁻², which improves device reliability and yield. Scalability of epitaxial processes in CMOS manufacturing enables integration for 3D integrated circuits (ICs), where sequential epitaxial stacking of active layers reduces interconnect delays and footprint. CMOS-compatible epitaxial bonding of III-V materials on silicon hosts supports heterogeneous 3D architectures, achieving up to 50% area savings while maintaining thermal budgets below 400°C. Doping during epitaxy further tunes carrier concentrations, influencing mobility as detailed in related material properties discussions.

Optoelectronic and Photonic Devices

Epitaxy plays a pivotal role in the fabrication of optoelectronic and photonic devices by enabling precise control over heterostructures that manipulate light-matter interactions through bandgap engineering. In light-emitting diodes (LEDs) and lasers, techniques such as and are used to grow quantum well structures, where thin layers of material confine charge carriers to enhance radiative recombination. A landmark example is the development of , which rely on MBE or MOCVD to form InGaN/GaN multiple quantum wells on sapphire substrates, achieving high-brightness emission in the visible spectrum. This breakthrough, recognized by the 2014 Nobel Prize in Physics awarded to Isamu Akasaki, Hiroshi Amano, and Shuji Nakamura, demonstrated how epitaxial growth overcame challenges in wide-bandgap III-nitride materials to enable efficient blue light emission, paving the way for white LEDs used in solid-state lighting. Photodetectors, essential for fiber optic communications, benefit from epitaxial growth of layers on substrates, which provides lattice-matched heterostructures with tunable absorption properties. By adjusting the indium composition in (typically In_{0.53}Ga_{0.47}As for 1.55 \mum detection), the absorption coefficient can be optimized—reaching values around 0.67 \mum^{-1} at telecom wavelengths—to maximize quantum efficiency while minimizing dark current. These epitaxial layers form the active region in p-i-n photodiodes, where defect-free interfaces ensure high responsivity, often exceeding 0.8 A/W, supporting high-speed data transmission in optical networks. Photonic crystals leverage periodic epitaxial multilayers to achieve strong confinement via photonic bandgap effects, directing and trapping photons for enhanced device performance. In GaAs/AlGaAs systems grown by MBE, alternating layers create one-dimensional distributed Bragg reflectors (DBRs) that form vertical- surface-emitting lasers (VCSELs), where reflectivity >99% confines axially for low-threshold operation. This epitaxial approach allows integration of quantum wells within the , boosting and enabling compact photonic devices. The historical impact of epitaxy in these devices stems from its ability to produce defect-free heterointerfaces in wide-bandgap materials, which dramatically improves internal quantum efficiency (IQE) to over 90% in InGaN LEDs by reducing non-radiative recombination. Such efficiencies arise from smooth interfaces that minimize threading dislocations, allowing visible light emission where bulk growth historically failed due to high defect densities. This advancement has transformed , enabling energy-efficient lighting and high-speed with widespread applications.

Advanced Materials and Nanostructures

Epitaxial growth techniques have enabled the fabrication of III-V semiconductor s and nanowires with precise control over size and density, leveraging the Stranski-Krastanov (SK) mode to form self-assembled nanostructures that enhance photovoltaic performance. In SK growth, a thin wetting layer of like InAs on GaAs transitions to three-dimensional formation due to lattice mismatch strain, producing uniform InAs/GaAs s with diameters around 10-20 nm and heights of 5-10 nm. These dots extend the absorption spectrum into the , increasing sub-bandgap photocurrent in GaAs-based solar cells by up to 20% through hot carrier extraction and multiple exciton generation. For nanowires, () facilitates axial heterostructures, such as InAs segments in GaAs nanowires, which serve as quantum dot arrays for tandem photovoltaic architectures, improving efficiency by capturing lower-energy photons. Epitaxial approaches have also advanced two-dimensional (2D) materials, particularly graphene grown on silicon carbide (SiC) or hexagonal boron nitride (hBN) substrates, enabling spintronic devices with long spin coherence lengths. Thermal decomposition of SiC(0001) surfaces via sublimation yields few-layer epitaxial graphene with high carrier mobility exceeding 10,000 cm²/V·s, where the buffer layer interacts covalently to maintain lattice alignment. This configuration supports efficient spin injection and detection, achieving spin transport lengths over 10 μm at room temperature due to minimal spin-orbit coupling and valley preservation. On hBN, van der Waals epitaxy of graphene via chemical vapor deposition or transfer aligns the lattices with a mismatch below 2%, forming moiré superlattices that enhance spin filtering in tunnel junctions, with magnetoresistance ratios up to 100% observed in graphene/hBN heterostructures. These systems are pivotal for spin valves and logic gates, where the topological protection of spin states reduces dissipation. Molecular beam epitaxy (MBE) has been instrumental in growing high-quality Bi₂Se₃ films as topological insulators, featuring robust helical edge states suitable for applications. Under Se-rich conditions at 300-400°C, MBE deposits quintuple layers of Bi₂Se₃ on sapphire or /SiC substrates, yielding films with thicknesses down to 5 nm and surface state mobilities above 1,000 cm²/V·s. The Dirac-like dispersion at the surface, protected by time-reversal , enables dissipationless spin currents, which are proposed for encoding Majorana fermions in topological qubits. Ultrathin films exhibit a crossover to two-dimensional topological behavior, suppressing bulk conduction and enhancing coherence times for processing. Such MBE-grown Bi₂Se₃ structures integrate with superconducting contacts to realize proximity-induced topological , a key step toward fault-tolerant quantum bits. Multifunctional oxide heterostructures, such as epitaxial BaTiO₃ on SrTiO₃, leverage ferroelectric properties for devices with high endurance. Pulsed laser deposition or at 600-700°C grows c-axis-oriented BaTiO₃ films on (001) SrTiO₃, achieving tetragonal phase stabilization through 1-2% compressive strain, which enhances remnant to over 50 μC/cm². This strain-induced enables 64-level multilevel states in resistive switching memory, with retention exceeding 10¹⁰ cycles and switching speeds below 10 ns. The forms a in SrTiO₃, coupling ferroelectric domains to conductive channels for low-power operation in . Post-2020 advances in hybrid epitaxy have propelled solar cells toward efficiencies above 25% by combining vapor and solution methods for high-crystallinity films. Quasi-epitaxial growth initiates on lattice-matched substrates like Au(001), followed by solution infiltration to form oriented MAPbI₃ layers with grain sizes over 1 μm and defect densities below 10¹⁶ cm⁻³. This hybrid approach yields power conversion efficiencies of 25.5% in single-junction devices, attributed to reduced non-radiative recombination and improved charge extraction. For inorganic variants like CsPbBr₃, ensures epitaxial alignment, boosting open-circuit voltages to 1.5 V and stabilities over 1,000 hours under illumination. These techniques address phase instability, paving the way for scalable exceeding 30% efficiency.

References

  1. [1]
    Epitaxy - an overview | ScienceDirect Topics
    Epitaxy is crystal growth where new layers form with order given by a substrate, replicating its orientation, and is a single-crystal film formation.
  2. [2]
    An Introduction to Epitaxy - AZoM
    Feb 14, 2019 · Epitaxy is a crystallography technique where crystals grow on a substrate, forming thin films that lock into crystallographic orientations.What Is Epitaxy? · Types Of Epitaxy · Uv-Vis-Nir Microspectroscopy...
  3. [3]
    [PDF] Chapter 3: Epitaxy
    Epitaxy is a process to grow a thin crystalline layer on a crystalline substrate, where the seed crystal is the substrate.
  4. [4]
    1960: Epitaxial Deposition Process Enhances Transistor Performance
    In 1951 Gordon Teal and Howard Christensen at Bell Labs developed a process, now called epitaxial deposition, to grow a thin layer of material on a substrate ...
  5. [5]
    The Epitaxy (Epi) Process in Semiconductor Fabrication | Cadence
    Sep 29, 2025 · The epitaxy (epi) process in semiconductor fabrication aims to deposit a fine layer of single crystal, usually around 0.5 to 20 microns, on a single crystal ...Key Takeaways · The Epitaxy (epi) Process In... · Types Of Epi Processes In...Missing: fundamentals - - | Show results with:fundamentals - -
  6. [6]
    Dissertation — Heteroepitaxy and Selective Epitaxial Growth - IuE
    2.1 Fundamentals of Epitaxy. The term epitaxy refers to the ordered deposition of well-defined crystal layers on top of a bulk crystal (substrate). The incoming ...2.1 Fundamentals Of Epitaxy · 2.1. 2 Crystal Facets And... · 2.2 Modeling Epitaxy
  7. [7]
    Review Fundamental aspects of organometallic vapor phase epitaxy
    The bonding at the surface during OMVPE growth is found to be determined mainly by thermodynamic factors. However, features such as steps and kinks are ...
  8. [8]
    [PDF] Understanding thin film formation through Molecular Beam epitaxy ...
    Determining how the impinging atoms in the vapor phase interact with the substrate surface is a key component to understand the nucleation and growth mechanism ...
  9. [9]
    Numerical simulations of island formation in a coherent strained ...
    ... shear modulus and Poisson's ratio, and ε0 is the misfit strain. The stress field must be in mechanical equilibrium, which requires σij,j=0. The surface is ...
  10. [10]
    Introduction | SpringerLink
    Royer established in 1928 the conditions for oriented overgrowth, defining the term epitaxy (“arrangement on”), too [1.5]. He formulated the following rule ...
  11. [11]
    Epitaxial Growth of Semiconductors (Chapter 1) - Low-Dimensional ...
    Royer carried out an extensive study of the growth of ionic crystals on one another and on mica, mainly from aqueous solution and, using optical microscopy, ...Missing: early | Show results with:early
  12. [12]
    Milestones:Molecular Beam Epitaxy, 1968–1970
    Oct 24, 2025 · In 1968–1970, Molecular Beam Epitaxy (MBE) techniques using reflection high-energy electron diffraction for growing epitaxial compound ...
  13. [13]
    III–V semiconductor devices grown by metalorganic chemical vapor ...
    Thus, the First International Conference on Metal-Organic Vapor Phase Epitaxy” (ICMOVPE-I) was held in Ajaccio, Corsica, France in early May 1981. This ...
  14. [14]
    The Nobel Prize in Physics 2000 - NobelPrize.org
    Alferov and Herbert Kroemer "for developing semiconductor heterostructures used in high-speed- and opto-electronics" and the other half to Jack S. Kilby ...
  15. [15]
    A Short History of Atomic Layer Deposition: Tuomo Suntola's Atomic ...
    Oct 15, 2014 · Tuomo Suntola decided to call the new growth technology “atomic layer epitaxy”. Epitaxy originates from Greek, and means on-arrangement. It was ...
  16. [16]
    Graphene and Beyond: Recent Advances in Two-Dimensional ...
    We aim to highlight the most recent discoveries in the following topics: theory-guided synthesis for enhanced control of 2D morphologies, quality, yield, as ...
  17. [17]
    New synthesis technology for single-crystal 2D semiconductors ...
    Mar 5, 2025 · A research team has successfully developed a new synthesis technology for 2D semiconductors. This technique enables the direct growth of ...
  18. [18]
    Homoepitaxial Growth - an overview | ScienceDirect Topics
    Homoepitaxial growth is defined as the deposition of an epitaxial layer of the same material as the substrate, allowing for the creation of smooth, ...
  19. [19]
    Molecular Beam Epitaxy: Principals, Advantages and Challenges
    7.1 Homoepitaxy. In the first approach, which is called homoepitaxy, the grown layer and the substrate are of the same material. In other words, the thin ...
  20. [20]
    High-quality and high-purity homoepitaxial diamond (100) film ...
    Sep 16, 2015 · Thus, dislocation density in the homoepitaxial layer can be decreased drastically by the oxygen plasma etching after the mechanical polishing.Missing: advantages | Show results with:advantages
  21. [21]
    Silicon Based Epitaxial Thin Films - MKS Instruments
    Epitaxial films can be formed using a number of different methodologies, including evaporation, sputtering, molecular beam epitaxy, liquid phase epitaxy, and ...
  22. [22]
    (PDF) Growth kinetics of Si-molecular beam epitaxy - ResearchGate
    linearly dependent on Si-flux density and independant of temperature and orientation — indicates a condensation coefficient ...
  23. [23]
    Autodoping during homoepitaxy of silicon (Journal Article) | OSTI.GOV
    Sep 1, 1985 · Antimony, arsenic, phosphorus, and boron autodoping in silicon epitaxy · Thu Aug 01 00:00:00 EDT 1985 · J. Electrochem. Soc.; (United States) ...
  24. [24]
    Morphological instabilities during homoepitaxy on vicinal GaAs(110 ...
    Depending on the growth conditions, a striking variety of morphological instabilities have been found that range from step bunching in the Ga-supply limited ...Missing: challenges | Show results with:challenges
  25. [25]
    [PDF] The Influence of the Early Relaxation Phase on the ... - OPUS
    2.2 Heteroepitaxial Growth. Lattice Mismatch. Heteroepitaxy is defined as the epitaxial growth of a crystalline material onto an- other, where the two ...
  26. [26]
    [PDF] Elastic relaxation during 2D epitaxial growth: a study of in-plane ...
    This regime is usually called the pseudomorphic regime, corresponding to the accumulation of elastic energy when the thickness is increased. When this ...
  27. [27]
    [PDF] Defect Reductions in Epitaxial Growth Using Superlattice Buffer Layers
    In the usual strained-layer superlattices the lattice strain is produced by the enforced lattice mismatch of two crystals with different lattice constants.
  28. [28]
    Heteroepitaxy on high-quality GaAs on Si for optical interconnection ...
    In this paper, we therefore pay our attention to threading dislocations, mainly from the viewpoint of how to suppress their density. We focus our discussion on ...
  29. [29]
    [PDF] Epitaxial growth of highly mismatched III-V materials on (001) silicon ...
    Nov 14, 2017 · A unique type of defect associated with III-V/Si hetero-epitaxy is antiphase-domains (APDs), which arise from the lack of inversion symmetry of ...
  30. [30]
    [PDF] Reduced thermal conductivity of epitaxial GaAs on Si due ... - Bowers
    Mar 11, 2019 · There is an additional in-plane biaxial residual thermal stress of 250 MPa in sample s-Si due to the mismatch of the thermal expansion ...Missing: challenges | Show results with:challenges
  31. [31]
  32. [32]
  33. [33]
  34. [34]
    Heteroepitaxial growth modes revisited - RSC Publishing
    Sep 28, 2023 · There are three different growth modes for heteroepitaxy and these are the Frank–van der Merwe (FM), Volmer–Weber (VW), and Stranski–Krastanov ...Missing: seminal | Show results with:seminal
  35. [35]
  36. [36]
    [PDF] Low Temperature Silicon Selective Epitaxial Growth(SEG) and ...
    Apr 1, 1992 · Though SEG was first reported in 1962 [45], it has only recently overcome problems with defects, se:lectivity, and growth uniformity by ...
  37. [37]
    [PDF] The Science and Practice of Metal-Organic Vapor Phase Epitaxy ...
    The typical temperature range for growth is in the mass transport or diffusion limited regime where all of the TMGa is depleted at the surface and results in.
  38. [38]
    [PDF] Organometallic Vapor-Phase Epitaxial Growth of - DSpace@MIT
    This chapter presents an overview of OMVPE and spectroscopic ellipsometry, explaining backgound material necessary to develop a process model. This process ...
  39. [39]
    [PDF] Epitaxy
    Molecular Beam Epitaxy. • non-CVD vapor phase epitaxy via evaporation of material in ultra- high vacuum environment. By utilizing very low growth rates (≈.
  40. [40]
    [PDF] Heteroepitaxial Growth of Vertical GaAs Nanowires on Si (111 ...
    Sep 28, 2008 · Once epitaxial GaAs NWs are nucleated on Si substrates, the growth temperature and precursor flow rates are key parameters to the NW growth.
  41. [41]
    [PDF] Selective Area Growth, Etching, and Doping of GaN by MOCVD for ...
    This dissertation explores selective area growth, etching using TBCl, and doping of GaN for power electronics, using MOCVD.
  42. [42]
    [PDF] Modeling and Analysis of a Continuous Hydride Vapor Phase ...
    to develop a continuous hydride vapor phase epitaxy growth system capable of fabricating ... adsorption of the gas species in the kinetically-limited growth ...
  43. [43]
  44. [44]
    [PDF] Growth of III-V compounds by liquid phase epitaxy
    AT is limited to 5-10°C to prevent any 'spontaneous' nucleation and growth in the solution. ii) Equilibrium or ramp coolino. The solution is brought into ...Missing: advantages | Show results with:advantages
  45. [45]
    [PDF] Liquid Phase Epitaxy - ICTP
    It is the aim of this work to provide an overview and describe the more recent devel- opments of the LPE technique and to critically compare it with competing ...
  46. [46]
    [PDF] 7 - Solid-Phase Epitaxy - Projects at Harvard
    Thin Films and Epitaxy: Basic Techniques edited by T.F. Kuech (Elsevier North-Holland, Boston, 2015). Print book ISBN: 978-0-444-63304-0; e-book ISBN ...
  47. [47]
    Molecular dynamics simulations of the solid phase epitaxy of Si
    Sep 23, 2009 · The solid phase epitaxy of an amorphous layer on crystalline silicon is studied by means of molecular dynamics. Three stacks of 5120, 4928, ...
  48. [48]
    Epitaxy of oligothiophenes on alkali metal hydrogen phthalates
    Aug 6, 2025 · After reviewing the basic properties of molecular solids, the attempts ... solid phase epitaxy (SPE) ... [Show full abstract] based on ...
  49. [49]
    Lateral Solid-Phase Epitaxy of Oxide Thin Films on Glass Substrate ...
    May 27, 2014 · This technique is lateral solid-phase epitaxy, where epitaxial crystallization of amorphous precursor is seeded with ultrathin oxide nanosheets sparsely.
  50. [50]
    Direct observation of intra-grain defect formation during local solid ...
    Jul 13, 2025 · In this study, we directly observed elementary processes during crystal growth at local interfaces between crystalline Si (c-Si) grains and ...
  51. [51]
    Arsenic doping of GeSi epitaxial layers grown in the dichlorosiiane
    incorporation of arsenic as an n-type dopant of these layers in the H-Si-Ge-Cl system.5'6 Jung et al.' demonstrated that doping concentrations up to 2 x 10 ...
  52. [52]
    Ex-situ doping of epitaxially grown Ge on Si by ion-implantation and ...
    Apr 15, 2020 · In fact, it induces an ultra-fast liquid phase epitaxial regrowth which enhances the active dopant incorporation, in some cases even above solid ...
  53. [53]
    Segregation coefficients in Te-rich Hg Cd Te systems - ScienceDirect
    The expressions for segregation coefficients of Te-rich Hg Cd Te systems are derived with respect to liquid compositions as well as to liquid phase ...
  54. [54]
    Lattice diffusion and surface segregation of B during growth of SiGe ...
    Analysis of B segregation during growth shows that (i) for layers in epitaxy on (100)Si, B segregation decreases with increasing Ge concentration, i.e. ...
  55. [55]
    [PDF] Breaking the doping limit in silicon by deep impurities - arXiv
    In conclusion, we have demonstrated that the long-standing n-type doping limit in Si can be overcome by using deep-level donors (e.g. Te) instead of ...
  56. [56]
    Review—Dopant Selection Considerations and Equilibrium Thermal ...
    Mar 4, 2016 · Dopant incorporation techniques which require thermal treatment steps to move dopants onto lattice sites like ion implantation and monolayer ...
  57. [57]
    Surfactant behavior and limited incorporation of indium during in situ doping of GeSn grown by molecular beam epitaxy
    - **Surface Mobility**: Indium acts as a surfactant during GeSn epitaxy, accumulating on the surface and forming mobile Sn-In liquid droplets, which enhances Sn segregation and alters local material composition.
  58. [58]
    Molecular beam epitaxy of silicon: Effects of heavy Sb doping
    A model describing the incorporation of thermal dopants into single crystal films grown by molecular beam epitaxy (MBE) is presented.
  59. [59]
    [PDF] Resonant scattering in delta-doped heterostructures
    When the thick- ness of the inserted layer approaches a single monolayer, the term ''delta doping'' has been adopted. In these structures, it is necessary to ...
  60. [60]
    Electrical properties of Si-doped GaAs layers grown on (411)A GaAs ...
    This result suggests that Si can be used as an n-type dopant in GaAs for GaAs/AlGaAs resonant tunnelling diodes grown on (411)A GaAs substrates with atomically ...
  61. [61]
    Characterization of epitaxial GaAs MOS capacitors using atomic ...
    The Zn-doped epi-GaAs with a doping concentration of approximately 1018 cm-3 converts the epi-GaAs layer into p-type since the Zn doping is relatively higher ...
  62. [62]
    [PDF] Molecular-beam epitaxial growth and characterization of silicon ...
    n-p-n heterojunction bipolar transistors grown by all Si doping exhibit excellent current voltage characteristics and a common emitter current gain /3=240.
  63. [63]
    Si incorporation and Burstein–Moss shift in n-type GaAs
    Another important phenomenon occurring in the heavily donor-doped semiconductors is an increase in the interband transition energy due to the filling of the ...
  64. [64]
    Observation of the Burstein—Moss shift in heavily Te-doped In 0.5 ...
    The Burstein—Moss shift was observed for the first time in heavily Te-doped In0.5Ga0.5P epilayers grown by liquid phase epitaxy.
  65. [65]
    [PDF] lattice contraction due to carbon doping of GaAs grown by ...
    Epitaxial layers of GaAs have been grown by metalorganic molecular beam epitaxy ... lattice parameter of heavily carbon-doped GaAs. Compen- sation of the CAS ...
  66. [66]
    Effects of doping on the lattice parameter of SrTiO 3 - AIP Publishing
    Jun 27, 2012 · One obvious mechanism is through the size mismatch between the dopant impurity and the host atom for which it substitutes. ... doping of the ...
  67. [67]
    Gettering Sinks for Metallic Impurities Formed by Carbon-Cluster Ion ...
    Gettering sinks for metallic impurities formed by carbon-cluster ion implantation in epitaxial silicon wafers have been investigated using technology ...
  68. [68]
    Electrical properties of Gallium Arsenide (GaAs)
    Electron Hall mobility versus temperature for different doping levels and degrees of compensation (high temperatures): Open circles: Nd=4Na=1.2·1017 cm-3; Open ...
  69. [69]
    [PDF] A study of minority carrier lifetime versus doping concentration in nâ
    A study of minority carrier lifetime versus doping concentration in n-type GaAs ... Comprehensive studies using high-quality epitaxial material grown by modern ...
  70. [70]
    In Situ Observation of Epitaxial Growth during Evaporative ... - MDPI
    For smaller differences between the lattice parameters, the interfacial energy will also be low, accounting for the stronger tendency among these crystals to ...
  71. [71]
    When epitaxy controls garnet growth - SPIESS - Wiley Online Library
    Apr 11, 2007 · Garnet crystallographic orientation was analysed with electron backscatter diffraction (EBSD): the obtained crystallographic dispersion patterns ...Microstructure · Interpretation Of Pole... · Discussion
  72. [72]
    Mineral Specimen: Anatase and epitaxial Rutile - Fabre Minerals
    6AT47Q6: Floater growths of Anatase crystals, many of them doubly terminated, in parallel growth. The Anatase has been completely covered by epitaxial Rutile ...
  73. [73]
    [PDF] Computational Approach for Epitaxial Polymorph Stabilization ...
    May 4, 2016 · epitaxial phase. To include an ... In natural minerals, TiO2 has three well- known polymorphs, i.e., anatase, rutile, and brookite.<|control11|><|separator|>
  74. [74]
    [PDF] Revision 1 Volcanic SiO2-Cristobalite: A natural product of Chemical ...
    tridymite, which is also found in many volcanic rocks (Kayama et al. ... epitaxial lattice matching to proceed during VPC deposition and the subsequent transition.
  75. [75]
    High-speed SiGe HBTs and their applications - ScienceDirect
    The SiGe HBT was fabricated by selective-epitaxial growth (SEG). As a result of the optimization, its cutoff frequency was increased to 130 GHz and its ECL gate ...
  76. [76]
    Epitaxial growth of SiGe layers for BiCMOS applications
    The performance of Si has been extended through higher speed devices using the heterojunction bipolar transistors (HBT) and better current drivability in CMOS ...<|separator|>
  77. [77]
    SiGe HBT technology with fT/fmax of 300GHz/500GHz and 2.0 ps ...
    A SiGe HBT technology featuring fT/fmax/BVCEO=300GHz/500GHz/1.6V and a minimum CML ring oscillator gate delay of 2.0 ps is presented.Missing: strain cutoff
  78. [78]
    A BiCMOS technology featuring a 300/330 GHz (fT/fmax) SiGe HBT ...
    The paper presents a 0.13 mum SiGe BiCMOS technology for millimeter wave applications. This technology features a high performance HBT (fT = 300 GHz /fmax ...
  79. [79]
    AlGaN/GaN-based multi-channel epitaxial structure with an ultra-low ...
    Apr 15, 2025 · This multi-channel structure with an ultra-low 2DEG density is highly suitable for enhanced-mode FinFETs, offering significant potential for radio-frequency ...
  80. [80]
    AlInGaN/GaN double-channel FinFET with high on-current and ...
    The fabricated AlInGaN/GaN double channel FinFETs exhibit considerably higher maximum drain current of 290 mA/mm than those of AlGaN/GaN FinFETs.
  81. [81]
    Modeling Electrostatics and Low-Field Electron Mobility of GaN ...
    Jul 22, 2022 · Gallium nitride (GaN) high electron mobility transistors (HEMTs) are currently being used for RF applications due to the intrinsically high ...Missing: AlGaN effect
  82. [82]
    Epitaxial lateral overgrowth of III-V semiconductors on Si for ...
    In this chapter we describe how epitaxial lateral overgrowth (ELOG) method can be employed to reduce the threading dislocation density in III-V layers on Si.Missing: ELO | Show results with:ELO
  83. [83]
    Bending of dislocations in GaN during epitaxial lateral overgrowth
    Nov 15, 2004 · Epitaxial lateral overgrowth (ELO) is a growth method used to reduce the density of dislocations in semiconductor films. ELO was used for the ...
  84. [84]
    Epitaxial Bonding and Transfer Processes for Large-Scale ...
    Nov 28, 2018 · A process flow for the heterogeneous integration of III-V epitaxial material onto a silicon host wafer using CMOS-compatible materials and methods
  85. [85]
    3D-Stacked CMOS Takes Moore's Law to New Heights
    Aug 11, 2022 · We've created experimental devices that stack atop each other, delivering logic that is 30 to 50 percent smaller.
  86. [86]
    [PDF] Shuji Nakamura - Nobel Lecture
    With the success of the developed high efficiency, high power blue LED,. Nichia Chemical Corporation commercialized the first white LEDs by combin- ing the blue ...Missing: original | Show results with:original
  87. [87]
    Nobel Lecture: Background story of the invention of efficient blue ...
    Oct 5, 2015 · Background story of the invention of efficient blue InGaN light emitting diodes. Shuji Nakamura. Shuji Nakamura. University of California, Santa Barbara, ...Missing: MBE Prize
  88. [88]
    Ultrahigh Responsivity-Bandwidth Product in a Compact InP ...
    Sep 23, 2016 · Direct bandgap III-V based compound semiconductors, particularly the InP/InGaAs system, have at least an order of magnitude higher absorption ...
  89. [89]
    High-Q photonic crystal cavities in all-semiconductor photonic ...
    Jun 14, 2017 · We investigate a heterostructure-based approach comprising a high refractive index core and lower refractive index cladding layers.
  90. [90]
    III-V quantum dot enhanced photovoltaic devices - SPIE Digital Library
    Aug 24, 2010 · Using strain-balanced Stranski-Krastanov QD formation, we have demonstrated sub-gap photon collection and increased current in QD-enhanced ...
  91. [91]
    Epitaxial growth of crystal phase quantum dots in III–V ...
    Mar 6, 2023 · Crystal phase quantum dots (QDs) are formed during the axial growth of III–V semiconductor nanowires (NWs) by stacking different crystal phases of the same ...Missing: photovoltaics | Show results with:photovoltaics
  92. [92]
    [1307.1555] Highly efficient spin transport in epitaxial graphene on SiC
    Jul 5, 2013 · We report here on highly efficient spin transport in two-terminal polarizer/analyser devices based on high-mobility epitaxial graphene grown on silicon carbide.
  93. [93]
    [PDF] GROWTH OF QUANTUM WELL FILMS OF TOPOLOGICAL ... - arXiv
    High quality quantum well films of. Bi2Se3, a typical three-dimensional topological insulator, have been grown on α-Al2O3 (sapphire). (0001) by molecular beam ...
  94. [94]
    Crossover of the three-dimensional topological insulator Bi 2 Se 3 to ...
    Jun 13, 2010 · The topological protection of the surface state could be useful for both low-power electronics and error-tolerant quantum computing. For a thin ...
  95. [95]
    Growth of ultrathin Bi2Se3 films by molecular beam epitaxy
    Dec 22, 2022 · We explore the growth of Bi 2 Se 3 films having thicknesses down to 4 nm on sapphire substrates using molecular beam epitaxy that were then characterized with ...
  96. [96]
    Lead-free epitaxial ferroelectric material integration on ... - Nature
    Jul 23, 2015 · We report lead-free ferroelectric based resistive switching non-volatile memory (NVM) devices with epitaxial (1-x)BaTiO 3 -xBiFeO 3 (x = 0.725) (BT-BFO) film ...
  97. [97]
    A review of molecular beam epitaxy of ferroelectric BaTiO3 films on ...
    We review the growth by MBE of the ferroelectric compound BaTiO 3 on silicon (Si), germanium (Ge) and gallium arsenide (GaAs) and we discuss the film ...
  98. [98]
    Epitaxial inorganic metal-halide perovskite films with controlled ...
    Mar 20, 2023 · We demonstrate the epitaxial growth of several nanometer thick purely (001)-orientated films of C s P b B r 3 and C s S n B r 3 MHPs on Au(001) by molecular ...
  99. [99]
    Perovskite solar cells with high-efficiency exceeding 25%: A review
    Feb 4, 2024 · The certified efficiency of perovskite solar cells has reached 26.1%, but only a few have exceeded 25%, with strategies to improve efficiency.