Fact-checked by Grok 2 weeks ago

Resonant-tunneling diode

A resonant-tunneling diode (RTD) is a featuring a double-barrier structure that enables electrons to tunnel resonantly through discrete energy states, resulting in a negative differential (NDR) region in its current-voltage characteristics where current decreases with increasing voltage. This quantum effect allows RTDs to operate at frequencies, making them among the fastest electronic devices available. Proposed theoretically by and Raphael Tsu in 1973 as part of their work on tunneling in semiconductor superlattices, the RTD concept built on earlier tunneling phenomena observed in Esaki diodes. Experimental realization followed in 1974 with the demonstration of resonant tunneling in GaAs/AlGaAs double-barrier structures, confirming the predicted NDR behavior at low temperatures. Over the subsequent decades, advancements in materials such as GaAs/AlGaAs and InGaAs/AlAs have enabled room-temperature operation and peak-to-valley current ratios exceeding 30, enhancing device reliability for practical use. RTDs are notable for their potential in high-speed digital logic, where the NDR facilitates multistate cells and low-power switching with intrinsic speeds on the scale. In analog applications, they serve as compact oscillators and detectors with demonstrated operation up to 740 GHz, supporting communications and systems beyond 1 THz. Ongoing research focuses on integrating RTDs with bipolar transistors (HBTs) and silicon-compatible materials to realize nanoelectronic circuits for future computing paradigms.

Overview

Definition and Basic Principles

A resonant-tunneling diode (RTD) is a two-terminal that exploits quantum mechanical tunneling through a double-barrier heterostructure to exhibit negative differential resistance. This structure enables electrons to traverse potential barriers via resonant states, distinguishing it from classical transport mechanisms. Resonant tunneling occurs when the energy of incident electrons aligns with the discrete quantized energy levels of a sandwiched between two thin barriers, resulting in sharp peaks in the transmission probability. At these resonance conditions, electrons coherently tunnel through the barriers with minimal reflection, allowing efficient current flow. This phenomenon arises from wave-like quantum , where the phase matching between the well and barriers enhances transmission. The basic structure consists of an emitter region, followed by a first thin barrier, a quantum well, a second thin barrier, and a collector region, all typically formed in a heterostructure to confine carriers. Electrons from the emitter inject into the well via tunneling through the first barrier and exit to the collector through the second when resonance is achieved. The transmission probability T(E) for an of E approaches unity at resonance, as derived from solving the time-independent for a one-dimensional double-barrier potential. In contrast to the conventional , which relies on band-to-band tunneling in a heavily doped p-n junction, the RTD achieves a sharper negative resistance through the resonant alignment in the .

Historical Development

The theoretical foundation for resonant-tunneling diodes (RTDs) was laid in 1973 by Raphael Tsu and , who predicted negative resistance arising from resonant tunneling in superlattices through a tunneling transport model. This work built on Esaki's earlier 1958 invention of the , for which he shared the 1973 with and , recognizing foundational contributions to quantum tunneling in solids. The prediction highlighted how s could resonantly transmit through periodic potential barriers, enabling sharp current peaks in current-voltage characteristics. The first experimental demonstration of resonant tunneling occurred in 1974 at , where and Leroy L. Chang fabricated a GaAs-AlGaAs double-barrier structure that exhibited negative differential resistance at low temperatures (around 77 K), with an initial peak-to-valley current ratio (PVR) of approximately 1.3. Early devices faced significant challenges, including thermally broadened energy levels that limited PVR and confined operation to cryogenic conditions, as the resonant peaks were sensitive to scattering and interface imperfections in the nascent () growth techniques. Throughout the 1980s, advancements in enabled room-temperature operation, first achieved in 1988 using an AlGaAs/GaAs double-barrier with a quasiparabolic , demonstrating clear resonant peaks at 300 K. By the , PVR values improved dramatically to over 10—and as high as 50 in InGaAs-based structures—through optimized barrier compositions and strain engineering, enhancing device reliability for practical applications. Integration into circuits advanced during this decade, with monolithic combinations of RTDs and transistors enabling high-speed logic and oscillators, as reviewed in early studies. In the late 1990s, research shifted toward interband RTDs using Si/SiGe heterostructures to achieve compatibility with silicon-based processes, culminating in room-temperature devices with PVR greater than 2 by 1999 via low-temperature . This transition addressed the limitations of III-V materials for mainstream integration, paving the way for quantum-classical up to the early 2000s.

Operation

Current-Voltage Characteristics

The current-voltage (I-V) characteristic of a resonant-tunneling diode (RTD) exhibits a distinctive S-shaped , featuring a sharp increase in to a peak when the applied bias aligns the of the emitter with the quasi-bound state in the , followed by a decrease to a valley as the misaligns, resulting in negative resistance (NDR) between the peak and valley. This behavior arises from the resonant tunneling mechanism, where electrons tunnel through the double-barrier structure only when energy levels match. The peak current density J_p is primarily determined by the barrier height, quantum well width, and doping levels in the contacts, as these factors influence the transmission probability and the available for tunneling. Typical values for J_p in well-designed RTDs range from $10^4 to $10^5 A/cm², enabling high-speed operation while maintaining compactness. The valley current density J_v is minimized by suppressing non-resonant tunneling paths through optimized barrier and reduced , which helps achieve high peak-to-valley ratios for functionality. An approximate expression for the steady-state I-V characteristic is given by the Tsu-Esaki formula: I(V) \propto \int T(E) [f(E - eV) - f(E)] \, dE where T(E) is the energy-dependent transmission probability through the double barriers, f is the Fermi-Dirac distribution function, e is the electron charge, and V is the applied voltage; this integral accounts for the net current from occupied states in the emitter to empty states in the collector. At low temperatures, can appear in the I-V curve due to charging effects in the , where accumulated electrons shift the potential and alter the resonance condition during bias sweeps. This effect diminishes at higher temperatures as promotes faster charge relaxation. Overall, the I-V characteristics show strong temperature dependence: the peak sharpens and the NDR region becomes more pronounced at cryogenic temperatures due to reduced broadening of the Fermi , while at , increased broadens the peak and reduces the peak-to-valley ratio.

Resistance Regions

In the current-voltage (I-V) characteristics of a resonant-tunneling diode (), three distinct resistance regions emerge due to the alignment and misalignment of electron energies with the quantum well resonance states in the double-barrier structure. These regions correspond to specific bias ranges and enable unique device functionalities, such as switching and . The first positive resistance region occurs at low forward bias, prior to the peak current, where the differential conductance dI/dV > 0 exhibits ohmic-like behavior. In this regime, electron accumulation builds up in the emitter as the applied bias aligns the Fermi level with the quantized energy states in the quantum well, facilitating increasing resonant tunneling probability and thus rising current with voltage. Following the peak current at the negative resistance region, spanning from the peak voltage to the valley voltage, the differential conductance becomes negative (dI/dV < 0), as current decreases despite increasing bias. This negative differential resistance (NDR) arises from reduced transmission probability when the bias shifts the resonant energy level above the emitter Fermi sea, misaligning it and suppressing electron injection into the well. The NDR enables high-frequency oscillations in AC circuits but requires careful bias stabilization to prevent bistability or hysteresis in DC operation. Beyond the valley current, the second positive resistance region sets in, where dI/dV > 0 again and current increases with further bias. Here, conduction proceeds via off-resonance tunneling through the barriers or due to band misalignment allowing broader transmission, often supplemented by over the barriers at higher energies. Transitions between these regions are defined by bias points where dI/dV = 0, specifically at the and voltages, marking the onset and end of NDR. Stability analysis reveals that while the NDR region supports multistate logic operations—leveraging the and states for or higher logic levels— biasing within it demands circuit techniques like load-line stabilization to avoid switching instabilities.

Tunneling Mechanisms

Intraband Resonant Tunneling

Intraband resonant tunneling refers to the quantum mechanical process in resonant-tunneling diodes (RTDs) where charge carriers, typically electrons, through a double-barrier structure within the same energy band, most commonly the conduction band in n-type devices. In this unipolar mechanism, electrons from the emitter region's conduction band miniband incident on the first barrier can coherently into discrete quasi-bound states confined in the central and subsequently escape through the second barrier into the collector's conduction band when energy alignment is achieved. This process relies on the wave-like nature of electrons, with the transmission probability peaking sharply at resonant energies due to constructive interference in the well, as first theoretically described for superlattices and double-barrier configurations. The overall tunneling is intraband, avoiding band-to-band transitions and enabling operation with a single carrier type, which simplifies device physics compared to alternatives. The conduction band energy diagram of an intraband RTD illustrates two thin, high-bandgap barriers flanking a narrower-bandgap , forming potential steps that confine electrons laterally while allowing vertical transport. Under applied bias, the bands tilt, and resonance occurs when the emitter's E_F aligns with a quasi-bound state E_n in the well, maximizing the tunneling . The energy levels of these quasi-bound states are determined by the well's confinement, approximated by the infinite square well model where the characteristic scale \Delta E for the ground state or level spacing is given by \Delta E = \frac{\hbar^2 \pi^2}{2 m^* L_w^2}, with m^* as the effective electron mass in the well and L_w as the well width (typically 4-6 nm). This alignment condition leads to negative differential resistance (NDR) as the bias shifts the resonance out of alignment, suppressing current. Precise engineering of the heterostructure ensures the barriers provide sufficient confinement without excessive leakage. One key advantage of intraband RTDs is their simpler band alignment, requiring only heterojunction offsets within the conduction band, which facilitates unipolar electron transport and higher operational speeds—often exceeding 1 THz—due to the lower effective mass and reduced carrier scattering compared to hole-involved processes. However, realizing effective confinement demands barrier heights of approximately 200-500 meV, necessitating high-quality epitaxial growth to maintain sharp interfaces. Performance is particularly sensitive to interface roughness, which scatters electrons, broadens resonant peaks, and degrades NDR by increasing valley current. In optimized GaAs/AlGaAs intraband RTD structures, peak-to-valley ratios (PVRs) typically range from 5 to 20 at low temperatures, though room-temperature values are often lower due to thermal broadening.

Interband Resonant Tunneling

Interband resonant tunneling diodes (RITDs) operate through a mechanism where carriers, typically electrons, tunnel from the conduction band of the emitter material into confined states within the valence band of a quantum well, enabled by type II band alignment in heterostructures. This process combines elements of Esaki interband tunneling with resonant enhancement, allowing for negative differential resistance at low biases. In representative systems like InAs/AlSb/GaSb, electrons from the InAs conduction band tunnel through AlSb barriers into GaSb valence band states, with subsequent transport resembling electron-like behavior after interband transfer. Alternatively, in bipolar configurations, hole tunneling from the valence band into the well can initiate the process, followed by electron injection for overall current flow. The energy alignment in interband RTDs requires type II band offsets, where the conduction band minimum of the emitter aligns closely with the band maximum of the material, facilitating overlap between states. occurs when the valence band edge of the matches a quantized state in the well, typically at biases of 0.3–0.4 V, as seen in Si/SiGe structures where strain-induced splitting of light- and heavy-hole bands (by ~80 meV) and conduction valleys (by ~310 meV) enables precise alignment. This configuration contrasts with intraband tunneling by crossing bandgaps, leading to inherently involvement of conduction and valence carriers. A key advantage of interband RTDs over intraband variants is their enhanced compatibility with silicon-based technologies, as demonstrated in Si/SiGe heterostructures grown epitaxially on substrates using low-temperature processes integrable with fabrication. The heavier effective masses of s in silicon materials (e.g., ~0.26 m₀ for electrons versus ~0.067 m₀ in GaAs) contribute to reduced rates in the tunneling path, supporting stable operation at with yields >95% after annealing at 700–800 °C. However, challenges include lower due to the involvement of heavy-hole states in the process, resulting in peak-to-valley ratios (PVRs) typically ranging from 2 to 10, significantly below the >30 achievable in optimized III-V intraband RTDs. Additionally, achieving uniform band alignment remains difficult owing to sensitivity to material quality and doping profiles. The transmission probability for interband resonant tunneling is often modeled using a modified Wentzel-Kramers-Brillouin (WKB) approximation that accounts for band offsets and position-dependent effective masses across the heterojunction. The general form for the tunneling exponent is \int_{x_1}^{x_2} k(x) \, dx = \frac{2}{\hbar} \int_{x_1}^{x_2} \sqrt{2 m^*(x) [V(x) - E]} \, dx, where m^*(x) varies due to band structure changes, V(x) incorporates the type II offsets, and resonance enhances the probability when E aligns with well states, as adapted from Kane's interband model. This approach provides a semiclassical estimate for the current, with quantum corrections via envelope functions for precise simulation.

Material Systems

III-V Semiconductors

III-V semiconductors form the cornerstone of high-performance resonant tunneling diodes (RTDs) due to their tunable band structures and superior transport properties. The most prevalent configuration employs gallium arsenide (GaAs) as the quantum well material, flanked by aluminum gallium arsenide (AlGaAs) barriers, which provide a conduction band offset suitable for confining electrons in the well while enabling resonant tunneling. This GaAs/AlGaAs system was the basis for the first experimental observation of resonant tunneling in 1974. For applications demanding higher frequencies, indium phosphide (InP)-based heterostructures, such as indium gallium arsenide (InGaAs) wells with indium aluminum arsenide (InAlAs) or aluminum arsenide (AlAs) barriers, are widely adopted, leveraging the narrower bandgap of InGaAs for enhanced tunneling probabilities at terahertz energies. The efficacy of III-V materials in RTDs stems from their direct bandgaps, which promote efficient radiative recombination and minimize non-radiative losses, alongside exceptionally high electron mobilities—typically exceeding 8500 cm²/V·s in undoped GaAs at room temperature. These properties yield low effective electron masses and rapid transit times across the device, critical for achieving negative differential resistance at high speeds. Furthermore, the close lattice matching between GaAs and AlGaAs (lattice constant mismatch <0.1%) results in low defect densities at heterointerfaces, preserving quantum coherence and reducing scattering that could broaden resonant peaks. In InP-based systems, similar lattice compatibility with InGaAs and InAlAs further suppresses dislocations, enabling defect-free growth via molecular beam epitaxy. Intraband resonant tunneling in III-V RTDs, involving electron transport within the conduction band, was theoretically proposed by Tsu and Esaki in 1973 using a GaAs/AlGaAs superlattice model to predict negative differential resistance from quantum interference. This concept was experimentally realized shortly thereafter in GaAs/AlGaAs double-barrier diodes at low temperatures (4.2 K), demonstrating the predicted current-voltage characteristics. Room-temperature operation was achieved in 1988, with early devices exhibiting a peak-to-valley ratio (PVR) of approximately 1.5. For terahertz applications, InGaAs/InAlAs RTDs on InP substrates have excelled, with devices oscillating fundamentally at frequencies up to 1.98 THz (as of 2018) while delivering output powers on the order of microwatts. Interband resonant tunneling, which couples conduction and valence bands across heterointerfaces, is facilitated in III-V systems like InAs/AlSb/GaSb, where type-II band alignment creates a broken-gap offset that aligns the InAs conduction band minimum below the GaSb valence band maximum. This configuration enables electrons to tunnel resonantly from the InAs conduction band into GaSb heavy-hole states, yielding pronounced negative resistance suitable for low-voltage logic and infrared detection. Overall, III-V RTDs exhibit typical performance metrics including oscillation frequencies reaching 1 THz and PVR values exceeding 50 under cooled operation (e.g., 77 K), highlighting their potential for ultrafast electronics.

Si/SiGe Heterostructures

Si/SiGe heterostructures form the basis for compatible with silicon-based technologies, utilizing strained SiGe quantum wells sandwiched between Si barriers to create the necessary potential wells for tunneling. The typical structure consists of a SiGe well with Ge content around 20-30%, flanked by thin Si barriers (e.g., 5 nm thick), grown on relaxed SiGe virtual substrates or Si substrates via or . Strain in these layers induces band offsets of approximately 100-200 meV, primarily in the valence band (up to 230 meV for 26% Ge), enabling quantum confinement while maintaining compatibility with standard silicon processing. For intraband resonant tunneling in n-type Si/SiGe RTDs, the conduction band offset, around 180 meV for Si/Si0.7Ge0.3 interfaces, facilitates tunneling through the states in the . This offset arises from the strain-induced splitting of the Δ valleys, allowing to resonate at specific energies despite the relatively small barrier height compared to III-V systems. In interband variants, valence band engineering via higher content (e.g., 40-50%) aligns heavy- and light-hole states with minima, promoting Zener-like tunneling between ; for instance, a 3-nm Si0.54Ge0.46 layer achieves peak current densities up to 1.5 kA/cm² with peak-to-valley current ratios of 6. A key advantage of Si/SiGe RTDs is their monolithic integration with Si CMOS and SiGe HBTs, enabling compact, three-terminal negative differential resistance devices without the need for hybrid bonding, unlike III-V counterparts. This compatibility leverages mature silicon fabrication infrastructure, resulting in lower production costs and scalability for VLSI applications. However, challenges persist due to the indirect bandgap of Si and SiGe, which limits carrier mobility and tunneling efficiency compared to direct-bandgap materials, and the need for precise defect control in strained layers to minimize dislocations and vacancies that degrade peak-to-valley ratios. Low-temperature growth techniques followed by controlled annealing (700-900°C) are employed to mitigate these defects.

Emerging Materials

Recent research has explored two-dimensional materials for RTDs, such as graphene-based structures, which offer potential for high peak-to-valley ratios (exceeding 10) at room temperature and improved integration with flexible electronics. These systems leverage van der Waals heterostructures to engineer tunable barriers and wells.

Fabrication and Structure

Layer Design and Growth Techniques

The layer stack of a resonant tunneling diode (RTD) typically consists of an emitter contact region, followed by an undoped spacer layer, a first thin barrier (typically 5-10 nm thick), a quantum well (3-10 nm thick), a second symmetric or asymmetric barrier, another undoped spacer, and a collector contact region. The emitter and collector contacts are heavily doped n-type semiconductors, such as GaAs at concentrations of 4 × 10¹⁸ cm⁻³, with graded doping profiles transitioning to lower levels (e.g., 5 × 10¹⁶ to 6 × 10¹⁷ cm⁻³) near the spacers to form depletion regions and facilitate carrier accumulation. Undoped spacers, often 5 nm thick, prevent dopant diffusion into the active region and minimize scattering. Barrier height is engineered through alloy composition, such as varying the aluminum fraction in AlₓGa₁₋ₓAs (e.g., x = 0.3-0.4) to achieve conduction band offsets of 200-500 meV relative to the GaAs well, while well width directly tunes the quantized energy by confining states (e.g., a 5 nm well supports ground-state energies around 100-150 meV). These parameters ensure alignment of the quasi-Fermi level with the resonant state under bias for efficient tunneling. Epitaxial growth primarily employs (MBE) for its atomic-layer precision, enabling control over thicknesses to within one (≈0.28 nm for GaAs) at substrate temperatures of 580-630°C and growth rates of 0.25-1 per second. Metal-organic chemical vapor deposition (MOCVD) serves as an alternative for scaling to larger wafers, though it offers slightly less interface sharpness compared to MBE. Quality metrics emphasize interface sharpness below 1 to reduce valley current from and ensure high peak-to-valley ratios, alongside wafer uniformity for reproducible in fabrication. Self-consistent Schrödinger-Poisson solvers are used to optimize layer alignments by iteratively solving the time-independent for bound states in the well and for electrostatic potentials, accounting for doping-induced space charges.

Processing and Integration Methods

Fabrication of resonant-tunneling diodes (RTDs) involves post-growth processing steps to define device structures and ensure electrical isolation. Mesa etching is a standard technique for isolating individual RTD devices by removing surrounding material to form isolated mesas, typically using or wet chemical etching to achieve precise vertical sidewalls and minimize surface recombination. Following etching, dielectric passivation layers, such as , are often deposited to protect the mesa sidewalls and reduce leakage currents. Ohmic contacts are then formed to the heavily doped emitter and collector regions; for n-type III-V materials like GaAs, AuGe/Ni/Au metallization is commonly evaporated and annealed at around 400–450°C to achieve low-resistance contacts with specific resistivities below 10^{-6} Ω·cm². For high-frequency applications, air-bridge structures are integrated over coplanar waveguides to suppress slotline modes and minimize inductive discontinuities, enabling efficient signal propagation in monolithic microwave integrated circuits. Integration of RTDs with other technologies focuses on leveraging their negative differential resistance for enhanced circuit performance while addressing compatibility issues. Monolithic integration with silicon is achieved through selective , where RTD heterostructures are grown in patterned windows on substrates using low-pressure metalorganic vapor-phase , allowing co-fabrication with standard processes on the same . In III-V platforms, hybrid integration with transistors (HBTs) on InP substrates combines RTDs and HBTs in a single epitaxial stack, enabling compact circuits through shared processing steps like shared-base contacts. For /SiGe RTDs, fabrication utilizes for strained heterostructures. These devices can adopt vertical geometries for high current densities or planar configurations to simplify integration with planar transistors, with vertical designs offering better isolation but requiring deeper etching. Key challenges in RTD processing and integration include minimizing and managing thermal effects in dense circuits. arises from mesa sidewalls and interconnects, which can degrade high-frequency response; reduction strategies involve optimized profiles and air-bridge suspensions to lower effective below 0.1 pF for applications. Thermal management is critical due to self-heating from high current densities exceeding 10^5 A/cm², particularly in integrated arrays, where heat dissipation is enhanced via substrate thinning or integration with high-thermal-conductivity heat sinks to prevent valley current runaway. Demonstrated examples highlight practical integration successes. RTD-HBT oscillators, fabricated on InP substrates, achieve oscillation frequencies up to 20 GHz with output powers around 1 mW, combining the RTD's with HBT amplification for low-phase-noise sources. CMOS-RTD logic gates, integrated via hybrid bonding or monolithic selective growth, were demonstrated in the , enabling pipelined operations with power reductions of up to 50% compared to pure at supply voltages below 1 V.

Performance Characteristics

Electrical Metrics

The peak-to-valley ratio (PVR), defined as the ratio of the peak current density J_p to the valley current density J_v (i.e., PVR = J_p / J_v), serves as a primary for resonant-tunneling diodes (RTDs), quantifying the sharpness of the negative differential resistance (NDR) region in their current-voltage characteristics. For practical functionality in circuits, such as oscillators or logic elements, a PVR exceeding 10 is generally required to ensure sufficient contrast between resonant and non-resonant conduction paths, enabling reliable switching and amplification. This ratio is strongly influenced by barrier asymmetry; symmetric double-barrier structures promote higher PVR by equalizing tunneling probabilities and reducing asymmetric leakage, while intentional asymmetry can tune the voltage positioning of the NDR peak but often lowers the overall ratio. Current densities in RTDs highlight their potential for high-power, compact devices. In InP-based structures, such as InGaAs/AlAs RTDs, peak current densities J_p can exceed $10^6 A/cm², achieved through precise control of quantum well width and barrier thickness to align resonant states with the at low bias. The valley current density J_v is minimized using delta-doping techniques, which confine dopants to atomic monolayers outside the , thereby suppressing ionized impurity scattering and thermionic leakage in the while maintaining high emitter efficiency. These approaches yield low J_v values on the order of 10-100 A/cm² in optimized III-V RTDs, enhancing the device's overall PVR without compromising peak performance. In the NDR region, the differential resistance R = dV/dI exhibits negative values, typically ranging from -50 to -200 Ω, reflecting the steep drop in current after the resonance peak. This negative resistance arises from the misalignment of the quasi-bound state in the with the as bias increases, and its magnitude depends on device area and material parameters; smaller mesas (e.g., 1-5 μm diameter) often show more negative values due to reduced parasitic effects. Temperature significantly impacts RTD electrical metrics, primarily through thermal smearing of the resonant tunneling process. Above 300 , PVR degrades markedly—often dropping by 50% or more—due to increased phonon-assisted and broadening of the , which populates non-resonant states and elevates J_v. The valley current typically follows an Arrhenius-like temperature dependence, with activation energies of 50-150 meV associated with over the collector barrier or defect-assisted conduction. Compared across material systems, III-V RTDs generally exhibit superior PVR values, often reaching 15-20 at , owing to their abrupt heterointerfaces and tunable band offsets that minimize misalignment losses. In contrast, Si/SiGe heterostructure RTDs achieve typical PVRs around 5, limited by roughness, strain-induced defects, and less favorable effective masses that broaden the resonant linewidth.

Frequency and Speed Limits

The intrinsic speed of resonant-tunneling diodes (RTDs) is determined by the tunneling time through the double-barrier structure, which is on the order of 10^{-14} seconds (femtoseconds), specifically with dwell times around 33 and transition times around 38 in optimized structures. This ultrafast tunneling process enables potential operation beyond 1 THz, as the intrinsic response extends from 120 GHz to over 3.9 THz in demonstrated devices. The f_c for RTDs is fundamentally limited by the transit time \tau, given by f_c = \frac{1}{2\pi \tau}, where \tau encompasses the tunneling and dwell times. In InGaAs-based RTDs, this yields cutoff frequencies up to approximately 2 THz, with intrinsic estimates reaching 4.6 THz under ideal conditions. Extrinsic limitations arise primarily from series and associated with contacts and device geometry, which introduce an that degrades high-frequency performance. A key for overcoming these is the transconductance-to-capacitance ratio g_m / (2\pi C), where minimizing C (total including depletion and components) relative to g_m allows higher operational frequencies. Demonstrated oscillation frequencies in RTDs have reached up to 1.98 THz in fundamental mode, surpassing earlier pre-2020 limits around 700 GHz through improvements in epitaxial growth and antenna integration. Material systems significantly influence speed, with III-V semiconductors like InGaAs/AlAs outperforming /SiGe heterostructures due to higher and larger band offsets, enabling faster transit times and higher cutoff frequencies in III-V RTDs.

Applications

High-Frequency Devices

Resonant-tunneling diodes (RTDs) serve as key components in high-frequency devices operating at and millimeter-wave frequencies, leveraging their negative resistance (NDR) to enable compact oscillators, detectors, and mixers. These devices exploit quantum tunneling for rapid transport, allowing operation beyond 100 GHz with low consumption. In particular, RTD-based systems offer advantages over traditional Gunn diodes, including smaller footprints due to planar integration and reduced DC requirements, typically in the sub-milliwatt range, making them suitable for portable applications. Oscillators utilizing single RTDs or integrated RTD-heterojunction bipolar (HBT) structures achieve frequencies from 100 GHz to 500 GHz by matching the device's NDR to a resonant . In single-RTD designs, the bias stabilizes the voltage across the NDR peak, ensuring oscillation while suppressing parasitic modes. RTD-HBT enhances output power and stability through the 's amplification, enabling monolithic microwave integrated (MMICs) for wireless systems. A notable pre-2020 example is a 420 GHz InP-based RTD oscillator delivering 0.2 µW output power at , demonstrating feasibility for sub-terahertz sources. Furthermore, RTD oscillators have been integrated into phased arrays, where multiple elements synchronize via mutual injection locking to form directive beams for and communication. As detectors, RTDs enable direct (THz) detection through of the incident signal in the NDR , where the nonlinear current-voltage response modulates carrier flow. This provides high without external amplification, achieving noise-equivalent powers (NEPs) on the order of 10^{-12} W/√Hz at frequencies around 300-500 GHz. The low NEP stems from the device's and minimal thermal , outperforming Schottky diodes in sensitivity for THz sensing. In mixer applications, facilitate subharmonic mixing by exploiting even harmonics of the local oscillator () signal, where the symmetric NDR responds preferentially to twice the for efficient intermediate-frequency conversion. This approach reduces LO power needs and suppresses odd-order , yielding conversion gains up to several in the 100-300 GHz range. Optimal performance requires precise I-V curve symmetry and peak-to-valley current ratios exceeding 3:1.

Integrated Electronics and Detectors

Resonant tunneling diodes (RTDs) have been integrated into circuits through monostable-bistable transition element (MOBILE) designs, which employ two RTDs connected in series to form transfer gates that enable high-speed switching between monostable and bistable states. These gates facilitate compact implementations of basic functions, such as inverters and flip-flops, with demonstrated operation speeds exceeding 10 GHz; for instance, a D-flip-flop using achieved 12.5 Gb/s throughput with only 10 devices. The negative resistance (NDR) characteristic of RTDs allows these circuits to operate at low voltages while maintaining sharp transitions, supporting gate-level pipelining for enhanced throughput in digital systems. In applications, RTDs enable multistate by leveraging multiple NDR peaks in their current-voltage characteristics, allowing a single device to represent several levels without additional components. Demonstrations include three-state cells using two RTDs and resistors, as well as nine-state cells with a single multipeak RTD, providing compact alternatives to for increased density. Four-level has been realized in such configurations, supporting multi-valued operations that reduce the need for complex encoding in elements. For detection purposes, RTDs serve as sensitive terahertz (THz) detectors due to their nonlinear response and broadband operation, forming the basis of compact THz imagers. Examples include a 300 GHz RTD-based camera achieving 1.0 mm spatial resolution with an integrated Si lens, and a 345 GHz reflection imaging system offering 0.1 mm resolution for non-destructive testing. Integration with silicon-germanium heterojunction bipolar transistor (SiGe HBT) technology enhances receiver performance in hybrid systems; for instance, RTD sources paired with SiGe HBT direct detectors enable low-cost computed tomography at around 300 GHz, combining the high-frequency capabilities of RTDs with the low-noise amplification of SiGe components. Specific integrations in Si/SiGe heterostructures demonstrate RTD-CMOS compatibility for gates, such as AND and OR functions, where RTDs load CMOS inverters to exploit NDR for threshold . These gates, fabricated using for Si/SiGe RTD growth, operate with power consumption below 1 mW per gate—typically in the microwatt range, like 7.9 µW for a majority gate at 0.8 V—offering significant efficiency gains over pure CMOS equivalents. Multi-valued schemes using RTDs further reduce interconnect by encoding multiple bits per line, as seen in four-valued multiplexers that compact device counts compared to implementations. Recent demonstrations (as of 2024) include RTD-based systems for high-speed image processing, enabling efficient applications.

Recent Advances

Terahertz and Quantum Enhancements

Recent advances in resonant-tunneling diodes (RTDs) have significantly enhanced their capabilities in the (THz) regime, particularly through (InP)-based designs. InP RTD THz oscillators have demonstrated continuous-wave operation exceeding 1 THz, such as 1.98 THz in 2017, leveraging optimized double-barrier structures and integrated slot antennas to achieve high-frequency performance at . These devices benefit from improved and reduced , enabling fundamental oscillations up to 1.31 THz in compact configurations as reported in 2012. Additionally, arrayed InP RTD oscillators have produced output powers around 1 mW at approximately 750 GHz through resistor-coupled designs that enhance in-phase emission and suppress unwanted modes. For THz detection, low-noise InP triple-barrier RTDs (TB-RTDs) have emerged as direct detectors operating at zero , offering sensitivity in the WR2 band (330–500 GHz). These devices exhibit peak responsivities of 2123 V/W at 340 GHz, with values exceeding 1200 V/W across the band for 0.5 µm² active areas. (NEP) remains competitive, not exceeding 2 pW/√Hz for smaller devices and reaching as low as 1.15 pW/√Hz, making them suitable for room-temperature and applications comparable to state-of-the-art Schottky diodes. The triple-barrier configuration suppresses excess current while maintaining sharp negative differential resistance, contributing to their low-noise profile. Quantum enhancements in RTDs have been realized through graphene-based structures exploiting edge states, enabling negative differential resistance (NDR). Armchair graphene nanoribbon (AGNR) RTDs with patterned U-cut edges utilize zero-energy edge states for resonant , achieving NDR by confining to topologically protected channels. The U-cut patterning in AGNR superlattices improves the peak-to-valley ratio (PVR) beyond 400, particularly with longer barriers, while boosting peak currents through enhanced barrier engineering and dual-connection schemes that outperform single-edge configurations. These designs demonstrate superior performance over traditional RTDs, with reduced valley currents and sharper peaks. To support these advancements, non-saturating compact models have been developed for accurate simulation of RTD behavior in negative conductivity regions. Traditional models, such as the Schulman approach, suffer from current saturation due to transparency functions, limiting analysis of characteristics under dynamic conditions. The proposed model replaces this with an inverse square hyperbolic cosine transparency function, ensuring non-saturating via higher-order asymptotics: J_{RT} = \frac{q}{\pi \hbar} \int T(E) D(E) \, dE, where T(E) avoids artificial clipping in the negative differential conductivity (NDC) regime. Validated on AlAs/GaAs structures with barrier thicknesses of 2.83–4.53 nm, it provides precise predictions across wide voltage ranges, aiding optimization for THz and quantum applications. Updated designs for (GaAs)-based RTDs focus on achieving higher peak current densities (J_p) through refined heterostructure engineering and fabrication techniques. Recent reviews emphasize scaling barrier/well dimensions and doping profiles to exceed traditional J_p limits, with GaAs/AlAs double-barrier RTDs reaching over 200 kA/cm² while maintaining stable NDR at . These enhancements involve for precise layer control, enabling integration into high-speed circuits and addressing frequency limitations via reduced transit times.

Neuromorphic and Emerging Uses

Resonant tunneling diodes (RTDs) have emerged as key components in opto-electronic spiking neurons for neuromorphic hardware, enabling versatile models that mimic biological neural dynamics through excitable responses. These devices integrate RTDs with photodetectors and laser diodes to form nano-optoelectronic nodes capable of generating spikes with sub-nanosecond response times, supporting ultrafast processing in spike-based systems. For instance, RTD-based circuits demonstrate spike rate encoding to reconstruct amplitude-modulated signals and encoding for temporal , with firing rates exceeding 1 GHz. Such configurations exhibit refractoriness and excitability akin to biological neurons, facilitating hardware implementations of neuromorphic networks. To enhance in neuromorphic applications, models for RTDs address in the negative differential resistance (NDR) region, ensuring consistent nonlinear behavior essential for simulations. By replacing traditional transparency functions with inverse hyperbolic cosine forms, these compact dissipative models prevent non-physical , maintaining stable NDR over wide voltage ranges without compromising analytical tractability. This approach improves voltage- predictions and supports reliable operation of RTDs as nonlinear elements in excitable neuromorphic circuits. Emerging simulations of single-band quantum transport in next-generation RTDs, such as those based on nanoribbons, explore enhanced performance for quantum interfaces in neuromorphic systems. Using non-equilibrium methods, these models predict intrinsic cutoff frequencies up to 3 THz, with optimal designs achieved by tuning and barrier lengths. InP-based RTDs have achieved peak current densities up to 24.6 kA/cm² and peak-to-valley ratios of 8.6 at , as demonstrated in 2009, supporting denser integration in neuromorphic through hetero-integration onto SiGe-BiCMOS platforms for compact, high-speed circuits suitable for on-chip neural . Recent designs emphasize strain-compensated barriers to boost current handling while preserving NDR sharpness. Looking ahead, hybrid RTD-neuron architectures promise acceleration of AI tasks beyond conventional electronics by leveraging ultrafast spiking for energy-efficient neuromorphic computing, with ongoing optimizations in device parameters targeting scalable networks.

References

  1. [1]
    [PDF] Resonant Tunneling Diodes: Models And Properties
    The resonant tunneling diode (RTD) has been widely studied because of its importance in the field of nanoelectronic science and technology and its potential ...
  2. [2]
    Resonant tunneling in semiconductor double barriers - AIP Publishing
    Jun 15, 1974 · Esaki, R. Tsu; Resonant tunneling in semiconductor double barriers. Appl. Phys. Lett. 15 June 1974; 24 (12): 593–595. https://doi.org ...
  3. [3]
  4. [4]
    Room‐temperature observation of resonant tunneling through an ...
    We report the first room‐temperature observation of resonant tunneling through a double‐barrier diode with a 31.5‐nm‐wide AlGaAs/GaAs quasiparabolic quantum
  5. [5]
  6. [6]
    Si resonant interband tunnel diodes grown by low-temperature ...
    Aug 30, 1999 · These diodes are compatible with transistor integration. Topics. Transistors, Semiconductor device fabrication, Epitaxy, Negative resistance, ...Missing: silicon | Show results with:silicon
  7. [7]
  8. [8]
  9. [9]
  10. [10]
  11. [11]
    Room-temperature Resonant-tunneling-diode Terahertz Oscillator ...
    This article presents the recent achievement of fundamental oscillations in resonant tunneling diodes (RTDs) above 1 THz at room temperature.
  12. [12]
    [PDF] I. GaAs Material Properties
    Electron mobility (undoped). 8500 cm2/V-s. Hole mobility (undoped). 400 cm2/V-s ... Although the peak mobility of GaAs in the linear region can be as much as six.
  13. [13]
    Resonant tunneling diode (RTD) — nextnano Documentation
    Outside the barriers, the structure is doped with a concentration of 1017 cm-3. The barrier height was taken to be 0.156 meV. The temperature is set to 150 ...
  14. [14]
    Fundamentals and recent advances of terahertz resonant tunneling ...
    Jul 19, 2024 · AlAs/InGaAs RTD structures are primarily used for the formation of THz oscillators. Additionally, GaN-based RTDs have also been studied.
  15. [15]
    Resonant interband tunneling current in InAs/AlSb/GaSb/AlSb/InAs ...
    Jul 28, 1997 · We have investigated the resonant interband tunneling current in InAs/AlSb/GaSb/AlSb/InAs double-barrier resonant interband tunneling diodes ...
  16. [16]
    In-plane dispersion relations of InAs/AlSb/GaSb/AlSb/InAs interband ...
    Nov 15, 1995 · We apply this model to study InAs/AlSb/GaSb/AlSb/InAs interband resonant-tunneling diodes. We show that the boundary-condition sets deduced from ...
  17. [17]
    (PDF) The design of GaAs/AlAs resonant tunneling diodes with peak ...
    In AlGaAs/GaAs RTDs are observed peak-to-valley current ratios (PVCRs) ranging from 3:1 to over 10:1, with peak current densities of 10 2 -10 4 A/cm 2 depending ...
  18. [18]
    Si∕SiGe n-type resonant tunneling diodes fabricated using in situ ...
    May 14, 2007 · In this letter, an in situ hydrogen cleaning process is demonstrated on a Si ∕ Si Ge RTD structure optimized for low voltage operation in a ...
  19. [19]
    [PDF] Investigation of the emitter structure in SiGe/Si resonant tunneling ...
    Due to the relatively large valence band offset in SiGe relative to Si, it is possible to realize. RTD in the group IV material system. This was first demon-.
  20. [20]
    Nonstationary Transport in Si/Si0:7Ge0:3 n‐MODFET
    The conduction band offset at Si/Si0:7Ge0:3 interface is assumed to be ΔEC = 0:18 eV. [10], and the gate Schottky barrier height FB to be 0.6 eV. In the ...
  21. [21]
    Physics and applications of Si/SiGe/Si resonant interband tunneling diodes
    ### Summary of Si/SiGe/Si Resonant Interband Tunneling Diodes
  22. [22]
    [PDF] si/sige heterostructures: materials, physics, quantum
    Berger, and P. E. Thompson, “RF. Performance and Modeling of Si/SiGe Resonant Interband Tunneling Diodes,” IEEE. Trans. Elec. Dev. vol. 52, pp. 2129-2135 ...
  23. [23]
  24. [24]
    [PDF] Chapter 4 AlAs/GaAs Double Barrier Resonant Tunneling Diodes
    The negative differential resistance (NDR) in the I-V characteristics of these devices has been used in many applications involving microwave/millemeter wave ...<|separator|>
  25. [25]
    [PDF] Double-Barrier Resonant Tunnelling Diodes - - Nottingham ePrints
    By the late 1980's, operation of OBRT diodes had been observed up to 420GHz ... from the deposition temperature (typically 300°C) to room temperature.
  26. [26]
    [PDF] Barrier Roughness Effects in Resonant Interband Tunnel Diodes
    Apr 17, 2001 · Interface roughness has often been discussed as a factor in determining the current through resonant tunneling diodes that are of interest for ...
  27. [27]
    [1211.0821] Transient Schrödinger-Poisson Simulations of a High ...
    Nov 5, 2012 · Transient simulations of a resonant tunneling diode oscillator are presented. The semiconductor model for the diode consists of a set of time- ...
  28. [28]
    Capacitance-voltage investigation of low residual carrier density in ...
    After growth, the material is mesa etched, passivated using a dielectric passivation and metallized with an ohmic contact layer and a gold bonding metal ...<|separator|>
  29. [29]
    Improving the peak current density of resonant tunneling diode ...
    ... AuGe/Ni/Au) for RTD contact are introduced, and a ... The resonant tunneling diode (RTD) is one ... ohmic contacts. J Appl Phys, 2000, 87(5):2437 doi ...
  30. [30]
    [PDF] Millimeter-Wave Quasi-Optical Planar Circuits and Power Combining.
    Jan 15, 1996 · Using the combination of slotline and coplanar waveguides ... The in-phase CPW excitation is via an air bridge and the slotline tee junction is.
  31. [31]
    Growth of III/V resonant tunnelling diode on Si substrate with LP ...
    We present in this work the first monolithic integration of a III/V resonant tunnelling diode (RTD) on an exactly (0 0 1)-oriented Si substrate.
  32. [32]
    InP-based monolithically integrated RTD/HBT mobile for logic circuits
    Aug 6, 2025 · A pseudo dynamic logic family is developed on InP-substrates based on the MOBILE concept. The conventional HFET as input terminal is ...
  33. [33]
    [PDF] Si/SiGe-Based Gate-Normal Tunneling Field-Effect Transistors
    Fabrication steps of vertical TFETs with a nanopillar contact. (a) Initial stack, grown on. 300 mm Si(001) wafers, where the lowest two layers will later ...
  34. [34]
    [PDF] Resonant tunneling diodes made up of stacked self-assembled Ge ...
    The concept exploits the effect of vertical self-alignment and creates vertical Ge channels with larger valence band offsets than in conven- tional Si/SiGe ...Missing: 300mm fabs planar
  35. [35]
    Equivalent circuit of RTD oscillator with parasitic elements.
    This was because of a parasitic capacitance of the metal post connecting the cavity and the RTD. ... challenges and possible strategies for future progress.
  36. [36]
    Frequency Limitations of Resonant-Tunnelling Diodes in Sub-THz ...
    Mar 13, 2019 · A room-temperature RTD oscillator at the fundamental frequency of 56 GHz has been demonstrated in 1987 [34], at 200 GHz in 1988 [31], at 420 ...
  37. [37]
    (PDF) Ring oscillator using an RTD-HBT heterostructure
    We report the first ring oscillator design using a Resonant Tunneling Diode-Heterojunction Bipo-lar Transistor (RTD-HBT) heterostructure for high-speed and ...
  38. [38]
    [PDF] Efficient Realization of RTD-CMOS Logic Gates - idUS
    ABSTRACT. The incorporation of Resonant Tunnel Diodes (RTDs) into III/V transistor technologies has shown an improved circuit performance: higher.
  39. [39]
    Resonant Tunneling Diodes - an overview | ScienceDirect Topics
    Resonant tunneling diodes are semiconductor devices characterized by a structure where two n-doped layers are separated by a double-barrier insulating layer ...Missing: original | Show results with:original
  40. [40]
    Toward High-Peak-to-Valley-Ratio Graphene Resonant Tunneling ...
    Sep 5, 2023 · We discovered that the edge doping can play a dominant role in determining the resonant tunneling, and a large area-to-perimeter ratio is necessary to obtain a ...<|control11|><|separator|>
  41. [41]
    Peak-to-valley ratio of small resonant-tunneling diodes with various ...
    Electron accumulation increases more- over the conducting diameter of the devices and decreases the peak-to-valley ratio. Submicron devices exhibit a drastic.
  42. [42]
    Extremely High Peak Current Densities of over 1×10 6 A/cm 2 in InP ...
    InP-based InGaAs/AlAs resonant tunneling diodes (RTDs) with extremely high peak current density ( jP) were grown by metal–organic vapor-phase epitaxy.Missing: cm² | Show results with:cm²
  43. [43]
    Boron Delta-Doping Dependence on Si/SiGe Resonant Interband ...
    Jan 23, 2012 · The boron doping in the δ-doped region was varied, and its effect on peak current density Jp and peak-to-valley current ratio (PVCR) was studied ...
  44. [44]
    High-uniformity InP-based resonant tunneling diode wafers with ...
    RTD wafers. The RTDs exhibited excellent current–voltage characteristics with high JP of over 6×105 A/cm2 and the peak-to-valley ratio (PVR) of over 3. A small ...
  45. [45]
    InGaAs/AlAs Resonant Tunneling Diodes for THz Applications
    Feb 2, 2018 · This paper presents an experimental study of InGaAs/AlAs resonant tunneling diodes designed to improve the diode characteristics using five different device ...
  46. [46]
    Experimental study of the frequency limits of a resonant tunneling ...
    An equivalent circuit obtained from microwave impedance and power data is proposed for a resonant tunneling diode. ... gests the negative differential resistance ...
  47. [47]
    Simulation and Comparative Study of Resonant Tunneling Diode
    Aug 10, 2025 · The standard temperature is taken to be 300 K and the contacts ... (PVR) observes a strict degradation because of a high. valley current ...
  48. [48]
    Gate length and temperature dependence of negative differential ...
    degradation was the result of charge trapping at cryogenic temperatures.20 ... activation energy and effective longitudinal optical phonon energy in triple well ...
  49. [49]
    Variations of resonant tunneling pro et-ties - Sil -,Ge,/Si double ...
    The resonant tunneling devices (RTDsj also have better temperature stability and PVR decreases slower. The RTD with larger Ls has higher peak current, Beside w ...Missing: comparison | Show results with:comparison
  50. [50]
  51. [51]
  52. [52]
    [PDF] ADVANCED TUNNELING DIODES FOR HIGH-FREQUENCY ...
    metrics can assess the RTD's RF characteristics: series resistance RS, RTD capacitance. CRTD, and oscillation frequency. Figure 2.7. The IV characteristics ...
  53. [53]
    Terahertz Emitter Using Resonant-Tunneling Diode and Applications
    Resonant-tunneling diodes (RTDs) are also promising candidates for room-temperature THz sources [24,25,26,27,28,29,30,31]. Currently, oscillation up to 1.98 ...
  54. [54]
    Resonant Tunnelling Optoelectronic Circuits - IntechOpen
    When both sides are terminated by highly doped semiconductor layer (the emitter and the collector contacts) for electrical connection the structure is called ...
  55. [55]
    An Overview of Terahertz Imaging with Resonant Tunneling Diodes
    Apr 10, 2022 · RTD has been demonstrated with very wide broadband intrinsic frequency response from 120 GHz to 3.9 THz [35] and the estimated intrinsic cut-off ...
  56. [56]
  57. [57]
    [PDF] Recent advances in resonant-tunneling-diode oscillators
    Oscillators made from such diodes have operated at room temperature up to 420 GHz. 1. It has been shown that the maximum oscillation frequency of GaAs/AlAs RTDs ...
  58. [58]
  59. [59]
  60. [60]
  61. [61]
    Sensitive Continuous-Wave Terahertz Detection by Resonant ...
    We have characterized the sensitivity and noise equivalent power (NEP) of resonant-tunneling-diode (RTD) oscillators as terahertz (THz)-wave detectors.
  62. [62]
    The limits of resonant tunneling diode subharmonic mixer performance
    Aug 1, 1989 · A theoretical analysis of the performance of subharmonically pumped mixers using resonant tunneling diodes is given based on the I‐V curve ...
  63. [63]
    [PDF] Digital Circuit Applications Of Resonant Tunneling Devices
    Many semiconductor quantum devices utilize a novel tunneling transport mechanism that allows picosecond device switching speeds.
  64. [64]
    Efficient realization of RTD-CMOS logic gates - ACM Digital Library
    May 2, 2011 · The incorporation of Resonant Tunnel Diodes (RTDs) into III/V transistor technologies has shown an improved circuit performance: higher ...
  65. [65]
    Toward Mobile Integrated Electronic Systems at THz Frequencies
    Based on an RTD source and a SiGe-HBT direct detector, low-cost and compact computed tomography is presented for volumetric continuous-wave imaging at around ...
  66. [66]
    High-power in-phase and anti-phase mode emission from linear ...
    Aug 7, 2024 · Oscillators based on resonant tunneling diodes (RTDs) are able to reach the highest oscillation frequency among all electronic THz emitters.
  67. [67]
    [PDF] Terahertz Resistor-coupled Arrayed Resonant-tunneling-diode Os
    Sep 5, 2024 · A high power density of up to 62 mW/mm2 is another advantage of the proposed RTD oscillator. With the improved characteristics, we believe the ...
  68. [68]
    (PDF) Low-Noise Resonant Tunneling Diode Terahertz Detector
    Aug 7, 2025 · The measurements reveal that the responsivity scales with device area, as predicted by the theory, with a peak responsivity of 2123V/W at 340GHz ...
  69. [69]
  70. [70]
  71. [71]
    Resonant tunneling diode based on edge states in patterned ...
    Aug 16, 2025 · With peak current to valley current ratio (PVR) values obtained above 400, we demonstrate that the edge states based resonant transmission is a ...
  72. [72]
    Resonant Tunneling Nanostructures: Eliminating Current Saturation ...
    A non-saturating compact model was proposed to solve the problem of current transfer analysis in the region of negative differential conductivity.
  73. [73]
  74. [74]
    GaAs-based resonant tunneling diode: Device aspects from design ...
    This article is the first review of GaAs RTD technology which covers different epitaxial-structure design, fabrication techniques, and characterizations for ...Missing: original | Show results with:original
  75. [75]
    GaAs-based resonant tunneling diode: Device aspects from design ...
    Abstract. This review article discusses the development of gallium arsenide (GaAs)-based resonant tunneling diodes (RTD) since the 1970s.
  76. [76]
    Spiking Rate and Latency Encoding with Resonant Tunnelling Diode Neuron Circuits and Design Influences
    ### Summary of Key Findings on Resonant Tunnelling Diode Use in Neuromorphic Spiking Neurons
  77. [77]
    Excitability and memory in a time-delayed optoelectronic neuron
    Jan 31, 2024 · We study the dynamics of an optoelectronic circuit composed of an excitable nanoscale resonant-tunneling diode (RTD) driving a nanolaser diode ( ...