Fact-checked by Grok 2 weeks ago

Tunnel diode

A tunnel diode, also known as an Esaki diode, is a two-terminal formed by a heavily doped p-n junction that exhibits negative differential resistance (NDR) in its forward current-voltage (I-V) characteristics due to quantum mechanical electron tunneling across a thin potential barrier. This tunneling occurs in degenerate semiconductors where the conduction band of the n-side overlaps with the valence band of the p-side, allowing charge carriers to pass directly through the forbidden energy gap without thermal excitation. The device was invented in 1957 by Japanese physicist while working at Tokyo Tsushin Kogyo (now Sony Corporation), marking the first experimental demonstration of electron tunneling in solids. The I-V curve of a tunnel diode under forward bias initially rises sharply to a peak current (I_p), then decreases to a valley current (I_v), creating the NDR region where resistance is negative, before rising again like a conventional diode at higher voltages. In reverse bias, it behaves as a highly conductive rectifier due to continued tunneling. Common materials include germanium for early devices, as well as gallium arsenide (GaAs) and indium phosphide (InP) for higher performance, with junction widths on the order of nanometers to enable the tunneling effect. Esaki's discovery provided crucial evidence for quantum tunneling in semiconductors and earned him a share of the 1973 Nobel Prize in Physics, alongside Ivar Giaever for tunneling in superconductors and Brian David Josephson for the Josephson effect. Tunnel diodes offer exceptional speed, with switching times in the tens of picoseconds and operation up to millimeter-wave frequencies, along with low power consumption and noise, making them suitable for high-frequency applications. Key uses include oscillators, low-level detectors, and amplifiers, particularly in and communication systems, as well as in early computer logic circuits for fast switching. Despite their promise as alternatives to transistors in the late and , their two-terminal design limited integration into complex logic, leading to niche rather than widespread adoption; however, they remain relevant in modern quantum and high-speed electronics research.

History

Invention

In the post-World War II era, semiconductor research in Japan advanced rapidly, driven by the need for innovative electronics in consumer devices. At Tokyo Tsushin Kogyo (later Sony Corporation), physicist Leo Esaki pursued experiments on heavily doped p-n junctions to narrow the depletion region sufficiently for quantum mechanical effects to emerge. In July 1957, motivated by theoretical predictions of electron tunneling through thin barriers, Esaki fabricated germanium junctions with impurity concentrations exceeding 10^{19} cm^{-3}, aiming to observe interband tunneling currents. During current-voltage measurements at around 200 , Esaki first observed a novel region of negative differential resistance in these heavily doped p-n junctions, where increasing voltage led to decreasing current—a hallmark of tunneling-dominated transport. This phenomenon, later termed the Esaki diode or , represented the initial experimental of solid-state quantum tunneling in semiconductors, enabling potential applications in high-speed switching and . The underlying quantum tunneling mechanism allowed electrons to penetrate the potential barrier without classical thermal activation, distinguishing it from conventional behavior. Esaki reported his findings initially at the Physical Society of Japan fall meeting in 1957 and presented them internationally at a conference in Brussels in June 1958. The seminal results were published in 1958 as "New Phenomenon in Narrow Germanium p-n Junctions" in Physical Review, marking the formal introduction of the device to the global scientific community. For his pioneering work on tunneling phenomena in semiconductors, Esaki shared the 1973 Nobel Prize in Physics with Ivar Giaever, who extended similar studies to superconductors, and Brian Josephson, recognized for theoretical predictions in superconductivity. The prize citation highlighted their experimental discoveries regarding tunneling in solids, underscoring the tunnel diode's role as the first quantum electronic device.

Development and Recognition

Following the invention, at Tokyo Tsushin Kogyo (now ), the company played a pivotal role in prototyping the device through experimental fabrication and testing of heavily doped junctions. 's efforts enabled the first demonstrations of the diode's property, which facilitated its transition from theoretical concept to functional prototype by 1958. In the , the tunnel diode saw rapid industry adoption for high-speed electronics research, with companies like and leading development efforts. explored its use in computer circuits and systems, achieving switching speeds suitable for advanced calculations in applications requiring up to 70,000 operations per second. produced commercial tunnel diodes and published comprehensive manuals on their integration into and switching circuits, emphasizing their low-power, high-frequency capabilities for emerging electronic systems. Esaki's contributions gained international recognition in 1973 when he shared the with and Brian David Josephson for their experimental discoveries on tunneling phenomena in semiconductors and superconductors, underscoring the tunnel diode's demonstration of quantum tunneling as a foundational advance in . During the era, the pursued early military applications of tunnel diodes, particularly in radar and communications technologies, with research focused on radar systems to enhance speeds for defense needs.

Physical Principles

Quantum Tunneling Effect

The quantum tunneling effect arises from the wave-like nature of electrons, as described by , which allows them to penetrate barriers that would be impassable in . According to the time-independent , the probability density of finding an in a forbidden region decays exponentially but remains nonzero if the barrier is sufficiently thin, typically on the order of 10 or less in semiconductors. This penetration enables electrons to "tunnel" from one side of the barrier to the other without acquiring the energy to surmount it classically. The probability P of tunneling through a potential barrier is approximated using the Wentzel-Kramers-Brillouin (WKB) method for slowly varying potentials, particularly derived for rectangular barriers as: P \approx \exp\left(-2 \int_{x_1}^{x_2} \sqrt{\frac{2m}{\hbar^2} (V(x) - E)} \, dx \right), where m is the electron mass, \hbar is the reduced Planck's constant, V(x) is the potential barrier height, E is the electron energy, and the integral is over the barrier width from turning points x_1 to x_2. This exponential dependence highlights that tunneling probability decreases rapidly with increasing barrier width or height, making it significant only for nanoscale barriers. In the context of semiconductors like those used in tunnel diodes, heavy doping at levels around $10^{19} cm^{-3} creates degenerate conditions where the lies within the conduction band on the n-side and the valence band on the p-side, leading to overlapping filled and empty states across the junction. This degeneracy narrows the to a few nanometers, facilitating substantial tunneling currents. Unlike conventional p-n junctions, where transport relies on classical of minority carriers or over the barrier, the heavily doped tunnel junction suppresses these mechanisms due to the absence of minority carriers and the thin barrier, with band-to-band tunneling dominating the current flow.

Energy Band Structure

In a conventional undoped p-n junction, the conduction band and valence band are separated by a forbidden bandgap, preventing direct carrier transport across the junction without thermal excitation or high bias. In contrast, the tunnel diode relies on degenerate doping, where both the p-type and n-type regions are heavily doped (typically 10¹⁹ to 10²⁰ cm⁻³), filling the conduction band on the n-side and the valence band on the p-side with carriers up to energies within the bands themselves. This degenerate condition positions the deep within the conduction band of the n-type material and within the valence band of the p-type material, creating a high of occupied states on the n-side and unoccupied states (holes) on the p-side. The heavy doping induces significant across the junction, resulting in a narrow , approximately 10 nm wide, with a strong built-in on the order of 10⁶ V/cm. At zero , this bending causes an overlap in the such that occupied states in the conduction of the n-side align energetically with unoccupied states in the of the p-side, enabling quantum tunneling of electrons directly across the forbidden gap without requiring equal to the bandgap. In a representative , the conduction edge on the n-side lies at a similar to the valence edge on the p-side within the , forming a thin potential barrier that electrons can penetrate via wavefunction overlap. This band structure is material-dependent, with tunneling efficiency influenced by the semiconductor's bandgap energy, as narrower bandgaps allow for greater state alignment and higher tunneling probabilities. Common materials include (bandgap 0.66 eV at 300 K), (1.42 eV), and (1.11 eV), where germanium's smaller bandgap facilitates more efficient tunneling in early devices. The degenerate nature ensures that the Fermi level positioning supports substantial carrier densities available for tunneling, distinguishing the tunnel diode from standard junctions.

Device Structure

Materials and Fabrication

Tunnel diodes are primarily fabricated using heavily doped semiconductors to enable the quantum tunneling effect, with early devices relying on as the base material due to its suitable bandgap and ease of achieving degenerate doping levels. tunnel diodes typically incorporate acceptor concentrations (N_A) and donor concentrations (N_D) exceeding 10^{19} cm^{-3}, often using impurities such as for p-type and or for n-type doping to create the narrow essential for tunneling. For higher-frequency applications, has become a preferred material, offering improved and operation up to millimeter-wave frequencies, while is utilized in optoelectronic and high-speed devices for its lattice compatibility with compound semiconductors. variants, particularly silicon-germanium (SiGe) heterostructures, provide compatibility with processes, enabling integration in modern integrated circuits despite challenges in achieving uniform heavy doping. Recent advances as of 2023 include the fabrication of nanowire tunnel diodes, which offer a platform for investigating band-to-band tunneling in one-dimensional structures, addressing challenges in axial doping and junction uniformity. Additionally, as of 2024, two-dimensional materials such as MoS_2 have been explored in metal-insulator-metal (MIM) tunnel diode configurations for ultrafast and applications. The fabrication process begins with epitaxial growth techniques to establish precise doping profiles across the p-n junction, where (MBE) or metal-organic (MOCVD) allows atomic-level control over layer thickness and impurity incorporation, typically resulting in depletion widths of 5-10 nm. Following growth, mesa etching is employed to isolate the junction, defining a small active area—often 10-100 \mu m^2—to minimize and enhance high-frequency performance. Ohmic contacts are then formed using metal and annealing, ensuring low-resistance access to the heavily doped regions without introducing additional defects. Achieving uniform heavy doping remains a key challenge, as concentrations above 10^{19} cm^{-3} can induce defects and , particularly in heterostructures, potentially degrading tunneling efficiency and increasing series resistance. In GaAs and InP devices, autodoping and compensation effects during growth must be mitigated to maintain sharp band overlaps, while silicon-based variants face solubility limits for dopants like . Fabrication techniques have evolved from 1950s alloy methods, where a dopant-containing wire or pellet was fused to a substrate to form the , to planar processes by the 1970s that incorporated oxide masking and for reproducible, integrated structures. This shift enabled scalability and reliability, with modern epitaxial approaches further refining control over heavy doping to alter the energy band structure for optimal tunneling.

Junction Design

The Esaki junction in a tunnel diode consists of an abrupt p+-n+ transition, achieved through heavy degenerate doping on both sides of the p-n interface to create a degenerate semiconductor regime. This abrupt profile ensures a minimal transition region between the p-type and n-type materials, distinguishing it from graded junctions in conventional diodes. The depletion width W at the junction is extremely narrow, typically 5-10 nm, which is critical for enabling the tunneling mechanism. This width is determined by the formula W = \sqrt{\frac{2\epsilon (V_{bi} - V)}{q N_{eff}}}, where \epsilon is the permittivity of the semiconductor, V_{bi} is the built-in potential, V is the applied bias, q is the elementary charge, and N_{eff} = \frac{N_A N_D}{N_A + N_D} is the effective doping concentration; for symmetric doping with N_A \approx N_D = N, N_{eff} \approx N/2, and the high value of N (often exceeding $10^{19} cm^{-3}) significantly minimizes W. Electrode configurations are designed to minimize parasitic effects, with point-contact structures commonly used to define very small areas (often <1 \mum^2 ) and thereby reduce shunt . Alternatively, mesa structures involve to isolate the , providing precise control over the active area while maintaining low . Ohmic contacts are typically formed using Au-Ge alloys for n-type regions or Ti/Pt/Au multilayer stacks for improved and low resistance. The peak-to-valley (PVCR) depends on the of the doping profile across the and the overall junction area, where balanced p+ and n+ doping concentrations enhance the ratio by optimizing tunneling efficiency, and smaller areas help maintain high ratios by limiting excess current contributions. Degenerate doping levels, as selected from the materials, are essential to support this symmetry and narrow width. Packaging for tunnel diodes emphasizes hermetic seals, often using or enclosures, to protect the sensitive from environmental factors and ensure reliable performance. Lead inductance is carefully managed through short, low-inductance connections or mesh screens to avoid resonances that could degrade high-frequency .

Operation

Forward Bias Behavior

Under forward bias, the tunnel diode exhibits a tunneling-dominated flow that distinguishes it from conventional p-n diodes. At low forward voltages (typically V < V_p ≈ 0.05–0.1 V), the applied bias shifts the energy s such that the valence band of the p-region aligns with the conduction band of the n-region, enabling electrons to tunnel directly from occupied states in the valence band to empty states in the conduction band. This interband tunneling increases exponentially with voltage due to the progressively better alignment of available quantum states across the narrow . The peak current I_p is reached when the bias voltage V_p aligns the maximum number of occupied and unoccupied states for tunneling, maximizing the current flow. The tunneling current density J can be expressed as J \approx \frac{q}{2\pi \hbar} \int T(E) [f_p(E) - f_n(E)] \, dE, where q is the , ħ is the , T(E) is the through the barrier at energy E, and f_p(E) and f_n(E) are the Fermi-Dirac functions for the p- and n-regions, respectively. This accounts for the energy-dependent probability of tunneling and the availability of charge carriers. As the forward bias exceeds V_p, the energy bands misalign, reducing the overlap between tunneling states and causing the tunneling current to decrease sharply. The device then transitions to normal diffusion current mechanisms, where minority carriers are injected across the junction, leading to an overall current rise at higher voltages but with much lower efficiency than the initial tunneling regime. This profile results in the emergence of negative differential resistance in the vicinity of the peak. The forward bias behavior is highly temperature-dependent, with tunneling effects being strongest at low temperatures (optimally around 77 K) where thermal energy minimally smears the Fermi distributions and preserves sharp band alignments. At , thermal broadening reduces the tunneling probability, weakening the peak and the region, though the effect remains observable in properly designed devices.

Reverse Bias Behavior

Under reverse bias, the tunnel conducts a significant primarily through interband quantum tunneling, where electrons tunnel from the filled valence band states in the p-region to empty conduction band states in the n-region across the narrow , akin to the but initiated at very low voltages due to heavy doping. This mechanism results in a high reverse that is weakly dependent on temperature, distinguishing it from diffusion-dominated currents in conventional diodes. As the reverse bias voltage increases modestly (e.g., from 0 to several volts), the tunneling current saturates at a relatively constant level because the increasing misaligns the energy bands, depleting the number of available quantum states for tunneling and reducing the tunneling probability. The valley current in this regime remains low, typically on the order of a few milliamperes, limited by the scarcity of carriers beyond initial tunneling, without exhibiting a distinct peak-valley transition as seen in forward for context on the overall I-V curve shape. At higher reverse biases exceeding the V_{br}, usually 5-10 V depending on material (e.g., ~5 V for GaAs variants), the current rises abruptly due to the onset of avalanche multiplication, where the intensified accelerates carriers to generate additional electron-hole pairs through . Reverse operation offers limited utility compared to forward mode and poses challenges, as the valley current increases with temperature—exacerbating power dissipation—and can lead to without external series resistance to limit current and stabilize the device.

Electrical Characteristics

Current-Voltage Relationship

The current-voltage (I-V) relationship of the tunnel diode exhibits a unique N-shaped curve in forward bias, characterized by a rapid increase in current due to quantum tunneling, reaching a peak current I_p at a low peak voltage V_p, followed by a decrease to a valley current I_v at valley voltage V_v, and then a subsequent rise resembling a conventional p-n diode. For germanium-based tunnel diodes, typical parameters include V_p \approx 0.055 V, V_v \approx 0.35 V, and an I_p / I_v ratio of 3 to 10, with higher ratios indicating better performance for applications relying on . These values arise from the heavy doping that enables band overlap, allowing tunneling at low biases, though exact parameters vary with material and fabrication; for example, diodes show V_p \approx 0.15 V and I_p / I_v \approx 15. Due to the degenerate doping, the measured peak voltage V_p is significantly lower than the bandgap voltage, typically 40-70 mV for . The I-V curve is nearly symmetric near zero due to tunneling in both directions, with the forward displaying the pronounced peak-valley , while in reverse the increases with voltage due to continued tunneling, resulting in high and Zener at low voltages typically near 0 V. The small-signal equivalent circuit for the in the negative region consists of a series R_s (typically 5-10 \Omega), a C_j (10-50 pF, voltage-dependent), a small series L_s (often negligible below frequencies), and a negative conductance -G_n (where G_n = 1 / |R_n| and R_n \approx -50 to -100 \Omega). This model captures the device's behavior for circuit analysis, with C_j dominating the parasitic effects. I-V characteristics are measured using curve tracers (e.g., 571) or oscilloscopes applying a slow voltage ramp to trace the curve without inducing oscillations, often with stabilizing resistors to suppress effects. The operational frequency is limited by the , where R is the magnitude of the negative resistance and C includes C_j plus external , typically yielding cutoff frequencies up to several GHz for small devices.

Negative Resistance Region

In the negative resistance region of a tunnel diode, the device exhibits negative differential resistance (NDR), defined as a condition where the differential conductance \frac{dI}{dV} < 0 over a voltage range between the peak voltage V_p and the valley voltage V_v. This phenomenon arises because the tunneling current, which dominates at lower forward biases, decreases as the applied voltage increases, while the competing diffusion current component rises more gradually, resulting in an overall net decrease in total current with increasing voltage. The magnitude of the negative resistance, given by |R_n| = \frac{1}{|\frac{dI}{dV}|}, typically falls in the range of approximately -50 to -200 Ω for conventional or tunnel diodes. This value determines the device's ability to sustain oscillations when coupled with an external resonant circuit, as the inherent instability of the NDR region requires careful biasing and stabilization to prevent spontaneous switching between stable states. From a small-signal perspective, the tunnel diode in this region can be modeled as a negative conductance G_n = -\frac{1}{R_n} in with the junction C, augmented by series R_s and lead L_s. At frequencies above the series resonant f_r = \frac{1}{2\pi \sqrt{L_s C}}, the of this combination exhibits inductive , enabling applications in high-frequency and without additional inductors. The quantum nature of the tunneling process contributes to exceptionally low performance in this region, with tunnel diode amplifiers achieving noise figures of 1-3 , significantly better than many contemporary solid-state devices of the era. This low arises primarily from the shot noise associated with coherent tunneling, rather than or excess noise mechanisms, providing a key advantage for sensitive applications.

Applications

Microwave and Oscillator Uses

Tunnel diodes have been employed in oscillator s since their invention, leveraging the region to enable simple, low-power signal generation without complex feedback networks. These oscillators operate on principles similar to devices but with simpler construction, often using series or shunt-tuned resonant circuits where the diode's and an external form the tank circuit. The self-oscillation frequency is approximately given by f \approx \frac{1}{2\pi \sqrt{LC}}, where L is the and C is the junction , allowing tunability across bands. In germanium-based tunnel diodes, oscillation frequencies typically reach up to 12 GHz, while variants extend capabilities to 100 GHz, as demonstrated in configurations for high-frequency applications. A key advantage of tunnel diode oscillators is their low performance and minimal power requirements, often delivering output powers in the microwatt range without needing a separate , which simplifies integration into compact systems. For instance, C-band oscillators (around 5-6 GHz) have been tuned over 3.5-6.1 GHz with power levels of about 2 µW at 10.8 GHz, making them suitable for low-level signal sources in early and communication equipment during the . In satellite communications, X-band (8-12 GHz) tunnel diode oscillators provided reliable, low- sources with gains up to 15 dB and noise figures as low as 5.5 dB over bandwidths of 480 MHz, enhancing in and space systems. Beyond oscillation, tunnel diodes serve as microwave detectors, particularly in square-law operation for detecting low-level signals where the output current is proportional to the square of the input voltage. This mode was widely used in 1960s radar receivers and superheterodyne systems, achieving noise figures of 12-15 in S-band (2-3 GHz) converters and L-band mixers tunable over 500 MHz bandwidths. The backward diode variant, a specialized tunnel diode with asymmetric doping, excels in zero-bias detection, eliminating the need for forward bias and reducing 1/f , which is beneficial for and low intermediate-frequency applications at frequencies up to 9 GHz. These diodes offer tangential sensitivity comparable to commercial detectors, with improvements of 4 over types, and cutoff frequencies around 2 GHz, supporting reliable and imaging.

Switching and Logic Circuits

Tunnel diodes leverage their negative differential resistance (NDR) region to achieve exceptionally fast switching speeds, with transition times on the order of tens of picoseconds, far surpassing the switching times of early transistors prevalent in the . This rapid response stems from the quantum tunneling mechanism, enabling the device to switch between high- and low-conductance states almost instantaneously without the carrier storage delays inherent in transistor-based switches. The bistable points on the current-voltage (I-V) curve further support reliable operation in switching applications by providing distinct stable voltage levels. In logic applications, tunnel diodes form the basis of early high-speed circuits, including monostable multivibrators capable of generating pulses with widths as short as 5 s and repetition rates up to 70 MHz. These multivibrators operate by triggering the diode from its stable low-voltage state to a temporary high-voltage state before recovery, offering faster recovery times and lower triggering currents compared to equivalents. AND and OR gates are implemented using principles, where multiple input signals sum analogically to control the diode's firing , enabling binary decision-making with minimal components. Such gates exhibit delays around 1 and support repetition rates exceeding 200 MHz in integrated configurations. Extensions to resonant tunneling diodes (RTDs), which build on the original tunnel diode structure with double-barrier quantum wells, enable multi-valued logic by exploiting multiple NDR peaks in the I-V characteristic for or states. Tunnel diode memory (TDM) cells utilize the self-latching bistable behavior of series-connected diodes to store data, providing flip-flop stability through the NDR region's that maintains state without continuous power. These cells achieve cycle times as low as 25 nanoseconds, with the flip-flop configuration supporting set-reset operations at repetition rates up to 100 MHz. Historically, tunnel diodes were integrated into circuits for , including IBM's explorations in and memory applications, where their low power dissipation—on the order of microwatts per gate—promised significant energy savings over logic. These circuits dissipated hundreds of times less power than contemporary gates, typically operating at levels below 100 μW while maintaining high-speed performance in (SLT) hybrids.

Modern Applications

In recent years as of 2025, tunnel diodes have seen renewed applications in low-power and high-frequency electronics. For example, fully tunnel-diode-based backscattering RFID tags operate at 5.8 GHz with 20 μW power and 48 dB gain (2022). They are also used in mm-wave rectifiers for perpetual IoT devices and non-contact sensing systems like TunnelSense for motion detection (2024).

Comparisons and Limitations

Versus Conventional Diodes

Tunnel diodes differ fundamentally from conventional p-n diodes in their operational mechanism and electrical behavior. While conventional p-n diodes rely on of minority carriers across a moderately doped , leading to exponential current increase above a forward threshold of approximately 0.7 V for , tunnel diodes operate via quantum mechanical in heavily doped junctions (doping levels exceeding 10^{19} cm^{-3}), enabling significant current flow at much lower forward voltages (around 0.05-0.1 V). This results in a characteristic negative differential resistance (NDR) region where current decreases with increasing voltage, absent in conventional p-n diodes that exhibit purely positive resistance and monotonic forward conduction. In comparison to Schottky diodes, which achieve fast switching through majority transport across a metal-semiconductor without minority storage, tunnel diodes offer superior speed in the picosecond range (typically 10-100 ) due to the tunneling process, compared to Schottky's nanosecond-scale response. However, Schottky diodes lack the NDR feature that enables and in tunnel diodes, and they maintain a low forward (0.2-0.5 V) without the quantum tunneling dependency. Both types provide low-noise operation, but tunnel diodes are more prone to variations, particularly in early germanium devices where rising temperatures broaden the energy bands and reduce tunneling probability, potentially eliminating the NDR region above 200-300 K, whereas GaAs-based tunnel diodes maintain NDR at ; Schottky diodes show greater thermal stability in tasks. Tunnel diodes and Zener diodes both leverage band-to-band tunneling, but in distinct regimes: Zener diodes employ it for reverse-bias breakdown at low voltages (typically <5-6 V) with heavy doping to regulate voltage, exhibiting sharp reverse conduction without forward NDR, while higher voltage Zeners (>5-6 V) primarily use with moderate doping. In contrast, tunnel diodes utilize forward-bias tunneling at low voltages (<0.5 V) due to degenerate doping on both sides, producing the signature NDR for low-power, high-speed applications, while operating as a low-barrier in reverse bias similar to a conventional . This forward-oriented tunneling makes tunnel diodes unsuitable for high-voltage regulation, where Zener diodes excel, though tunnel diodes demonstrate greater temperature sensitivity, with NDR diminishing faster than Zener's stable breakdown characteristics.

Decline and Modern Relevance

By the 1970s and 1980s, tunnel diodes experienced a significant decline in mainstream adoption, primarily due to advancements in alternative technologies that offered better performance, integration, and cost-effectiveness. For and oscillator applications, they were largely supplanted by Gunn diodes and IMPATT diodes, which provided higher power handling capabilities and more stable operation at elevated frequencies without the stringent fabrication requirements of heavily doped tunnel structures. In logic and switching circuits, silicon bipolar junction transistors (BJTs) emerged as superior alternatives, enabling easier monolithic integration, lower production costs, and scalability in integrated circuits, which diminished the niche for discrete tunnel diode components. Contributing to their limited longevity were inherent challenges in fabrication and operational stability. Achieving the precise heavy doping levels necessary for quantum tunneling often resulted in low yields due to material impurities and control issues, complicating and increasing costs. Additionally, tunnel diodes exhibited temperature instability, with peak currents degrading under , which further restricted their reliability in commercial environments. Despite their decline, tunnel diodes retain niche relevance through evolutions like resonant tunneling diodes (RTDs), which have seen renewed interest in the for (THz) sources. RTD-based oscillators have demonstrated fundamental operation up to 1.98 THz at , with output powers reaching 0.7 mW at 1 THz using array configurations, positioning them as compact alternatives for and imaging. RTDs also serve as low-noise mixers in subharmonic configurations, achieving conversion losses below 10 dB at cryogenic temperatures, with potential extensions to requiring precise signal mixing. Recent developments emphasize integration with for mm-wave applications in / communications. Graphene-based RTDs have been engineered for high peak-to-valley ratios, enabling negative differential resistance at for ultrafast switching. Similarly, III-V compound RTDs, such as InAs/AlSb structures, support THz detection and generation up to 1 THz, with ongoing research into monolithic integration for enhanced efficiency. Their legacy persists in , where heavy doping confers superior radiation hardness against ionizing and , making them suitable for oscillators and radiation-tolerant circuits.

References

  1. [1]
  2. [2]
    [PDF] InAs/GaSb Tunnel Diodes - RIT Digital Institutional Repository
    Jul 8, 2015 · Figure 2.8: Example I-V characteristic of a tunnel diode. This plot shows the components of total current: tunneling current (Jt), excess ...<|control11|><|separator|>
  3. [3]
    Leo Esaki – Facts - NobelPrize.org
    Leo Esaki demonstrated a previously unknown type of tunneling in semiconductor material, material that is a cross between electrical conductors and insulators.
  4. [4]
    Robert Noyce and the Tunnel Diode - IEEE Spectrum
    May 2, 2005 · Noyce no longer worked for Shockley when he read Leo Esaki's Physical Review article in 1958. In September 1957, he, Moore, and six other ...<|control11|><|separator|>
  5. [5]
    1958: Tunnel Diode Promises a High-Speed Semiconductor Switch
    The Esaki or tunnel diode, named for the quantum-mechanical tunneling effect it exploited, offered fast switching speed with very low power consumption.
  6. [6]
    [PDF] APPLICATIONS OF TUNNELING TO ACTIVE DIODES - DTIC
    The low power consumption and small size of tunnel diodes offer advantages in these applications. For these reasons development effort should be continued ...
  7. [7]
    [PDF] 1957 Invention of the Esaki diode
    In July 1957, Esaki of Sony observed a tunnel current component expected in quantum theory during his experiment to measure current-voltage characteristics ...
  8. [8]
    Invention of tunnel diode - Esaki diode [Sony]
    In 1957, Leo Esaki of Sony (then Tokyo Tsusin Kogyo -Tokyo Communication Iindustry) discovered a negative resistance characteristics in the current-voltage ...
  9. [9]
    Chapter9 The Model 2T7 Transistor - Sony History
    Sony announced the successful testing of the Esaki Diode, also known as the tunnel diode, which Sony researcher Leona Esaki invented.
  10. [10]
  11. [11]
    [PDF] RCA Tunnel Diode Manual
    The high switching speed of tunnel diodes, coupled with their simplicity and stability, makes them par- ticularly suitable for high-speed computer applications.Missing: IBM | Show results with:IBM
  12. [12]
    TUNNEL DIODES FOR PHASED ARRAY RADAR. - DTIC
    Summaries and conclusions are given regarding the work performed and the results obtained in the device and circuits programs during the past year.
  13. [13]
    WKB Approximation - Richard Fitzpatrick
    The WKB approximation only applies to situations in which there is very little chance of a particle tunneling through the potential barrier in question.
  14. [14]
    The Tunneling Transistor - IEEE Spectrum
    Sep 30, 2013 · When an energy barrier has a thickness below about 10 nanometers, there is a small but nonzero probability that an electron that starts on one ...Missing: nm | Show results with:nm
  15. [15]
    [PDF] Silicon-based tunnel diode technology - Seabaugh Nanoelectronics
    Following Esaki's discovery, tunnel diodes have received interest because of their remarkable multivalued I-V characteristic and inherent high switching speeds.Missing: adoption | Show results with:adoption
  16. [16]
    [PDF] Characterization of Novel III-V Semiconductor Devices - DSpace@MIT
    Aug 1, 2025 · The tunnel diode is a typical p-n junction with the exception that both the p-side and n-side of the device are degenerately-doped. As a result, ...
  17. [17]
    Semiconductor Band Gaps - HyperPhysics
    Material. Energy gap (eV). 0K. 300K. Si. 1.17. 1.11. Ge. 0.74. 0.66. InSb. 0.23. 0.17. InAs. 0.43. 0.36. InP. 1.42. 1.27. GaP. 2.32. 2.25. GaAs. 1.52. 1.43.
  18. [18]
    Tunnel Diode - an overview | ScienceDirect Topics
    The tunnel diode is basically a very highly doped pn-junction (around 10 19 to 10 20 cm −3 ) that makes use of a quantum mechanical effect called tunneling.
  19. [19]
    [PDF] Leo Esaki - Nobel Lecture
    Current-voltage characteristics in a Si tunnel diode at 4.2, 80 and 298 K. ... Esaki, L., Phys. Rev. 109, 603 (1958);. Esaki, L., Solid State Physics in ...
  20. [20]
  21. [21]
  22. [22]
    [PDF] Silicon/silicon-germanium heterostructure tunnel diodes
    The interband tunnel diode was discovered by Leo Esaki [2] in 1957. He ... Fluid Analysis,” provisional patent filing U.S. Patent Office (30 July 2002).
  23. [23]
    Tunnel Diode: Structure & Fabrication - Electronics Notes
    The tunnel diode is created by using a wire coated with an alloy containing the required dopants. This is pressed hard onto the heavily doped substrate, and ...
  24. [24]
    Doping challenges and pathways to industrial scalability of III–V ...
    Jan 12, 2021 · This scheme requires current matching between the two junctions through a tunnel diode, i.e., a junction between degenerate semiconductors.Missing: heavy | Show results with:heavy
  25. [25]
    Gallium-arsenide deep-level tunnel diode with record negative ...
    Our large negative conductance results from a large density of deep-states within a narrow energy range in LTG-GaAs: 1020 cm−3 of arsenic-antisite AsGa [3] deep ...
  26. [26]
    Design principles and construction of planar Ge Esaki diodes
    The design and fabrication of Ge Esaki diodes are discussed with particular emphasis on the development of a planar process using conventional oxide masking ...Missing: evolution | Show results with:evolution
  27. [27]
    [PDF] AD L1 4 340 - DTIC
    The tunnel diode is an abrupt junction diode made from materials so heavily doped as to be degenerate in both the n and p regions. Because of this construction, ...
  28. [28]
    [PDF] PSpice Implementation of a New Esaki Tunnel Diode Macro- Model
    Esaki, “New phenomenon in narrow germanium p-n junctions,” Physical Review, vol. 109, No. 2,. (1958),pp. 603–604. [9] E. Warde, S. Sakr, M. Tchernycheva and ...
  29. [29]
    [PDF] A FORMING STUDY OF POINT-CONTACT TUNNEL DIODES - DTIC
    diodes by the application of brief electrical overloads are studied. As a test of the theory, a diode fabricated from N-doped germanium and an aluminum ...
  30. [30]
    [PDF] Low current density silicon tunnel diodes - University of Notre Dame
    A mesa etch is formed by reactive ion etching (Plasma Therm RIE 790) to isolate the devices, using aluminum as the mask (CF4/O2: 26.2/3.8 sccm, 100 W, 40 mT), ...
  31. [31]
    [PDF] Au/Ge/Ni/Au and Pd/Ge/Ti/Au Ohmic Contacts to AlxGa1−x As ...
    In this work, we developed an optimal Ohmic con- tact for the AlxGa1−xAs/InGaAs (x = 0.75) PHEMT structure from a comparative study of Au/Ge/Ni/Au and Pd/Ge/Ti/ ...
  32. [32]
    Computational design of NDR tunnel diodes with high peak ... - arXiv
    Dec 13, 2023 · Existing semiconductor-based NDR tunnel diodes suffer from low peak-to-valley current ratio (PVCR) values, which is attributed to the band tails ...
  33. [33]
    [PDF] Development of tunnel diode devices and models for circuit design ...
    Field effect transistor circuits augmented with tunnel diodes lead to decreased circuit footprints, decreased device count, improved operating speeds, and ...
  34. [34]
    Tunnel Diode: Construction, Working, Advantages, & Applications
    The barrier potential is directly proportional to the junction width. The tunnel diode has a very small barrier width in nanometers as compared to a normal PN ...
  35. [35]
  36. [36]
    [PDF] Polarization-induced Zener tunnel diodes in GaN/InGaN/GaN ...
    Oct 21, 2015 · The tunnel junction, on the other hand, shows a strong backward-diode behavior, conducting much higher current at reverse bias than forward bias ...
  37. [37]
    None
    Summary of each segment:
  38. [38]
    Tunnel Diode Equivalent Circuit | Behaviour | Construction
    The junction capacitance of the Tunnel Diode Equivalent Circuit is highly dependent on the bias voltage and temperature. Connecting a tuned circuit directly ...
  39. [39]
    [PDF] Tunnel Diode Manual
    dependent on the junction temperature produced by the current flowing through it, the physical size of the junction area and the efficiency with which heat ...
  40. [40]
    Tunnel diodes - TekWiki
    Jun 22, 2025 · Tunnel diodes (Esaki diodes) exhibit negative differential resistance in part of their U/I curve, ie increasing voltage results in a decreasing current in that ...
  41. [41]
    [PDF] Tunnel Diode Circuits at Microwave Frequencies - DTIC
    APPLICATIONS FOR THE TUNNEL DIODE. The negative resistance of the tunnel diode can be put to use in a va- riety of circuits with many useful properties. Two ...
  42. [42]
  43. [43]
    [PDF] ADVANCED TUNNELING DIODES FOR HIGH-FREQUENCY ...
    use of the diode's negative differential resistance feature, a 100 GHz oscillator with an output power of 83 μW was realised. Advanced design system ...
  44. [44]
    [PDF] TUNNEL-DIODE MICROWAVE AMPLIFIERS - Johns Hopkins APL
    Tunnel diodes provide a means of low-noise microwave amplification, with the amplifiers using the negative resistance of the tunnel diode to.Missing: definition | Show results with:definition
  45. [45]
    [PDF] Fabrication and characteristics of microwave backward diodes in ...
    These diodes, which are valuable for certain applications, are complemented by a special form of tunnel diode, the back diode microwave detector, which has also ...
  46. [46]
    Monostable circuit with tunnel diode has fast recovery
    A monostable multivibrator circuit using a tunnel diode makes it possible for the MSMV to exceed the performance of present multivibrators in two respects.
  47. [47]
    Resonant tunneling diodes for multi-valued digital applications
    Resonant tunneling diodes (RTD) have unique folding V-I characteristics, which lend themselves to multi-valued applications, The negative differential ...Missing: historical | Show results with:historical
  48. [48]
    Tunnel Diode - an overview | ScienceDirect Topics
    Japanese Journal of Applied Physics 40(10A): L1005–L1007. A 10-GHz four ... Leo Esaki was awarded the Nobel prize in physics in 1973. For an Esaki type ...
  49. [49]
    INVESTIGATION OF A TUNNEL-DIODE FLIP-FLOP - DTIC
    A nonsynchronous tunnel-diode flip-flop is described and analyzed. All of the detailed analysis deals with the case of an unloaded circuit.
  50. [50]
    Leo Esaki - IBM
    The IBM researcher earned the Nobel Prize for his discoveries in electron tunneling and changed the consumer electronics industry.
  51. [51]
    History Of The Diode | Hackaday
    Aug 15, 2016 · They were replaced in the 1960s by silicon rectifiers which have a lower voltage drop. In 1961 IBM tried to develop computer logic using ...
  52. [52]
    [PDF] IBM Research Report
    Oct 29, 2007 · The ability of the tunnel diode to carry current at very high density compelled attention. Tiny tunnel diodes had a small capacitance and ...
  53. [53]
    [PDF] PN and Metal–Semiconductor Junctions
    A PN junction is formed by converting a P-type semiconductor to N-type, or vice-versa, and acts as a rectifier or diode.
  54. [54]
    Tunnel Diode vs. P-N Junction Diode: Key Differences
    This article compares the tunnel diode and the p-n junction diode, highlighting the key differences between them. Both are semiconductor diodes used as ...
  55. [55]
    The pn‐Junction | part of Semiconductor Basics - IEEE Xplore
    The Schottky diode replaces one of the semiconductors by a metal, making it a faster and lower power diode. The tunnel diode, which can be explained only in ...
  56. [56]
    A Comparison of Tunnel Diode and Schottky Diode in Rectifier at 2.4 ...
    Apr 6, 2020 · Abstract: This paper presents a comparison of tunnel diode to Schottky diode in rectifier at 2.4 GHz under −10dBm input power region.
  57. [57]
    Negative resistance devices - Microwave Encyclopedia
    Specifically, IMPATT diodes, Gunn diodes, ISIS diodes and tunnel (Esaki) diodes have regions of negative resistance that are exploited. On a Smith chart ...Missing: declined | Show results with:declined
  58. [58]
    6.11: Diode and Vacuum Sources - Engineering LibreTexts
    Oct 23, 2024 · Three representative devices are the tunnel diode, the IMPATT diode, and the Gunn diode. Tunnel Diode. A tunnel diode is a pn junction diode ...
  59. [59]
  60. [60]
    Future Looks Bright For Tunnel Diodes, Promising Faster, More ...
    Oct 2, 1998 · When Nobel Prize winner Leo Esaki discovered the tunnel diode in 1957, the super-fast, current-switching device was touted as a kind of Holy ...<|separator|>
  61. [61]
    Thermal instability of peak current in tunnel diodes - SpringerLink
    It is shown that thermal stresses developed at the boundary between the electrode alloy and the semiconductor lead to deformations (plastic and creep), ...Missing: fabrication yield
  62. [62]
    Terahertz Emitter Using Resonant-Tunneling Diode and Applications
    By optimizing the areas of the varactor diode and RTD, a frequency change of about 100 GHz was obtained. Moreover, using an array configuration of frequency- ...
  63. [63]
    GaAs/AlAs resonant-tunneling diode for subharmonic mixers
    Aug 10, 2025 · The operation of a subharmonic mixer incorporating the resonant-tunneling diode is simulated for a number of values of its quantum-well width.
  64. [64]
    Toward High-Peak-to-Valley-Ratio Graphene Resonant Tunneling ...
    Sep 5, 2023 · For applications in high frequency electronics, multivalued logic gates, and memory devices, resonant tunneling diodes (RTDs) have drawn ...<|control11|><|separator|>
  65. [65]
    Advanced Terahertz Sensing and Imaging Systems Based on ...
    In this paper, we first review the development of heterostructure backward tunnel diodes (HBDs) in the InAs/AlSb/AlGaSb material system that have been ...
  66. [66]
    [PDF] TUNNEL DIODE
    This occurs up to a point known as the. “reverse breakdown voltage” when conduction begins (often accompanied by destruction of the device). In the tunnel diode ...Missing: datasheet | Show results with:datasheet