Fact-checked by Grok 2 weeks ago

Quantum capacitance

Quantum capacitance refers to the intrinsic capacitance arising from the finite density of electronic states (DOS) in a material, which limits the ability to store additional charge at a given Fermi level without a corresponding change in chemical potential. It is defined mathematically as C_Q = e^2 \rho(E_F), where e is the elementary charge and \rho(E_F) is the DOS at the Fermi energy E_F, and it manifests as an additional capacitive element in series with the classical geometric or electrostatic capacitance. This phenomenon becomes prominent in low-dimensional systems, such as two-dimensional materials like graphene or one-dimensional structures like carbon nanotubes, where the DOS is low and quantum effects dominate charge accumulation. The concept of quantum capacitance was first theoretically introduced by Serge Luryi in 1988, who described it as a consequence of the in quantum wells, requiring additional energy to fill discrete electronic states and thereby introducing a phase delay between voltage and current. Earlier observations by H. Gerischer in 1985 had hinted at reduced capacitance in materials with low , but Luryi's work formalized it within the framework of nanoelectronic devices. Theoretically, quantum capacitance originates from kinetic, exchange-correlation, and electron-phonon interactions in the electrode or channel material, and it can be positive or exhibit minima near zero bias due to thermal broadening effects. In practical applications, quantum capacitance plays a critical role in enhancing in supercapacitors, where it contributes to the total device alongside the electric double-layer , particularly in high-surface-area . For instance, in graphene-based electrodes, it can reach values around 2.55 μF/cm² at zero voltage, and modifications like nitrogen doping can significantly boost it to improve overall performance. In , such as field-effect transistors, quantum capacitance limits the gate control over the channel and influences switching speeds, while recent studies have revealed its quantum geometric origins in insulators through virtual interband transitions. Experimental measurement typically involves electrochemical impedance spectroscopy or to isolate its contribution from other capacitive elements.

Introduction

Definition and Basic Principles

Quantum capacitance refers to the fundamental limitation on charge storage in quantum-confined systems due to the finite density of states available for electrons. It was first theoretically introduced in the context of two-dimensional electron gases in semiconductor heterostructures. Formally, quantum capacitance C_q is defined as C_q = e^2 \frac{dn}{d\mu}, where e is the elementary charge, n is the carrier density, and \mu is the chemical (electrochemical) potential. This arises because adding or removing charge in such systems requires a change in the electrochemical potential to accommodate the carriers into available quantum states, rather than solely relying on electrostatic separation. In , the total capacitance C_\mu between a gate and the channel is given by the series combination C_\mu = \left( \frac{1}{C_g} + \frac{1}{C_q} \right)^{-1}, where C_g is the geometric associated with the physical separation of charges. When the is low, C_q becomes small and acts as a bottleneck, limiting the overall charge storage capacity below classical expectations. The \mu (or \bar{\mu}) differs from the electrostatic potential by incorporating quantum mechanical effects, such as the energy required to fill bands or account for Pauli exclusion; in contrast, the electrostatic potential describes only the classical . In semiconductors, band-filling effects further modulate C_q, as the potential must shift to populate higher levels, leading to non-constant that varies with . For instance, in insulators or semiconductors at low carrier densities, the sparse results in a very small C_q, which dominates the total and reduces the effective compared to classical insulators where is purely geometric. This effect is particularly pronounced in materials like III-V semiconductors, where the inherently low inherently limits C_q.

Historical Development

The of quantum capacitance, while implicitly connected to earlier theories of screening such as the Thomas-Fermi model in semiconductors, was first explicitly proposed by Serge Luryi in 1988. Earlier, in 1985, H. Gerischer had observed reduced in electrodes and attributed it to the low near the . Luryi introduced it in the context of mesoscopic systems, demonstrating its role in modifying the of short-channel metal-oxide-semiconductor field-effect transistors (MOSFETs) due to the finite in quantum-confined gases. This theoretical framework highlighted how quantum capacitance arises in parallel with geometric , affecting charge accumulation in low-dimensional structures. In 2003, Juan Bisquert extended the idea to electrochemical interfaces, interpreting the measured in dye-sensitized nanocrystalline TiO₂ solar cells as a chemical capacitance originating from the in nanostructured semiconductors. Bisquert emphasized its importance for understanding charge storage and recombination in systems, distinguishing it from traditional electrostatic . The first direct experimental confirmation of quantum capacitance occurred in 2006, reported by Shahal Ilani and colleagues in carbon nanotubes. Using a combination of charge sensing and scanning gate microscopy, they measured the thermodynamic capacitance of interacting electrons, revealing its dependence on electron-electron interactions and providing evidence for the predicted non-classical behavior in one-dimensional systems. Subsequent advancements after 2010 incorporated quantum capacitance into research on two-dimensional materials, particularly , where studies linked it to the linear dispersion of the . For instance, measurements of quantum capacitance in disordered allowed mapping of the average and fluctuating , underscoring its sensitivity to the relativistic-like band structure. This period marked the transition from isolated theoretical and experimental validations to broader recognition in , filling conceptual gaps from pre-1988 screening theories while enabling insights into quantum transport in emerging materials.

Theoretical Framework

Classical Capacitance Comparison

Classical capacitance, as defined in traditional , is a purely geometric quantity given by C = \epsilon \frac{A}{d}, where \epsilon is the of the medium, A is the area of the plates, and d is the separation between them. This formulation assumes an infinite (DOS) in the electrodes, allowing charge to be added or removed without any change in the chemical potential, and maintains a uniform potential across the conductor surfaces. Under these conditions, the behaves as an ideal electrostatic device with complete screening of electric fields by the metallic electrodes. In contrast, quantum capacitance arises from the finite DOS in quantum systems, such as two-dimensional electron gases (2DEGs), where adding charge requires an energy shift in the Fermi level to accommodate electrons into available states. Classical capacitance overlooks this kinetic energy cost associated with Fermi level modulation, treating the electrodes as having unlimited states at the same potential, whereas quantum capacitance accounts for the discrete nature of energy levels and the resulting resistance to charge accumulation. This difference becomes prominent in systems where the classical approximation of infinite DOS fails, leading to an effective capacitance that includes both electrostatic and quantum contributions. Quantum effects dominate when the quantum capacitance C_q is smaller than the geometric capacitance C_g, causing the total capacitance—given briefly by the series combination C_\text{total} = (C_g^{-1} + C_q^{-1})^{-1}—to be limited by C_q. Such conditions prevail in nanoscale or low carrier density systems, including high-mobility 2DEGs, where the classical model breaks down due to the sparse DOS near the , significantly altering charge storage behavior. For instance, in these 2DEGs, the quantum capacitance manifests as an additional "kinetic" contribution to charge storage in circuits, analogous to the inertial response of electrons but focused on static energy requirements rather than dynamic inductance.

Derivation and Key Equations

The quantum capacitance arises from the finite density of states in a material, which limits the change in carrier density for a given shift in the chemical potential. To derive it from first principles, consider the electron density n in a system as n = \int_0^\infty \rho(E) f(E - \mu) \, dE, where \rho(E) is the density of states, f(E - \mu) = [1 + \exp((E - \mu)/kT)]^{-1} is the Fermi-Dirac distribution, \mu is the chemical potential, k is Boltzmann's constant, and T is temperature. The charge Q = -e n (per unit area or volume), and the relevant voltage is the shift in electrochemical potential \Delta V = \Delta \mu / e. Thus, the quantum capacitance is defined as C_q = \frac{\partial Q}{\partial V} = e^2 \frac{\partial n}{\partial \mu}, which connects directly to the electronic compressibility of the system. At zero temperature (T = 0), the Fermi-Dirac function becomes a , so n = \int_0^\mu \rho(E) \, dE and \frac{\partial n}{\partial \mu} = \rho(\mu), yielding the key result C_q = e^2 \rho(\mu). This expression highlights that C_q is proportional to the at the , vanishing in gapped insulators where \rho(\mu) = 0 and diverging in systems with singular features. For a (2DEG) with a parabolic band dispersion, the is : \rho(E) = \frac{g_v g_s m^*}{2\pi \hbar^2}, where g_v is the valley degeneracy, g_s = 2 is the spin degeneracy, and m^* is the effective mass. Substituting into the zero-temperature formula gives C_q = \frac{g_v m^* e^2}{\pi \hbar^2}, which is independent of density and thus over a wide range of gate voltages. In typical 2DEGs like those in GaAs (g_v = 1), this simplifies to C_q = \frac{m^* e^2}{\pi \hbar^2}. In practical devices, the quantum capacitance combines in series with other capacitances to form the total electrochemical capacitance C_\mu, given by \frac{1}{C_\mu} = \frac{1}{C_g} + \frac{1}{C_q} + \frac{1}{C_{int}}, where C_g is the geometric (oxide) capacitance and C_{int} accounts for interface traps or other effects that can reduce the effective capacitance. This series model explains why C_q becomes observable when it is comparable to or smaller than C_g. The derivations assume non-interacting electrons at zero temperature; interactions like effects are neglected, though they can be incorporated perturbatively. For finite temperatures, the general form C_q = e^2 \frac{\partial n}{\partial \mu} holds, but \frac{\partial n}{\partial \mu} is computed using the Sommerfeld expansion to approximate the thermal smearing of the Fermi function, yielding corrections of order (kT/\mu)^2 to the zero-temperature DOS value.

Physical Underpinnings

Density of States Influence

The density of states (DOS), denoted as \rho(E), plays a central role in determining the quantum capacitance C_q, which is fundamentally proportional to the DOS at the Fermi energy, C_q = e^2 \rho(E_F) at zero temperature. This relationship arises because the DOS quantifies the number of available electronic states per unit energy, directly influencing the change in carrier density with respect to the chemical potential. In regions of low DOS, such as near band edges, the quantum capacitance is suppressed, limiting the ability to add or remove charge without significant shifts in the Fermi level. Conversely, in systems with a constant DOS, such as a two-dimensional electron gas with parabolic dispersion, C_q remains independent of carrier density, providing a stable capacitive response. In one-dimensional systems, features like Van Hove singularities—sharp peaks in the at subband edges—lead to enhanced local quantum capacitance. These singularities occur due to the flattening of energy bands at critical points, dramatically increasing the availability of states and causing abrupt rises in C_q as the aligns with these peaks. For instance, in carbon nanotubes, the subband manifests as distinct capacitance steps, reflecting the one-dimensional of the . This enhancement is crucial for understanding capacitance variations in quasi-1D conductors. Screening effects further highlight the DOS's influence, as described by the Thomas-Fermi approximation, where the screening length \lambda_{TF} \propto 1/\sqrt{\rho(E_F)} determines how effectively charges shield external fields. In the classical Thomas-Fermi limit, infinite DOS implies perfect screening with zero penetration depth, but finite DOS introduces quantum corrections that make \lambda_{TF} non-negligible, thereby linking C_q directly to the screening properties and allowing partial field penetration into the electron gas. These corrections become prominent in low-DOS scenarios, altering device electrostatics. At finite temperatures, thermal smearing of the Fermi edge broadens the effective over an energy window of order k_B T, increasing the integrated state availability and thus modulating C_q. This temperature-induced enhancement smooths sharp DOS features, such as those at band edges, and can raise C_q by factors depending on the system dimensionality and base DOS profile. For example, in semiconductors, this effect becomes noticeable above a few , influencing capacitance measurements. Ultimately, the governs the thermodynamic relation \frac{d\mu}{dn} = \frac{1}{e^2 \rho(E_F)} at low temperatures, representing the inverse slope of the versus carrier density curve and serving as the foundational link between microscopic state availability and macroscopic capacitive behavior.

Behavior in Low-Dimensional Systems

In systems, such as a confined in a , the is constant due to the parabolic , resulting in a constant quantum capacitance C_q = e^2 D, where D = m / (\pi \hbar^2) and m is the effective . This constancy arises because the DOS per unit area remains independent of energy in the absence of subband mixing. However, in finite-width quantum wells with multiple subbands, the overall DOS exhibits step-like features at subband edges, leading to corresponding steps in C_q as the crosses these thresholds during charging. In one-dimensional (1D) systems, like nanowires or carbon nanotubes, the confinement quantizes the energy into subbands, producing a that diverges at van Hove singularities near subband bottoms. This causes C_q to exhibit oscillatory behavior, with sharp peaks at these singularities corresponding to enhanced charge accumulation when the aligns with subband onsets. For instance, in ballistic 1D channels, the subband structure modulates C_q periodically with gate tuning, reflecting the inverse square-root energy dependence of the 1D away from singularities. Zero-dimensional (0D) systems, such as quantum dots, feature discrete energy levels to strong confinement in all directions, leading to a highly peaked C_q at the addition energies where electrons are sequentially added to the dot. In the Coulomb blockade regime, charge addition is suppressed between these energies, resulting in near-zero C_q in the blockade valleys, while C_q diverges sharply at the peaks to the delta-function-like DOS at discrete levels. This behavior stems from the interplay of single-particle level spacing and charging energy, with C_q maxima occurring precisely at the electrochemical potentials for integer occupancy changes. The gate voltage dependence of C_q in low-dimensional field-effect transistors (FETs) directly modulates the g_m, as g_m \propto C_q in the linear response regime, with C_q reaching minima in band-gap regions where the vanishes. For example, in 1D FETs, subband filling under gate bias causes C_q oscillations that imprint on g_m, enhancing device sensitivity to voltage but also introducing nonlinearity near singularities. The interplay between quantum capacitance and becomes pronounced in low-dimensional systems under ballistic transport conditions, where low C_q in DOS gaps amplifies across the channel. In such regimes, the reduced screening from finite C_q shifts more of the applied gate voltage drop onto the quantum system, steepening potential gradients and enhancing velocity overshoot in ballistic carriers. This effect is particularly evident in narrow, edge-disordered 2D ballistic devices, where nonmonotonic potential profiles emerge at low densities due to the series combination of geometric and quantum capacitances.

Experimental Aspects

Measurement Techniques

One established method for extracting quantum capacitance involves AC impedance spectroscopy, where an signal is applied to the device, and the -dependent impedance is analyzed to determine the total . The quantum capacitance C_q is isolated by subtracting the geometric capacitance C_g of the system, typically in a range spanning 1 Hz to 1 MHz to capture both low-frequency electrochemical responses and higher-frequency electronic contributions. This technique is particularly useful for two-dimensional materials like , where the impedance data allows separation of series capacitances without direct contact. DC capacitance-voltage (C-V) profiling serves as another key approach, employing voltage sweeps to measure as a function of applied , often visualized through Mott-Schottky plots that the reciprocal of the squared against voltage. In systems, these plots enable the separation of quantum capacitance from depletion capacitance by identifying the linear region dominated by space-charge effects, providing insights into carrier density and . The method requires careful selection of measurement frequencies to avoid diffusion capacitance dominance at high biases. Scanning probe methods, such as scanning impedance microscopy or scanning microscopy, offer non-contact, spatially resolved measurements of local quantum capacitance, ideal for heterogeneous structures like carbon nanotubes or sheets. A conductive probe tip is raster-scanned over the sample surface while an bias modulates the local , allowing detection of capacitance variations on the nanoscale through changes in reflected signals or tip-sample impedance. These techniques probe effective screening lengths and density-of-states modulations without invasive electrodes. In gating configurations, quantum capacitance is probed using a three-electrode electrochemical setup, where the charge Q accumulated on the (the low-dimensional material) is differentiated with respect to the gate voltage V to yield C_q = \frac{dQ}{dV}. This method leverages ionic gating to achieve high charge densities while minimizing parasitic effects from solid dielectrics, commonly applied to gated or nanotube devices immersed in an solution. Accurate extraction of quantum capacitance necessitates rigorous calibration to account for parasitic capacitances and quantum corrections during . Parasitic contributions from leads, contacts, or interfaces are subtracted using models fitted to baseline measurements, while quantum corrections—such as those from finite —are incorporated in post-processing to align experimental data with theoretical expectations for the material's band structure.

Observed Phenomena and Challenges

In , experimental measurements of quantum capacitance C_q as a function of carrier density reveal a characteristic V-shaped profile, with a minimum value at the neutrality point corresponding to the Dirac point, where the vanishes linearly. This minimum C_q is non-zero due to residual doping or , and C_q increases linearly away from the Dirac point on both and sides, consistent with the linear . In quantum dots, quantum capacitance exhibits oscillatory behavior linked to discrete shell filling as electrons are added one by one, manifesting as peaks or steps in capacitance-voltage traces that reflect charging events and shell structures analogous to orbitals. These oscillations arise from the quantization of energy levels, with periodic variations in C_q corresponding to the addition of electrons to successive shells, observed in few-electron regimes. Temperature effects on quantum capacitance in two-dimensional electron gases (2DEGs) demonstrate thermal broadening, where increasing temperature smears the sharp features in the , leading to an overall increase in C_q due to enhanced occupation of states near the . In GaAs-based 2DEGs, capacitance structures associated with energy gaps or broaden with temperature, reducing peak heights while confirming the temperature-dependent rise in average C_q. A primary challenge in quantum capacitance measurements is separating C_q from parasitic contributions like , which can dominate in nanoscale devices and obscure the intrinsic signal, particularly in high-mobility systems where low contact resistances are essential for accurate extraction. becomes prominent in low-capacitance regimes below 1 , limiting sensitivity and requiring low-temperature, high-impedance setups to mitigate thermal and effects. Sample inhomogeneity in , such as charge puddles in , further complicates measurements by introducing spatial variations in carrier density that broaden or distort C_q profiles. Quantum capacitance shows high sensitivity to disorder, necessitating ultra-clean samples with minimal impurities to observe pristine features like the Dirac minimum, as even low levels of scattering can suppress C_q or induce negative compressibility effects. Discrepancies between theory and experiment often stem from many-body interactions, such as excitons in quantum dots, which enhance binding energies and shift charging peaks beyond non-interacting predictions, requiring inclusion of correlation effects for accurate modeling. Cryogenic measurements in topological insulators have revealed quantum corrections to , including linear C_q behavior from surface Dirac fermions in the bulk-depleted regime. Recent advances as of 2024 highlight ongoing developments in 2D topological insulators, including edge-state contributions. For instance, as of 2025, dispersive gate-sensing measurements of quantum in InAs–Al devices have enabled interferometric single-shot detection at cryogenic temperatures, demonstrating applications in topological . Additionally, photonic has been proposed as a non-invasive probe for quantum geometric in insulators.

Applications

In Nanostructured Materials

In , the () exhibits a linear dependence on , \rho(E) \propto |E|, arising from its linear near the Dirac points. This results in a quantum capacitance C_q \propto \sqrt{n}, where n is the carrier density, with a characteristic minimum occurring at the charge neutrality point due to the vanishing at the Dirac . Experimental measurements confirm this behavior, showing C_q values on the order of 1–10 \muF/cm² near neutrality, increasing with doping away from the Dirac point. Carbon nanotubes display quantum capacitance influenced by their one-dimensional structure and chirality-dependent electronic subbands, leading to periodic oscillations in C_q as the crosses subband edges. In metallic nanotubes, such as armchair types, the features constant values between van Hove singularities, yielding relatively steady C_q modulated by interband transitions. Semiconducting nanotubes, like zigzag types, exhibit suppressed C_q within the bandgap, with sharp increases upon band filling. These chirality-specific variations, observed in gate-dependent measurements, highlight differences in filling across subbands for metallic versus semiconducting tubes. In other two-dimensional materials, quantum capacitance reflects band structure peculiarities. For monolayer MoS₂, the direct bandgap of approximately 1.8 eV suppresses C_q near the in the undoped state, as the low in the gap limits charge accumulation, with C_q dipping to near-zero values before rising sharply in the conduction or bands. In black phosphorus, structural anisotropy leads to direction-dependent C_q, with higher values along the armchair direction due to enhanced from puckered lattice geometry, compared to the direction; this directional variation stems from anisotropic effective masses. Doping in these nanostructured materials, whether chemical (e.g., or substitution) or electrostatic (via gating), modulates C_q by shifting the and altering the . In and carbon nanotubes, such tuning can enhance C_q by 20–50% through introduced midgap states or band filling, facilitating sensitive detection in chemical sensors where adsorbates induce local doping changes. Compared to , where parabolic dispersion yields a more constant but lower effective C_q in two-dimensional systems due to higher effective masses, 's relativistic linear bands enable inherently higher C_q values, often exceeding 10 \muF/cm² at moderate doping. In , such as Ti₃C₂, structural engineering like vacancies enhances quantum capacitance for improved performance.

In Electronic and Energy Devices

In , quantum capacitance plays a critical role in field-effect transistors (FETs), where its relatively low value—stemming from the material's linear band structure and minimal near the charge neutrality point—limits the total and thereby constrains device switching speeds. This limitation arises because the effective is determined by the series combination of the geometric () capacitance and quantum capacitance, with the latter often becoming the in high-frequency operation, capping cutoff frequencies below theoretical ballistic limits. To address this, dual-gate architectures have been developed, employing both and bottom gates to independently modulate the channel potential and effectively boost the overall through parallel electrostatic coupling, enabling improved and higher operational frequencies up to 50 GHz in scaled devices. Such designs enhance charge induction efficiency without altering the intrinsic quantum capacitance, providing a pathway for FETs to approach performance parity with counterparts in logic applications. In devices like supercapacitors, in nanostructured electrodes, such as TiO₂ nanotubes, significantly augments total by facilitating pseudocapacitive charge storage via density-of-states variations, surpassing classical electrostatic limits and increasing . For instance, in TiO₂-based systems, the quantum contribution enables efficient accumulation in the conduction band, yielding specific capacitances exceeding 150 F/g at moderate scan rates when combined with conductive additives like reduced graphene oxide. This mechanism is particularly beneficial in hybrid electrodes, where quantum effects bridge the gap between double-layer and faradaic processes, allowing devices to store more charge per unit volume while maintaining fast charge-discharge kinetics. Quantum capacitance-based sensors leverage fluctuations in C_q induced by analyte-induced shifts in the density of states for high-sensitivity detection of biomolecules and gases. In graphene or transition metal dichalcogenide FETs, biomolecule adsorption—such as DNA or proteins—alters the local Fermi level, causing measurable C_q variations that translate to conductance changes with sub-ppb detection limits for species like NO₂ or NH₃. Similarly, for biomolecules, quantum capacitance-limited MoS₂ biosensors detect pH shifts or protein binding with 75-fold enhanced sensitivity over traditional ionic-gated devices, enabling remote, label-free monitoring in physiological environments. In dye-sensitized solar cells (DSSCs) employing TiO₂ photoanodes, quantum capacitance governs charge separation at the dye-electrolyte interface and influences recombination rates, directly impacting (V_{oc}) . By modeling C_q as a function of quasi-Fermi level, optimizations such as plasmonic incorporation enhance C_q by up to 20%, reducing recombination losses and boosting V_{oc} to values approaching 0.9 V while improving overall power conversion . This approach allows for fine-tuned band alignment, minimizing energy barriers for charge injection and extraction in nanostructured TiO₂ networks. Looking to future implications, quantum capacitance holds promise for capacitors, where precise control of charge states in low-dimensional systems like graphene-hexagonal heterostructures enables readout of Rydberg transitions via capacitance oscillations, potentially stabilizing operations. In neuromorphic devices, discrete steps in C_q arising from quantized density-of-states features in materials like MoS₂ facilitate memcapacitor designs that emulate , with state-dependent capacitance variations mimicking learning-forgetting cycles and enabling energy-efficient analog computing at picajoule scales. Recent advances include harnessing quantum capacitance in material/molecular layer junctions for novel electronic device functionality as of 2024.

References

  1. [1]
    Quantum capacitance: The large but hidden ... - ScienceDirect.com
    Capacitance is defined as the ability of a material to hold electric charges. This ability classically depends on the surface area and distance between the ...
  2. [2]
    Quantum Capacitance of Two-Dimensional-Material-Based ...
    This review critically examines the supercapacitor design strategies, including choosing the right materials and electrolytes.
  3. [3]
    Quantum capacitance devices | Applied Physics Letters
    Feb 8, 1988 · S. Luryi and S. M. Sze, in Silicon Molecular Beam Epitaxy, edited by E. Kasper and J. C. Bean (CRC Uniscience, Boca Raton, FL, 1987) ...
  4. [4]
    The quantum geometric origin of capacitance in insulators - Nature
    May 30, 2024 · This capacitance is a purely quantum phenomenon that arises from coherent, virtual interband transitions between the full valence band and the ...Gapped Dirac Hamiltonians · Kubo Formula For... · Dielectric Constant Of...
  5. [5]
  6. [6]
    its origin and significance for nanocomposite solar cells
    The capacitance measured in dye -sensitised nanocrystalline TiO 2 solar cells (DSSC) is interpreted in terms of a chemical capacitance.Missing: quantum | Show results with:quantum
  7. [7]
    Measurement of the quantum capacitance of interacting electrons in ...
    Sep 24, 2006 · In this work we present the first direct capacitance measurements of carbon nanotubes (NT). A NT strongly coupled to a metallic gate forms a ...
  8. [8]
    Density of States and Its Local Fluctuations Determined by ... - Nature
    May 3, 2013 · By measuring the quantum capacitance of disordered graphene we simultaneously determine both the ADOS and its local fluctuations of the system.
  9. [9]
    Nanoscale origins of super-capacitance phenomena - ScienceDirect
    Feb 28, 2019 · We review the origins of capacitive phenomena at the nanoscale to demonstrate that electrochemistry cannot be understood without a critical ...
  10. [10]
    [PDF] Theory of single-electron charging of quantum wells and dots
    current-voltage characteristics of the quantum well may exhibit a Coulomb staircase type structure similar to that observed in its metallic counterpart. However ...<|separator|>
  11. [11]
    [PDF] Quantum Capacitance in Nanoscale Device Modeling
    Aug 16, 2004 · The concept of “quantum capacitance” was used by Luryi1 in order to develop an equiv- alent circuit model for devices that incorporate a highly ...
  12. [12]
    Role of quantum capacitance in coupled low-dimensional electron ...
    Mar 27, 2006 · Based on the quantum capacitance of the involved layers we develop a general model to determine the charging state of coupled low-dimensional ...
  13. [13]
    [PDF] ELECTRON TRANSPORT IN QUANTUM DOTS.
    The addition energy µdot(N+1) − µdot(N) is large for a small capacitance and/or a large energy splitting ∆E = EN+1 − EN between 0D-states. It is important to.
  14. [14]
    [PDF] How Do Quantum Effects Influence the Capacitance and Carrier ...
    Quantum effects, including quantum and centroid capacitance, reduce gate capacitance in MoS2 transistors, limiting gate control. At 0.5 nm EOT, MoS2  ...<|control11|><|separator|>
  15. [15]
    Gate electrostatics and quantum capacitance in ballistic graphene ...
    May 6, 2019 · We experimentally investigate the charge induction mechanism across gated, narrow, ballistic graphene devices with different degrees of edge ...Missing: bending | Show results with:bending
  16. [16]
    [PDF] Gate electrostatics and quantum capacitance in ballistic graphene ...
    This profile is nonmonotonic at low carrier densities due to the interplay between classical and quantum capacitance contributions across the gated device.Missing: bending | Show results with:bending
  17. [17]
  18. [18]
  19. [19]
    Quantum capacitance and density of states of graphene
    The density of states, which is directly proportional to the quantum capacitance, is found to be significantly larger than zero at and around the charge ...
  20. [20]
    Shell structures in self-assembled InAs quantum dots probed by ...
    Nov 9, 2005 · In this work, we have investigated electron filling in single InAs QDs grown on the surfaces of GaAs by measuring tunneling conductances in a ...
  21. [21]
    Shell Filling and Spin Effects in a Few Electron Quantum Dot
    Oct 21, 1996 · We study atomiclike properties of artificial atoms by measuring Coulomb oscillations in vertical quantum dots containing a tunable number of electrons starting ...
  22. [22]
    [PDF] Energy gaps of the two-dimensional electron gas explored with ...
    Aug 15, 1993 · capacitance structure is a strong function of temperature through the temperature range down to below ... washed out due to thermal broadening.
  23. [23]
    [PDF] Chapter 13 - Quantum Capacitance of Graphene Sheets and ... - arXiv
    Graphene has an atomically thin body so that its quantum capacitance can dominate the device's electrostatics. Moreover, its DOS is a strong function of Fermi ...
  24. [24]
    rf Quantum Capacitance of the Topological Insulator in the Bulk ...
    Feb 22, 2018 · Allowing this contact resistance to vary yields negligible variation compared to the fivefold increase observed in the channel resistance. We ...
  25. [25]
    Large array of low-frequency readout quantum capacitance detectors
    Jan 6, 2021 · The QCD has demonstrated photon shot noise-limited performance for 1.5 THz radiation under loading conditions between 10 − 20 and 10 − 18 W and ...
  26. [26]
    Influence of density inhomogeneity on the quantum capacitance of ...
    Our numerical results show that the multi-peaks in the non-monotonic quantum capacitance–voltage characteristics are broadened as the potential fluctuation ...
  27. [27]
    Negative compressibility observed in graphene containing resonant ...
    Negative compressibility has been suggested to exist in an ultraclean electron system with a low density of impurities.1–3 But much to our surprise, we observed ...
  28. [28]
    (PDF) Direct Quantitative Electrical Measurement of Many-Body ...
    Jan 29, 2014 · By this we demonstrate an electrical measurement of excitonic states in quantum dots. Magnetocapacitance-voltage spectroscopy reveals that the ...
  29. [29]
    [PDF] RF-Quantum Capacitance of the Topological Insulator Bi 2 Se ... - HAL
    Dec 1, 2020 · We are able to simultaneously measure both the quantum capacitance and the channel resistance of the device. The capacitance exhibits a linear ...
  30. [30]
    2024 roadmap on 2D topological insulators - IOPscience
    Mar 5, 2024 · 2D topological insulators promise novel approaches towards electronic, spintronic, and quantum device applications. This is owing to unique ...
  31. [31]
    Quantum Capacitance Extraction for Carbon Nanotube Interconnects
    In this paper we have quantified the quantum capacitance (Cq) of a variety of carbon nanotubes including different chiralities (armchair and metallic zigzag ...
  32. [32]
    Probing the electron states and metal-insulator transition ... - Nature
    Jan 14, 2015 · (d) The quantum capacitance Cq of MoS2 plotted as a function of surface potential Vs at 300 K, which yields a band gap width of around 1.14 eV.
  33. [33]
    Improving the Quantum Capacitance of Graphene-Based ...
    Aug 2, 2019 · We explore the stability, electronic properties, and quantum capacitance of doped/co-doped graphene with B, N, P, and S atoms based on first-principles methods.
  34. [34]
    Assessment of High-Frequency Performance Limits of Graphene ...
    Feb 7, 2011 · The intrinsic cut-off frequency is close to the LC characteristic frequency set by graphene kinetic inductance and quantum capacitance, which ...
  35. [35]
    [PDF] Mobility Extraction and Quantum Capacitance Impact in High ... - arXiv
    The field-effect mobility of graphene devices is discussed. We argue that the graphene ballistic mean free path, Lball can only be extracted by taking into ...
  36. [36]
    Equivalent Circuit Modeling of a Dual-Gate Graphene FET - MDPI
    The quantum capacitance and surface potential dependence on the top-gate-to-source voltage were studied for monolayer and bilayer graphene channel by using ...<|separator|>
  37. [37]
    RGO nanocomposite as high specific capacitance electrode for ...
    Aug 12, 2024 · CV studies highlight that the TiO2-RGO composite has a high specific capacitance of 152 F g−1 at a substantially faster scan rate of 25 mV s−1 ...
  38. [38]
    High Charge-Storage Performance of Morphologically Modified ...
    Mar 4, 2021 · We report the influence of morphological changes on the process of charge storage for electrode materials based on T i ⁢ O 2 nanotubes and provide theoretical ...
  39. [39]
    Quantum Capacitance Model for Graphene FET-Based Gas Sensor
    Feb 1, 2019 · In this paper, a new quantum capacitance model for the gas sensor employing the graphene field effect transistor platform is proposed. Hence, a ...Missing: biomolecules | Show results with:biomolecules
  40. [40]
    Quantum Capacitance-Limited MoS2 Biosensors Enable Remote ...
    Channel noise could further limit SNR when the devices are operated in the inversion-depletion regime, for example as demonstrated with ultra-thin body double- ...
  41. [41]
    Dye versus Quantum Dots in Sensitized Solar Cells - ACS Publications
    Nov 16, 2011 · In opposition to the conventional dye-sensitized solar cell, here we show that inorganic QDs play a direct role in the recombination process.
  42. [42]
    Efficient quantum capacitance enhancement in DSSC by gold ...
    Aug 9, 2025 · An optimization work on dye-sensitized solar cells (DSSCs) based on both artificial and natural dyes was carried out by a fine synthesis ...
  43. [43]
    MoS 2 quantum capacitance for memcapacitor - AIP Publishing
    Oct 3, 2024 · In this work, we proposed a unique memcapacitor based on MoS2 quantum capacitance, capable of simulating synaptic learning-forgetting ability.