Fact-checked by Grok 2 weeks ago

Quantum Hall effect

The Quantum Hall effect (QHE) is a quantum mechanical manifestation of the classical observed in two-dimensional systems (2DES) at low temperatures and under strong perpendicular , where the Hall conductance quantizes into discrete plateaus at values \sigma_{xy} = \nu \frac{[e^2](/page/elementary_charge)}{[h](/page/H+)}, with \nu as the filling factor (an integer or rational fraction), [e](/page/elementary_charge) the , and [h](/page/H+) Planck's , of impurities or sample . The integer quantum Hall effect (IQHE) was discovered in 1980 by Klaus von Klitzing during measurements on silicon metal-oxide-semiconductor field-effect transistors (MOSFETs), revealing Hall resistance plateaus at R_H = \frac{h}{i e^2} for integer i, with precision on the order of 1 part in 10^6 (3 ppm). This quantization arises from the filling of discrete formed by orbits of electrons in the , with extended states at the centers of these levels contributing to conduction while localized states at the edges cause the plateaus. Von Klitzing's finding, which linked the effect directly to fundamental constants, earned him the 1985 and established the QHE as a cornerstone for precise , enabling the international standard for resistance (the ) based on \frac{h}{e^2}. The fractional quantum Hall effect (FQHE), an even more profound extension, was observed in 1982 by Daniel C. Tsui and Horst L. Störmer in high-mobility gallium arsenide heterostructures at temperatures below 4 K and magnetic fields exceeding 12 T, showing plateaus at fractional filling factors like \nu = 1/3, corresponding to Hall resistance R_H = 3 \frac{h}{e^2}. This counterintuitive behavior, defying simple single-particle models, was theoretically explained by Robert B. Laughlin in 1983 through a many-body wavefunction describing an incompressible quantum fluid of electrons, where collective interactions produce quasiparticles with fractional charge (e.g., e/3) and anyonic statistics, forming a novel state of matter. Tsui, Störmer, and Laughlin shared the 1998 Nobel Prize in Physics for this discovery, highlighting the role of strong electron correlations in the FQHE. Beyond its foundational role in , the QHE illuminates topological phases of matter, where robustness against disorder stems from global geometric properties rather than local symmetries, influencing fields from to exotic braiding for fault-tolerant qubits. The effect's precision has redefined electrical units in the system since 2019, tying the and to exact values of h and e, and continues to drive into higher-order topological insulators and non-Abelian states.

Theoretical Background

Classical Hall Effect

The Hall effect refers to the generation of a voltage difference across an , transverse to an applied , when subjected to a perpendicular to the current flow. This phenomenon arises from the exerted on moving charge carriers, which deflects them toward one side of the conductor, creating a charge accumulation and an opposing that balances the magnetic in steady state. Discovered by Edwin Hall in 1879 using a thin sample, the effect provides a means to determine properties such as carrier type and density in materials. In the standard experimental setup, a thin, flat —often a metallic strip or sample—is oriented such that flows along the x-direction, driven by an applied voltage. A uniform is applied along the z-direction, to the plane of the sample, while the transverse Hall voltage is measured between contacts placed along the y-direction on opposite edges of the sample. Hall's original apparatus employed a approximately 2 cm wide and 9 cm long, mounted on glass and positioned between the poles of an , with supplied by a Bunsen cell and the voltage detected using a sensitive . The observed deflection in the was proportional to the strength and reversed upon inverting the field, confirming the transverse nature of the . Within the classical , the Hall resistivity \rho_{xy}, defined as the ratio of the transverse to the , is given by \rho_{xy} = -\frac{B}{n e}, where B is the strength, n is the of charge carriers (electrons for n-type materials), and e is the . This relation emerges from the equilibrium condition where the \mathbf{F} = -e (\mathbf{v} \times \mathbf{B}) on drifting carriers (with negative velocity for positive current in n-type) balances the Hall , yielding a linear dependence on B and inverse proportionality to carrier (in magnitude). The formula assumes isotropic scattering and neglects quantum mechanical effects, allowing direct extraction of n from measured \rho_{xy}. The classical Hall effect accurately describes observations at room temperature and moderate magnetic fields, where thermal energy dominates over cyclotron motion. However, at low temperatures and high magnetic fields—particularly in thin, high-mobility samples approaching two-dimensional confinement—the linear relationship deviates, signaling the onset of quantum phenomena where carrier motion quantizes into discrete levels.

Two-Dimensional Electron Gas

The (2DEG) forms at the interface between two semiconductors with different band gaps, where electrons are confined to a thin layer due to the created by the band offset. In metal-oxide-semiconductor field-effect transistors (MOSFETs), the 2DEG arises in the inversion layer beneath the when a positive gate voltage depletes holes and accumulates electrons at the Si-SiO₂ interface, typically achieving densities around 10¹¹–10¹² cm⁻². This configuration was pivotal in the initial observation of the quantum Hall effect, as it provides a controllable planar system for magnetotransport measurements. A more advanced realization occurs in modulation-doped heterostructures, such as GaAs/AlGaAs, where the 2DEG resides at the hetero separated from ionized donors in the wider-band-gap AlGaAs layer. This spatial separation, introduced via with intentional undoping near the interface, minimizes from impurities and enables densities of 2–5 × 10¹¹ cm⁻². The transfer from the doped AlGaAs to the GaAs triangular well formed by the conduction band discontinuity (approximately 0.3 for x ≈ 0.3 in AlₓGa₁₋ₓAs), resulting in a degenerate confined to roughly 10 nm thickness. Key properties of these 2DEGs include high , often exceeding 10⁶ cm²/V·s at cryogenic temperatures in optimized GaAs/AlGaAs structures, due to reduced from remote doping. Low disorder is essential, as residual impurities or interface roughness can broaden the wavefunctions and degrade ; for instance, mean free paths can reach hundreds of micrometers in high-quality samples. The carrier density is tunable via gate electrodes deposited on the surface, which modulate the and thus the occupancy, spanning ranges from 10¹⁰ to 10¹² cm⁻² without significant loss in capacitively coupled designs. In a magnetic field B, electrons in the 2DEG undergo cyclotron motion, characterized by the frequency ω_c = eB / m^, where e is the and m^ is the effective mass (0.067 m_e for GaAs). This classical orbital motion, with r_c = ℏk_F / (eB) scaling inversely with B, sets the stage for quantized levels but requires minimal for clear Hall plateaus. Observing the quantum Hall plateaus demands exceptionally clean samples with above ~10⁵ cm²/V·s to ensure localization of states between and suppress , as lower quality leads to smoothed transitions rather than sharp quantized features.

Landau Levels and Quantization

In a two-dimensional electron gas subjected to a perpendicular magnetic field \mathbf{B} = B \hat{z}, the quantum mechanical description begins with the Schrödinger equation for a single electron of charge -e (where e > 0) and effective mass m^*. The Hamiltonian is H = \frac{1}{2m^*} (\mathbf{p} + e \mathbf{A})^2, where \mathbf{p} = -i \hbar \nabla is the canonical momentum and \mathbf{A} is the vector potential satisfying \nabla \times \mathbf{A} = \mathbf{B}. In the Landau gauge \mathbf{A} = (0, B x, 0), the equation separates into independent motion along the y-direction (free particle) and oscillatory motion along x. This yields a set of discrete energy eigenvalues known as Landau levels. The energy spectrum consists of equally spaced levels given by E_n = \hbar \omega_c \left( n + \frac{1}{2} \right), \quad n = 0, 1, 2, \dots where \omega_c = e B / m^* is the cyclotron frequency. Each Landau level is infinitely degenerate in the absence of boundaries or disorder, with the degeneracy per unit area g = e B / h, corresponding to one state per magnetic flux quantum \Phi_0 = h / e. The wavefunctions for the nth level take the form of plane waves in the y-direction, \exp(i k y), combined with harmonic oscillator functions in the x-direction shifted by the guiding center coordinate x_0 = - l_B^2 k, where l_B = \sqrt{\hbar / (e B)} is the magnetic length. This structure reflects the quantization of cyclotron orbits into discrete rings in the semiclassical limit. The occupation of these levels is characterized by the filling factor \nu = n_e h / (e B), where n_e is the areal electron density. When \nu is an integer, the Fermi level lies in a gap between filled and empty Landau levels, stabilizing the system against low-energy excitations. This quantization underpins the integer quantum Hall effect observed in high-mobility two-dimensional systems at low temperatures and strong fields.

Integer Quantum Hall Effect

Density of States

In the absence of disorder, the density of states (DOS) in a two-dimensional electron gas subjected to a strong perpendicular magnetic field exhibits a series of infinitely narrow delta-function peaks corresponding to the discrete Landau levels, located at energies E_n = \hbar \omega_c (n + 1/2), where \omega_c = eB / m^* is the cyclotron frequency, n = 0, 1, 2, \dots, e is the electron charge, B is the magnetic field strength, and m^* is the effective mass. Each Landau level possesses a high degeneracy, with the number of states per unit area given by eB / h, where h is Planck's constant; this degeneracy originates from the quantization of the orbital motion, where each state corresponds to one magnetic flux quantum \Phi_0 = h/e threading the sample. Consequently, the filling factor \nu = n h / (e B), with n the electron density, determines how many Landau levels are occupied, leading to sharp steps in transport properties when \nu is integer. In realistic systems, such as heterostructures, disorder arising from impurities, interface roughness, and background scatterers broadens these ideal delta-function peaks, transforming the into finite-width distributions with tails extending into the inter-level gaps. The broadening is commonly modeled as Gaussian in shape, with a width \Gamma typically on the order of a few meV, depending on sample quality and strength; this arises from the random potential fluctuations that mix states within and between nearby , particularly through short-range scatterers. The tails in the , often exponential or Gaussian-like, allow for a small but nonzero density of states in the nominally forbidden gaps, though the extended states remain localized near the center of each level due to effects. At zero temperature, the \mu positions itself within the broadened Landau levels when partially filled, pinning to the DOS peak and enabling metallic conduction. However, as the or is tuned such that \nu reaches an integer value, \mu undergoes discontinuous jumps between adjacent Landau levels, residing in the depleted regions of the DOS within the gaps. These gaps between fully filled levels correspond to insulating states, where the vanishing DOS at \mu suppresses bulk conduction, while the finite DOS tails can influence localization lengths and the sharpness of quantum Hall plateaus at finite temperatures.

Quantized Hall Conductivity

In the integer quantum Hall effect, the Hall conductivity \sigma_{xy} exhibits precise quantization, taking values \sigma_{xy} = \nu \frac{e^2}{h}, where \nu is an known as the filling factor, e is the , and h is Planck's constant. This quantization arises in two-dimensional electron systems subjected to strong perpendicular magnetic fields and low temperatures, where the filling factor \nu represents the number of filled . Experimentally, this was first observed by Klaus von Klitzing in 1980 using silicon metal-oxide-semiconductor field-effect transistors, where plateaus in the Hall resistance R_H = \rho_{xy} = \frac{h}{\nu e^2} were measured with high precision, independent of sample details like mobility or geometry. The quantized Hall resistance corresponds to the von Klitzing constant R_K = \frac{[h](/page/H+)}{[e](/page/E!)^2} \approx 25812.807\, \Omega, such that R_H = \nu R_K. This constant provides a universal standard for resistance , as its value depends solely on fundamental constants and has been verified to better than 1 part in $10^{10} through comparisons with conventional resistance standards. The quantization persists over wide ranges of and , with deviations occurring only near the transitions between plateaus. Theoretically, the quantization can be derived using the Kubo formula, which expresses the linear response conductivity in terms of current-current correlation functions. For a non-interacting in a , the Hall conductivity is given by \sigma_{xy} = \frac{e^2}{\hbar} \sum_n f_n \Omega_n, where f_n is the Fermi occupation of the n-th Landau level and \Omega_n is its Berry curvature, but in the clean limit with filled levels, it simplifies to \sigma_{xy} = \nu \frac{e^2}{h} due to the degeneracy and gap structure. In the presence of disorder or a weak periodic potential, the Kubo formula yields an multiple of \frac{e^2}{h} when the lies in a gap between levels, as the off-diagonal response terms cancel for partially filled states but accumulate integrally for complete fillings. A deeper topological understanding comes from the thermodynamic argument involving the Chern number, or TKNN invariant, introduced by Thouless, Kohmoto, Nightingale, and den Nijs in 1982. This invariant, defined as the integral of the curvature over the for each filled , equals an integer C_n for the n-th Landau level, leading to \sigma_{xy} = \left( \sum_n C_n \right) \frac{e^2}{h}. For isolated Landau levels in a periodic potential, each contributes C_n = 1, ensuring robust quantization protected by the of the filled states, even against weak disorder or lattice effects, as long as gaps remain open. This framework explains the universal nature of the observed plateaus without relying on dissipationless edge transport.

Edge States and Resistivities

In the integer quantum Hall effect, transport occurs predominantly through chiral edge states that propagate unidirectionally along the boundaries of the sample. These states carry dissipationless current due to the absence of backscattering, a consequence of the strong confining trajectories into skipping orbits near the edges, where cyclically bounce off the boundary without reversing direction. The arises from the , ensuring that edge states on opposite boundaries propagate in opposite directions, with the direction determined by the orientation. This edge-dominated transport contrasts with bulk conduction, as the wavefunctions of filled extend to the sample edges, forming these conducting channels while the interior remains insulating. The longitudinal resistivity \rho_{xx} vanishes (\rho_{xx} = 0) within the Hall plateau regions because electron states in the bulk are localized by , preventing or that would contribute to longitudinal voltage drops. Current flows exclusively along the chiral edge channels without energy loss, as elastic or inelastic backscattering between counter-propagating edges is suppressed by the and spatial separation. In contrast, during transitions between plateaus—when the crosses extended states at the center of a Landau level—\rho_{xx} exhibits peaks due to partial equilibration and between edge channels. The transverse Hall resistivity \rho_{xy} forms quantized plateaus at values \rho_{xy} = \frac{h}{\nu e^2}, where \nu is the filling corresponding to the number of occupied edge channels. Each fully transmitted chiral channel contributes a e^2/h, leading to precise quantization as long as the chemical potentials at the contacts are equilibrated without inter-channel mixing. This behavior is rigorously described by the Landauer-Büttiker formalism, which models multi-terminal conductance in terms of transmission probabilities between edge channels, predicting exact quantization for ideal samples where edge states remain uncoupled. Dissipation occurs only in the transitional regions, where partial backscattering or inter-channel scattering broadens the plateaus' boundaries.

Fractional Quantum Hall Effect

Laughlin Wavefunction

In 1983, Robert Laughlin proposed a trial wavefunction to describe the of a at fractional filling factors \nu = 1/m, where m is an odd positive , providing a theoretical framework for the observed fractional quantization in the quantum Hall effect. This builds upon the filled Landau level wavefunction for the case but incorporates strong correlations through a Jastrow-like factor to account for the incompressible nature of the state at these fillings. The explicit form of the Laughlin wavefunction in the symmetric gauge is given by \psi_m(z_1, \dots, z_N) = \prod_{i < j} (z_i - z_j)^m \exp\left( -\sum_{k=1}^N \frac{|z_k|^2}{4 \ell_B^2} \right), where z_k = x_k + i y_k are the complex coordinates of the electrons, N is the number of electrons, and \ell_B = \sqrt{\hbar / eB} is the magnetic length. The wavefunction serves as a variational ansatz, and its energy is minimized within the lowest Landau level by projecting onto the single-particle orbitals, demonstrating that it yields the lowest variational energy at filling \nu = 1/m compared to other trial states. This minimization reveals the incompressibility of the state: the system resists changes in particle density without a significant energy cost, analogous to a quantum liquid with short-range correlations that prevent clustering or crystallization. Calculations show that for m=3 (corresponding to \nu = 1/3), the pair correlation function exhibits a deep minimum at short distances, suppressing probability for electrons to occupy the same position, which stabilizes the fractional state against perturbations. Laughlin further interpreted the modulus squared of the wavefunction, |\psi_m|^2, as the Boltzmann factor of a classical two-dimensional one-component plasma (a ) at finite temperature T = 1/m, where electrons interact via a logarithmic potential in the plane. This plasma analogy, derived by mapping the quantum probability density to a classical equilibrium distribution, predicts a fluid phase for the gas at the relevant densities, with no crystalline order, confirming the liquid-like ground state. Monte Carlo simulations of this plasma validate the ansatz's stability, showing phase transitions only at densities far from the quantum Hall regime. On closed surfaces of higher genus, such as a , the Laughlin wavefunction predicts a degenerate ground state with degeneracy m^{2g-2} for genus g > 1, arising from the topological properties of the state and the need for single-valued wavefunctions under non-contractible loops. This degeneracy, exact within the model, underscores the robustness of the incompressible fluid against geometric deformations and provides a topological characterizing the . For the simplest non-trivial case of a (g=1), the degeneracy is m, reflecting the flux insertion sectors.

Composite Fermions and Quasiparticles

In the composite fermion model developed by , electrons in a strong perpendicular are treated as composite objects formed by binding each to an even number $2p of magnetic flux quanta, where p is a positive . These s behave as fermions in an effective magnetic field B^* = B - 2p n \frac{h c}{e}, with B the external field, n the two-dimensional , h Planck's constant, c the , and e the . The composite fermions then occupy in this reduced field, leading to quantized Hall states at filling factors \nu = p/(2p+1), which map the onto a series of quantum Hall states of the composite fermions. This framework unifies the observed principal sequence of fractions, such as \nu = 1/3, $2/5, $3/7, with the Laughlin state at \nu = 1/m (for odd m = 2p + 1) corresponding to the p=1 case. To account for the full hierarchy of observed fractional states, the model extends to higher levels where composite fermions themselves undergo a , condensing into their own Laughlin-like states. This generates a sequence of filling factors \nu = p / (2 p q \pm 1), with q a positive integer, producing daughter states that explain both odd- and even-denominator fractions, such as \nu = 4/11 or $1/2. The excitations in these fractional states are exotic quasiparticles known as anyons, which obey fractional statistics intermediate between bosons and fermions. For quasiholes in the Laughlin states at \nu = 1/m, exchanging two quasiholes around each other produces a braiding phase e^{i \theta} with statistical angle \theta = \pi / m. These quasiholes also carry a fractional electric charge e/m, as theoretically derived from the topological properties of the ground state wavefunction and confirmed through adiabatic continuity arguments.

Experimental Verification

The fractional quantum Hall effect was first experimentally observed in 1982 by Daniel C. Tsui, Horst L. Störmer, and Arthur C. Gossard in a high-mobility formed in a GaAs-AlGaAs heterostructure. They reported a Hall resistance plateau at R_{xy} = 3 h / e^2 corresponding to the filling factor \nu = 1/3, observed under magnetic fields around 10 T and temperatures around 0.4 K, marking a departure from integer quantization. Subsequent experiments confirmed the fractional nature of the quasiparticles through measurements. In 1997, researchers at , including Roberto de Picciotto, demonstrated the existence of quasiparticles with charge e/3 at \nu = 1/3 by observing the crossover from to Johnson-Nyquist noise in a GaAs sample, where the effective charge was extracted from the noise power as a function of bias voltage and temperature. Even-denominator fractional states, such as at \nu = 5/2, were identified in the half-filled spin-up Landau level of GaAs quantum wells, indicating paired states or non-Abelian anyons. These were first observed in 1987 by Richard Willett and collaborators, who measured quantized Hall resistance at R_{xy} = (2/5) h / e^2 in samples with electron densities around $2.5 \times 10^{11} cm^{-2} and fields up to 12 T, highlighting the role of electron interactions in the second Landau level. Recent advances in have provided direct evidence for braiding statistics in fractional quantum Hall states. In 2023, experiments using Fabry-Pérot interferometers in GaAs devices at \nu = 2/5 revealed phase shifts consistent with Abelian braiding, as the patterns exhibited e^{i 4\pi /5} statistical phases upon encirclement, observed through conductance oscillations as a function of gate voltage and .

Applications and Metrology

Resistance Standards

The quantum Hall effect (QHE) provides an intrinsic for electrical , realized through the quantized Hall resistance R_H = \nu^{-1} R_K, where \nu is the filling factor and R_K = h/e^2 is the von Klitzing constant. With the 2019 redefinition of the SI units, R_K was fixed exactly at 25 812.807 Ω using the defined values of the h and e, eliminating the previous conventional value and its associated uncertainty of about 2 parts in 10^7. This makes the QHE-based independent of material properties or artifact standards, relying solely on fundamental constants for reproducibility. Practical realizations employ cryogenic setups to achieve the low temperatures and high magnetic fields required for observing multiple quantized plateaus. Traditional systems use (GaAs)/aluminum gallium arsenide (AlGaAs) heterostructures cooled to around 1.5–4.2 in baths with perpendicular magnetic fields of 5–12 T, enabling integer filling factors up to \nu = 10 or more for broad resistance ranges. More recent advancements incorporate epitaxial or exfoliated devices, which offer wider plateau widths and operate effectively at slightly higher temperatures (up to 5 ) and lower fields (around 5–6 T) in table-top cryocoolers or closed-cycle systems, facilitating easier integration into labs. These graphene-based setups, often with molecular doping for carrier stability, support parallel array configurations of multiple Hall bars to scale resistances while maintaining quantization. Recent developments as of 2024–2025 have advanced (QAHE) materials for , enabling quantized resistance standards without external magnetic fields using magnetic topological insulators. These QAHE devices offer potential for simplified, zero-field primary resistance standards with traceable measurements. The accuracy of QHE resistance standards reaches relative precisions of a few parts in 10^{10}, as demonstrated in direct comparisons between graphene and GaAs devices using cryogenic current comparators. This level of precision underpins global ohm calibrations at national metrology institutes, where QHE standards bridge primary realizations to working resistance artifacts like 1 Ω or 100 Ω shunts, ensuring traceability to the SI with minimal uncertainty propagation.

Sensing and Quantum Devices

The quantum Hall effect (QHE) enables high-sensitivity detectors by exploiting the quantized Hall plateaus, which offer a stable, temperature-independent response for precise measurements. These plateaus, observed in two-dimensional systems under strong , allow sensors to detect minute variations with minimal , achieving sensitivities as low as ~80 nT/√Hz at 4.2 K in cryogenic setups. For instance, ultraclean Hall sensors have demonstrated robust performance in multi-tesla fields, where the QHE ensures linearity and low , making them suitable for applications in and . In topological quantum computing, the fractional QHE at filling factor ν=5/2 hosts non-Abelian anyons manifested as Majorana zero modes (MZMs) along the sample edges, providing a platform for fault-tolerant qubits through braiding operations. These MZMs, theoretically arising from Moore-Read Pfaffian pairing of composite fermions, encode quantum information non-locally, rendering it robust against local errors and decoherence. Experimental signatures, such as even-odd conductance peaks in quantum point contacts, support the presence of these modes in high-mobility GaAs heterostructures, paving the way for topological qubits that could outperform conventional superconducting ones in scalability. The fractional states at ν=5/2 exhibit anyonic braiding statistics essential for universal quantum computation. Advances in the 2020s have leveraged 's superior mobility to realize QHE-based prototypes operable at elevated temperatures, overcoming traditional cryogenic limitations. Graphite-gated devices have exhibited quantized Hall conductivity up to 150 , enabling compact sensors for industrial magnetic mapping without cooling. Complementing this, variants in III-V semiconductors and heterostructures have demonstrated dissipationless currents at elevated temperatures, with quantum Hall effects observed up to higher temperatures in InAs/GaSb quantum wells as of 2025, facilitating magnet-free quantum spintronic devices for information processing. Despite these progresses, significant challenges persist in deploying QHE-based quantum devices, particularly regarding and maintenance. Realizing large arrays of MZMs requires ultra-high sample mobilities exceeding 10^7 cm²/V·s and precise control over disorder, while poisoning from the soft topological gap limits times to seconds at best, hindering reliable braiding. Ongoing efforts focus on material engineering, such as hybrid superconductor-semiconductor interfaces, to extend and integrate thousands of qubits for practical computation.

Historical Development

Discovery and Early Experiments

The discovery of the quantum Hall effect began with experiments on two-dimensional electron systems under strong perpendicular magnetic fields, where electrons occupy discrete . In 1980, Klaus von Klitzing investigated magnetotransport in a metal-oxide-semiconductor () at low temperatures and high magnetic fields. He observed that the Shubnikov–de Haas oscillations in the longitudinal resistivity, which typically arise from the filling of , transitioned into well-defined plateaus in the Hall resistivity at integer filling factors ν = 2 and ν = 4. These plateaus corresponded to quantized Hall resistance values of h/(ν e²), with h the and e the . Von Klitzing's measurements were conducted at temperatures below 4 K, specifically around 1.5 K, and magnetic fields up to 15 T, using samples with densities controlled via a gate voltage. The plateaus appeared unexpectedly wide and independent of sample geometry or impurity levels, marking a departure from prior oscillatory behaviors. In 1982, Daniel C. Tsui and Horst L. Störmer extended these studies to higher-mobility samples, observing the . Using modulation-doped GaAs-AlGaAs heterostructures, they reported a quantized Hall plateau at ν = 1/3, with Hall resistance ρ_xy = 3 h/e², accompanied by a pronounced minimum in the longitudinal resistivity ρ_xx. This occurred in the extreme , at even higher magnetic fields around 12 T. These early experiments required cryogenic setups, initially refrigerators for temperatures down to about 0.3 K, but soon advanced to dilution refrigerators achieving milliKelvin ranges (e.g., 20–50 mK) to suppress thermal excitations and minimize disorder scattering from impurities, which would otherwise broaden the and obscure the quantization. The observed plateaus puzzled researchers, as they deviated sharply from the classical prediction of a linearly increasing Hall with , ρ_xy = B/(n e), instead showing constant values over wide field ranges.

Theoretical Advances and Recognition

Following the experimental discovery of the integer quantum Hall effect, developed a theoretical framework in to explain the robustness of the observed quantization in the Hall conductance. In his gauge invariance argument, Laughlin considered a cylindrical for the two-dimensional system and analyzed the effect of adiabatically threading a through the system. He demonstrated that gauge invariance requires the Hall conductance to remain unchanged upon the addition of a flux quantum, leading to precise quantization in units of e^2 / [h](/page/h), where e is the charge and h is Planck's constant; this robustness holds even in the presence of impurities or , as long as the system returns to a unique after the flux insertion. A significant advancement came in 1982 with the work of David J. Thouless, Mahito Kohmoto, M. P. Nightingale, and R. den Nijs (TKNN), who provided a topological interpretation of the integer quantum Hall effect. They showed that the quantized Hall conductance in a periodic potential is given by the Chern number, a topological invariant associated with the filled Bloch bands in the presence of a magnetic field. This integer-valued Chern number, computed as an integral over the of the Berry curvature, directly equals the number of filled and ensures the quantization's topological protection against perturbations. The TKNN framework unified the effect with broader concepts in , highlighting its deep topological origins. These theoretical insights were recognized with prestigious awards. In 1985, Klaus von Klitzing received the for his discovery of the quantized Hall effect, underscoring the experimental foundation that spurred these developments. The , observed in early experiments at filling factors like \nu = 1/3, prompted further theory; in 1998, Daniel C. Tsui and Horst L. Störmer shared the with Laughlin for their discovery of this phenomenon and his theoretical explanation, which extended gauge arguments to fractional states. Building on Laughlin's ideas, J. K. Jain introduced the theory in , offering a unified description of both and fractional quantum Hall effects. By attaching an even number of magnetic flux quanta to each , forming that experience an effective , Jain explained the of fractional filling factors (e.g., \nu = p / (2p \pm 1)) as quantum Hall states of these quasiparticles, where p is an . This approach resolved discrepancies in earlier models and predicted new fractional states, later verified experimentally, providing a powerful for strongly correlated two-dimensional systems.

Topological Invariants

The topological classification of quantum Hall states relies on invariants that characterize the global properties of the electronic wavefunctions in the presence of a , distinguishing these states from conventional band insulators. These invariants, rooted in the geometry of the Bloch bundle over the , provide a robust framework for understanding the quantized Hall even in the absence of perfect . In the integer quantum Hall effect, which serves as the simplest example, the topological invariants manifest as integers that dictate the number of filled contributing to the transport properties. The primary topological invariant is the Chern number C, defined for an isolated filled band as the integral of the F over the two-dimensional : C = \frac{1}{2\pi} \int_{\text{BZ}} F \, d^2 k, where F = \nabla_k \times \mathbf{A} and \mathbf{A} is the . This number is always an integer for gapped insulators, reflecting the topological winding of the wavefunction phase. For a system with multiple filled bands, the total Chern number is the sum over individual band contributions, \sum_n C_n. The Hall conductivity \sigma_{xy} is then directly given by \sigma_{xy} = \frac{e^2}{h} \sum_n C_n, linking the bulk topological property to the quantized . This relation, known as the , holds for non-interacting electrons in a periodic potential under a . A key consequence of these invariants is the bulk-boundary correspondence, which asserts that the Chern number of the bulk predicts the existence and of gapless edge modes. Specifically, for a sample with an open boundary, the number of chiral edge states equals the of the total Chern number, ensuring dissipationless current propagation along the edges in the direction determined by the . This principle explains the robustness of edge in quantum Hall systems and has been rigorously established through microscopic models on lattices. In realistic disordered systems, where translational invariance is broken by impurities or irregularities, the standard TKNN invariants require generalization using tools from non-commutative geometry. Here, the Chern number is reformulated in terms of projections in the of the disordered , yielding a non-commutative analog that remains quantized in the presence of strong as long as a mobility gap persists. This approach, developed through cyclic and the Chern-Connes character, extends the topological protection to experimentally relevant inhomogeneous samples and underpins the precise quantization observed in disordered quantum Hall setups.

Photonic and Relativistic Analogs

The photonic quantum Hall effect (PQHE) arises from engineered structures that mimic the topological properties of electronic systems without requiring external magnetic fields, enabling robust unidirectional light propagation along edges. Seminal theoretical proposals by Haldane and Raghu in 2005 demonstrated how photonic crystals with broken time-reversal symmetry, achieved via nonreciprocal media like gyromagnetic materials, could support chiral edge states analogous to those in the integer quantum Hall effect, with topological protection ensuring backscattering immunity. Experimental realizations followed, such as in 2019 using arrays of coupled ring resonators to observe the photonic anomalous quantum Hall effect, where lattice modulation induces effective magnetic fields, leading to quantized edge modes with Chern numbers up to 2. These platforms, including strained graphene or lattice defects, have advanced topological photonics for applications in lossless waveguides and robust optical devices. Relativistic analogs of the quantum Hall effect emerge in materials hosting massless Dirac fermions, where charge carriers behave like relativistic particles, altering the quantization sequence. In , the linear dispersion near Dirac points and a Berry phase of π result in a quantum Hall effect, with plateaus at filling factors ν = ±2, ±6, ±10, and so on, reflecting the fourfold degeneracy from and . This contrasts with the integer sequence in conventional semiconductors and was first observed experimentally in 2005 at low temperatures and up to 45 T, confirming the relativistic nature through the sequence of plateaus, including a prominent ν=0 plateau. A related phenomenon is the (QAHE), observed in thin films of magnetic topological insulators like Cr-doped (Bi,Sb)₂Te₃ in 2013, exhibiting quantized Hall conductance σ_xy = e²/h at zero due to intrinsic opening a gap in the Dirac surface states, enabling dissipationless edge transport without external fields. In Weyl semimetals, three-dimensional topological materials with Weyl nodes, the —a effect—manifests as an anomalous without net , driven by momentum-space monopoles and Berry curvature. Theoretical frameworks predict a quantized anomalous proportional to the separation of Weyl nodes in momentum space, enabling dissipationless charge transport. Experimental evidence includes negative and transverse in materials like TaAs, attributed to the chiral anomaly pumping charge between Weyl nodes under parallel electric and magnetic fields. Recent experiments in the have extended these analogs to non-electronic waves, realizing robust edge states in acoustic and systems. In , researchers observed a three-dimensional acoustic quantum Hall effect in Weyl acoustic crystals fabricated via , where synthetic gauge fields induce chiral surface modes protected by , with quantized pumping along edges. Similarly, metamaterials using nonreciprocal gyroscopic arrays demonstrated topological edge waves in 2020, emulating quantum Hall physics through broken Newton's third law via active , supporting robust . These analogs leverage topological invariants for backscattering-immune propagation in classical wave systems.

References

  1. [1]
    [PDF] THE QUANTIZED HALL EFFECT - Nobel lecture, December 9, 1985
    The pioneering work by Fowler, Fang, Howard and. Stiles [1] has shown that new quantum phenomena become visible if the electrons of a conductor are confined ...
  2. [2]
  3. [3]
    Nobel Prize in Physics 1985
    ### Summary of the Quantum Hall Effect from the 1985 Nobel Prize
  4. [4]
    Nobel Focus: Current for a Small Charge - PHYSICS - APS.org
    Oct 16, 1998 · The integer quantum Hall effect had been explained by assuming independent electrons that did not interact with one another, but the FQH effect ...
  5. [5]
    The Nobel Prize in Physics 1998 - NobelPrize.org
    The Nobel Prize in Physics 1998 was awarded jointly to Robert B. Laughlin, Horst L. Störmer and Daniel C. Tsui for their discovery of a new form of quantum ...
  6. [6]
    The quantum Hall effect continues to reveal its secrets to ... - Nature
    Jul 29, 2020 · Because of this, the quantum Hall effect is used to confirm the accuracy of the ohm, the unit of electrical resistance. Von Klitzing received ...
  7. [7]
    On a New Action of the Magnet on Electric Currents - jstor
    On a New Action of the JIa19flet onz Electric Currents. BY E. H. HALL, Fellow of the Johns Hopkins University. SOMETIME during the last University year, wlhile ...
  8. [8]
    [PDF] The Quantum Hall Effect - DAMTP
    As the title suggests, this book focuses on the composite fermion approach as a lens through which to view all aspects of the quantum Hall effect.
  9. [9]
    High-mobility capacitively-induced two-dimensional electrons in a ...
    Feb 11, 2016 · We demonstrate that the device hosts a 2DEG with a wide tunable density range with limited mobility degradation. We observe commensurability ...
  10. [10]
    Cyclotron resonance in the high mobility GaAs/AlGaAs 2D electron ...
    Feb 20, 2019 · The angular frequency of this circular motion ω = 2πf = qB/m* is independent of the radius of the circular orbit and the electron velocity. Thus ...
  11. [11]
    Recent progress on quantum Hall effect in unconventional material ...
    Jan 13, 2025 · This requires the material to be very clean with high mobility (typically over 10 000 cm2/V s) and a long electron mean free path. A common ...
  12. [12]
    Electronic properties of two-dimensional systems | Rev. Mod. Phys.
    Apr 1, 1982 · The review covers electronic properties of 2D systems, including energy levels, transport, and optical properties, especially at the (100)  ...
  13. [13]
    Quantized Hall conductivity in two dimensions | Phys. Rev. B
    It is shown that the quantization of the Hall conductivity of two-dimensional metals which has been observed recently by Klitzing, Dorda, and Pepper and by ...
  14. [14]
    Exact density of states of the lowest Landau level in a spin
    Feb 15, 1997 · The density of states of this system illustrates the interplay between the Zeeman splitting of Landau levels and the disorder-induced broadening ...
  15. [15]
    von Klitzing constant - CODATA Value
    Concise form, 25 812.807 45... Omega ; Click here for correlation coefficient of this constant with other constants ; Source: 2022 CODATA
  16. [16]
    The quantized Hall effect | Rev. Mod. Phys.
    Jul 1, 1986 · The quantized Hall effect. Klaus von Klitzing. Klaus von Klitzing Max-Planck-Institut für Festkörperforschung, D-7000 Stuttgart 80, West Germany.
  17. [17]
    [PDF] 2. The Integer Quantum Hall Effect - DAMTP
    The TKNN formula is the statement that the Hall conductivity is a topological invariant of the system. It's important because it goes some way to explaining the ...
  18. [18]
    Quantized Hall Conductance in a Two-Dimensional Periodic Potential
    The Hall conductance of a two-dimensional electron gas has been studied in a uniform magnetic field and a periodic substrate potential.
  19. [19]
    Quantized Hall conductance, current-carrying edge states, and the ...
    Quantized Hall conductance, current-carrying edge states, and the existence of extended states in a two-dimensional disordered potential.Missing: TKNN | Show results with:TKNN
  20. [20]
    Absence of backscattering in the quantum Hall effect in multiprobe ...
    We find that the quantum Hall effect occurs only if the sample exhibits at least two sets of equilibrated edge states which do not interact via elastic or ...
  21. [21]
    Composite-fermion approach for the fractional quantum Hall effect
    Jul 10, 1989 · In this paper a new possible approach for understanding the fractional quantum Hall effect is presented. ... J. K. Jain (to be published).Missing: seminal | Show results with:seminal
  22. [22]
    Fractional Statistics and the Quantum Hall Effect | Phys. Rev. Lett.
    Aug 13, 1984 · The statistics of quasiparticles entering the quantum Hall effect are deduced from the adiabatic theorem. These excitations are found to obey fractional ...
  23. [23]
    Ampere: The Quantum Metrology Triangle | NIST
    May 15, 2018 · The quantum Hall effect is the world standard for resistance calibration. ... In the 2019 SI redefinition, however, that changed immediately.
  24. [24]
    [PDF] SI Brochure - 9th ed./version 3.02 - BIPM
    May 20, 2019 · that a laboratory reference standard of resistance based on the quantum Hall effect would be stable and reproducible,. • that a detailed ...<|control11|><|separator|>
  25. [25]
    Essay: Quantum Hall Effect and the New International System of Units
    May 20, 2019 · The quantum Hall effect could similarly be used as a resistance standard more stable than any wire resistor. Indeed, a conventional value for ...
  26. [26]
    Accurate graphene quantum Hall arrays for the new International ...
    Nov 14, 2022 · Here we present experiments on quantization accuracy of a 236-element quantum Hall array (QHA), demonstrating RK/236 ≈ 109 Ω with 0.2 part-per- ...
  27. [27]
    Quantum Hall resistance standards from graphene grown ... - Nature
    Apr 20, 2015 · Our results show that graphene-based QHRS can replace their GaAs counterparts by operating in as-convenient cryomagnetic conditions, but over an ...
  28. [28]
    Graphene-based quantum Hall effect metrology | MRS Bulletin
    Nov 23, 2012 · This article reports on metrological investigations of the QHE in graphene, with accuracy down to 10−10, demonstrating that a quantum resistance ...
  29. [29]
    The Quantum Hall Effect in the Era of the New SI - PMC - NIH
    In the context of the quantum Hall effect (QHE), the von Klitzing constant underwent a change from its conventional value set in 1990 (RK-90 = 25 812.807 Ω) to ...
  30. [30]
    Magnetic field detection limits for ultraclean graphene Hall sensors
    Aug 20, 2020 · Hall-effect sensors are attractive for a variety of magnetic field sensing applications ranging from position detection in robotics and tracking ...
  31. [31]
    Majorana zero modes and topological quantum computation - Nature
    Oct 27, 2015 · This scenario is believed to be realized in the ν=5/2 fractional quantum Hall states, where charge-e/4 excitations are hypothesised to support ...
  32. [32]
    High-Temperature Quantum Hall Effect in Graphite-Gated Graphene ...
    These findings show that graphite-gated high-mobility graphene devices hold great potential for high-sensitivity Hall sensors and resistance metrology ...Missing: 2020s | Show results with:2020s
  33. [33]
    Quantum spin Hall effect in III-V semiconductors at elevated ...
    Oct 24, 2025 · The quantum spin Hall effect (QSHE), a hallmark of topological insulators, enables dissipationless, spin-polarized edge transport and has ...
  34. [34]
    New Method for High-Accuracy Determination of the Fine-Structure ...
    The quantum Hall effect, discovered unexpectedly 35 years ago, is now the basis for defining the unit of electrical resistance.
  35. [35]
    Two-Dimensional Magnetotransport in the Extreme Quantum Limit
    May 31, 1982 · Two-Dimensional Magnetotransport in the Extreme Quantum Limit. D. C. Tsui*,†, H. L. Stormer*, and A. C. Gossard.
  36. [36]
    Quantized Hall conductance as a topological invariant | Phys. Rev. B
    Mar 15, 1985 · The new formulation generalizes the earlier result by Thouless, Kohmoto, Nightingale, and den Nijs to the situation where many-body interaction ...Missing: PRL | Show results with:PRL
  37. [37]
    Chern number and edge states in the integer quantum Hall effect
    Nov 29, 1993 · One is the Thouless–Kohmoto–Nightingale–den Nijs (TKNN) integer in the infinite system and the other is a winding number of the edge state. In ...Missing: paper | Show results with:paper
  38. [38]
    Disordered Topological Insulators: A Non-Commutative Geometry ...
    Aug 6, 2025 · This review deals with strongly disordered topological insulators and covers some recent applications of a well established analytic theory ...
  39. [39]
    Possible Realization of Directional Optical Waveguides in Photonic ...
    We show how, in principle, to construct analogs of quantum Hall edge states in “photonic crystals” made with nonreciprocal (Faraday-effect) media.Missing: paper | Show results with:paper
  40. [40]
    Photonic Anomalous Quantum Hall Effect | Phys. Rev. Lett.
    We experimentally realize a photonic analogue of the anomalous quantum Hall insulator using a two-dimensional (2D) array of coupled ring resonators.
  41. [41]
    [2208.00605] Photonic quantum Hall effects - arXiv
    Aug 1, 2022 · This article reviews the development of photonic analogues of quantum Hall effects, which have given rise to broad interest in topological phenomena in ...
  42. [42]
    Non-Newtonian Topological Mechanical Metamaterials Using ...
    Dec 18, 2020 · We introduce a method to design topological mechanical metamaterials that are not constrained by Newtonian dynamics.