Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2021 May 17;118(21):e2100736118. doi: 10.1073/pnas.2100736118

Terahertz detection based on nonlinear Hall effect without magnetic field

Yang Zhang a, Liang Fu a,1
PMCID: PMC8166175  PMID: 34001618

Significance

Despite the enormous potential of terahertz technology, fast, sensitive, and broadband terahertz detection at room temperature has been a challenge. We propose a method for long-wavelength photodetection using topological quantum materials without invoking any diode or junction region. The inherently quadratic relation between transverse current and input voltage at zero magnetic field is used to rectify the incident terahertz or infrared electric field into a direct current. We identify the Weyl semimetal NbP and ferroelectric semiconductor GeTe for terahertz/infrared photodetection with large current responsivity without external bias.

Keywords: topological materials, terahertz, infrared, Berry curvature, nonlinear Hall effect

Abstract

We propose a method for broadband long-wavelength photodetection using the nonlinear Hall effect in noncentrosymmetric quantum materials. The inherently quadratic relation between transverse current and input voltage at zero magnetic field is used to rectify the incident terahertz or infrared electric field into a direct current, without invoking any diode. Our photodetector operates at zero external bias with fast response speed and has zero threshold voltage. Remarkably, the intrinsic current responsivity due to the Berry curvature mechanism is a material property independent of the incident frequency or the scattering rate, which can be evaluated from first-principles electronic structure calculations. We identify the Weyl semimetal NbP and ferroelectric semiconductor GeTe for terahertz/infrared photodetection with large current responsivity without external bias.


Quantum material is one of the most important fields in condensed-matter physics. While at a fundamental level the properties of all materials are governed by quantum mechanics, quantum materials are distinctive in that they exhibit novel electrical and optical properties originating from the quantum nature of electron wavefunction. The complexity and richness of quantum wavefunction underlie new phases of matter and emergent properties, as revealed through the lens of geometry (1), topology (2), and entanglement (3). The rise of quantum materials not only advances our fundamental understanding of solids but also opens exciting opportunities for inventing new technologies.

Terahertz technology is a fast-growing field with wide-ranging applications (4, 5). Since terahertz waves can transmit through many materials, terahertz spectroscopy and imaging can be used to identify concealed objects for security check, detect chemical composition and material defects for quality control, and examine biological tissues in medical imaging. Terahertz sensors can help autonomous vehicles to identify remote objects under foggy conditions. Terahertz-band communication is regarded as a key enabling technology for the next-generation wireless networks. Despite the enormous potential, fast, sensitive, and broadband terahertz detection at room temperature has been a technology challenge. Conventional terahertz detectors, such as Golay cells, bolometers, and pyroelectric sensors, rely on thermal response and therefore usually require cryogenic cooling and have slow operation speed. Alternatively, Schottky diodes in conjunction with antennas are used to convert high-frequency electric field signals to direct current (DC). However, they typically have cutoff frequencies below 1 THz.

In this work, we propose a method of terahertz detection based on the intrinsic nonlinear Hall effect in quantum materials. This effect was predicted (6) and observed (7, 8) in inversion-breaking crystals without the need of magnetic field, where an applied electric field induces a transverse current to the second order (9, 10). Our idea here is to use this quadratic transverse current–voltage characteristic of quantum materials to rectify and detect terahertz radiation directly, without invoking any junction region. This “Hall rectifier” (11) has a number of important advantages. It is an intrinsic full-wave rectifier with a large attainable responsivity at zero bias and a high cutoff frequency that ranges from subterahertz to tens of terahertz. It features a very short response time down to ps. Due to its simple device architecture, Hall rectifiers can be easily fabricated, integrated on chip, and potentially mass produced.

Remarkably, we show that the current responsivity due to the intrinsic second-order Hall effect is independent of the frequency and intensity of incident radiation, and it is an intrinsic material property determined by electronic band structure and the Berry curvature dipole (6). This finding enables us to predict the right materials for high-sensitivity terahertz detection using first-principles calculations. We identify a number of topological semimetals with anisotropic energy dispersion, in which the intrinsic current responsivity can reach 1 A/W without external bias and over a very broad range of frequency.

Unlike photogalvanic effects at optical frequencies due to interband transitions (1226), the photoresponse of metals and degenerate semiconductors in the terahertz frequency range is usually governed by an intraband process involving electrons near the Fermi surface only. Importantly, since the dispersion relation of the Bloch electron in nonmagnetic materials necessarily satisfies ϵk=ϵk due to time-reversal symmetry, the second-order response of noncentrosymmetric materials is an inherently quantum-mechanical effect arising from the inversion asymmetry of the electron wavefunction within the unit cell. Therefore, quantum materials having complex Bloch wavefunctions, such as topological surface states and inversion-breaking topological metals, generally exhibit a large second-order response.

We start with a brief review of the nonlinear Hall effect due to the Berry curvature dipole (6). As a type of second-order response, this effect can be viewed as a current-induced anomalous Hall effect. The anomalous current due to Berry curvature Ωk is given by

jA=e2ddk(2π)df(k)E×Ωk, [1]

where E is the applied electric field and f(k) is the distribution function in the presence of the field. To first order in E, the anomalous current is obtained by replacing f with the equilibrium Fermi–Dirac distribution f0(k)=1/(eβ(ϵkμ)+1). In time-reversal-invariant and inversion-breaking systems, Berry curvature can be nonzero, but the condition Ωk=Ωk leads to zero linear-response anomalous current due to the cancellation of contributions from Bloch electrons at ±k.

On the other hand, the anomalous current can be nonzero when we extend the analysis to the second order in E. In the current-carrying state, the distribution function f differs from equilibrium and, to first order in E, is given by

f(k)=f0(k)+eEτ(kf0), [2]

where τ is the scattering rate in Drude transport. Then, the anomalous current at second order is obtained as

jaA=ϵabλe3τBλc2EbEcχabcEbEc, [3]

where the nonlinear Hall conductivity χabc is a third-rank tensor that is antisymmetric in indexes a,b. Bλc is the Berry curvature dipole, an intrinsic property of Bloch electron wavefunctions defined by ref. 6:

Bλc=ddk(2π)df0(k)Ωλkc. [4]

B is allowed by symmetry in noncentrosymmetric crystals and surfaces having certain point groups (see below). For example, in materials with a polar axis, large B has been found due to the presence of tilted Dirac or Weyl points in the band structure (6, 2735). When B is nonzero and in the presence of an applied electric field, the imbalance in the probability of occupation f(k)f(k) leads to a net Berry curvature, which then induces a second-order anomalous current in the transverse direction.

The inherently quadratic relation between the transverse current and input voltage due to the nonlinear Hall effect is ideal for rectifying alternating current (AC) input E(t)=Ecos(ωt), including both the positive and negative half cycles. At very low frequency ω1/τ, the time-dependent transverse current simply follows the AC voltage, jA(t)=jA|E(t)cos2(ωt), which is a sum of DC and 2ω components with equal weight. Thus, the DC output j0, or the time average of jA(t) over a full cycle, is given by j0=12jA=12χEE. Unlike diodes, our Hall rectifier does not involve any junction or barrier structure, but relies on the inherent nonlinearity of a homogeneous quantum material. Since the carriers do not need to overcome a barrier to conduct a current, the Hall rectifier has a zero-threshold voltage and operates without the need of a finite bias, which is ideally suited for detecting small signals in terahertz frequency. Since both positive and negative voltages generate transverse current in the same direction, dark current is eliminated.

Provided that the incident frequency is well below the threshold for interband transition, the second-order direct current due to the Berry curvature mechanism shows a Drude-like response (6, 9, 10)

ja0=11+ω2τ2χabc2EbEc, [5]

where the frequency dependence comes from the change of distribution function to the first order of the oscillating electric field. Note that for time-reversal-invariant systems, the DC photocurrent effect and, more generally, the linear photogalvanic effect vanishes in the absence of dissipation τ10, because time-reversal symmetry forbids a direct current (37). At frequencies ωτ1, the photocurrent j0 from the intraband process decreases as 1/ω2. However, in reality, even in the case of good metals, the scattering rate at room temperature is at least a few tens of terahertz. Therefore, the nonlinear Hall effect of a homogeneous quantum material can be used as a replacement of diodes to rectify incident waves in a very broad range of frequency up to at least 10 THz, and the output direct current is linearly proportional to the incident power.

An important parameter for terahertz detectors is current responsivity, which measures the electrical output per optical input. In our device shown in Fig. 1, the antennas collect the incident radiation and create a strongly enhanced terahertz field in the active region of the Hall rectifier, which generates the photoresponse current in the transverse direction. The power absorbed by the Hall rectifier P=σab(ω)EaEbS, where S=LW is the area of the Hall rectifier with length L and width W. σ is the linear-response AC conductivity given by

σab(ω)=e2Dab2τ(1+ω2τ2) [6]

within Drude transport theory, where the Drude weight D is defined by

Dab=ddk(2π)dEkaEkbf0E. [7]

The output current per absorbed power defines the current responsivity of our Hall rectifier:

R=IHP=1Wj0σabEaEb. [8]

Note that R is inversely proportional to the width of the Hall rectifier, so it is advantageous to use a narrow channel.

Fig. 1.

Fig. 1.

Terahertz photodetection based on the nonlinear Hall effect in quantum materials. (A) Schematic diagram of an antenna-coupled topological semimetal or ferroelectric semiconductor terahertz detector element with substrate resonant cavity length of L. The thickness of substrate is given by the resonant condition L=2n+14εrλ0(n=0,1,2) with εr as the dielectric constant of the Si substrate and λ0 as the wavelength of incoming terahertz radiation in free space (36). (B) The oscillating terahertz field enhanced by a bowtie antenna induces a direct current in the transverse direction via the intrinsic second-order response of the quantum material itself. (C) Current–voltage characteristic of the nonlinear Hall rectifier. The quadratic relation enables full-wave rectification of AC input at zero external bias, with zero threshold voltage and zero dark current.

Due to the quadratic dependence of transverse current j0 on the terahertz field, R is inherently independent of incident power; i.e., our terahertz detector exhibits ideal linearity over a large range of optical power. Moreover, since both the nonlinear Hall conductivity and linear response conductivity exhibit Drude behavior with the same τ/(1+ω2τ2) dependence, R is independent of frequency and scattering rate, regardless of the value of ωτ. Therefore, the current responsivity due to the intrinsic nonlinear Hall effect is entirely determined by the Berry curvature dipole B and the Drude weight D. B is a property of Bloch electron wavefunction, and D is a property of energy dispersion. This is an important result of our work. As both B and D are material properties attainable from first-principles electronic structure calculations, the intrinsic current responsivity R of quantum materials can be predicted without any tuning parameter. This enables computational materials design of long-wavelength photodetectors.

Concerning practical application, the responsivity of a photodetector measures the electrical output per incident optical power, which is reduced from the intrinsic responsivity R defined above by the absorption coefficient, i.e., the ratio between absorbed and incident power. Resonant optical cavity and antenna with impedance match (Fig.1) can be used to maximize the absorption of incident terahertz radiation and optimize the photodetector responsivity toward the fundamental limit R set by the intrinsic second-order nonlinearity of quantum materials.

To search for quantum materials suitable for the Hall rectifier, we note that a large Berry curvature dipole appears in the vicinity of band crossings with anisotropic energy dispersion. Therefore inversion-breaking Weyl semimetals such as TaAs, TaP, NbP, and NbAs (3841) are good candidates. Our first-principles calculations reveal that among these four materials, NbP has the largest current responsivity (Fig. 2). The NbP family of Weyl semimetals has the point group C4v with a fourfold rotation axis in the z direction and two vertical mirror planes Mx, My. Thus, the only nonzero elements of the Berry curvature dipole tensor are Byx=Bxy. Hence, the photoresponse is largest when the incident electric field lies within the xy plane and the photocurrent along the z direction is measured. In this configuration, the intrinsic current responsivity is given by

R=1WeBxy2Dxx. [9]

Interestingly, we find that large contributions to the Berry curvature dipole in NbP come from regions near two extended loops in momentum space, rather than the vicinity of Weyl points as in TaAs. These loops are located on a Mx-invariant plane and correspond to a nodal line of band crossings that would occur due to band inversion if spin–orbit coupling were turned off. The introduction of spin–orbit coupling leads to strong momentum-dependent band hybridization and therefore creates anisotropic energy dispersion and a large Berry curvature dipole near the Fermi level. As a result, a large current responsivity R is found for a range of Fermi energy within ±20 meV. Our calculation shows that the responsivity of the NbP-based long-wavelength photodetector at zero external bias can reach about 0.3 A/W for a realistic device width W=1μm. Reducing the width to W=0.1μm leads to R=3 A/W.

Fig. 2.

Fig. 2.

(A) Lattice structure of NbP. (B) Band structure of NbP, which shows a narrow band inversion around Γ1. (C) Three-dimensional momentum resolved Berry curvature dipole. (D) Current responsivity as a function of chemical potential.

We further calculate the current responsivity for other type I Weyl semimetal TaAs, TaP, and NbAs with point group C4v (NbAs shows a similar magnitude to NbP; see SI Appendix for details) and type II Weyl semimetal MoTe2 and WTe2 with point group C2v. In general, type I Weyl semimetals have a larger current responsivity because of the reduced Drude weight from small electron (hole) pockets.

Besides the topological semimetal, ferroelectric semiconductors are also promising candidates where the current responsivity can be optimized by doping. The ferroelectric XTe (X = Ge, Sn) monochalcogenides share rhombohedrally distorted rock salt structure (space group R3m). In particular, GeTe has a remarkably high Curie temperature up to 700 K. Here we calculate the Berry curvature dipole and current responsivity in p-doped GeTe. While the global bandgap of GeTe is 0.32 eV, the Berry curvature dipole Bxy reaches 0.5, which is much larger than that of the Rashba semiconductor BiTeI at the topological phase transition (30). As shown in Fig. 3, the resulting current responsivity of GeTe is sizable over a wide range of doping, which is attractive for room-temperature terahertz photodetection.

Fig. 3.

Fig. 3.

(A) Schematic illustration of rhombohedral distorted rock salt crystal structure of ferroelectric semiconductor GeTe with polarization along the c axis. (B) Fully relativistic band structure of GeTe along high-symmetry lines. The dotted line stands for the Fermi level at −0.3 eV. (C) Chemical potential-dependent Berry curvature dipole Bxy and Drude weight Dxx (in the unit of Ve(Ωcm)1). (D) current responsivity as a function of chemical potential in the valence band side of GeTe.

All of the materials identified here possess a polar axis. To use these materials for terahertz detection, it is important to consider the orientation of the polar axis. When the polar axis lies within the plane, terahertz radiation at normal incidence will produce a photocurrent along the polar axis. When the polar axis is out of plane (z axis), there is no in-plane photocurrent with normal incidence. However, the nonlinear Hall effect from the Berry curvature dipole results in χxzx=χzxx and χyzy=χzyy, so that terahertz radiation at oblique incidence will produce an in-plane photocurrent, which enables photodetection in the device configuration shown in Fig. 1.

In Fig. 4, we compare the performance of our nonlinear Hall long-wavelength photodetector with that of other detectors and rectifiers. To this end, we plot their current responsivity versus operating frequency (4253). In the long-wavelength region below 37 THz, the current responsivity is typically limited to less than 1 A/W and varies dramatically with operating frequency. As the Drude half-width of NbP at room temperature determined from optical conductivity is around 30 meV (54), we expect our nonlinear Hall photodetector can operate up to 15 THz with a large current responsivity reaching 3 A/W. At higher frequency, the interband linear photogalvanic effect (LGPE) starts to dominate in the TaAs family of Weyl semimetals (21, 55). As shown in SI Appendix, we find a large current responsivity due to LGPE for TaAs, NbP, and GeTe, which provides a promising platform for infrared photodetection.

Fig. 4.

Fig. 4.

Comparison of current responsivity to that in previous studies in the frequency range 106 to 1015 Hz. The reference data points are adopted and updated from a terahertz detector review article (42); labels 1 to 11 correspond to a series of works on two-dimensional materials and topological semimetals (4353).

Another important figure of merit for photodetection is the noise-equivalent power (NEP), defined as the noise power density over the responsivity. At room temperature, thermal noise is the dominant noise source compared to the shot noise from discrete incident photons. We compute the NEP from the root-mean square of the noise current as NEP=4kBT/r0R, where T is the operating temperature and r0 is the sample resistance. With sample width W=0.1μm and room temperature resistance r0=300Ω (56), the NEP of the NbP device at peak responsivity is estimated to be 2.5 pW/Hz0.5. For comparison, the NEP of the typical Schottky diode-based terahertz detector is around 100 pW/Hz0.5 at the subterahertz (0.11 THz) range and quickly increases at the higher frequency.

In addition to the Berry curvature dipole, other mechanisms including skew scattering and side jump also contribute to the nonlinear Hall effect (10, 11, 5763). For crystals with certain point groups such as C3v, although the Berry curvature dipole cannot contribute to the in-plane nonlinear Hall effect, theory predicts that second-order nonlinear transport in diffusive systems arises from skew scattering and side jump induced by the inherent chirality of Bloch electron wavefunction (11). Moreover, in the ballistic transport regime, nonlinear Hall conductance is entirely determined by the integral of Berry curvature over half of the Fermi surface (64), which is also an intrinsic property of Bloch electron wavefunction. Regardless of the microscopic origins, the inherent second-order nonlinearity of quantum materials can be utilized for long-wavelength photodetection by rectifying oscillating electric field into direct current.

On the experimental side, a second-order response was observed in recent electrical transport (65) and terahertz photocurrent experiments (66, 67) on topological insulators. This effect arises from the Dirac surface states with hexagonal warping (68), which enables second-order response due to Berry curvature and skew scattering. The dependence of photocurrent on incident frequency between 0.5 and 4 THz is consistent with Drude behavior expected from intraband processes. Very recently, a nonlinear Hall effect at room temperature was observed in type II Weyl semimetal TaIrTe4 and enabled radiofrequency rectification (69). Moreover, room temperature terahertz detection and imaging have been achieved using the second-order nonlinear response of topological surface states in Dirac semimetal PdTe2 (49). The photocurrent as a function of incident power displays excellent linearity. Remarkably, the current responsivity without bias reaches as large as 0.2 A/W. These exciting results on many fronts encourage future development of terahertz/infrared technology based on Hall rectifiers.

Methods

For TaAs, TaP, NbAs, NbP, MoTe2, WTe2, and GeTe, we first perform density functional theory (DFT) Bloch wave functions from the full-potential local-orbital (FPLO) program (70) within the generalized gradient approximation (GGA) (71). By projecting the Bloch wavefunction to localized Wannier functions, we construct the tight-binding model with 40 bands from d and s orbitals for transition metal Nb (Ta) and s and p orbitals for pnicogen P (As). For MoTe2 and WTe2, the tight-binding model is constructed with 88 bands from d orbitals for Mo (W) and p orbitals for Te. For ferroelectric GeTe, the tight-binding model is constructed with 52 bands from 4s,4p,5s,5p,4d orbitals for Ge and 5s,5p,6s,5d,6p orbitals for Te as the tight-binding model with reduced basis has a large error fitting to the DFT band structure. The Berry curvature dipole and Drude conductivity integration are both done with a 500×500×500 momentum space mesh.

Supplementary Material

Supplementary File

Acknowledgments

This work was supported by the US Army Research Laboratory and the US Army Research Office through the Institute for Soldier Nanotechnologies, under Collaborative Agreement W911NF-18-2-0048. Y.Z. and L.F. were supported in part by a Simons Investigator Award from the Simons Foundation.

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2100736118/-/DCSupplemental.

Data Availability

All study data are included in this article and/or SI Appendix.

References

  • 1.Nagaosa N., Sinova J., Onoda S., MacDonald A. H., Ong N. P., Anomalous Hall effect. Rev. Mod. Phys. 82, 1539 (2010). [Google Scholar]
  • 2.Hasan M. Z., Kane C. L., Colloquium: Topological insulators. Rev. Mod. Phys. 82, 3045 (2010). [Google Scholar]
  • 3.Savary L., Balents L., Quantum spin liquids: A review. Rep. Prog. Phys. 80, 016502 (2016). [DOI] [PubMed] [Google Scholar]
  • 4.Tonouchi M., Cutting-edge terahertz technology. Nat. Photonics 1, 97–105 (2007). [Google Scholar]
  • 5.Sizov F., Rogalski A., THz detectors. Prog. Quant. Electron. 34, 278–347 (2010). [Google Scholar]
  • 6.Sodemann I., Fu L., Quantum nonlinear Hall effect induced by Berry curvature dipole in time-reversal invariant materials. Phys. Rev. Lett. 115, 216806 (2015). [DOI] [PubMed] [Google Scholar]
  • 7.Ma Q., et al. , Observation of the nonlinear Hall effect under time-reversal-symmetric conditions. Nature 565, 337–342 (2019). [DOI] [PubMed] [Google Scholar]
  • 8.Kang K., Li T., Sohn E., Shan J., Mak K. F., Observation of the nonlinear anomalous Hall effect in 2d wte2. Nonlinear anomalous Hall effect in few-layer WTe2. Nat. Mater. 18, 324–328 (2019). [DOI] [PubMed] [Google Scholar]
  • 9.Moore J. E., Orenstein J., Confinement-induced Berry phase and helicity-dependent photocurrents. Phys. Rev. Lett. 105, 026805 (2010). [DOI] [PubMed] [Google Scholar]
  • 10.Deyo E., Golub L., Ivchenko E., Spivak B., Semiclassical theory of the photogalvanic effect in non-centrosymmetric systems. arXiv [Preprint] (2009). https://arxiv.org/abs/0904.1917 (Accessed 14 March 2021).
  • 11.Isobe H., Xu S. Y., Fu L., High-frequency rectification via chiral Bloch electrons. Sci. Adv. 6, eaay2497 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Sipe J., Shkrebtii A., Second-order optical response in semiconductors. Phys. Rev. B 61, 5337 (2000). [DOI] [PubMed] [Google Scholar]
  • 13.de Juan F., Grushin A. G., Morimoto T., Moore J. E., Quantized circular photogalvanic effect in Weyl semimetals. Nat. Commun. 8, 1–7 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Morimoto T., Nagaosa N., Topological nature of nonlinear optical effects in solids. Sci. Adv. 2, e1501524 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Chan C. K., Lee P. A., Burch K. S., Han J. H., Ran Y., When chiral photons meet chiral fermions: Photoinduced anomalous Hall effects in Weyl semimetals. Phys. Rev. Lett. 116, 026805 (2016). [DOI] [PubMed] [Google Scholar]
  • 16.Ishizuka H., Hayata T., Ueda M., Nagaosa N., Emergent electromagnetic induction and adiabatic charge pumping in noncentrosymmetric Weyl semimetals. Phys. Rev. Lett. 117, 216601 (2016). [DOI] [PubMed] [Google Scholar]
  • 17.Yang X., Burch K., Ran Y., Divergent bulk photovoltaic effect in Weyl semimetals. arXiv [Preprint] (2017). https://arxiv.org/abs/1712.09363 (Accessed 14 March 2021).
  • 18.Young S. M., Rappe A. M., First principles calculation of the shift current photovoltaic effect in ferroelectrics. Phys. Rev. Lett. 109, 116601 (2012). [DOI] [PubMed] [Google Scholar]
  • 19.Parker D. E., Morimoto T., Orenstein J., Moore J. E., Diagrammatic approach to nonlinear optical response with application to Weyl semimetals. Phys. Rev. B 99, 045121 (2019). [Google Scholar]
  • 20.Xu S. Y., et al. , Electrically switchable Berry curvature dipole in the monolayer topological insulator wte 2. Nat. Phys. 14, 900–906 (2018). [Google Scholar]
  • 21.Osterhoudt G. B., et al. , Colossal mid-infrared bulk photovoltaic effect in a type-I Weyl semimetal. Nat. Mater. 18, 471–475 (2019). [DOI] [PubMed] [Google Scholar]
  • 22.Rees D., et al. , Helicity-dependent photocurrents in the chiral Weyl semimetal RhSi. Sci. Adv. 6, eaba0509 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Ni Z., et al. , Giant topological longitudinal circular photo-galvanic effect in the chiral multifold semimetal CoSi. Nat. Commun. 12, 154 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Ma J., et al. , Nonlinear photoresponse of type-II Weyl semimetals. Nat. Mater. 18, 476–481 (2019). [DOI] [PubMed] [Google Scholar]
  • 25.Ma Q., et al. , Direct optical detection of Weyl fermion chirality in a topological semimetal. Nat. Phys. 13, 842–847 (2017). [Google Scholar]
  • 26.König E., Xie H. Y., Pesin D., Levchenko A., Photogalvanic effect in Weyl semimetals. Phys. Rev. B 96, 075123 (2017). [Google Scholar]
  • 27.Zhang Y., Sun Y., Yan B., Berry curvature dipole in Weyl semimetal materials: An ab initio study. Phys. Rev. B 97, 041101 (2018). [Google Scholar]
  • 28.Dzsaber S., et al. , Giant spontaneous Hall effect in a nonmagnetic Weyl-Kondo semimetal. arXiv [Preprint] (2018). https://arxiv.org/abs/1811.02819 (Accessed 14 March 2021). [DOI] [PMC free article] [PubMed]
  • 29.Zhang Y., Van Den Brink J., Felser C., Yan B., Electrically tuneable nonlinear anomalous Hall effect in two-dimensional transition-metal dichalcogenides WTe2 and MoTe2. 2D Mater. 5, 044001 (2018). [Google Scholar]
  • 30.Facio J. I., et al. , Strongly enhanced Berry dipole at topological phase transitions in BiTeI. Phys. Rev. Lett. 121, 246403 (2018). [DOI] [PubMed] [Google Scholar]
  • 31.You J. S., Fang S., Xu S. Y., Kaxiras E., Low T., Berry curvature dipole current in the transition metal dichalcogenides family. Phys. Rev. B 98, 121109 (2018). [Google Scholar]
  • 32.Son J., Kim K. H., Ahn Y., Lee H. W., Lee J., Strain engineering of the Berry curvature dipole and valley magnetization in monolayer MoS2. Phys. Rev. Lett. 123, 036806 (2019). [DOI] [PubMed] [Google Scholar]
  • 33.Zhou B. T., Zhang C. P., Law K. T., Highly tunable nonlinear Hall effects induced by spin-orbit couplings in strained polar transition-metal dichalcogenides. Phys. Rev. Appl. 13, 024053 (2020). [Google Scholar]
  • 34.Pantaleon P. A., Low T., Guinea F., Tunable large Berry dipole in strained twisted bilayer graphene. arXiv [Preprint] (2020). https://arxiv.org/abs/2010.11086 (Accessed 14 March 2021).
  • 35.Tsirkin S. S., Puente P. A., Souza I., Gyrotropic effects in trigonal tellurium studied from first principles. Phys. Rev. B 97, 035158 (2018). [Google Scholar]
  • 36.Tu X., et al. , Investigation of antenna-coupled Nb5N6 microbolometer THz detector with substrate resonant cavity. Opt. Express. 26, 8990–8997 (2018). [DOI] [PubMed] [Google Scholar]
  • 37.Morimoto T., Nagaosa N., Nonreciprocal current from electron interactions in noncentrosymmetric crystals: Roles of time reversal symmetry and dissipation. Sci. Rep. 8, 1–16 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Bansil A., Lin H., Das T., Colloquium: Topological band theory. Rev. Mod. Phys. 88, 021004 (2016). [Google Scholar]
  • 39.Hasan M. Z., Xu S. Y., Belopolski I., Huang S. M., Discovery of Weyl fermion semimetals and topological Fermi arc states. Annu. Rev. Condens. Matter Phys. 8, 289–309 (2017). [Google Scholar]
  • 40.Weng H., Fang C., Fang Z., Bernevig B. A., Dai X., Weyl semimetal phase in noncentrosymmetric transition-metal monophosphides. Phys. Rev. X 5, 011029 (2015). [Google Scholar]
  • 41.Armitage N., Mele E., Vishwanath A., Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 90, 015001 (2018). [Google Scholar]
  • 42.Rogalski A., Kopytko M., Martyniuk P., Two-dimensional infrared and terahertz detectors: Outlook and status. Appl. Phys. Rev. 6, 021316 (2019). [Google Scholar]
  • 43.Yao Y., et al. , High-responsivity mid-infrared graphene detectors with antenna-enhanced photocarrier generation and collection. Nano Lett. 14, 3749–3754 (2014). [DOI] [PubMed] [Google Scholar]
  • 44.Pospischil A., et al. , Cmos-compatible graphene photodetector covering all optical communication bands. Nat. Photonics 7, 892–896 (2013). [Google Scholar]
  • 45.Xia F., Mueller T., Lin Ym., Valdes-Garcia A., Avouris P., Ultrafast graphene photodetector. Nat. Nanotechnol. 4, 839–843 (2009). [DOI] [PubMed] [Google Scholar]
  • 46.Schall D., et al. , 50 Gbit/s photodetectors based on wafer-scale graphene for integrated silicon photonic communication systems. ACS Photonics 1, 781–784 (2014). [Google Scholar]
  • 47.Gan X., et al. , Chip-integrated ultrafast graphene photodetector with high responsivity. Nat. Photonics 7, 883–887 (2013). [Google Scholar]
  • 48.Cakmakyapan S., Lu P. K., Navabi A., Jarrahi M., Gold-patched graphene nano-stripes for high-responsivity and ultrafast photodetection from the visible to infrared regime. Light Sci. Appl. 7, 1–9 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Guo C., et al. , Anisotropic ultrasensitive PdTe2-based phototransistor for room-temperature long-wavelength detection. Sci. Adv. 6, eabb6500 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Auton G., et al. , Terahertz detection and imaging using graphene ballistic rectifiers. Nano Lett. 17, 7015–7020 (2017). [DOI] [PubMed] [Google Scholar]
  • 51.Viti L., et al. , Plasma-wave terahertz detection mediated by topological insulators surface states. Nano Lett. 16, 80–87 (2016). [DOI] [PubMed] [Google Scholar]
  • 52.Viti L., et al. , Black phosphorus terahertz photodetectors. Adv. Mater. 27, 5567–5572 (2015). [DOI] [PubMed] [Google Scholar]
  • 53.Bandurin D. A., et al. , Resonant terahertz detection using graphene plasmons. Nat. Commun. 9, 5392 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Neubauer D., et al. , Optical conductivity of the Weyl semimetal NbP. Phys. Rev. B 98, 195203 (2018). [Google Scholar]
  • 55.Zhang Y., et al. , Photogalvanic effect in Weyl semimetals from first principles. Phys. Rev. B 97, 241118 (2018). [Google Scholar]
  • 56.Shekhar C., et al. , Extremely large magnetoresistance and ultrahigh mobility in the topological Weyl semimetal candidate NbP. Nat. Phys. 11, 645–649 (2015). [Google Scholar]
  • 57.Du Z., Wang C., Li S., Lu H. Z., Xie X., Disorder-induced nonlinear Hall effect with time-reversal symmetry. Nat. Commun. 10, 1–6 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Du Z., Wang C., Sun H. P., Lu H. Z., Xie X., Quantum theory of the nonlinear Hall effect. arXiv [Preprint] (2020). https://arxiv.org/abs/2004.09742 (Accessed 14 March 2021). [DOI] [PMC free article] [PubMed]
  • 59.Xiao C., Du Z., Niu Q., Theory of nonlinear Hall effects: Modified semiclassics from quantum kinetics. Phys. Rev. B 100, 165422 (2019). [Google Scholar]
  • 60.Xiao C., Zhou H., Niu Q., Scaling parameters in anomalous and nonlinear Hall effects depend on temperature. Phys. Rev. B 100, 161403 (2019). [Google Scholar]
  • 61.Golub L., Ivchenko E., Spivak B., Semiclassical theory of the circular photogalvanic effect in gyrotropic systems. Phys. Rev. B 102, 085202 (2020). [Google Scholar]
  • 62.Matsyshyn O., Sodemann I., Nonlinear Hall acceleration and the quantum rectification sum rule. Phys. Rev. Lett. 123, 246602 (2019). [DOI] [PubMed] [Google Scholar]
  • 63.Rudner M. S., Song J. C., Self-induced Berry flux and spontaneous non-equilibrium magnetism. Nat. Phys. 15, 1017–1021 (2019). [Google Scholar]
  • 64.Papaj M., Fu L., Magnus Hall effect. Phys. Rev. Lett. 123, 216802 (2019). [DOI] [PubMed] [Google Scholar]
  • 65.He P., et al. , Quantum frequency doubling in the topological insulator Bi2Se3. Nat Commun. 12, 698 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Olbrich P., et al. , Room-temperature high-frequency transport of Dirac fermions in epitaxially grown Sb2Te3- and Bi2Te3-based topological insulators. Phys. Rev. Lett. 113, 096601 (2014). [DOI] [PubMed] [Google Scholar]
  • 67.Plank H., et al. , Photon drag effect in (Bi1−xSbx)2Te3 three-dimensional topological insulators. Phys. Rev. B 93, 125434 (2016). [Google Scholar]
  • 68.Fu L., Hexagonal warping effects in the surface states of the topological insulator Bi2Te3. Phys. Rev. Lett. 103, 266801 (2009). [DOI] [PubMed] [Google Scholar]
  • 69.Kumar D., et al. , Room-temperature nonlinear Hall effect and wireless radiofrequency rectification in Weyl semimetal TaIrTe4. Nat. Nanotechnol. 16, 421–425 (2021). [DOI] [PubMed] [Google Scholar]
  • 70.Koepernik K., Eschrig H., Full-potential nonorthogonal local-orbital minimum-basis band-structure scheme. Phys. Rev. B 59, 1743 (1999). [Google Scholar]
  • 71.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File

Data Availability Statement

All study data are included in this article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES