Abstract
Many non-equilibrium phenomena have been discovered or predicted in optically-driven quantum solids1. Examples include light-induced superconductivity2,3 and Floquet-engineered topological phases4–8. These are short lived effects that should lead to measurable changes in electrical transport, which can be characterized using an ultrafast device architecture based on photoconductive switches9. Here, we report the observation of a light-induced anomalous Hall effect in monolayer graphene driven by a femtosecond pulse of circularly polarized light. The dependence of the effect on a gate potential used to tune the Fermi level reveals multiple features that reflect a Floquet-engineered topological band structure4,5, similar to the band structure originally proposed by Haldane10. This includes an approximately 60 meV wide conductance plateau centered at the Dirac point, where a gap of equal magnitude is predicted to open. We find that when the Fermi level lies within this plateau, the estimated anomalous Hall conductance saturates around 1.8±0.4 e2/h.
Optical driving has been proposed as a means to engineer topological properties in topologically trivial systems4–8. One proposal for such a ‘Floquet topological insulator’ is based on breaking time-reversal symmetry in graphene through a coherent interaction with circularly polarized light4. In this theory, the light field drives electrons in circular trajectories through the band structure (Fig. 1a). Close to the Dirac point, these states are predicted to acquire a non-adiabatic Berry phase with each optical cycle, and this contribution is equal and opposite for the upper and lower band. This time-averaged extra phase accumulation amounts to an energy shift that lifts the degeneracy of the Dirac point and opens a topological gap in the effective Floquet band structure (Fig. 1b).
The non-trivial topology of the Floquet bands forming this gap arises from their nonzero Berry curvature distribution4,5, which integrated over the Brillouin zone defines a topological invariant, called the Chern number11–14. Topologically protected transport is predicted to develop if the Fermi level (EF) lies inside the gap, exhibiting an anomalous Hall effect carried by edge states in the absence of an applied magnetic field5,10–15. This corresponds to the formation of a light-induced Chern insulator, equivalent to the phase originally proposed by Haldane10 and distinct from topological phases induced by spin-orbit interaction12–14,16,17. While quantum simulation experiments have validated aspects of this proposal in synthetic physical settings18,19, and Floquet-Bloch states have been detected in a topological insulator20, anomalous Hall currents originating from such a photon-dressed topological band structure have not been observed in a real material.
This proposal is unique because the anomalous Hall effect arises from Berry curvature that is coherently induced by light in a material where none is present in equilibrium. This is in contrast to previous observations of photo-induced anomalous Hall effects in semiconductor quantum wells21, monolayer transition metal dichalcogenides22 or Weyl semimetals23, which originate from Berry curvature intrinsic to the equilibrium band structure due to broken inversion symmetry.
Inducing and detecting anomalous Hall currents in graphene presents multiple experimental challenges. The laser electric field strength required to open an observable topological gap is estimated to be of the order of 107-108 V/m, even at mid-infrared wavelengths where the effect is enhanced4,5,15. Hence, to avoid material damage while still providing sufficient field strength, ultrafast laser pulses must be used. Consequently, the resulting Hall conductance changes are too short-lived to be probed with conventional transport techniques.
In this work, ultrafast anomalous Hall currents were detected on-chip by using a laser-triggered photoconductive switch9. A schematic of our device architecture is shown in Fig. 1c. An exfoliated monolayer graphene flake was transferred onto a doped silicon wafer with an oxide layer and contacted in a four-probe Hall geometry using standard lithography procedures24,25. The metallic leads formed microstrip transmission lines in conjunction with the oxide layer and silicon wafer. These directed ultrafast anomalous Hall currents generated in the graphene to a photoconductive switch for detection. The switch consisted of a resistive amorphous silicon patch that bridged the main transmission line and a probing line. When excited with a visible ultrafast laser pulse, the switch became highly conductive and detected currents flowing in the main transmission line with a time resolution set by the silicon carrier lifetime (less than 1 ps). By adjusting the time delay between the graphene laser drive pulse (pump) and the switch trigger pulse (probe), the temporal profile of ultrafast anomalous Hall currents could be characterized. The amplitude of the detected currents were determined by calibrating the photovoltaic response of the switch using a DC bias field, which is possible in our microstrip geometry because only the fundamental quasi-TEM mode propagates up to THz frequencies.
Graphene was driven using an approximately 500 fs laser pulse at a frequency of 46 THz (so that the photon energy ℏω ≈ 191 meV, wavelength ≈ 6.5 μm). Unless otherwise noted, a peak laser pulse fluence of 0.23 mJ/cm2 was used, corresponding to a peak intensity of 4.3 × 1012 W/m2 and peak electric field strength of 4.0 × 107 V/m (in free space, for circular polarization). The pulses were focused to a spot size of roughly 80 μm (FWHM), ensuring homogeneous illumination of the graphene flake and the contacts. A second ultrafast laser pulse centered at 520 nm was used to operate the photoconductive switch. The device was mounted in a microscopy cryostat designed for high-frequency transport measurements and cooled to a base temperature of 80 K. A global backgate formed by the silicon wafer and the oxide layer controlled the Fermi level (EF) in the graphene flake. The graphene field-effect mobility was measured to be μ ≈ 10,000 cm2/Vs in the vicinity of the Dirac point. The results presented here are from a single device, and consistent results have been obtained using five different devices.
Anomalous Hall currents induced by circularly polarized light are expected to exhibit the following traits: (1) They should be generated in the transverse direction (Ix) with respect to an applied DC voltage bias (Vy). (2) They should reverse polarity upon reversing the light helicity. (3) They should reverse polarity upon reversing Vy, with a linear functional dependence. To probe (1) and (2), we directly detected the difference between currents Ix generated with right- (clockwise from the perspective of the light source) versus left-circular polarization (henceforth referred to as Ix [⟳ - ⟲]), utilizing an optical polarization chopping technique.
Figure 2a displays the measured Ix [⟳ - ⟲] signal as a function of pump-probe time delay for a positive and negative Vy, with the Fermi level gated to the graphene Dirac point (EF = 0). The time-resolved signal exhibits a fast rise time followed by an exponential decay, and reverses polarity upon reversing Vy. From the polarity of the signals, we determine that for right circularly polarized light propagating along +z (in a right-handed coordinate system), a DC field applied to the graphene in the -y direction induces a Hall current in the +x direction.
The measured data in Fig. 2a contains a convolution between the intrinsic dynamics of anomalous Hall currents in the graphene and the response function of the circuit, which includes the contact resistance, graphene capacitance, microstrip impedance and dispersion in the transmission line. As detailed in supplementary S3, we directly calibrated these system parameters to determine the deconvolved current profiles, shown in Fig. 2b. The signals have a duration of approximately 3 ps, which exceeds that of the driving laser pulse (500 fs). We believe that this difference in timescales is mainly caused by the finite size of the graphene, which leads to arrival-time differences of currents generated in different parts of the flake, as well as the sublinear dependence of the Hall current on the drive laser intensity (Fig. 4a).
Having established the presence of ultrafast anomalous Hall currents in graphene, we investigated the functional dependence of Ix [⟳ - ⟲] on Vy at EF = 0. We did so by fixing the pump-probe time delay at the maximum of the Ix [⟳ - ⟲] signal in Fig. 2a, which we refer to as Îx [⟳ - ⟲], and measured the signal amplitude as a function of Vy (Fig. 3a). The data exhibit the expected linear dependence. We also investigated helicity-dependent currents generated in response to a longitudinal voltage bias Vx (Fig. 3b), which are not expected because graphene has D6h point group symmetry26. The data in Fig. 3b confirm that no helicity-dependent longitudinal currents were generated.
We define the peak anomalous Hall conductance of the non-equilibrium state as Gxy = Îg [⟳ - ⟲] / 2Vy, where Îg is the peak of the deconvolved signal in Fig. 2b. Figure 4a displays Gxy as a function of the laser drive pulse fluence, measured for EF = 0. The data show a sublinear dependence that saturates at high fluence. At the highest achievable fluence, we estimate Gxy = (1.8 ± 0.4) e2/h, consistent with recent numerical simulations of optically-driven graphene for our laser pulse parameters27.
To characterize the predicted topological Floquet bands, we varied EF (defined as the chemical potential without light excitation) using the backgate28 and compared the variation in Gxy (Fig. 4b) with calculations of the expected effective band structures based on Floquet theory for our laser pulse parameters (Fig. 4c) (see supplementary S7).
We observed that for low drive fluence (red circles), Gxy was independent of the backgate potential for |EF| ≲ 100 meV ≈ ℏω/2. In this low excitation regime, the Floquet-gap at the Dirac point (Δ0) is too small to be observable. However, significant gaps Δ1 appear precisely at ±ℏω/2 as a consequence of resonant Rabi splitting. Because these resonant gaps Δ1 are opened non-adiabatically, a non-equilibrium electron distribution can be expected in a ring of momenta surrounding the Dirac point where the photon resonance occurs, for |EF| ≲ ℏω/2. We believe that the observed plateau at low fluence thus results from the resonant excitation of electrons in conjunction with the predicted Berry curvature at the Δ1 band edges4,15,29–33. For |EF| ≳ ℏω/2, where first order resonant excitations are forbidden, Gxy decreased to zero and eventually changed sign when approaching the dielectric breakdown threshold of the device.
At higher fluences (blue and black circles), we resolved additional features near ±ℏω/2 in the Gxy vs. EF spectrum. They are closely aligned with the band edges of the calculated gaps Δ1, which are larger in size.
An additional feature appeared at the highest fluence (black circles): close to the Dirac point and away from the photon resonance, a conductance plateau with a width of approximately 60 meV was observed, where Gxy = (1.8 ± 0.4) e2/h (Fig. 4c right). Remarkably, the width of this plateau is very close to the calculated width of the light-induced topological gap at the Dirac point Δ0 (69 meV, red shading). This suggests that at high fluence there is a second source of anomalous Hall currents carried by a distribution of electrons occupying the lower dressed band beneath Δ0. For these momenta, the driving frequency is large compared to the splitting of the bands and a quasi-adiabatic transfer between the static and dressed bands is therefore possible. The effective Fermi level-dependent transverse transport then behaves similarly to what one would expect for gapped bands with non-zero total Berry curvature, given a Fermi-Dirac distribution, in that Gxy shows a plateau when EF lies inside the band gap11–14. The sharpness of the decrease in Gxy as EF is moved outside of Δ0 may depend on multiple factors, including the Berry curvature and non-equilibrium electron distributions in the Floquet bands, which is influenced by the frequency, amplitude and pulse-shape of the laser, as well as disorder, interactions and dissipation4,15,29–33.
Methods
All methods can be found in the supplementary information.
Supplementary Material
Acknowledgements
The authors acknowledge H. Aoki, L. Mathey, M. Nuske, A. Rubio, S.A. Sato, M.A. Sentef and P. Tang for fruitful discussions and B. Fiedler, B. Höhling, E. König and M. Volkmann for technical support. The research leading to these results received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC Grant Agreement no. 319286 (QMAC). J.W. McIver received funding from the Alexander von Humboldt Foundation.
Footnotes
Data availability.
The data represented in Figures 2,3, and 4 are available with the online version of this paper. All other data that supports the plots within this paper and other findings of this study are available from the corresponding author upon reasonable request.
Author contributions.
J.W. McIver conceived the experiment together with A. Cavalleri. A.C. and G. Meier supervised the project. J.W.M. and F.U. Stein designed and built the experimental setup. J.W.M., F.U.S., B. Schulte, T. Matsuyama and G.M. developed the on-chip circuitry. B.S. fabricated the graphene devices. B.S., J.W.M. and F.U.S. performed the measurements. B.S. and J.W.M. analyzed the data with support from T.M., G. Jotzu and G.M. Custom measurement electronics and circuit simulations were provided by T.M. and G.M. Floquet calculations were performed by G.J. The manuscript was written by J.W.M., G.J. and A.C. with contributions from all other authors.
References
- 1.Basov DN, Averitt RD, Hsieh D. Towards properties on demand in quantum materials. Nat Mat. 2017;16:1077–1088. doi: 10.1038/nmat5017. [DOI] [PubMed] [Google Scholar]
- 2.Fausti D, et al. Light-induced superconductivity in a stripe-ordered cuprate. Science. 2011;331:189–191. doi: 10.1126/science.1197294. [DOI] [PubMed] [Google Scholar]
- 3.Mitrano M, et al. Possible light-induced superconductivity in K3C60 at high temperature. Nature. 2016;530:461–464. doi: 10.1038/nature16522. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Oka T, Aoki H. Photovoltaic Hall effect in graphene. Phys Rev B. 2009;79 081406(R) [Google Scholar]
- 5.Kitagawa T, Oka T, Brataas A, Fu L, Demler E. Transport properties of nonequilibrium systems under the application of light: Photoinduced quantum Hall insulators without Landau levels. Phys Rev B. 2011;84 235108. [Google Scholar]
- 6.Lindner NH, Refael G, Galitski V. Floquet topological insulator in semiconductor quantum wells. Nat Phys. 2011;7:490–495. [Google Scholar]
- 7.Sie E, et al. Valley-selective optical Stark effect in monolayer WS2. Nat Mat. 2015;14:290–294. doi: 10.1038/nmat4156. [DOI] [PubMed] [Google Scholar]
- 8.Bukov M, D’Alessio L, Polkovnikov A. Universal high-frequency behavior of periodically driven systems: from dynamical stabilization to Floquet engineering. Adv In Phys. 2015;64(2):139–226. [Google Scholar]
- 9.Auston DH. Picosecond optoelectronic switching and gating in silicon. Appl Phys Lett. 1975;26:101–103. [Google Scholar]
- 10.Haldane FDM. Model for a Quantum Hall Effect without Landau Levels: Condensed-Matter Realization of the "Parity Anomaly". Phys Rev Lett. 1988;61:2015–2018. doi: 10.1103/PhysRevLett.61.2015. [DOI] [PubMed] [Google Scholar]
- 11.Xiao D, Chang M-C, Niu Q. Berry phase effects on electronic properties. Rev Mod Phys. 2010;82:1959–2007. [Google Scholar]
- 12.Hasan MZ, Kane CL. Colloquium: Topological insulators. Rev Mod Phys. 2010;82:3045–3067. [Google Scholar]
- 13.Qi X-L, Zhang S-C. Topological insulators and superconductors. Rev Mod Phys. 2011;83:1057–1110. [Google Scholar]
- 14.Bernevig BA, Hughes TL. Topological insulators and topological superconductors. Princeton University Press; Princeton, New Jersey: 2013. [Google Scholar]
- 15.Foa Torres LEF, Perez-Piskunow PM, Balseiro CA, Usaj G. Multiterminal Conductance of a Floquet Topological Insulator. Phys Rev Lett. 2014;113:266801. doi: 10.1103/PhysRevLett.113.266801. [DOI] [PubMed] [Google Scholar]
- 16.Chang C-Z et al. Experimental Observation of the Quantum Anomalous Hall Effect in a Magnetic Topological Insulator. Science. 2013;340:176–170. doi: 10.1126/science.1234414. [DOI] [PubMed] [Google Scholar]
- 17.König M, et al. Quantum Spin Hall Insulator State in HgTe Quantum Wells. Science. 2007;318:766–770. doi: 10.1126/science.1148047. [DOI] [PubMed] [Google Scholar]
- 18.Rechtsman MC, et al. Photonic Floquet topological insulator. Nature. 2013;496:196–200. doi: 10.1038/nature12066. [DOI] [PubMed] [Google Scholar]
- 19.Jotzu G, et al. Experimental realization of the topological Haldane model with ultracold fermions. Nature. 2014;515:237–240. doi: 10.1038/nature13915. [DOI] [PubMed] [Google Scholar]
- 20.Wang YH, Steinberg H, Jarillo-Herrero P, Gedik N. Observation of Floquet-Bloch States on the Surface of a Topological Insulator. Science. 2013;342:453–457. doi: 10.1126/science.1239834. [DOI] [PubMed] [Google Scholar]
- 21.Yin CM, et al. Observation of the photoinduced anomalous Hall effect in GaN-based heterostructures. Appl Phys Lett. 2011;98:122104. [Google Scholar]
- 22.Mak KF, McGill KL, Park J, McEuen PL. The valley Hall effect in MoS2 transistors. Science. 2014;344:1489–1492. doi: 10.1126/science.1250140. [DOI] [PubMed] [Google Scholar]
- 23.Seifert P, et al. In-plane anisotropy of the photon-helicity induced linear Hall effect in few-layer WTe2. Phys Rev B. 2019;99 161403(R) [Google Scholar]
- 24.Geim AK, Novoselov KS. The rise of graphene. Nat Mat. 2007;6:183–191. doi: 10.1038/nmat1849. [DOI] [PubMed] [Google Scholar]
- 25.Zhang Y, Tan Y-W, Stormer HL, Kim P. Experimental observation of the quantum Hall effect and Berry's phase in graphene. Nature. 2005;438:201–204. doi: 10.1038/nature04235. [DOI] [PubMed] [Google Scholar]
- 26.Glazov MM, Ganichev SD. High frequency electric field induced nonlinear effects in graphene. Phys Rep. 2014;535:101–138. [Google Scholar]
- 27.Sato S, et al. Microscopic theory for the light-induced anomalous Hall effect in graphene. Phys Rev B. 2019;99 214302. [Google Scholar]
- 28.Wang F, et al. Gate-Variable Optical Transitions in Graphene. Science. 2008;320:206–209. doi: 10.1126/science.1152793. [DOI] [PubMed] [Google Scholar]
- 29.Usaj G, Perez-Piskunow PM, Foa Torres LEF, Balseiro CA. Irradiated graphene as a tunable Floquet topological insulator. Phys Rev B. 2014;90 doi: 10.1103/PhysRevLett.113.266801. 115423. [DOI] [PubMed] [Google Scholar]
- 30.Mikami T, et al. Brillouin-Wigner theory for high-frequency expansion in periodically driven systems: Application to Floquet topological insulators. Phys Rev B. 2016;93 144307. [Google Scholar]
- 31.Dehghani H, Oka T, Mitra A. Out-of-equilibrium electrons and the Hall conductance of a Floquet topological insulator. Phys Rev B. 2015;91 155422. [Google Scholar]
- 32.Sentef MA, et al. Theory of Floquet band formation and local pseudospin textures in pump-probe photoemission of graphene. Nat Comm. 2015;6 doi: 10.1038/ncomms8047. 7047. [DOI] [PubMed] [Google Scholar]
- 33.Morimoto T, Nagaosa N. Topological nature of nonlinear optical effects in solids. Sci Adv. 2016;2(5):e1501524. doi: 10.1126/sciadv.1501524. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.