Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

All-optical valley switch and clock of electronic dephasing

Open Access Open Access

Abstract

2D materials with broken inversion symmetry posses an extra degree of freedom, the valley pseudospin, that labels in which of the two energy-degenerate crystal momenta, K or K’, the conducting carriers are located. It has been shown that shining circularly-polarized light allows to achieve close to 100% of valley polarization, opening the way to valley-based transistors. Yet, switching of the valley polarization is still a key challenge for the practical implementation of such devices due to the short valley lifetimes. Recent progress in ultrashort laser technology now allows to produce trains of attosecond pulses with controlled phase and polarization between the pulses. Taking advantage of such technology, we introduce a coherent control protocol to turn on, off and switch the valley polarization at faster timescales than electron-hole decoherence and valley depolarization, that is, an ultrafast optical valley switch. We theoretically demonstrate the protocol for hBN and MoS2 monolayers calculated from first principles. Additionally, using two time-delayed linearly-polarized pulses with perpendicular polarization, we show that we can extract the electronic dephasing time T2 from the valley Hall conductivity.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The synthesis of 2D materials with structure similar to graphene has opened the way to use and manipulate a new degree of freedom, the valley pseudospin, with potential applications in information processing and storage [13]. Valleys are energy minima of the electronic band structure of the crystal, which may be degenerate at different crystal momenta. In particular, for two-dimensional materials with a honeycomb structure and broken inversion symmetry, such as hexagonal boron nitride (hBN) or transition metal dichalcogenides (TMDs), the lowest conduction bands have energy-degenerate valleys located at the K and K’ points of the Brillouin zone (see Fig. 1(a), 1(d), 1(e)). The $C_3$ symmetry of the Bloch wavefunctions at the valleys leads to optical valley selection rules at $K$ and $K' \equiv -K$: $m_v - m_c \pm 1 = 3N$, where $N$ is an integer, the $\pm$ sign is determined by the the direction of rotation of the circular field, and the effective magnetic quantum numbers $m_{c,v}$ can be related to the Berry curvature [4].

 figure: Fig. 1.

Fig. 1. Band structure and optical valley selection rules. (a) Energy dispersion and Berry curvature (color map) for the $p_z$ bands in hBN. Arrows indicate the one-photon valley selection rules from each valley. (b) Few-cycle $\sigma _-$ circularly-polarized pulse. (c) Electron population in the first Brillouin zone (red-dashed hexagon) of the $p_z$ conduction band of hBN after the interaction with the pulse in (b). The population is normalized to 1 at each $k$-point. (d) Energy dispersion for MoS$_2$, including the $d$ orbitals of molybdenum and the $p$ orbitals of sulfur. (e) Expanded green region of panel (d), showing the relevant spin-split valence and conduction bands, their Berry curvature (color code) and the optical valley selection rules.

Download Full Size | PDF

Figure 1(a) and 1(e) indicates the value of the Berry curvature of the relevant bands of hBN and MoS$_2$, respectively, calculated from first principles [5], along with the valley selection rules. The selection rule is opposite in hBN and MoS$_2$ due to the different orbital character of the bands. Moreover, the strong-spin orbit coupling in MoS$_2$ lifts the spin degeneracy at the valleys, allowing for a coupling between spin and valley pseudospin [4]. The light field can be only few-cycles long, as that in Fig. 1(b), allowing to induce valley polarization at few-femtosecond timescales (Fig. 1(c)) [6].

One of the big obstacles for valleytronics are the short valley lifetimes, i.e., the time it takes the population to be scattered away from the valleys, and the short coherence times, i.e., the phase coherence of a particle in a superposition of two valleys [1]. Previous works have estimated that the valley lifetime decays after 200 fs and the valley coherence after 100 fs for excitons in WSe$_2$ monolayers [7]. Practical implementation of valleytronic devices therefore requires to induce valley polarization, switch it and read it at shorter timescales.

On the one hand, femtosecond control of the phase difference between the two valleys, i.e., valley coherence, has been achieved using the optical Stark effect [8]. On the other hand, femtosecond valley polarization, i.e., inducing population at $K$ or $K'$, has been achieved using strong single-cycle pulses [5,6], strong THz streaking fields [9], and strong tailored light fields [10]. In these methods, however, ultrafast switching of valley polarization, i.e., moving population from one valley to the other, is only possible while the material is dressed by the high-intense laser field. The high non-linearity of the process makes controlling the degree of valley polarization and its switching rate extremely challenging.

For a two-level system, population in the higher energy state can be switched on and off by two $\pi$-shifted resonant laser pulses delayed by $\tau = 2\pi /{\Delta E}$, where $\Delta E$ is the energy spacing between the two levels [11]. This is the simplest example of coherent control, which has seen a dramatic explosion thanks to the advances in high intensity and ultrafast laser technology, which has allowed control and imaging of real-time motion of electrons in atoms, molecules and solids [1217]. The production of trains of ultrashort pulses, down to the attosecond timescale, with a controllable phase and polarization relation between them is now possible [1820].

Here, we theoretically demonstrate a coherent control protocol that uses a sequence of ultrashort pulses to completely control the valley polarization, i.e., turn it on and off as well as switch between the valley states, on timescales shorter than 100 fs. The valley polarization is sensitive to the time of electronic coherence between bands, additionally providing a tool to measure such times.

2. Results

Throughout the paper, we will be considering weak-field, resonant processes between the upper valence bands and lowest conduction bands of 2D materials. First, we illustrate our method in hBN described by the two $p_z$ bands closest to the band gap, as shown in Fig. 1(a). It will become clear that the arguments are general, and later we will apply the method to MoS$_2$ including 22 bands (Fig. 1(d), 1(e)). The light-matter interaction is simulated using the density matrix equations in the independent particle approximation using ab initio equilibrium band structures and dipolar couplings, as described in previous works [5,21]. An exponentially-decaying electronic dephasing with lifetime $T_2$ is included to account for the loss of coherence between free electrons and holes in the conduction and valence bands, which we choose to be much shorter than the reported valley decoherence times [7]. To avoid confusion, we stress that we will focus on the ultrafast control of valley polarization (i.e., valley populations), not of valley coherence (i.e., the phase relation of a particle in a superposition of $K$ and $K'$ valleys). While valley coherence has been demonstrated in excitons [22], ultrafast control of this property is out of the scope of this work. To simplify our treatment, we will consider free carriers in initial thermal equilibrium, such that carriers generated from different crystal momenta are not mutually coherent. We note, however, that the results and physics behind our scheme is equally applicable to excitonic states, which will dominate the optical response in 2D materials.

To illustrate our method, let us first consider the problem where two linearly-polarized pulses with perpendicular polarizations and non-overlapping in time interact with hBN, as shown in Fig. 2(a) and 2(b). Besides their polarization direction, the two pulses in the sequence are identical. They are carried at the resonant band gap frequency $\omega =6$ eV, have a Gaussian envelope of 1.15 fs full width at half maximum and field strength of $F_0 = 0.1$ V/Å.

 figure: Fig. 2.

Fig. 2. Measurement of electronic dephasing. (a,b) Sequence of two linearly-polarized pulses with perpendicular polarizations separated by a time delay of (a) $\tau = 4.88$ fs and (b) $\tau = 5.16$ fs. (c,d) Electron populations in the first Brillouin zone of the $p_z$ conduction of hBN for the pulses in (a) and (b), respectively. (e,f) Valley Hall conductivity as a function of the time delay between the two pulses, calculated for a dephasing time of (e) $T_2 \to \infty$ and (f) $T_2 = 10$ fs. The orange curves show the full time-dependent calculation while the black curves show the calculation using lowest order perturbation theory (LOPT). In (f), the LOPT calculation is multiplied additionally by a decaying exponential with lifetime $T_2=10$ fs (see text).

Download Full Size | PDF

Figure 2(c) and 2(d) shows the electron populations in the conduction band which appear localized at $K$ and $K'$, respectively for the pulses in panels (a) and (b). The optical valley selection rules dictates that a linearly-polarized pulse does not induce valley polarization, while Fig. 2(c) and 2(d) appears at first to show the opposite: the degree of valley polarization of the two linear, non-overlapping pulses is similar to that produced by a circularly-polarized field. This is possible thanks to the sustained electronic coherence between the two few-cycle pulses: any time delay $\tau = \frac {1}{\omega _{\mathcal {F}}} (\pm \pi /2 + 2\pi N)$, where $N$ is an integer, between a first $y$-polarized linear pulse and a second $x$-polarized linear pulse will produce a $\sigma _\mp$ circularly-polarized field at the frequency $\omega _\mathcal {F}$ [11]. The dynamics at $\omega _\mathcal {F}$ will be equivalent for any chosen $N$ as long as the temporal separation of the pulses is smaller than the electronic coherence between the valence and conduction bands created by the first pulse, $\tau << T_2$. The valley polarization hence oscillates for different time delays between the two pulses (cf. panels c and d).

Due to the short duration of the pulses, the resulting bandwidth populates not only the $K$ and $K'$ points, but also many other nearby crystal momenta with different energies (see Fig. 2(c), 2(d)). Therefore, a time delay $\tau = \frac {1}{\Delta }\ (\pi /2 + 2\pi N)$, with $\Delta$ the band gap at $K$ and $K'$, does not guarantee maximum valley polarization since one cannot measure the latter at a single k-point. In order to obtain the optimal delay that yields the maximum valley polarization, a time delay scan must be performed first. Figure 2(e) and 2(f) shows the valley Hall conductivity (VHC) as a function of the time delay $\tau$, which is closely related to the valley polarization

$$\sigma (\tau) \propto \sum_n \int_{\text{BZ}} f_{n}(\mathbf{k},\tau) \Omega_{n} (\mathbf{k}) d\mathbf{k},$$
where $n$ is the band index, $f_{n}(\mathbf {k},\tau )$ is the electron population at the $\mathbf {k}$ point after the two-pulse sequence ($t\to \infty$) for the time delay $\tau$, $\Omega _n (\mathbf {k})$ is the field-free Berry curvature, and the integral runs over the first Brillouin zone. Since $\Omega _n (\mathbf {k}) = -\Omega _n (-\mathbf {k})$, the VHC is zero for zero valley polarization, and changes sign for opposite valley polarization. Figure 2(e), and 2(f) reveals that in the region where the pulses are overlapping ($\tau < 2$ fs), the VHC peaks at $\tau = \frac {1}{\Delta }\ (\pm \pi /2 + 2\pi N)$, i.e., where the pulse is circular at the $K$ and $K'$ points. For longer delays, however, the VHC peaks at $\tau = \frac {\pi }{\Delta }\ N$, reflecting the dominant contributions from other crystal momenta. The valley polarization drops as a function of $\tau$ until it reaches a stable value for the case where there is no loss of electronic coherence, i.e., $T_2 \to \infty$ (panel e). These dynamics can be perfectly reproduced using lowest order perturbation theory (cf. orange and black curves in panel e) and are a consequence of the broadband pulse. As the time delay $\tau$ between the pulses increases, the phase of the electronic wavepacket varies more strongly between neighbouring energies, $\Delta \phi = \Delta E \tau$. This leads to different interference patterns at close-by energies thus decreasing the VHC. This decrease stabilizes at some point if $\tau << T_2$. However, for $\tau$ on the order of $T_2$, the loss of electronic coherence between the valence and conduction bands adds an additional decay mechanism of the VHC, as shown in Fig. 2(f), which completely cancels the signal for $\tau \,{\gtrapprox }\,2T_2$. This presents an opportunity to accurately retrieve the dephasing lifetime $T_2$.

Let us suppose we perform a time-delay measurement of the VHC on hBN and we obtain the orange curve in Fig. 2(f). For our purposes, we have simulated this curve from the full time-dependent calculation setting $T_2=10$ fs. To obtain the unknown $T_2$, we first need to obtain the reference function $\sigma _{\text {ref}}(\tau )$ that includes the drop of the VHC due to the wavepacket bandwidth but no electronic dephasing, which is given in Fig. 2(e). This is obtained numerically, either through a full-time dependent calculation (orange curve in panel e), or simply by using lowest-order perturbation theory (black curve in panel e). Once $\sigma _{\text {ref}}(\tau )$ is known, the lifetime $T_2$ can be extracted from the fit of our measured time-delay scan of the VHC (orange curve in panel f) to the function $\sigma (\tau,T_2) = \sigma _{\text {ref}}(\tau ) e^{-\tau /T_2}$. Figure 2(f) shows the excellent agreement between the full time-dependent calculation using $T_2=10$ fs (orange curve) and $\sigma (\tau,T_2=10$ fs$)$ obtained through lowest order perturbation theory (black curve).

Let us now use the concepts above to generate an optical valley switch. The protocol consists of a sequence of four identical linearly-polarized pulses resonant with the band gap, but polarized along different directions (Fig. 3(a)–3(d). First, we drive conduction band population using a field polarized along the y direction, although any other polarization direction is valid at this stage. The conduction band population in the Brillouin zone after the interaction with the pulse is shown in Fig. 3(e). As expected from excitation by a linearly-polarized field, the electron population shows no valley polarization. This first pulse creates coherence between the valence and conduction bands.

 figure: Fig. 3.

Fig. 3. Optical valley switch in a two band description of hBN. (a-d) Sequence of few-cycle pulses generating different values of the valley polarization. (e-h) Electron populations in the first Brillouin zone of the $p_z$ conduction band of hBN after the interaction with the pulses in panels (a-d), respectively. (i) Valley Hall conductivity as a function of time for the four-pulse sequence in panel (d), demonstrating the optical valley switch. Simulations were performed for different values of the dephasing time $T_2$, and are shown in different colors.

Download Full Size | PDF

At an appropriate later time $\tau$, obtained through a time-delay scan of the VHC $\sigma (\tau )$, we make the system interact with a second linearly-polarized pulse polarized perpendicular to it (Fig. 3(b)). As we discussed earlier, the maximum valley polarization at $K$ is obtained for $\tau = N\times 2\pi /\Delta$, and we choose $\tau = 7 \times 2\pi /\Delta = 4.8$ fs, which guarantees both a reasonably high value of valley polarization and non-overlapping pulses. The valley polarization at $K$ is clear (Fig. 3(f)). Until now, we have simply induced valley polarization; no more than what we would have achieved with a single circularly-polarized pulse. Let us now switch off this valley polarization.

For this, we add a third pulse, linearly-polarized along the same direction of the first, as shown in Fig. 3(c). In this way, we are able to choose a time delay for the third pulse which completely switches off the valley polarization (Fig. 3(g)). The time delay chosen is $\tau = 14 \times 2\pi /\Delta = 9.6$ fs, again obtained through a time-delay scan of the VHC. We have now managed to switch on and off the valley polarization on a timescale of 10 fs, more than one order of magnitude shorter than the estimated valley lifetime [3] and electronic dephasing [23].

To finalize the scheme, we seek to induce the opposite valley polarization as that created by the second pulse. We add a fourth pulse linearly-polarized along the same direction as the second pulse (Fig. 3(d)). In this case, the optimal time delay for this operation is $\tau = 21.5 \times 2\pi /\Delta = 14.8$ fs. Figure 3(h) shows the switch of the valley polarization with respect to that induced by the second pulse. All of the protocol is performed on timescales one order of magnitude shorter than valley lifetimes.

Figure 3(i) shows the time-dependent VHC, $\sigma (t)$, for the four-pulse sequence in Fig. 3(d) calculated using electronic dephasing times ranging from $T_2=100$ fs, which is the estimated order of magnitude reported in recent work [23,24], to more than one order of magnitude shorter. Note that $\sigma (t)$ is slightly different to $\sigma (\tau )$ defined in Eq. (1), since now $f_n (\mathbf {k},t)$ refers to the population during the interaction with the four-pulse sequence (from $t=-2$ fs to $t=17.5$ fs). The VHC at the end of the scheme ($t \simeq 17.5$) drops by about one third between a dephasing time of 100fs and 10fs, but the valley switching is still clearly visible even for a separation between the first and last pulse of $\tau = 2.5\,T_2$, demonstrating the robustness of our method.

The method described above is general. While the large band gap of hBN allows to use resonant few-cycle pulses with very short duration, it also makes hBN challenging to be used as an optically-pumped valleytronic material. Indeed, producing few-cycle pulses of 6eV has, to the best of our knowledge, not been demonstrated. However, monolayer TMDs are perfect candidates for this scheme; their direct band gaps lie in the visible spectrum (between 400 nm and 700 nm) [1], a range of wavelengths for which several experimental groups have reported few-cycle pulse production with controlled shape and polarization [20,25]. TMDs also display strong dipole couplings, most prominently to excitonic states. Here, we will not consider excitonic transitons, but instead the resonant electronic transition between the valence bands and the conduction bands, as we did with hBN, and as shown in Fig. 1(e). We note, however, that the protocol would be equivalent for excitonic states, with the additional advantage that the wavepacket will be more localized in energy. This means that the decay of the valley polarization in the conduction band due to pulse bandwidth effects discussed in Fig. 2(e) will be less pronounced.

Figure 4 shows the protocol in MoS$_2$, where we included the 22 bands shown in Fig. 1(d) in the time-dependent simulations. The pulse parameters, shown in Fig. 4(a)–4(d), are in this case: $\omega = 1.58$ eV, matching the ab-initio minimum band gap of MoS$_2$ (see Fig. 1(d), 1(e)), field strength $F_0 = 0.01$ V/Å, and Gaussian pulse envelope with $6.8$ fs full width at half maximum. Figure 4(e)–4(h) shows that the valley polarization is turned on, off and switched analogously to hBN, apart from a sign change due to the opposite valley selection rules (cf. Figure 1(a), 1(e)). Figure 4(i) reveals that the VHC drops by a factor of 4 between $T_2=100$ fs and $T_2=20$ fs. The switch of the sign in the VHC is nonetheless still clearly visible for $T_2=20$ fs. Both the electron populations in panels (e)–(h) and the valley Hall conductivity in panel (i) were computed by adding up the populations of the two lower spin-split conduction bands (see Fig. 1(e)), where 99$\%$ of the excited population resided.

 figure: Fig. 4.

Fig. 4. Optical valley switch in a 22 band description of MoS$_2$. (a-d) Pulse sequences generating different values of the valley polarization. (e-h) Sum of the electron populations in the first two conduction bands of MoS$_2$ (see Fig. 1(e)) after the interaction with the pulses in panels (a-d), respectively. (i) Valley Hall conductivity as a function of time for the four pulse sequence in panel (d), computed from the sum of the electron populations in the two lowest conduction bands. Results for dephasing times $T_2=100$ fs and $T_2=20$ fs are shown in blue and orange, respectively.

Download Full Size | PDF

3. Conclusion

In summary, by using a sequence of ultrashort pulses with controlled phase delay and polarization, we have demonstrated an all-optical protocol to turn on, off and switch the valley polarization on timescales shorter than reported valley lifetimes and even valley decoherence times, and well within experimental reach. We have applied our method to hBN, described by two bands, and MoS$_2$, described by 22 bands, calculated from first principles, obtaining equally good results. We have also shown that by measuring the valley Hall conductivity as a function of the time delay between two perpendicularly-polarized pulses, the electronic dephasing time $T_2$ can be retrieved. Our work gives important steps towards the practical implementation of ultrafast all-optical valley switches and towards the determination of electronic coherence times in 2D materials.

Funding

“la Caixa” Foundation (100010434); Horizon 2020 Framework Programme (899794); H2020 Marie Skłodowska-Curie Actions (101028938).

Acknowledgments

R.E.F.S. acknowledges support from the fellowship LCF/BQ/PR21/11840008 from "La Caixa" Foundation (ID 100010434). M.I. and A.J.G acknowledge support from the Horizon 2020 Framework Programme (899794). A.J.G. acknowledges support from the H2020 Marie Skłodowska-Curie Actions (101028938).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. J. R. Schaibley, H. Yu, G. Clark, P. Rivera, J. S. Ross, K. L. Seyler, W. Yao, and X. Xu, “Valleytronics in 2D materials,” Nat. Rev. Mater. 1(11), 16055 (2016). [CrossRef]  

2. H. Yu, X. Cui, X. Xu, and W. Yao, “Valley excitons in two-dimensional semiconductors,” Natl. Sci. Rev. 2(1), 57–70 (2015). [CrossRef]  

3. S. A. Vitale, D. Nezich, J. O. Varghese, P. Kim, N. Gedik, P. Jarillo-Herrero, D. Xiao, and M. Rothschild, “Valleytronics: Opportunities, Challenges, and Paths Forward,” Small 14(38), 1801483 (2018). [CrossRef]  

4. D. Xiao, M.-C. Chang, and Q. Niu, “Berry phase effects on electronic properties,” Rev. Mod. Phys. 82(3), 1959–2007 (2010). [CrossRef]  

5. Á. Jiménez-Galán, R. E. F. Silva, O. Smirnova, and M. Ivanov, “Sub-cycle valleytronics: control of valley polarization using few-cycle linearly polarized pulses,” Optica 8(3), 277–280 (2021). [CrossRef]  

6. S. A. Oliaei Motlagh, J.-S. Wu, V. Apalkov, and M. I. Stockman, “Femtosecond valley polarization and topological resonances in transition metal dichalcogenides,” Phys. Rev. B 98(8), 081406 (2018). [CrossRef]  

7. K. Hao, G. Moody, F. Wu, C. K. Dass, L. Xu, C.-H. Chen, L. Sun, M.-Y. Li, L.-J. Li, A. H. MacDonald, and X. Li, “Direct measurement of exciton valley coherence in monolayer wse2,” Nat. Phys. 12(7), 677–682 (2016). [CrossRef]  

8. Z. Ye, D. Sun, and T. F. Heinz, “Optical manipulation of valley pseudospin,” Nat. Phys. 13(1), 26–29 (2017). [CrossRef]  

9. F. Langer, C. P. Schmid, S. Schlauderer, M. Gmitra, J. Fabian, P. Nagler, C. Schüller, T. Korn, P. G. Hawkins, J. T. Steiner, U. Huttner, S. W. Koch, M. Kira, and R. Huber, “Lightwave valleytronics in a monolayer of tungsten diselenide,” Nature 557(7703), 76–80 (2018). [CrossRef]  

10. Á. Jiménez-Galán, R. E. F. Silva, O. Smirnova, and M. Ivanov, “Lightwave control of topological properties in 2D materials for sub-cycle and non-resonant valley manipulation,” Nat. Photonics 14(12), 728–732 (2020). [CrossRef]  

11. W. S. Warren, H. Rabitz, and M. Dahleh, “Coherent control of quantum dynamics: the dream is alive,” Science 259(5101), 1581–1589 (1993). [CrossRef]  

12. A. Baltuška, T. Udem, M. Uiberacker, M. Hentschel, E. Goulielmakis, C. Gohle, R. Holzwarth, V. S. Yakovlev, A. Scrinzi, T. W. Hänsch, and F. Krausz, “Attosecond control of electronic processes by intense light fields,” Nature 421(6923), 611–615 (2003). [CrossRef]  

13. V. Gruson, L. Barreau, A. Jiménez-Galan, F. Risoud, J. Caillat, A. Maquet, B. Carré, F. Lepetit, J.-F. Hergott, T. Ruchon, L. Argenti, R. Taïeb, F. Martín, and P. Salières, “Attosecond dynamics through a fano resonance: Monitoring the birth of a photoelectron,” Science 354(6313), 734–738 (2016). [CrossRef]  

14. M. Kotur, D. Guenot, Á. Jiménez-Galán, D. Kroon, E. W. Larsen, M. Louisy, S. Bengtsson, M. Miranda, J. Mauritsson, C. Arnold, S. Canton, M. Gisselbrecht, C. T M. Dahlström, E. Lindroth, A. Maquet, L. Argenti, F. Martín, and A. L’Huillier, “Spectral phase measurement of a fano resonance using tunable attosecond pulses,” Nat. Commun. 7(1), 10566 (2016). [CrossRef]  

15. A. Baltuška, T. Fuji, and T. Kobayashi, “Controlling the carrier-envelope phase of ultrashort light pulses with optical parametric amplifiers,” Phys. Rev. Lett. 88(13), 133901 (2002). [CrossRef]  

16. R. Lopez-Martens, K. Varjú, P. Johnsson, J. Mauritsson, Y. Mairesse, P. Salieres, M. B. Gaarde, K. J. Schafer, A. Persson, S. Svanberg, C.-G. Wahlström, and A. L’Huillier, “Amplitude and phase control of attosecond light pulses,” Phys. Rev. Lett. 94(3), 033001 (2005). [CrossRef]  

17. N. Dudovich, O. Smirnova, J. Levesque, Y. Mairesse, M. Y. Ivanov, D. Villeneuve, and P. B. Corkum, “Measuring and controlling the birth of attosecond xuv pulses,” Nat. Phys. 2(11), 781–786 (2006). [CrossRef]  

18. K. M. Dorney, L. Rego, N. J. Brooks, J. San Román, C.-T. Liao, J. L. Ellis, D. Zusin, C. Gentry, Q. L. Nguyen, J. M. Shaw, A. Picón, L. Plaja, H. C. Kapteyn, M. M. Murnane, and C. Hernández-García, “Controlling the polarization and vortex charge of attosecond high-harmonic beams via simultaneous spin–orbit momentum conservation,” Nat. Photonics 13(2), 123–130 (2019). [CrossRef]  

19. I. Pupeza, D. Sánchez, J. Zhang, N. Lilienfein, M. Seidel, N. Karpowicz, T. Paasch-Colberg, I. Znakovskaya, M. Pescher, W. Schweinberger, V. Pervak, E. Fill, O. Pronin, Z. Wei, F. Krausz, A. Apolonski, and J. Biegert, “High-power sub-two-cycle mid-infrared pulses at 100 mhz repetition rate,” Nat. Photonics 9(11), 721–724 (2015). [CrossRef]  

20. V. Stummer, T. Flöry, G. Krizsán, G. Polónyi, E. Kaksis, A. Pugžlys, J. Hebling, J. A. Fülöp, and A. Baltuška, “Programmable generation of terahertz bursts in chirped-pulse laser amplification,” Optica 7(12), 1758–1763 (2020). [CrossRef]  

21. R. E. F. Silva, F. Martín, and M. Ivanov, “High harmonic generation in crystals using maximally localized wannier functions,” Phys. Rev. B 100(19), 195201 (2019). [CrossRef]  

22. A. M. Jones, H. Yu, N. J. Ghimire, S. Wu, G. Aivazian, J. S. Ross, B. Zhao, J. Yan, D. G. Mandrus, D. Xiao, W. Yao, and X. Xu, “Optical generation of excitonic valley coherence in monolayer WSe2,” Nat. Nanotechnol. 8(9), 634–638 (2013). [CrossRef]  

23. I. Kilen, M. Kolesik, J. Hader, J. V. Moloney, U. Huttner, M. K. Hagen, and S. W. Koch, “Propagation induced dephasing in semiconductor high-harmonic generation,” Phys. Rev. Lett. 125(8), 083901 (2020). [CrossRef]  

24. I. Floss, C. Lemell, G. Wachter, V. Smejkal, S. A. Sato, X.-M. Tong, K. Yabana, and J. Burgdörfer, “Ab initio multiscale simulation of high-order harmonic generation in solids,” Phys. Rev. A 97(1), 011401 (2018). [CrossRef]  

25. A. Wirth, M. T. Hassan, I. Grguras, J. Gagnon, A. Moulet, T. T. Luu, S. Pabst, R. Santra, Z. a. Alahmed, a. M. Azzeer, V. S. Yakovlev, V. Pervak, F. Krausz, and E. Goulielmakis, “Synthesized Light Transients,” Science 334(6053), 195–200 (2011). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Band structure and optical valley selection rules. (a) Energy dispersion and Berry curvature (color map) for the $p_z$ bands in hBN. Arrows indicate the one-photon valley selection rules from each valley. (b) Few-cycle $\sigma _-$ circularly-polarized pulse. (c) Electron population in the first Brillouin zone (red-dashed hexagon) of the $p_z$ conduction band of hBN after the interaction with the pulse in (b). The population is normalized to 1 at each $k$-point. (d) Energy dispersion for MoS$_2$, including the $d$ orbitals of molybdenum and the $p$ orbitals of sulfur. (e) Expanded green region of panel (d), showing the relevant spin-split valence and conduction bands, their Berry curvature (color code) and the optical valley selection rules.
Fig. 2.
Fig. 2. Measurement of electronic dephasing. (a,b) Sequence of two linearly-polarized pulses with perpendicular polarizations separated by a time delay of (a) $\tau = 4.88$ fs and (b) $\tau = 5.16$ fs. (c,d) Electron populations in the first Brillouin zone of the $p_z$ conduction of hBN for the pulses in (a) and (b), respectively. (e,f) Valley Hall conductivity as a function of the time delay between the two pulses, calculated for a dephasing time of (e) $T_2 \to \infty$ and (f) $T_2 = 10$ fs. The orange curves show the full time-dependent calculation while the black curves show the calculation using lowest order perturbation theory (LOPT). In (f), the LOPT calculation is multiplied additionally by a decaying exponential with lifetime $T_2=10$ fs (see text).
Fig. 3.
Fig. 3. Optical valley switch in a two band description of hBN. (a-d) Sequence of few-cycle pulses generating different values of the valley polarization. (e-h) Electron populations in the first Brillouin zone of the $p_z$ conduction band of hBN after the interaction with the pulses in panels (a-d), respectively. (i) Valley Hall conductivity as a function of time for the four-pulse sequence in panel (d), demonstrating the optical valley switch. Simulations were performed for different values of the dephasing time $T_2$, and are shown in different colors.
Fig. 4.
Fig. 4. Optical valley switch in a 22 band description of MoS$_2$. (a-d) Pulse sequences generating different values of the valley polarization. (e-h) Sum of the electron populations in the first two conduction bands of MoS$_2$ (see Fig. 1(e)) after the interaction with the pulses in panels (a-d), respectively. (i) Valley Hall conductivity as a function of time for the four pulse sequence in panel (d), computed from the sum of the electron populations in the two lowest conduction bands. Results for dephasing times $T_2=100$ fs and $T_2=20$ fs are shown in blue and orange, respectively.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

σ ( τ ) n BZ f n ( k , τ ) Ω n ( k ) d k ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.