Abstract
Free full text
Vibronic Exciton–Phonon States in Stack-Engineered van der Waals Heterojunction Photodiodes
Abstract
Stack engineering, an atomic-scale metamaterial strategy, enables the design of optical and electronic properties in van der Waals heterostructure devices. Here we reveal the optoelectronic effects of stacking-induced strong coupling between atomic motion and interlayer excitons in WSe2/MoSe2 heterojunction photodiodes. To do so, we introduce the photocurrent spectroscopy of a stack-engineered photodiode as a sensitive technique for probing interlayer excitons, enabling access to vibronic states typically found only in molecule-like systems. The vibronic states in our stack are manifest as a palisade of pronounced periodic sidebands in the photocurrent spectrum in frequency windows close to the interlayer exciton resonances and can be shifted “on demand” through the application of a perpendicular electric field via a source-drain bias voltage. The observation of multiple well-resolved sidebands as well as their ability to be shifted by applied voltages vividly demonstrates the emergence of interlayer exciton vibronic structure in a stack-engineered optoelectronic device.
The vibrational motion of atoms in solids is ordinarily expected to dissipate energy from electronic excitations. Strong coupling of vibrational motion to electrons, however, can significantly transform the nature of accessible excited states, dramatically enriching light–matter interactions. In soft matter, such as photosynthetic light-harvesting complexes, the interplay between atomic motion and exciton dynamics enhances electronic energy transfer1−4 in spite of the fluctuating physical environment. In crystals, the presence of strong interactions between electronic excitations and phonons,5,6 which are the elementary excitations of the atomic lattice, enables the trapping of excitations,7−9 allows mechanical control of electron transport,10,11 and drives the formation of exotic exciton–phonon quasiparticles and excitonic complexes.12,13
A particularly striking manifestation of strong exciton–phonon coupling is the periodic vibronic structure that appears in molecular absorption spectra.14 The well-separated and multiple peaks in these systems correspond to distinct vibronic states; resonant excitation therefore enables us to directly address each individual state. Easy access to both optical and electronic control in transition-metal dichalcogenides (TMDs) offers a unique opportunity to electrically control vibronic states in a semiconductor device. Nevertheless, the voltage-tunable vibronic structure of periodic and well-separated peaks has not been realized in semiconductor TMD heterostructure devices.
Here we report the emergence of multiple periodic and well-separated photocurrent peaks when individual van der Waals (vdW) layers are stacked to form atomically thin heterostructure photodiodes. In particular, we find that the interlayer photocurrent exhibits a rich structure with numerous photocurrent sidebands as a function of incident photon energy EPH. Strikingly, photocurrent sidebands are manifested only for values of EPH close to the interlayer exciton energy; they are absent for the intralayer excitons. These resonances occur periodically with an energy spacing of approximately 30 meV and are observed for two distinct interlayer excitons within the same heterostructure. This energy spacing corresponds to the frequency of a group of prominent phonon modes observed in the Raman spectrum of the heterostructure. Importantly, we demonstrate that the manifold of photocurrent peaks can be controlled by the source-drain bias voltage.
As we discuss below, this multipeaked structure of the exciton resonance can be attributed to the appearance of a manifold of well-formed vibronic coupled exciton–phonon states (mirroring those typically found in molecular systems). We note that this vibronic structure of interlayer excitons in vdW heterostructures has so far been obscured from purely optical probes such as photoluminescence (PL);15 see also the discussion below. In contrast, the multicomponent photocurrent spectroscopy that we employed provides a means to directly address and control the rich vibronic structure of interlayer excitons that results from strong electron–phonon coupling in stacked vdW materials.
Experimentally, we studied vdW heterostructures composed of bilayer tungsten diselenide (WSe2) stacked on top of monolayer molybdenum diselenide (MoSe2), as shown in Figure Figure11. 2L-WSe2/MoSe2 serves as an important benchmark for photocurrent spectroscopy because it exhibits two distinct interlayer exciton resonances, one which has been observed in many PL studies of WSe2/MoSe2 (K → K) and another that has not been previously observed (Γ → K), as discussed in detail below. Owing to this newly observed interlayer exciton and the type II band alignment (see detailed band structure calculations in ref (25)), 2L-WSe2/MoSe2 also fulfills the technological need for semiconductor heterostructures with near-infrared band gaps, similar to silicon. Indeed, this work establishes it as an excellent material system with a clearly evident photocurrent near 1 eV.
Using an inverted fabrication process (Figure Figure11a, Supporting Information Section S1(16)), we first patterned multilayer graphene gate electrodes and conventional metal source and drain contacts. The WSe2–MoSe2 heterostructures, assembled and characterized independently, were laminated onto the prefabricated device patterns. Combining the vdW heterostructure in this way enabled complete protective encapsulation of the WSe2-MoSe2 interface region between hexagonal boron nitride layers (schematic in Figure Figure11b). We fabricated and studied three devices, each with a random orientation between constituent layers. The results described below were consistent across all devices.
Devices were characterized using Raman, photoluminescence (PL), and photocurrent (PC) spectroscopy. As shown in Figure Figure11d, the Raman spectrum of the heterostructure (black line) exhibits peaks that are also evident in the individual MoSe2 (green line) and WSe2 (blue line) layers. When comparing Raman peaks between the heterostructure and its constituent layers, we observed a negligible shift of the peak positions as a function of energy. Importantly, we observed pronounced Raman peaks at the WSe2-MoSe2 interface (black line), at energies near 30 meV (29.9, 31.0, and 31.9 meV, Figure Figure11d inset). In our previous work,17 Raman peaks near 30 meV were used to identify the layer thickness, confirming that the heterostructure was composed of bilayer WSe2 and monolayer MoSe2. In MoSe2, the A1g peak at 241 cm–1 (29.8 meV), E2g1 peak at 288 cm–1, and lack of B2g1 peak between 350 and 360 cm–1 are characteristic of monolayer thickness. The Raman spectrum of WSe2 exhibits an A1g mode at 250 cm–1 (30.9 meV), an E2g mode at 260 cm–1 (32.2 meV), and a B2g1 mode at 309 cm–1, indicating bilayer thickness.
Figure Figure11e compares the PL vs photon energy EPH from MoSe2 (green line), WSe2 (blue line), and the heterostructure (black line). The stacked vdW heterostructure exhibits PL spectral features similar to those of the individual layers.18−23 In addition to the observed excitonic resonances characteristic of the individual layers, low-energy bound interlayer e–h pairs (excitons) are known to form between the valence band of bilayer WSe2 and the conduction band of MoSe2 (Figure Figure11c). Direct access to interlayer excitons has proven to be challenging. For example, although signatures of the lowest-lying interlayer exciton near EI ≈ 1.0 eV (arising from carriers in the momentum mismatched Γ- and K-valleys) can be indirectly inferred,17 direct PL signatures are washed out as a result of very small oscillator strengths.
To overcome the small oscillator strength that prevents strong and direct PL signatures for these interlayer excitons, here we instead employed measurements of the interlayer photocurrent IPC with the laser focused on the WSe2-MoSe2 heterostructure. In the photocurrent setup, interlayer excitons that are generated by infrared laser illumination of the vdW heterostructure can be subsequently dissociated. Once dissociated, separated electrons and holes transit the device, resulting in a photocurrent that increases in reverse bias (Figure Figure11f inset; the main panel shows dark current characterization of the device, displaying typical diode behavior).23−27
Figure Figure22 examines the detailed dependence of the interlayer photocurrent on EPH and VG. Using sensitive current amplification and sweeping EPH, we find that IPC, which is the difference in current measured with the light on and light off, is particularly pronounced in two frequency windows occurring near EPH ≈ 1.3 and 0.9 eV. Photoexcitation energies in the vicinity of these two hot spots correspond to interlayer excitons hosted in the WSe2-MoSe2 heterostructure, namely, the K → K (~1.3 eV) and Γ → K (~0.9 eV) interlayer excitons.17,18,28
Strikingly, when EPH was tuned between 1.24 and 1.40 eV (Figure Figure22a), we observed a periodic sequence of photocurrent peaks that occur in a narrow range of gate voltages (near VG = −3.5 V). Although the strongest peak occurs at EPH = 1.32 eV, it is only slightly stronger than several equally spaced maxima at higher and lower EPH (Figure Figure22b). We observed an average peak separation of 30 meV. To clarify this periodic modulation, we calculated the Fourier transform of the second derivative of the photocurrent data, where we find a clear periodic component at 1/Δε = (30 meV)−1 (marked by the blue dashed line in Figure Figure22c). Interestingly, near the interlayer excitation from the WSe2K-valley to the MoSe2K-valley (Figure Figure22d), the discrete energy difference Δε = 30 meV between photocurrent peaks closely corresponds to the frequency at which a strong Raman signal is observed in the heterojunction, Ω ≈ 30 meV. Here, is Planck’s constant and Ω is the phonon frequency.
In the same fashion as above, optical excitation of the lowest-lying (Γ → K) interlayer exciton also results in a series of approximately equally spaced discrete sidebands with energy spacing Ω ≈ 30 meV. This behavior is highlighted in Figure Figure22e, which shows IPC vs EPH and VG at infrared photon energies. In the range EPH = 0.88–1.03 eV, we observed a set of evenly spaced photocurrent maxima, which increase in amplitude as EPH increases. The lowest-energy peak occurs at EPH = 0.90 eV, and line traces of IPC vs EPH (Figure Figure22f) show regularly spaced peaks that are superimposed on a photocurrent background that increases with EPH. Taking the Fourier transform of the second derivative of the interlayer photocurrent data (Figure Figure22g) reveals two periodic components: a dominant component at Δε = 30 meV and a weaker component at 22 meV. In sharp contrast to this periodic structure in Figure Figure22, photocurrent corresponding to the excitation of intralayer excitons does not produce such sidebands (Supporting Information Section S4.116).
The appearance of a periodic array of photocurrent sidebands measured at fixed source-drain bias VSD can be understood through the strong coupling of phonons and interlayer excitons, with each of the sideband peaks identified with a coupled (interlayer) exciton–phonon state of frequency ωvib,n (VSD), where n is an index that labels the vibronic states/peaks. In these vibronic states, electronic excitations are intertwined with lattice displacements and can be understood as Franck–Condon-type progressions. (See the detailed theoretical description in Supporting Information Section S5.16) We note, parenthetically, that the 30 meV periodicity in photocurrent spectroscopy (Figure Figure22) closely matches a window of narrow phonon branches expected for the interlayer heterostructure. (See Supporting Information Section S3(16) for discussion of phononic origins as well as phonon dispersion calculations in the heterostructure.) Remarkably, each of these vibronic states, ωvib,n, is addressable by tuning EPH. This contrasts with phonon-broadened peaks wherein the action of exciton–phonon interaction broadens the exciton transition without yielding individual well-defined exciton–phonon states.6
Although the EPH spectroscopy in Figure Figure22 revealed the vibronic structure (at fixed VSD), we can leverage the out-of-plane electric field of the interlayer p–n junction to electrically control the registry of vibronic states. To see this, we conducted detailed VSD-dependent measurements of the interlayer exciton photoresponse. Figure Figure33a shows photoconductance (dIPC/dVSD) maps as a function of VSD and VG for a reverse-biased p–n heterojunction (device 2) with EPH = 0.99 eV. In these dIPC/dVSD maps, multiple vertical stripes stand out, corresponding to prominent periodic oscillations (as a function of VSD) in the photoconductance (Figure Figure33b). Indeed, this periodic oscillation in photoconductance is confirmed via Fourier transform of the dIPC/dVSD maps and displays a well-defined VSD periodicity. These photoconductance oscillations, observed at room temperature, can be understood to arise from VSD shifting the entire registry of vibronic peaks so that each peak shifts in and out of resonance with the fixed EPH.
Electrical control of the vibronic peaks is consistent with the fact that interlayer excitons exhibit an electric dipole, p, that points out of plane. The exciton–phonon vibronic states can therefore naturally be sensitive to an out-of-plane electric field, Ei. Indeed, the electrical control of ωvib,n (VSD) can be understood through a Stark shift ΔE = p·Ei = edVSD/tα = eVSD/αη that uniformly shifts the registry of vibronic states. Here, η = t/d is a dimensionless quantity that compares the interlayer distance t to the effective dipole length d (assumed to be approximately the same for all states). The ideality factor α (obtained in fits of Figure Figure11f) appears in this expression to phenomenologically capture the reduction of VSD due to leakages such as nonideal contacts.29 In the vdW p–n junction photodiode, the Stark shift and thus the registry of vibronic states are directly controlled using VSD.
Utilizing the VSD-dependent photoresponse to probe interlayer vibronic states, as demonstrated here, differs strongly from previous experiments on vdW heterostructures. Recently, interlayer exciton dipole strengths were estimated by studying an electrically controlled Stark shift through purely optical signals,30,31 and no evidence of a vibronic registry was reported. In our p–n junction photodiode, we can utilize the VSD-controlled optoelectronic response to estimate the dipole strength. As an illustration, we note that the separation between periodic vibronic peaks in Figure Figure22f approximately corresponds to 30 meV, yielding a modest estimate η = t/d ≈ 2.5 (consistent with recent measurements of interlayer exciton dipole strengths34). This yields energy-dependent d2IPC/dΔE2 vs ΔE oscillations with a 30 meV increment (Figure Figure33d,f). In these (Figure Figure33e), each pronounced dip corresponds to a vibronic state ωvib,n (VSD).
These results also differ from previous experiments on individual monolayer vdW materials. Although the absorption spectra of monolayer TMDs generally display broad absorption peaks,31−34 optical measurements of some vdW materials have attributed sparse and isolated sidebands to individual vibrational modes.35−39 In MoSe240,41 and WSe2,42,43 optical spectroscopy measurements have resolved sideband features attributed to strong intralayer exciton–phonon coupling yet reported only at very low temperatures. Here, photoconductance oscillations are clearly observed at room temperature (Figure Figure33), indicating that the photocurrent is a sensitive probe of interlayer excitons, which are particularly susceptible to exciton–phonon coupling that emerges from the stacking of two vdW crystal layers. Combining the photocurrent with other advanced optical and optoelectronic techniques may allow the further elucidation of the fine details of the electron–phonon coupling in the heterostructure (e.g., distinguishing adiabatic5 and nonadiabatic/resonant couplings44) and the measurement of its strength (e.g., determining whether the vibronic coupling is strong enough to induce self-trapping of the excitons7,9). Such advanced techniques could include coherent optical spectroscopy such as 2D electronic spectroscopy or methods to enhance the observation and identification of relevant phonons (e.g., resonant Raman).
By stacking atomically thin semiconductors, we have demonstrated a new type of device that harnesses both vibrational and electronic energy to absorb near-infrared light, an important part of the solar light spectrum. In solid-state optoelectronics, atomic vibrations are often thought of as a loss mechanism. For example, the excess energy of photoexcited electrons above the band edge is typically lost to phonons, thus lowering the efficiency of solar cells. This paradigm is turned on its head in photosynthetic complexes, wherein atomic vibrations are instead harnessed to enhance energy transport through strong vibronic coupling. Our device, a rudimentary vibronic photodiode, exhibits vibronic effects that are often observed in photosynthesis yet have not been harnessed in solid-state devices. Stack engineering, exemplified here by stacking two vdW layers, MoSe2 and WSe2, gives rise to an entire registry of exciton–phonon vibronic states that can be individually addressed and electrically controlled in a solid-state setting. From a broader perspective, access to vibronic states in these vdW layers may enable excitonic phenomena more traditionally found in molecule-like systems, ranging from singlet fission45 and exciton dissociation46 in organic compounds to long-lived coherent dynamics and enhancements to exciton transport in photosynthetic complexes.1−4,47−49 We anticipate that stack engineering of strong exciton–phonon coupling will establish vdW heterojunctions as a versatile platform for controlling vibronic physics in 2D semiconductors devices.
Acknowledgments
The authors acknowledge valuable discussions with Vasili Perebeinos. This work was supported by the Army Research Office Electronics Division Award no. W911NF2110260 (N.M.G., V.A., and T.B.A.), the Presidential Early Career Award for Scientists and Engineers (PECASE) through the Air Force Office of Scientific Research (award no. FA9550-20-1-0097; N.M.G. and T.B.A.), through support from the National Science Foundation Division of Materials Research CAREER Award (no. 1651247; N.M.G. and F.B.), and through the United States Department of the Navy Historically Black Colleges, Universities and Minority Serving Institutions (HBCU/MI) award no. N00014-19-1-2574 (N.M.G. and F.B.). T.B.A. acknowledges support from the Fellowships and Internships in Extremely Large Data Sets (FIELDS) program, a NASA MUREP Institutional Research Opportunity (MIRO) program (grant no. NNX15AP99A). R.v.G. acknowledges institutional support from the Royal Netherlands Academy of Arts and Sciences (KNAW) and the Canadian Institute of Solar Energy Research (CEA). J.C.W.S. acknowledges support from the Singapore Ministry of Education under its MOE AcRF Tier 3 Award (MOE2018-T3-1-002). M.S.R is grateful for the support of the European Research Council (ERC) under the European Union Horizon 2020 Research and Innovation Programme (grant agreement no. 678862) and the Villum Foundation. R.K.L. and S.S. acknowledge support from the NSF (EFRI-1433395). This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant no. ACI-1053575 and allocation ID TG-DMR130081.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.nanolett.2c00944
Detailed description of the three experimental techniques used to identify vibronic exciton–phonon states; numerical calculations of the phonon dispersion within the heterostructure; several candidate modes and detailed microscopic descriptions of atomic vibrations; additional photocurrent spectroscopy and scanning photocurrent data analysis; and a review of the theory of Franck–Condon transitions (PDF)
Author Contributions
F.B. and T.B.A. contributed equally to this work, performing device fabrication, detailed experiments, analysis, and modeling. N.M.G. conceived the experiments and supervised the experimental, theoretical, and computational components with additional input from R.K.L., V.A., R.v.G., M.S.R., and J.C.W.S. S.S. and R.L. conducted detailed computations of the vibrational band structure to support the interpretation. V.A., M.S.R. and J.C.W.S. provided theoretical support, and R.v.G. advised on experimental concepts. All authors contributed to the writing of the manuscript. The code that generates Figures 1–3 and all supporting information figures that present additional data is published alongside this work. The code can be found at https://github.com/qmolabucr/wse2mose2, and this repository includes all relevant data such that the results can be fully replicated.50
References
- Lee H.; Cheng Y.; Fleming G. R. Coherence dynamics in photosynthesis: protein protection of excitonic coherence. Science 2007, 316, 1462–1465. 10.1126/science.1142188. [Abstract] [CrossRef] [Google Scholar]
- Mohseni M.; Rebentrost P.; Lloyd S.; Aspuru-Guzik A. Environment-assisted quantum walks in photosynthetic energy transfer. J. Chem. Phys. 2008, 129, 174106.10.1063/1.3002335. [Abstract] [CrossRef] [Google Scholar]
- Collini E.; Wong C. Y.; Wilk K. E.; Curmi P. M. G.; Brumer P.; Scholes G. D. Coherently wired light-harvesting in photosynthetic marine algae at ambient temperature. Nature 2010, 463, 644.10.1038/nature08811. [Abstract] [CrossRef] [Google Scholar]
- Ma F.; Romero E.; Jones M. R.; Novoderezhkin V. I.; Van Grondelle R. Both electronic and vibrational coherences are involved in primary electron transfer in bacterial reaction center. Nat. Commun. 2019, 10, 933.10.1038/s41467-019-08751-8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Toyozawa Y.. Optical Processes in Solids; Cambridge University Press: Cambridge, U.K., 2003. [Google Scholar]
- Perebeinos V.; Tersoff J.; Avouris P. Effect of Exciton-Phonon Coupling in the Calculated Optical Absorption of Carbon Nanotubes. Phys. Rev. Lett. 2005, 94, 027402.10.1103/PhysRevLett.94.027402. [Abstract] [CrossRef] [Google Scholar]
- Rashba E. I.. Excitons; North-Holland Publishing Company: Amsterdam, 1982; Chapter 13, pp 543–597. [Google Scholar]
- Grover M. K.; Silbey R. Exciton-Phonon interactions in molecular crystals. J. Chem. Phys. 1970, 52, 2099.10.1063/1.1673263. [CrossRef] [Google Scholar]
- Toyozawa Y. Phonon structures in the spectra of solids. J. Lumin. 1970, 1, 736–746. [Google Scholar]
- Koch J.; von Oppen F. Franck-Condon blockade and giant Fano factors in transport through single molecules. Phys. Rev. Lett. 2005, 94, 206804.10.1103/PhysRevLett.94.206804. [Abstract] [CrossRef] [Google Scholar]
- Leturcq R.; Stampfer C.; Inderbitzin K.; Durrer L.; Hierold C.; Mariani E.; Schultz M. G.; von Oppen F.; Ensslin K. Franck-Condon blockade in suspended carbon nanotube quantum dots. Nat. Phys. 2009, 5, 327–331. 10.1038/nphys1234. [CrossRef] [Google Scholar]
- Toyozawa Y.; Hermanson J. Exciton-phonon bound state: a new quasiparticle. Phys. Rev. Lett. 1968, 21, 1637.10.1103/PhysRevLett.21.1637. [CrossRef] [Google Scholar]
- Itoh T.; Nishijima M.; Ekimov A. I.; Gourdon C.; Efros AI. L.; Rosen M. Polaron and exciton-phonon complexes in CuCl Nanocrystals. Phys. Rev. Lett. 1995, 74, 1645.10.1103/PhysRevLett.74.1645. [Abstract] [CrossRef] [Google Scholar]
- Harris D. C.; Bertolucci M. D.Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic Spectroscopy; Dover Publications: New York, 1989. [Google Scholar]
- Rivera P.; Yu H.; Seyler K.; Wilson N. P.; Yao W.; Xu X. Interlayer valley excitons in heterobilayers of transition metal dichalcogenides. Nat. Nanotechnol. 2018, 13, 1004–1015. 10.1038/s41565-018-0193-0. [Abstract] [CrossRef] [Google Scholar]
- See the Supporting Information for a discussion of the device fabrication, experimental details, a discussion of phonon modes, as well as a brief theoretical description of Franck–Condon progressions. Replication Data can be found in ref (50). The Supporting Information also contain refs (51−64).
- Barati F.; Grossnickle M.; Su S.; Lake R. K.; Aji V.; Gabor N. M. Hot carrier-enhanced interlayer electron–hole pair multiplication in 2D semiconductor heterostructure photocells. Nat. Nanotechnol. 2017, 12, 1134–1139. 10.1038/nnano.2017.203. [Abstract] [CrossRef] [Google Scholar]
- Karni O.; Barré E.; Lau S. C.; Gillen R.; Ma E. Y.; Kim B.; Watanabe K.; Taniguchi T.; Maultzsch J.; Barmak K.; Page R. H.; Heinz T. F. Infrared Interlayer Exciton Emission in MoS2/WSe2 Heterostructures. Phys. Rev. Lett. 2019, 123, 247402.10.1103/PhysRevLett.123.247402. [Abstract] [CrossRef] [Google Scholar]
- Unuchek D.; Ciarrocchi A.; Avsar A.; Watanabe K.; Taniguchi T.; Kis A. Room-temperature electrical control of exciton flux in a van der Waals heterostructure. Nature 2018, 560, 340–344. 10.1038/s41586-018-0357-y. [Abstract] [CrossRef] [Google Scholar]
- Rivera P.; Seyler K. L.; Yu H.; Schaibley J. R.; Yan J.; Mandrus D. G.; Yao W.; Xu X. Valley-polarized exciton dynamics in a 2D semiconductor heterostructure. Science 2016, 351, 688–691. 10.1126/science.aac7820. [Abstract] [CrossRef] [Google Scholar]
- Rivera P.; Schaibley J. R.; Jones A. M.; Ross J. S.; Wu S.; Aivazian G.; Klement P.; Seyler K.; Clark G.; Ghimire N. J.; Yan J.; Mandrus D. G.; Yao W.; Xu X. Observation of long-lived interlayer excitons in monolayer MoSe2–WSe2 heterostructures. Nat. Commun. 2015, 6, 6242.10.1038/ncomms7242. [Abstract] [CrossRef] [Google Scholar]
- Rigosi A. F.; Hill H. M.; Li Y.; Chernikov A.; Heinz T. F. Probing interlayer interactions in transition metal dichalcogenide heterostructures by optical spectroscopy: MoS2/WS2 and MoSe2/WSe2. Nano Lett. 2015, 15, 5033–5038. 10.1021/acs.nanolett.5b01055. [Abstract] [CrossRef] [Google Scholar]
- Furchi M. M.; Pospischil A. A.; Libisch F.; Burgdörfer J.; Mueller T. Photovoltaic effect in an electrically tunable van der Waals heterojunction. Nano Lett. 2014, 14, 4785–4791. 10.1021/nl501962c. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Lee C.; Lee G.; van der Zande A. M.; Chen W.; Li Y.; Han M.; Cui X.; Arefe G.; Nuckolls C.; Heinz T. F.; Guo J.; Hone J.; Kim P. Atomically thin p–n junctions with van der Waals heterointerfaces. Nat. Nanotechnol. 2014, 9, 676–681. 10.1038/nnano.2014.150. [Abstract] [CrossRef] [Google Scholar]
- Ross J.; Rivera P.; Schaibley J.; Lee-Wong E.; Yu H.; Taniguchi T.; Watanabe K.; Yan J.; Mandrus D.; Cobden D.; Yao W.; Xu X. Interlayer exciton optoelectronics in a 2D heterostructure p–n junction. Nano Lett. 2017, 17, 638–643. 10.1021/acs.nanolett.6b03398. [Abstract] [CrossRef] [Google Scholar]
- Buscema M.; Island J. O.; Groenendijk D. J.; Blanter S. I.; Steele G. A.; van der Zant H. S. J.; Castellanos-Gomez A. Photocurrent generation with two-dimensional van der Waals semiconductors. Chem. Soc. Rev. 2015, 44, 3691–3718. 10.1039/C5CS00106D. [Abstract] [CrossRef] [Google Scholar]
- Frisenda R.; Molina-Mendoza A. J.; Mueller T.; Castellanos-Gomez A.; Van der Zant H. S. J. Atomically thin p–n junctions based on two-dimensional materials. Chem. Soc. Rev. 2018, 47, 3339–3358. 10.1039/C7CS00880E. [Abstract] [CrossRef] [Google Scholar]
- Kunstmann J.; Mooshammer F.; Nagler P.; Chaves A.; Stein F.; Paradiso N.; Strucnk Ch.; Schuller Ch.; Seifert G.; Reichman D. R.; Korn T. Momentum-space indirect interlayer excitons in transition-metal dichalcogenide van der Waals heterostructures. Nature Phys. 2018, 14, 801–805. 10.1038/s41567-018-0123-y. [CrossRef] [Google Scholar]
- Bosnick K. Transport in carbon nanotube p-i-n diodes. Appl. Phys. Lett. 2006, 89, 163121.10.1063/1.2360895. [CrossRef] [Google Scholar]
- Jauregui L. A.; Joe A. Y.; Pistunova K.; Wild D. S.; High A. A.; Zhou Y.; Scuri G.; De Greve K.; Sushko A.; Yu C.-H.; Taniguchi T.; Watanabe K.; Needleman D. J.; Lukin M. D.; Park H.; Kim P. Electrical control of interlayer exciton dynamics in atomically thin heterostructures. Science 2019, 366 (6467), 870–875. 10.1126/science.aaw4194. [Abstract] [CrossRef] [Google Scholar]
- Peimyoo N.; Deilmann T.; Withers F.; Escolar J.; Nutting D.; Taniguchi T.; Watanabe K.; Taghizadeh A.; Craciun M. F.; Thygesen K. S.; Russo S. Electrical tuning of optically active interlayer excitons in bilayer MoS2. Nat. Nano. 2021, 16, 888–893. 10.1038/s41565-021-00916-1. [Abstract] [CrossRef] [Google Scholar]
- Chernikov A.; Berkelbach T. C.; Hill H. M.; Rigosi A.; Li Y.; Aslan O. B.; Reichman D. R.; Hybertsen M. S.; Heinz T. F. Exciton binding energy and nonhydrogenic rydberg series in monolayer WS2. Phys. Rev. Lett. 2014, 113, 076802.10.1103/PhysRevLett.113.076802. [Abstract] [CrossRef] [Google Scholar]
- Wang G.; Chernikov A.; Glazov M. M.; Heinz T. F.; Marie X.; Amand T.; Urbaszek B. Colloquium: Excitons in atomically thin transition metal dichalcogenides. Rev. Mod. Phys. 2018, 90, 021001.10.1103/RevModPhys.90.021001. [CrossRef] [Google Scholar]
- Mueller T.; Malic E. Exciton physics and device application of two-dimensional transition metal dichalcogenide semiconductors. npj 2D Mater. Appl. 2018, 2, 29.10.1038/s41699-018-0074-2. [CrossRef] [Google Scholar]
- Cassabois G.; Valvin P.; Gil B. Hexagonal boron nitride is an indirect bandgap semiconductor. Nat. Photonics 2016, 10, 262.10.1038/nphoton.2015.277. [CrossRef] [Google Scholar]
- Cannuccia E.; Monserrat B.; Attaccalite C. Theory of phonon-assisted luminescence in solids: application to hexagonal boron nitride. Phys. Rev. B 2019, 99, 08110910.1103/PhysRevB.99.081109. [CrossRef] [Google Scholar]
- Paleari F.; Miranda H. P. C.; Molina-Sánchez A.; Wirtz L. Exciton-phonon coupling in the ultraviolet absorption and emission spectra of bulk hexagonal boron nitride. Phys. Rev. Lett. 2019, 122, 187401.10.1103/PhysRevLett.122.187401. [Abstract] [CrossRef] [Google Scholar]
- Chen H.-Y.; Sangalli D.; Bernardi M. Exciton-phonon interaction and relaxation times from first principles. Phys. Rev. Lett. 2020, 125, 107401.10.1103/PhysRevLett.125.107401. [Abstract] [CrossRef] [Google Scholar]
- Jin C.; Kim J.; Suh J.; Shi Z.; Chen B.; Fan X.; Kam M.; Watanabe K.; Taniguchi T.; Tongay S.; Zettl A.; Wu J.; Wang F. Interlayer electron-phonon coupling in WSe2/hBN heterostructures. Nat. Phys. 2017, 13, 127.10.1038/nphys3928. [CrossRef] [Google Scholar]
- Chow C.; Yu H.; Jones A.; Schaibley J.; Koehler M.; Mandrus D.; Merlin R.; Yao W.; Xu X. Phonon-assisted oscillatory exciton dynamics in monolayer MoSe2. npj 2D Materials and Applications 2017, 1, 33.10.1038/s41699-017-0035-1. [CrossRef] [Google Scholar]
- Shree S.; Semina M.; Robert C.; Amand T.; Balocchi A.; Manca M.; Courtade E.; Marie X.; Taniguchi T.; Watanabe K.; Glazov M.; Urbaszek B. Observation of exciton-phonon coupling in MoSe2 monolayers. Phys. Rev. B 2018, 98, 035302.10.1103/PhysRevB.98.035302. [CrossRef] [Google Scholar]
- Lin K.-Q.; Ong C.; Bange S.; Faria Junior P.; Peng B.; Ziegler J.; Zipfel J.; Bauml C.; Paradiso N.; Watanabe K.; Taniguchi T.; Strunk C.; Monserrat B.; Fabian J.; Chernikov A.; Qiu D.; Louie S.; Lupton J. Narrow-band high-lying excitons with negative-mass electrons in monolayer WSe2. Nat. Commun. 2021, 12, 5500.10.1038/s41467-021-25499-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Funk V.; Wagner K.; Wietek E.; Ziegler J.; Forste J.; Lindlau J.; Forg M.; Watanabe K.; Taniguchi T.; Chernikiv A.; Hogele A. Spectral asymmetry of phonon sideband luminescence in monolayer and bilayer WSe2. Physical Review Research 2021, 3, L042019.10.1103/PhysRevResearch.3.L042019. [CrossRef] [Google Scholar]
- Reichardt S.; Wirtz L. Nonadiabatic exciton-phonon coupling in Raman spectroscopy of layered materials. Science Adv. 2020, 6, eabb591510.1126/sciadv.abb5915. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Stern H. L.; Cheminal A.; Yost S. R.; Broch K.; Bayliss S. L.; Chen K.; Tabachnyk M.; Thorley K.; Greenham N.; Hodgkiss J. M.; Anthony J.; Head-Gordon M.; Musser A. J.; Rao A.; Friend R. H. Vibronically coherent ultrafast triplet-pair formation and subsequent thermally activated dissociation control efficient endothermic singlet fission. Nat. Chem. 2017, 9, 1205–1212. 10.1038/nchem.2856. [Abstract] [CrossRef] [Google Scholar]
- Bian Q.; Ma F.; Chen S.; Wei Q.; Su X.; Buyanova I. A.; Chen W. M.; Ponseca C. S. Jr; Linares M.; Karki K. J.; Yartsev A.; Inganäs O. Vibronic coherence contributes to photocurrent generation in organic semiconductor heterojunction diodes. Nat. Commun. 2020, 11, 617.10.1038/s41467-020-14476-w. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Novoderezhkin V. I.; Romero E.; Prior J.; van Grondelle R. Exciton-vibrational resonance and dynamics of charge separation in the photosystem II reaction center. Phys. Chem. Chem. Phys. 2017, 19, 5195–5208. 10.1039/C6CP07308E. [Abstract] [CrossRef] [Google Scholar]
- Romero E.; Auguli R.; Novoderezhkin V. I.; Faretti M.; Thieme J.; Zigmantas D.; van Grondelle R. Quantum coherence in photosynthesis for efficient solar-energy conversion. Nat. Phys. 2014, 10, 676–682. 10.1038/nphys3017. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Romero E.; Novoderezhkin V. I.; van Grondelle R. Quantum design of photosynthesis for bio-inspired solar energy conversion. Nature 2017, 543, 355–365. 10.1038/nature22012. [Abstract] [CrossRef] [Google Scholar]
- Arp T. B.. qmolabucr/wse2mose2: Publication Release, version v1.2; Zenodo, 2021; 10.5281/zenodo.5516680. [CrossRef]
- Barati F.. Optoelectronics investigations of electron dynamics in 2D-TMD semiconductor heterostructure photocells: from electron-hole pair multiplication to phonon assisted anti-Stokes absorption. Thesis, University of California, Riverside, CA, 2018. [Google Scholar]
- Arp T. B.; Gabor N. M. Multiple parameter dynamic photoresponse microscopy for data-intensive optoelectronic measurements of van der Waals heterostructures. Rev. Sci. Instrum. 2019, 90, 023702.10.1063/1.5085007. [Abstract] [CrossRef] [Google Scholar]
- Kresse G.; Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169.10.1103/PhysRevB.54.11169. [Abstract] [CrossRef] [Google Scholar]
- Kresse G.; Hafner J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558.10.1103/PhysRevB.47.558. [Abstract] [CrossRef] [Google Scholar]
- Kresse G.; Furthmüller J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15.10.1016/0927-0256(96)00008-0. [Abstract] [CrossRef] [Google Scholar]
- Blöchl P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953.10.1103/PhysRevB.50.17953. [Abstract] [CrossRef] [Google Scholar]
- Perdew J. P.; Chevary J. A.; Vosko S. H.; Jackson K. A.; Pederson M. R.; Singh D. J.; Fiolhais C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671.10.1103/PhysRevB.46.6671. [Abstract] [CrossRef] [Google Scholar]
- Wang Y.; Perdew J. P. Correlation hole of the spin-polarized electron gas, with exact small-wave-vector and high-density scaling. Phys. Rev. B 1991, 44, 13298.10.1103/PhysRevB.44.13298. [Abstract] [CrossRef] [Google Scholar]
- Kresse G.; Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758.10.1103/PhysRevB.59.1758. [CrossRef] [Google Scholar]
- Grimme S. Semiempirical gga-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787.10.1002/jcc.20495. [Abstract] [CrossRef] [Google Scholar]
- Togo A.; Tanaka I. First principles phonon calculations in materials science. Scripta Mater. 2015, 108, 1.10.1016/j.scriptamat.2015.07.021. [CrossRef] [Google Scholar]
- Grosso G.; Grosso; Parravicini G. P.Solid State Physics; Academic Press: London, 2000. [Google Scholar]
- Franck J. Elementary processes of photochemical reactions. Trans. Faraday Soc. 1926, 21, 536–542. 10.1039/tf9262100536. [CrossRef] [Google Scholar]
- Condon E. U. A Theory of Intensity Distribution in Band Systems. Phys. Rev. 1926, 28, 1182.10.1103/PhysRev.28.1182. [CrossRef] [Google Scholar]
Full text links
Read article at publisher's site: https://doi.org/10.1021/acs.nanolett.2c00944
Read article for free, from open access legal sources, via Unpaywall: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9335870
Citations & impact
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/130696410
Smart citations by scite.ai
Explore citation contexts and check if this article has been
supported or disputed.
https://scite.ai/reports/10.1021/acs.nanolett.2c00944
Article citations
Exciton-assisted electron tunnelling in van der Waals heterostructures.
Nat Mater, 22(9):1094-1099, 26 Jun 2023
Cited by: 1 article | PMID: 37365227 | PMCID: PMC10465355
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Funding
Funders who supported this work.
Air Force Office of Scientific Research (1)
Grant ID: FA9550-20-1-0097
Army Research Office (1)
Grant ID: W911NF2110260
European Research Council (1)
Grant ID: 678862
Ministry of Education - Singapore (1)
Grant ID: MOE2018-T3- 1-002
National Aeronautics and Space Administration (1)
Grant ID: NNX15AP99A
National Science Foundation (2)
Grant ID: EFRI-1433395
Grant ID: 1651247
Grant ID: N00014-19-1-2574