Strain-modulated excitonic gaps in mono- and bi-layer MoSe2
Ji Jianting, Zhang Anmin, Xia Tianlong, Gao Po, Jie Yinghao, Zhang Qian, Zhang Qingming†,
Department of Physics, Beijing Key Laboratory of Opto-Electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China

 

† Corresponding author. E-mail: qmzhang@ruc.edu.cn

Project supported by the National Basic Research Program of China (Grant No. 2012CB921701) and the National Natural Science Foundation of China (Grant Nos. 11474357 and 11004245). Qingming Zhang and Tianlong Xia were supported by the Fundamental Research Funds for the Central Universities of China and the Research Funds of Renmin University of China.

Abstract
Abstract

Photoluminescence (PL) and Raman spectra under uniaxial strain were measured in mono- and bi-layer MoSe2 to comparatively investigate the evolution of excitonic gaps and Raman phonons with strain. We observed that the strain dependence of excitonic gaps shows a nearly linear behavior in both flakes. One percent of strain increase gives a reduction of ∼ 42 meV (∼ 35 meV) in A-exciton gap in monolayer (bilayer) MoSe2. The PL width remains little changed in monolayer MoSe2 while it increases rapidly with strain in the bilayer case. We have made detailed discussions on the observed strain-modulated results and compared the difference between monolayer and bilayer cases. The hybridization between 4d orbits of Mo and 4p orbits of Se, which is controlled by the Se–Mo–Se bond angle under strain, can be employed to consistently explain the observations. The study may shed light into exciton physics in few-layer MoSe2 and provides a basis for their applications.

1. Introduction

Most transition metal disulfides MX2 (M = Mo, W, Ta, Re; X = S, Se, Te) are semiconductors with outstanding performance and have been widely investigated and applied. Two-dimensional MX2 has the similar intra-layer honeycomb lattice as graphene and the van der Waals interaction is responsible for the weak interlayer coupling. This allows one to obtain few-layer MX2 flakes by mechanical exfoliation from bulk crystals just like graphene. Compared to pristine graphene with zero energy gap, MX2 has a gap of 1–2 eV, which is a large advantage for applications.[13] Furthermore, its higher mobility[4] is particularly essential to semiconducting devices such as logical devices,[3,5] sensors,[6] photo-diodes,[7] and photo-detectors.[8] The unique valley polarization revealed in MX2 provides us with a new degree of freedom for device design.[3,913]

External strain is a conventional and convenient way to tune energy gap and lattice vibrations in low-dimensional semiconductors. The subtle changes induced by strain can be easily probed by the micro-Raman technique. In fact, the strain-modulated effects have been extensively investigated in graphene[14,15] and MoS2.[1619] First-principles calculations predicted that strain can substantially change the band structure of monolayer MX2 and drive a transition from direct gap to indirect gap.[2024] Interestingly, strain can also modulate magnetic properties of few-layer MX2[20,21] and provide possible scenarios for some optical devices or solar cells.[25] Among few-layer MX2 compounds, MoSe2 has attracted special attention since its energy gap under zero strain (∼ 1.55 eV), is exactly close to the optimum value balanced between the optimum utilization of the solar spectrum and the maximum efficiency of conversion into electricity.[26] This means that two-dimensional MoSe2 may play an important role in the future opto-electric applications.

In this work, we carried out careful Raman scattering studies of strain-induced effects in mono- and bi-layer MoSe2. The linear dependence of energy gaps on strain was observed and quantitatively determined. The strain effect can be well understood in term of the change of Se–Mo–Se bond angle. We have comparatively discussed the strain effects, which may help us get insight into exciton physics in MoSe2 and provide a solid basis for applications.

2. Experimental methods

MoSe2 bulk crystals were grown through vapor transport equilibration under strict conditions.[27] Similar to graphene, mono- and bi-layer MoSe2 flakes used in this work were mechanically exfoliated from bulk crystals but transferred to flexible polyethylene terephthalate (PET) substrate instead of conventional SiO2/Si substrate, to carry out measurements under strain. As shown in Fig. 1(a), the monolayer piece is a nearly isosceles triangular area with a side length of ∼ 8 microns. The bilayer piece is the neighboring parallelogram. Monolayer MoSe2 flake was determined by the A1g phonon frequency (240.5 cm− 1, Fig. 1(b)) and the PL peak position located at 1.55 eV. The extremely strong PL intensity under very low-intensity excitation light further confirmed its thickness since it is a direct-gap semiconductor only in the monolayer case. Then one can further determine the bilayer flake which becomes an indirect-gap semiconductor and its PL intensity is lower than the monolayer case by a factor of ∼ 1/20.

Fig. 1. (a) Optical image of monolayer and bilayer MoSe2. (b) Raman phonon in the monolayer case. (c) Home-made setup for applying strain on films. (d) Schematic of quantitatively estimated strain. The arc length of bended PET substrate is L and its thickness is d. If the straight-line distance between any two points on the PET substrate is taken as x, one can calculate the radius of curvature R. Tuning R by changing x gives variable strains e. For accuracy, the two points should be as close as possible to the arc top.

The home-made setup for applying strain is illustrated in Fig. 1(c). By changing the screw positions, one can adjust the curvature radius of the PET substrate.[18] External strain can be applied to the flakes through the mediation of van der Waals interaction between flake and substrate. The calculation details can be found in Fig. 1(d). The applied strain e is defined as e = d/2R in units of mm. The maximum applied strain was limited to 0.75%, since the flake may slip off from the substrate under a strain higher than 0.8%.[18]

PL and Raman scattering measurements were performed with a Jobin–Yvon HR800 system equipped with liquid-nitrogen cooled CCD and a He–Ne excitation laser of 633 nm (MellesGriot). The configuration of backscattering was adopted in our measurements. The laser power was monitored and controlled at a level of 60 μW to protect the flakes and the focus spot is around 2 microns in diameter. All the measurements were carried out at room temperature.

3. Results and discussion

PL spectra of mono- and bi-layer MoSe2 under various strains are shown in Fig. 2. The evolutions of PL peaks with strain are distinguished in two flakes. PL peak position or energy gap of A-exciton moves down almost linearly with increasing strain in the monolayer sample, and its width remains little changed. In bilayer MoSe2, PL intensities are significantly weakened because of its indirect gap. Its position has the similar linear strain dependence but its width is obviously reduced by strain. The PL peak positions and widths extracted from Fig. 2 are summarized in Fig. 3. The linear behaviors are clearly seen in both samples in Fig. 3. The fitting in the monolayer case gives a slope of −42 meV/% strain, which is well consistent with first-principles calculations using the GGA method.[28,29] The calculated rate of strain change is −45 meV/% strain for the direct gap. In the bilayer sample, the fitted slope is −35 meV/% strain, slightly smaller than that in the monolayer case. In comparison, the slope is −45 meV/% strain and −53 meV/% strain in mono- and bilayer MoS2,[17] respectively. Considering the different energy gaps (1.55 eV and 1.82 eV for monolayer MoSe2 and MoS2, respectively), we obtained the rate of energy-gap change with strain, ∼ 2.7% per 1% strain for MoSe2 and ∼ 2.4% per 1% strain for MoS2. This means that the excitonic gap of MoSe2 is even more sensitive to strain than MoS2. However, this is reversed in the bilayer case, i.e., 2.3% per 1% strain for MoSe2 and 2.9% per 1% strain for MoS2.

Fig. 2. PL spectra in monolayer (a) and bilayer (b) MoSe2 with applying strain from 0 to 0.75%.
Fig. 3. (a) Strain dependence of A-exciton PL peak positions in mono- and bi-layer MoSe2. (b) Strain dependence of A-exciton PL peak widths in mono- and bi-layer MoSe2.

The difference can be understood in terms of the different interlayer coupling in both samples. Generally the uniaxial strain stretches the Mo–Se bond and reduces the Se–Mo–Se bond angle. The valence bands of MoSe2 contain five d orbits of Mo and 4p orbits of Se. In detail, the conduction bands at K points and the valence bands at Γ points are dominated by orbits, while dx2y2 and dxy orbits contribute to the valence bands at K points. The rest dxz, dyz orbits and 4p orbits of Se have a smaller contribution. The application of in-plane uniaxial strain increases the distance between Mo and Se atoms and reduces the Mo–Se–Mo bond angle. Consequently, it brings an effective increase of the vertical distance between Mo and Se atoms, and hence a reduction of the hybridization between pz orbits of Se and orbits of Mo. This moves down the bottom of the conduction bands at K points and moves down the top of the valence bands at Γ points.

On the other hand, the decrease of the Mo–Se–Mo bond angle under strain reduces the distance between Mo and Se layers. This enhances the coupling between px/py orbits of Se and dx2y2 and dxy orbits of Mo. However, the absolute distance between Mo and Se is against the enhancement. The net effect of strain on the valence band top becomes very small. Taking all the factors discussed above, we will see that strain modulates the direct gap at K points by changing the conduction band bottom and keeping the valence band top fixed. Larger strain gives a smaller gap and vice versa. If strain is too large to move up the valence band top at Γ points over that at K points, monolayer MoSe2 will become an indirect-gap semiconductor.[2024,2830]

Figure 3(b) shows the evolution of PL peak width with strain. Under zero strain, the PL peak of the bilayer sample is about two times wider than that in the monolayer one. The width remains unchanged with strain in monolayer MoSe2, while it clearly becomes sharper with strain in the bilayer sample. This can also be explained by the reduction of Se–Mo–Se bond angle under strain.[30] The angle reduction effectively decreases the thickness of the MoSe2 layer or is equivalent to the increase of the interlayer distance, which weakens the interlayer coupling and decreases the A-exciton PL peak width.

4. Summary

We first observed the direct energy-gap evolution with strain in mono- and bilayer MoSe2. The measured rates of change are consistent with the calculated ones. We further discussed the possible mechanisms in both cases in terms of interlayer coupling. The reduction of the Mo–Se–Mo bond angle under strain modifies the hybridization between the d orbits and p orbits of Se, and further shifts the conduction and valence bands. This effectively reduces the band gap. The studies on excitonic energy gaps under strain are important to the applications and provide fundamental insight to exciton physics in this two-dimensional system.

Reference
1Mak K FLee CHone JShan JHeinz T F 2010 Phys. Rev. Lett. 105 136805
2Splendiani ASun LZhang YLi TKim JChim C YGalli GWang F 2010 Nano Lett. 10 1271
3Xiao DLiu G BFeng WXu XYao W 2012 Phys. Rev. Lett. 108 196802
4Radisavljevic BRadenovic ABrivio JGiacometti VKis A 2011 Nat. Nanotechnol. 6 147
5Radisavljevic BKis A 2013 Nat. Mater. 12 815
6Perkins F KFriedman A LCobas ECampbell P MJernigan G GJonker B T 2013 Nano Lett. 13 668
7Sundaram R SEngel MLombardo AKrupke RFerrari A CAvouris PSteiner M 2013 Nano Lett. 13 1416
8Yin ZLi HLi HJiang LShi YSun YLu GZhang QChen XZhang H 2012 ACS Nano 6 74
9Xiao DYao WNiu Q 2007 Phys. Rev. Lett. 99 236809
10Yao WXiao DNiu Q 2008 Phys. Rev. 77 235406
11Cao TWang GHan WYe HZhu CShi JNiu QTan PWang ELiu BFeng J 2012 Nat. Commun. 3 887
12Mak K FHe KShan JHeinz T F 2012 Nat. Nanotechnol. 7 494
13Zeng HDai JYao WXiao DCui X 2012 Nat. Nanotechnol. 7 490
14Huang MYan HChen CSong DHeinz T FHone J 2009 Proc. Natl. Acad. Sci. U.S.A. 106 7304
15Mohiuddin T M GLombardo ANair R RBonetti ASavini GJalil RBonini NBasko D MGaliotis CMarzari NNovoselov K SGeim A KFerrari A C 2009 Phys. Rev. 79 205433
16Rice CYoung R JZan RBangert UWolverson DGeorgiou TJalil RNovoselov K S 2013 Phys. Rev. 87 081307(R)
17Conley H JWang BZiegler J IHaglund R FJr.Pantelides S TBolotin K I 2013 Nano Lett. 13 3626
18Zhu C RWang GLiu B LMarie XQiao X FZhang XWu X XFan HTan P HAmand TUrbaszek B 2013 Phys. Rev. 88 121301(R)
19He KPoole CMak K FShan J 2013 Nano Lett. 13 2931
20Lu PWu XGuo WZeng X C 2012 Phys. Chem. Chem. Phys. 14 13035
21Pan HZhang Y W 2012 J. Phys. Chem. 116 11752
22Li T 2012 Phys. Rev. 85 235407
23Shi HPan HZhang Y WYakobson B I 2013 Phys. Rev. 87 155304
24Horzum SSahin HCahangirov SCudazzo PRubio ASerin TPeeters F M 2013 Phys. Rev. 87 125415
25Feng JQian XHuang C WLi J2012Nat. Photo.6865
26Loferski J J 1956 J. Appl. Phys. 27 777
27Bougouma MBatan AGuel BSegato TLegma J BReniers FDelplancke-Ogletree M PBuess-Herman CDoneux T 2013 J. Cryst. Growth 363 122
28Kumar AAhluwalia P K 2013 Physica B-Conden. Matt. 419 66
29Kumar AAhluwalia P K 2013 Modelling Simul. Mater. Sci. Eng. 21 065015
30Chang C HFan XLin S HKuo J L 2013 Phys. Rev. 88 195420