Defects Activated Photoluminescence in Two-Dimensional Semiconductors: Interplay Between Bound, Charged, and Free Excitons
Defects Activated Photoluminescence in Two-Dimensional Semiconductors: Interplay Between Bound, Charged, and Free Excitons
Defects Activated Photoluminescence in Two-Dimensional Semiconductors: Interplay Between Bound, Charged, and Free Excitons
SUBJECT AREAS:
two-dimensional semiconductors:
TWO-DIMENSIONAL
MATERIALS interplay between bound, charged, and
ELECTRONIC PROPERTIES AND
MATERIALS
APPLIED PHYSICS
free excitons
SYNTHESIS AND PROCESSING Sefaattin Tongay1,2*, Joonki Suh1,2*, Can Ataca3, Wen Fan1, Alexander Luce1,4, Jeong Seuk Kang1,
Jonathan Liu1, Changhyun Ko1, Rajamani Raghunathanan3, Jian Zhou1, Frank Ogletree4, Jingbo Li2,
Received Jeffrey C. Grossman3 & Junqiao Wu1,2,4
16 May 2013
1
Accepted Department of Materials Science and Engineering, University of California, Berkeley, California 94720, United States, 2Institute of
27 August 2013 Semiconductors, Chinese Academy of Sciences, P.O. Box 912, Beijing 100083, Peoples Republic of China, 3Department of
Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, United States,
Published 4
Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States.
13 September 2013
Point defects in semiconductors can trap free charge carriers and localize excitons. The interaction between
these defects and charge carriers becomes stronger at reduced dimensionalities, and is expected to greatly
Correspondence and influence physical properties of the hosting material. We investigated effects of anion vacancies in
requests for materials monolayer transition metal dichalcogenides as two-dimensional (2D) semiconductors where the vacancies
should be addressed to density is controlled by a-particle irradiation or thermal-annealing. We found a new, sub-bandgap emission
J.W. (wuj@berkeley. peak as well as increase in overall photoluminescence intensity as a result of the vacancy generation.
Interestingly, these effects are absent when measured in vacuum. We conclude that in opposite to
edu)
conventional wisdom, optical quality at room temperature cannot be used as criteria to assess crystal quality
of the 2D semiconductors. Our results not only shed light on defect and exciton physics of 2D
semiconductors, but also offer a new route toward tailoring optical properties of 2D semiconductors by
* These authors
defect engineering.
contributed equally to
this work.
I
n semiconductors, lattice point defects such as vacancies and interstitials can act as very efficient traps for
electrons, holes and excitons, and strongly influence transport and optical properties of the host material.
Excitons bound to defects, if recombine radiatively, lead to light emission at energies lower than the band-to-
band optical transition energy. Such interactions become stronger in reduced dimensionalities due to tighter
localization of the electron wavefunction. For example, in three dimensions (3D) within the hydrogenic defect
2
model, shallow defects bind electrons at a ground-state binding energy equal to 13:6 eV|m
er , where m* is the
2
effective mass and er is relative dielectric constant. This is increased to 54:4 eV|m er in two dimensions (2D)
simply due to the dimensionality effect1. Similarly, the binding energy and recombination dynamics of Wannier
and Frenkel excitons are expected to be drastically different going from 3D to 2D24. The escalated binding energy
means stabilization of bound excitons and their emission features at higher temperatures, pointing to a potentially
useful way to tailor optical properties of 2D semiconductors.
Although the understanding of point defects in conventional 3D semiconductors is well established and the
defects database is relatively complete, physics and behaviour of point defects in 2D semiconductors, such as the
newly emerging monolayer semiconducting transition metal dichalcogenides (TMDs), have remained an unex-
plored field. In this work we report the effects of anion vacancies, the dominant point defect species, on photo-
luminescence (PL) and Raman spectra of monolayer TMDs. Since these materials become direct-bandgap
semiconductors with relatively high PL intensity in the monolayer limit5, the defect effects can be easily mon-
itored optically. We find that irradiation with MeV a particles or thermal annealing at sub-decomposition
temperatures introduce anion vacancies in monolayer MoS2, MoSe2, and WSe2, where the vacancy density
can be controlled by the irradiation dose or annealing time. These defects introduce a new emission peak at
,0.15 to 0.25 eV below the free-exciton PL peak, and its intensity is enhanced as the defect density is increased.
Moreover, the overall PL intensity also increases at higher defect densities, and as such, the defective material
Figure 1 | (a). AFM images taken on monolayer MoS2, MoSe2, and WSe2. (b). Raman spectrum measured on monolayer MoS2, MoSe2, and WSe2, where
the solid and dashed curves correspond to monolayers and few-layers, respectively. (c). Room-temperature normalized PL for monolayer MoS2, MoSe2,
and WSe2.
becomes more luminescent compared to pristine monolayers. irradiation dose; and (2) the integrated intensity of the main PL line
Surprisingly, these effects are absent when measured in vacuum, at 1.90 eV increases by ,3 times while the PL peak position shifts to
suggesting that the interaction between ambient gas molecules and higher energy by ,20 meV. The Stopping and Range of Ions in
the defect sites play a significant role in the process. Our density Matter (SRIM) calculations estimate that 7.5 3 1023 vacancy/cm
functional theory (DFT) calculations show that these anion vacancies 3 ion is generated on MoS2 upon the a particle irradiation, and this
create energy levels approximately 0.2 eV below the band edge, and corresponds to 6 3 1011 cm22 defect density, or approximately one
gas molecules can be physically adsorbed at the defect sites with defect per 100 unit cells, for 8 3 1013 cm22 irradiation dose. We note
relatively large charge transfer, which electron-depletes the material. that here the PL was all measured in the presence of N2 gas
(,50 Torr), and the effect of N2 will be discussed later. The small
Results blueshift and enhancement in the main PL intensity bear much
Monolayer MoS2, MoSe2, and WSe2 flakes were mechanically exfo- resemblance with the previously reported transition from charged
liated from bulk crystals onto 90 nm SiO2 where a relatively high exciton (X2 or eeh) to neutral free exciton (X0 or eh) in 2D systems,
contrast can be observed at the flakes6. The monolayers were iden- such as 2D electron gas (2DEG) heterostructures13 and more recently
tified using atomic force microscopy (AFM), Raman, and PL mea- on monolayer TMDs24. In these cases, the charged to neutral exciton
surements. AFM measurements on monolayers yield a height of
,0.7 nm corresponding to the thickness of a single unit cell
(Figure 1a). From the bulk to monolayers, the out-of-plane Raman
mode (A1g) softens and the in-plane mode (E2g) stiffens for MoS27
and MoSe28,9, whereas for WSe210, the degenerate A1g and E2g modes
in the bulk split by 12 cm21 in the monolayer as a result of broken
degeneracy (Figure 1b). The PL signal is greatly enhanced by orders
of magnitude from the bulk to monolayer due to the indirect to direct
bandgap transition, consistent with previous reports5,8,11. At room
temperature, monolayer MoS2, MoSe2, and WSe2 show a strong PL
peak at 1.84 eV, 1.56 eV and 1.65 eV, respectively (Figure 1c). The
exfoliated monolayers were irradiated with 3 MeV a particles at
controlled doses to create different densities of point defects, and
the samples were cooled down to the lowest attainable temperature
in our system (77 K) for PL and Raman measurements. Before and
after the irradiation, the sulfur to molybdenum (S/Mo) atomic ratio
was monitored using nano-Auger electron spectroscopy (nano-
AES). We find that the S/Mo ratio decreases slightly (Figure 2a),
implying that the irradiation induces S vacancies in the 2D crystal,
consistent with earlier results12. After the irradiation, the full-width-
at-half maximum of the Raman peaks slightly broadens (Figure 2b)
due to relaxation of the Raman selection rule at the defects. Figure 2 | (a). Nano-Auger spectrum taken on a monolayer MoS2 before
At 77 K, the as-exfoliated MoS2 monolayers display a strong PL and after irradiation with a particles at a dose of 8 3 1013 cm22. (b). Raman
peak at 1.90 eV corresponding to the direct bandgap at the K sym- spectrum of the same. (c). PL spectrum for pristine and irradiated
metry point. Upon the a particle irradiation at different doses, the 77- monolayer MoS2 at the shown irradiation doses. The PL was taken at 77 K
K PL spectrum changes significantly as shown in Figure 2c and these in N2 (50 Torr) environment with a constant laser excitation power. The
changes are summarized as follows: (1) a new PL peak appears at irradiation-caused enhancement in bound exciton (XB) and free exciton
1.78 eV and the integrated intensity of this peak increases with the (X0) emission intensity is indicated.
transition is associated with charge depletion24 or charge local- temperature is increased and completely disappears above 250 K.
ization13 which stabilizes (destabilizes) neutral (charged) excitons. At room temperature, both pristine and defective monolayers exhibit
In accord with these studies, we attribute the observed blueshift to high optical quality with a strong, single PL peak associated with the
irradiation-induced defect sites interacting with N2 molecules, band-to-band optical transition at the K point. Therefore, we con-
resulting in depletion and localization of charge carriers in the mono- clude that in opposite to conventional wisdom believed in this new
layer TMDs. Such effects will be discussed more in detail later. field of 2D semiconductors16,17, optical quality at room temperature
Next, we focus on the new peak at 1.78 eV (Figure 1c). Since this (PL intensity and sharpness) cannot be used as criteria to assess the
PL peak appears after the irradiation and its intensity increases at crystal quality of the monolayers. Indeed, the defective monolayers
higher doses, we attribute it to radiative recombination of bound yield even stronger PL intensity at room temperature (Figure 4a).
excitons (XB), i.e. neutral excitons (X0) bound to defects. To probe However, the PL spectrum at 77 K immediately tells the difference
this more, we also introduced the defects by thermal annealing in between the pristine and defective monolayers (Figure 2c).
vacuum. Here, monolayers were annealed ,100uC below their ther-
The PL spectra discussed so far were all recorded in the presence of
mal decomposition temperature (,600uC), and this process is
N2 gas, regardless of the measurement temperature (300 K or 77 K).
known to create S vacancies in MoS214. After the annealing, the PL
Interestingly, when measured in vacuum, the aforementioned defect
spectrum also displays a defect-induced bound exciton peak at
peak (XB) in the PL spectrum disappears at both room and low
,1.78 eV (Figure 3a), except that the peak appears relatively stron-
temperatures (Figure 4ab). We also note that the occurring of the
ger and broader than in the irradiated samples, possibly due to a
defect PL peak was instantaneously reversible when the chamber is
higher density of point defects and defect clusters created by anneal-
purged with or pumped out of N2. This implies that the interaction
ing. It is expected that vacancies generation by particle irradiation is a
between the defect sites and the N2 gas molecules is weak (physi-
series of highly non-equilibrium, random and isolated events, while
sorbed), but dictates the optical emission of the material. The above
thermal annealing is much slower and may facilitate formation of
results are discussed in detail in conjunction with first-principles
vacancy clusters with different configurations. These defect com-
calculations shown below.
plexes with different clustering configurations may have different
exciton binding energies, thus broadening the observed defect PL
peak. Discussion
Further confirmation on the defect origin of the 1.78 eV peak To understand the physical origin of the defect-induced PL peak (XB)
comes from its excitation power dependence (Figure 3c). Since the and the intensity enhancement in the main PL (X0), we calculated the
XB is associated with excitons bound to defects, the PL intensity of XB band structure and the density of states (DOS) of defective mono-
is expected to saturate at high excitation power intensities when these layer MoS2. Since the a particle irradiation12,18 and thermal annealing
defects are fully populated with excitons15. Consistent with this, the may result in various types of sulfur vacancies in MoS2, the mono-
intensity of XB exhibits a sub-linear laser power dependence with a layer MoS2 was modelled in the presence of sulfur vacancies in dif-
tendency to saturate at high excitation powers, whereas the free ferent arrangements (Supplementary Information Fig. S1). However,
exciton intensity (X0) scales linearly without any sign of saturation our calculations show that di-sulfur vacancies are most relevant to
(Figure 3bc). As a result, the overall PL spectrum is mostly domi- the experiments presented here, so our discussions will be limited to
nated by the defect peak, XB, at low excitation intensities, and the X0 the di-S vacancies. In Figure 5a, we show the band structure of a
line becomes observable only at high excitation intensities monolayer MoS2 in the presence of di-S vacancies with and without
(Figure 3b). the N2 gas molecules around. We first note that once the di-S vacan-
Next, we consider the effects of temperature and interaction with cies are created, new states appear within the bandgap. Specifically,
ambient gas molecules on the PL spectrum of defective monolayers. states with energies 0.2 , 0.3 eV near the conduction/valence band
First, we note that the XB peak is prominent at low temperatures or edge (Figure 5a red) are of particular interest, as they are close to the
low excitation powers (Figure 3ab). It becomes weaker as the energy difference between XB and X0 (0.12 eV). However, since the
account in the spin-polarized calculations. The super-cell size, kinetic energy cut-off, 15. Schmidt, T., Lischka, K. & Zulehner, W. Excitation-power dependence of the
and Brillouin zone sampling of the calculations were determined after extensive near-band-edge photoluminescence of semiconductors. Phys. Rev. B 45,
convergence analyses. A large spacing of ,15 A between the 2D single layers was used 89898994 (1992).
to prevent interlayer interactions. A plane-wave basis set with kinetic energy cut-off of 16. Liu, K. K. et al. Growth of Large-Area and Highly Crystalline MoS2 Thin Layers on
300 eV was used. In the self-consistent field potential and total energy calculations, Insulating Substrates. Nano Lett. 12, 15381554 (2012).
the Brillouin zone was sampled by special k-points. The numbers of these k-points 17. Yu, Y. et al. Controlled Scalable Synthesis of Uniform, High-Quality Monolayer
were (25 3 25 3 1) for the primitive 1H-MoS2 and were scaled according to the size of and Few-layer MoS2 Films. Sci. Rep. 3, 1866 (2013).
the super cells. All atomic positions and lattice constants were optimized using the 18. Inoue, A., Komori, T. & Shudo, K. I. Atomic-scale structures and electronic states
conjugate gradient method, where the total energy and atomic forces were minimized. of defects on Ar-ion irradiated MoS2. J. Electron. Spec. Rel. Phen. http://
The convergence for energy were chosen to be 1026 eV between two consecutive dx.doi.org/10.1016/j.elspec.2012.12.005.
steps, and the maximum Hellmann-Feynman forces acting on each atom was less 19. Tongay, S. et al. Broad-range modulation of light emission of 2D semiconductors
than 0.01 eV/A upon ionic relaxation. The pressure in the unit cell was kept below with molecular gas adsorption. Nano Lett. 13, 28312836 (2013).
5 kbar. Numerical calculations were performed by using the VASP software24. 20. Yu, K. M. et al. Diluted IIVI Oxide Semiconductors with Multiple Band Gaps.
Phys. Rev. Lett. 91, 246403 (2003).
21. Blochl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 1795317979
1. Yang, X. L., Guo, S. H., Chan, F. T., Wong, K. W. & Ching, W. Y. Analytic solution (1994).
of a 2D hydrogen atom. I. Nonrelativistic theory. Phys. Rev. A 43, 11861196 22. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized Gradient Approximation
(1991). Made Simple. Phys. Rev. Lett. 77, 38653868 (1996).
2. Newaz, A. K. M. et al. Electrical control of optical properties of monolayer MoS2. 23. Grimme, S. Semiempirical GGA-type density functional constructed with a long-
Solid State Commun. 155, 4952 (2013). range dispersion correction. J. Comput. Chem. 27, 17871799 (2006).
24. Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B
3. Mak, K. F. et al. Tightly bound trions in monolayer MoS2. Nature Mater. 12,
47, 558561 (1993).
207211 (2012).
4. Ross, J. et al. Electrical Control of Truly Two Dimensional Neutral and Charged
Excitons in a Monolayer Semiconductor. Nature Commun. 4, 1474 (2012).
5. Mak, K., Lee, C., Hone, J., Shan, J. & Heinz, T. F. Atomically Thin MoS2: A New
Direct-Gap Semiconductor. Phys. Rev. Lett. 105, 136805 (2010).
Acknowledgements
This work was supported by the Office of Science, Office of Basic Energy Sciences, of the
6. Benameur, M. M. et al. Visibility of dichalcogenide nanolayers. Nanotech. 22, U.S. Department of Energy under Contract No. DE-AC02-05CH11231.
125706 (2011).
7. Lee, C. et al. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS
Nano. 4, 26952700 (2010). Author contributions
8. Tongay, S. et al. Thermally driven crossover from indirect toward direct bandgap S.T. and J.W. conceived the project. S.T., J.S., F.W., A.L., J.S.K., J.L. and J.Z. performed the
in 2D semiconductors: MoSe2 versus MoS2. Nano Lett. 12, 55765580 (2012). measurements. C.A., R.R. and J.C.G. performed the density functional theory calculations.
9. Horzum, S. et al. Phonon softening and direct to indirect band gap crossover in C.K. and F.O. performed nano-Auger measurements. S.T. and J.W. wrote the manuscript.
strained single-layer MoSe2. Phys. Rev. B 87, 125415 (2013). All authors have read the manuscript.
10. Sahin, H. et al. Anomalous Raman Spectrum and Dimensionality Effects in WSe2.
Phys. Rev. B 87, 165409 (2013).
11. Zhao, W. et al. Evolution of Electronic Structure in Atomically Thin Sheets of WS2
Additional information
Supplementary information accompanies this paper at http://www.nature.com/
and WSe2. ACS Nano 7, 791797 (2013).
scientificreports
12. Mathew, S. et al. Magnetism in MoS2 induced by proton irradiation. Appl. Phys.
Lett. 101, 102103 (2012). Competing financial interests: The authors declare no competing financial interests.
13. Finkelstein, G., Shtrikman, H. & Bar-Joseph, I. Optical Spectroscopy of a Two- How to cite this article: Tongay, S. et al. Defects activated photoluminescence in
Dimensional Electron Gas near the Metal-Insulator Transition. Phys. Rev. Lett. two-dimensional semiconductors: interplay between bound, charged, and free excitons. Sci.
74, 976979 (1995). Rep. 3, 2657; DOI:10.1038/srep02657 (2013).
14. Qiu, H. et al. Electrical characterization of back-gated bi-layer MoS2 field-effect
transistors and the effect of ambient on their performances. Appl. Phys. Lett. 100, This work is licensed under a Creative Commons Attribution 3.0 Unported license.
123104 (2012). To view a copy of this license, visit http://creativecommons.org/licenses/by/3.0