Vous êtes sur la page 1sur 6

Journal of Alloys and Compounds 613 (2014) 122127

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Effect of vacancies on structural, electronic and optical


properties of monolayer MoS2: A rst-principles study
Li-ping Feng , Jie Su, Zheng-tang Liu
State Key Lab of Solidication Processing, College of Materials Science and Engineering, Northwestern Polytechnical University, Xian, Shaanxi 710072, China

a r t i c l e

i n f o

Article history:
Received 7 January 2014
Received in revised form 22 May 2014
Accepted 3 June 2014
Available online 11 June 2014
Keywords:
Monolayer MoS2
Vacancy
Electronic structure
Optical properties

a b s t r a c t
Effects of vacancies on structural, electronic and optical properties of monolayer MoS2 were investigated
using rst-principles plane-wave pseudopotential method based on density functional theory. Results
show that the band structure and band gap of perfect monolayer MoS2 are in good agreement with
the available experimental and theoretical data. Structural analysis indicates that ions surrounding Mo
vacancies show an outward relaxation, while that ions surrounding S vacancies exhibit slightly inward
relaxation. Electronic analysis implies that the band gaps of defective monolayer MoS2 are smaller than
that of perfect one. After introduction of neutral S-vacancy, monolayer MoS2 has changed from direct to
indirect band gap. Mo vacancies bring about acceptor-like levels and p-type conductivities, whereas S
vacancies lead to donor-like levels and n-type conductivities. With the increasing charge states of vacancies, the band gaps get smaller and the defect energy levels become deeper. Moreover, as the charge
states of vacancies increase, the static dielectric constants of monolayer MoS2 with Mo vacancies
decrease, whereas the static dielectric constants of monolayer MoS2 with S vacancies increase.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Molybdenum disulde (MoS2) with hexagonal structure is an
important semiconductor material with unique physical and
chemical properties, and it has a variety of applications including
hydrogen production, solar cells, catalysis, biomedicine, sensing,
photovoltaic and super-lubricity [110]. Most recently, the monolayer MoS2 has been used to construct eld-effect transistors
(FETs), which can offer lower power consumption than classical
transistors [11,12]. New phototransistor based on monolayer
MoS2 has been fabricated and exhibits a better photoresponsivity
as compared with the graphene-based device [13]. Except for
experimental studies, a lot of theoretical work has been performed
to investigate the properties of monolayer MoS2. Ataca et al.
[14,15] have studied the lattice dynamics and structure stability
of monolayer MoS2, indicating that MoS2 can form stable and
suspended monolayer with honeycomb structures. Moreover, a
great number of literatures had predicted the electronic, elastic,
mechanical, and optical properties of monolayer MoS2 [1621].
Dhas and Suslick [22] have found that vacancies exist in monolayer MoS2 when they synthesized the monolayer MoS2 through
sonochemical deposition method. It is well known that vacancy

Corresponding author. Tel.: +86 29 88488013; fax: +86 29 88492642.


E-mail address: lpfeng@nwpu.edu.cn (L.-p. Feng).
http://dx.doi.org/10.1016/j.jallcom.2014.06.018
0925-8388/ 2014 Elsevier B.V. All rights reserved.

defects have a strong inuence on geometric structures, electronic


structures, magnetic properties, optical characteristics and so on
[2328]. However, few literatures have reported the effect of
vacancies on properties of monolayer MoS2. Makarova et al. [29]
have investigated selective adsorption of thiol molecules at
S-vacancies on MoS2 (0 0 0 1). Ataca et al. [30,31] have calculated
the formation energy of neutral vacancies in monolayer MoS2
and studied the inuence of vacancies on magnetic properties of
monolayer MoS2, implying that vacancy creation appears to be a
promising way to extend the applications of MoS2.
To the best of our knowledge, the effects of charged vacancies
on electronic structure and optical properties of monolayer MoS2
are not well understood yet. It is well known that the knowledge
of vacancy defects is very important for the practical applications
of monolayer MoS2 as well as for the designing and analyzing of
optoelectronic devices. Therefore, this work is focused on investigating the effect of charged vacancies on structure relaxation, electronic structure and optical properties of monolayer MoS2 using
rst-principles calculations.

2. Computational details
In the present calculations, the exchange correlation of the generalized gradient
approximation (GGA) with the PerdewWang 1991 (PW-91) functional [32] as
implemented in CASTEP code [33] was employed. The ionic cores are represented
by ultrasoft pseudopotential for Mo and S atoms. The valence electron congurations

123

L.-p. Feng et al. / Journal of Alloys and Compounds 613 (2014) 122127
include Mo 4p6 4d5 5s1 electrons and S 3s2 3p4 electrons. The plane-wave cutoff
energy was set to be 380 eV after extensive convergence analysis. The Brillouin-zone
integration was performed over the 5  5  2 grid sizes using the MonkorstPack
method, where the self-consistent convergence of the total energy is
5.0  106 eV/atom. The optimized primitive lattice constants a = 3.17 and
c = 12.32 for bulk MoS2 are in good agreement with other theoretical data
(a = 3.17 , c = 12.58 [34], a = 3.16 , c = 12.29 [35]) and experimental values
(a = 3.20 , c = 12.29 [36]). A 4  4 supercell of 48 atoms was constructed for
defect-free monolayer MoS2 (shown in Fig. 1(a)). A 12 vacuum region was used
to separate the two dimensional single layers of MoS2 along the c axis to hinder interlayer coupling [37]. To introduce an isolated vacancy, an Mo or S atoms is removed
from the host supercell, respectively. The large supercell size separates the two adjacent defects over 10 , which is sufcient to neglect the articial Coulomb interaction between the defects [38,39]. The charge states of vacancies were controlled by
adjusting the background charge of crystals. Defective monolayer MoS2 with single
Mo and S vacancy was presented in Fig. 1(b) and (c), respectively.

3. Results and discussion


3.1. Structural properties
Vacancies in monolayer MoS2 can lead to its structural relaxation. Table 1 lists the distances from charged vacancy to the neighboring ions for monolayer MoS2 before and after structural
relaxation. For Mo vacancy, ions surrounding the Mo-vacancy
show an outward relaxation because of the ionic size and charge
effects [40]. Additionally, the outward relaxations of the rst nearest neighbor (NN) S ions are larger than those of the second NN Mo
ions. For example, the outward relaxation of the rst NN S ions and
the second NN Mo ions for V 0Mo is about 1.6% and 0.6%, respectively.
When the charge states of Mo vacancies increase, the second NN
Mo ions exhibit more outward relaxations due to the increasing
electrostatic interactions between the vacancy and the second
NN Mo ions [41]. In the case of V 4
Mo , the second NN Mo ions
(1.3%) undergo almost twice higher relaxation than that of V 0Mo
(0.6%). However, the distances from the rst NN S ions to Movacancy almost maintain 2.45 with the increasing charge states
of Mo-vacancy because of the signicant electrostatic repulsions
between the rst NN S ions [40]. For S vacancy, the ions surrounding the S-vacancy exhibit slightly inward relaxation. That was also
observed in many materials when whose big size ion is removed to
form vacancy [4043]. In addition, the inward relaxations of the
second NN S ions are larger than those of the rst NN Mo ions
due to the decreased electrostatic repulsions between the second
NN S ions and the S-vacancy [43]. For example, the inward relaxation of the rst NN Mo ions and the second NN S ions for V 0S is
about 0.4% and 1.5%, respectively. Moreover, both the rst and second NN ions undergo more inward relaxation with the increasing
charge states of S vacancies. In the case of V 2
S , the relaxation of
the rst and second NN ions reaches 1.2% and 2.2%, respectively.

Table 1
Distances from charged vacancy (V yx , x is Mo or S atom, y represents the charge states
of vacancy) to the neighboring atoms for monolayer MoS2 before and after the
structural relaxation. Neighboring atomic species, their coordination numbers and the
relaxations in % are also shown in parentheses.
Vacancies

Distance in (atomic species  coordination number)


First NN

Second NN

V 0Mo

2.41 (S  6)
2.45 (1.6%)

3.16 (Mo  6)
3.18 (0.6%)

V 1
Mo

2.45 (1.6%)

3.18 (0.6%)

V 2
Mo

2.46 (1.9%)

3.18 (0.6%)

V 3
Mo

2.45 (1.6%)

3.19 (0.9%)

V 4
Mo

2.45 (1.6%)

3.20 (1.3%)

S (perfect)
V 0S

2.41 (Mo  3)
2.40 (0.4%)

3.16 (S\caxis  6, Skcaxis  1)


3.11 (1.5%)

V 1
S

2.38 (1.2%)

3.10 (1.8%)

V 2
S

2.38 (1.2%)

3.09 (2.2%)

Mo (perfect)

The above phenomena are further conrmed by the following Mulliken atomic population analysis.
3.2. Electronic properties
Fig. 2 shows the band structures and density of states (DOS) of
perfect and defective monolayer MoS2. As shown in Fig. 2(a), the
calculated direct band gap for perfect monolayer MoS2 is about
1.78 eV, which is in good agreement with other theoretical values
(1.80 eV [35,16], 1.70 eV [44]) and experimental data (1.98 eV [45],
1.90 eV [37]). Additionally, bottom of conduction bands and top of
valence bands for perfect monolayer MoS2 mainly consist of the
hybridization from Mo 4d and S 3p orbitals, which is consistent
with previous calculational results [30,46,47]. The band structures
for monolayer MoS2 with neutral S- and Mo-vacancy are shown in
Fig. 2(b) and (c), respectively. The conduction and valence bands of
monolayer MoS2 with vacancies are composed predominantly of
the strong hybridization from Mo 4d and S 3p orbitals. The calculated band gaps for monolayer MoS2 with neutral Mo and S
vacancy are 1.63 and 1.74 eV, respectively. It should be noted that
monolayer MoS2 has changed from direct to indirect band gap after
the introduction of neutral S-vacancy because the relaxation of
adjacent surrounding atoms of S-vacancy induces the variation of
the Mo 4d symmetry and the trigonal crystal eld around Mo
[39,48,49]. In contrast, monolayer MoS2 with neutral Mo-vacancy
still remains direct band gap. Fig. 3 presents the calculated band
gaps for monolayer MoS2 with neutral and charged vacancies. It
can be seen that the band gaps of defective monolayer MoS2 are

Fig. 1. Atomic congurations of monolayer MoS2. (a) Defect-free monolayer MoS2, (b) monolayer MoS2 with Mo vacancy and (c) monolayer MoS2 with S vacancy.

124

L.-p. Feng et al. / Journal of Alloys and Compounds 613 (2014) 122127

Fig. 3. The calculated band gaps for perfect monolayer MoS2 and defective
monolayer MoS2 with charged vacancies.

Fig. 2. The band structures and DOS for perfect monolayer MoS2 (a), monolayer
MoS2 with neutral Mo-vacancy (b), and monolayer MoS2 with neutral S-vacancy (c).

smaller than that of perfect monolayer MoS2. With the increasing


charge states of vacancies, the band gap of monolayer MoS2 with
Mo-vacancy decreases rapidly, while the band gap of monolayer
MoS2 with S-vacancy decreases comparatively slowly.
Moreover, as shown in Fig. 2(b) and (c), the vacancy defects also
introduce defect energy levels in the band gap. The calculated
defect energy levels for the neutral and charged vacancies in
monolayer MoS2 are shown in Fig. 4. For Mo vacancies, the defect
energy levels of 0, 1-, 2-, 3-, and 4- charge states lie at 0.11, 0.23,
0.37, 0.46, and 0.69 eV, respectively. It is clear that the defect
energy levels become deeper with the increasing charge states of
Mo vacancies. All the defect energy levels generated by Mo vacancies locate near the valence band maximum, suggesting that Mo
vacancies induce acceptor-like levels in the band gap. Hence, Mo
vacancies might trap the electrons from the valence bands. For S
vacancies, the defect energy levels of 0, 1+, and 2+ charge states
lie at 0.40, 0.60, and 0.67 eV, respectively. As the charge states of
S vacancies increase, the defect energy levels become deeper. Additionally, all the defect energy levels caused by S vacancies are close
to the conduction band minimum, implying that donor-like levels
are formed and that the electrons of S vacancies might tunnel into
the conduction bands.

To further investigate the electronic structure, Mulliken population of SMo bonding in monolayer MoS2 was analyzed. Mulliken
population can determine the type of bond and its magnitude, a
high positive value of bond population means the type of the bond
is dominated by covalency, and a value of zero presents a perfect
ionic bond [50]. Mulliken population of SMo bonding in bulk
2HMoS2 was reported to be about 0.25, which indicates a covalent character of SMo bonding [51]. For perfect monolayer
MoS2, the calculated Mulliken population of SMo bonding is
about 0.37 which is bigger than the value of 0.25 of bulk 2H
MoS2 [51], showing that the SMo bonding also has a covalent
character. Moreover, the excess charge on each S atom and the
depletion of electrons on each Mo atom were calculated to be
0.17 and 0.34e, respectively, implying that the SMo bonding in
perfect monolayer MoS2 also exhibits partially ionic character,
which is consistent with previous theoretical report [30]. The
Mulliken population and bond length of SMo bonding around
vacancies are shown in Fig. 5. The overlap population becomes
greater and the bond length becomes shorter for the rst NN
SMo bonding around vacancies compared with those of SMo
bonding in perfect monolayer MoS2, indicating that vacancies
enhance the covalent character of the SMo bonding. Obviously,
the variations for SMo bondings around Mo-vacancy are more signicant than those for SMo bondings around S-vacancy, which
may relate to the large relaxation of geometric structure when
the Mo-vacancy forms. Nevertheless, as the distance from SMo

Fig. 4. Calculated defect energy levels for the neutral and charged vacancies in
monolayer MoS2. The positions of the energy levels are given with respect to the
VBM in the case of VMo, while those from the CBM in the case of VS. The value of the
perfect monolayer MoS2 is set to zero.

L.-p. Feng et al. / Journal of Alloys and Compounds 613 (2014) 122127

125

Fig. 5. Mulliken population and bond length of SMo bonding around Mo-vacancy (a) and S-vacancy (b). The atoms serial number is indexed in Fig. 1(b) and (c).

bonding to vacancy increases, the variations of the SMo bonding


become weak.
Charge densities can further examine the change of chemical
bonding around vacancies. Fig. 6 presents the charge densities of
SMo bonding in perfect and defective monolayer MoS2. Compared
with the charge densities of SMo bonding in perfect monolayer
MoS2, the charge densities of the rst NN SMo bonding around
vacancies become higher, suggesting that the covalent characters
of the SMo bondings are strengthened when vacancies formed
[52]. Additionally, it can be seen that the charge densities of
SMo bondings around Mo-vacancy have more signicant changes
than those of SMo bondings around S-vacancy. Furthermore, the
charge densities of the SMo bondings decrease when the distance
from SMo bonding to vacancy increases. These results are consistent with the above Mulliken population analysis.

3.3. Optical properties


The optical properties of perfect and defective monolayer MoS2
have been calculated in the polarization direction of (1 0 0). 12
unoccupied states and the Gaussian broadening with a width of
0.5 eV were used for the optical spectra calculations after extensive
analysis. The imaginary parts of complex dielectric function of
monolayer MoS2 with Mo and S vacancies are shown in Fig. 7(a)
and (b), respectively. For the imaginary part of perfect monolayer
MoS2, there are three peaks in the energy regions from 0 to
20 eV, which are the absorptive transitions from the valence bands
to the conduction bands. These peaks are labeled A, B and C,
respectively. According to the analysis of the electronic structure,
peak A originates mainly from the transitions of S 3p into Mo 4d
conduction bands [53]. Peak B originates mainly from the

Fig. 6. The charge density contour prole of SMo bondings. (a) The SMo bonding in perfect monolayer MoS2, (bd) the adjacent SMo bondings around Mo-vacancy and (e
g) the adjacent SMo bondings around S-vacancy. The atoms serial number is indexed in Fig. 1(b) and (c).

126

L.-p. Feng et al. / Journal of Alloys and Compounds 613 (2014) 122127

Fig. 7. The imaginary parts of complex dielectric function of monolayer MoS2 with Mo vacancies (a) and S vacancies (b).

Fig. 8. The real parts of complex dielectric function of monolayer MoS2 with Mo vacancies (a) and S vacancies (b). Inset is the static dielectric constant for perfect monolayer
MoS2 and defective monolayer MoS2 with different charge states.

hybridization orbitals between S 3p and Mo 4d into Mo 4d conduction bands [21,54]. And peak C originates mainly from r bonding
between S 3p and Mo 5s into Mo 4d conduction bands [55]. The
imaginary parts of defective monolayer MoS2 have similar proles
with that of perfect one, but they move slightly toward lower
energies because of the localized effects of vacancies [28,56]. Additionally, extra peaks appear at low energy in the imaginary parts of
defective monolayer MoS2, which may be ascribed to the dangling
bonds and defective states formed after the removal of atom.
The real parts of complex dielectric function and static dielectric constants of prefect and defective monolayer MoS2 are shown
in Fig. 8. The calculated static dielectric constant of perfect monolayer MoS2 is about 1.50, which is consistent with other theoretical
value of 1.26 [14] calculated by GGA but a little smaller than the
theoretical value of 3.0 [57] calculated by LDA. It is obvious from
Fig. 8(a) that the static dielectric constants of monolayer MoS2 with
Mo vacancies are larger than that of perfect one. Moreover, monolayer MoS2 with neutral Mo vacancy has the largest static dielectric
constant of about 2.45. As the charge states of Mo vacancies
increase, the static dielectric constants of defective monolayer
MoS2 decrease, and the variation of the static dielectric constants
is satised to Penns model [58]. In Fig. 8(b), the static dielectric
constant of monolayer MoS2 with neutral S vacancy is found to
be about 1.50, which is the same as that of perfect monolayer
MoS2. Nevertheless, the static dielectric constants of monolayer
MoS2 with charged S vacancies are higher than that of perfect
one. In contrast to monolayer MoS2 with Mo vacancies, the static
dielectric constants of monolayer MoS2 with S vacancies show a

growing tendency with the increasing positive charge states of S


vacancies.
4. Conclusion
In summary, the structural, electronic and optical properties of
monolayer MoS2 with charged vacancies have been investigated
using the rst-principles calculations. Results show that the band
structure and band gap of perfect monolayer MoS2 consist well
with experimental and previous calculational data. Structural
relaxation is found in the monolayer MoS2 with vacancies. Ions
surrounding Mo vacancies show an outward relaxation while ions
surrounding S vacancies exhibit slightly inward relaxation. Band
structure, band gap, Mulliken population, and charge density of
perfect and defective monolayer MoS2 were analyzed. The band
gaps of monolayer MoS2 with vacancies are smaller than that of
perfect monolayer MoS2. After introduction of neutral S-vacancy,
monolayer MoS2 has changed from direct to indirect band gap.
Mo vacancies induce acceptor-like levels and p-type conductivities, whereas S vacancies introduce donor-like levels and n-type
conductivities. With the increasing charge states of vacancies, the
band gaps become smaller and the defect energy levels become
deeper. The complex dielectric functions of perfect and defective
monolayer MoS2 were obtained. Vacancies induce extra peaks at
low energy in the real and imaginary parts of complex dielectric
functions of defective monolayer MoS2. The calculated static
dielectric constant of perfect monolayer MoS2 is about 1.50. The
monolayer MoS2 with neutral Mo vacancy has the largest static

L.-p. Feng et al. / Journal of Alloys and Compounds 613 (2014) 122127

dielectric constant of about 2.45, while the monolayer MoS2 with


neutral S vacancy has the same static dielectric constant with perfect monolayer MoS2. Furthermore, with the increasing charge
states of vacancies, the static dielectric constants of monolayer
MoS2 with Mo vacancies decrease, whereas the static dielectric
constants of monolayer MoS2 with S vacancies increase.
Acknowledgements
We acknowledge the National Natural Science Foundation of
China under grant No. 61376091, the Natural Science Foundation
of Shaanxi Province under grant No. 2012JM6012, the Fundamental Research Funds for the Central Universities under grant No.
3102014JCQ01033 and the 111 Project under grant No. B08040.
Reference
[1] B. Hinnemann, P. Moses, J. Bonde, K. Jorgensen, J. Nielsen, S. Horch, I.
Chorkendorff, J. Norskov, J. Am. Chem. Soc. 127 (2005) 5308.
[2] T.F. Jaramillo, K.P. Jorgensen, J. Bonde, J.H. Nielsen, S. Horch, I. Chorkendorff,
Science 317 (2007) 100.
[3] G. Kline, K. Kam, R. Ziegler, B. Parkinson, Sol. Energy Mater. 6 (1982) 337.
[4] Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong, H. Dai, J. Am. Chem. Soc. 133 (2011)
7296.
[5] H. Wu, R. Yang, B. Song, Q. Han, J. Li, Y. Zhang, Y. Fang, R. Tenne, C. Wang, ACS
Nano 5 (2011) 1276.
[6] M. De, S. Rana, H. Akpinar, O.R. Miranda, R.R. Arvizo, U.H.F. Bunz, V.M. Rotello,
Nat. Chem. 1 (2009) 461.
[7] N. Yaacobi-Gross, M. Soreni-Harari, M. Zimin, S. Kababya, A. Schmidt, N.
Tessler, Nat. Mater. 10 (2011) 974.
[8] H. Holscher, D. Ebeling, U.D. Schwarz, Phys. Rev. Lett. 101 (2008) 246105.
[9] C.G. Lee, Q.Y. Li, W. Kalb, X.Z. Liu, H. Berger, R.W. Carpick, J. Hone, Science 328
(2010) 76.
[10] E. Gourmelon, O. Lignier, H. Hadouda, G. Couturier, J.C. Bermede, J. Tedd, J.
Pouzet, J. Salardenne, Sol. Energy Mater. 46 (1997) 115.
[11] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Nat. Nanotechnol.
6 (2011) 147.
[12] W.Z. Bao, X.H. Cai, D. Kim, K. Sridhara, M.S. fuhrer, Appl. Phys. Lett. 102 (2013)
042104.
[13] Z.Y. Yin, H. Li, H. Li, L. Jiang, Y.M. Shi, Y.H. Sun, G. Lu, Q. Zhang, X.D. Chen, H.
Zhang, ACS Nano 6 (2012) 74.
[14] C. Ataca, M. Topsakal, E. Akturk, S. Ciraci, J. Phys. Chem. C 115 (2011) 16354.
[15] C. Ataca, H. Sahin, S. Ciraci, J. Phys. Chem. C 116 (2012) 8983.
[16] S. Lebegue, O. Eriksson, Phys. Rev. B 79 (2009) 115409.
[17] E.S. Kadantsev, P. Hawrylak, Solid State Commun. 152 (2012) 909.
[18] X.D. Li, S. Yu, S.Q. Wu, Y.H. Wen, S. Zhou, Z.Z. Zhu, J. Phys. Chem. C 117 (2013)
15347.
[19] X.D. Li, J.T. Mullen, Z.H. Jin, K.M. Borysenko, M.B. Nardelli, K.W. Kim, Phys. Rev.
B 87 (2013) 115418.

127

[20] Q. Yue, J. Kang, Z.Z. Shao, X.A. Zhang, S.L. Chang, G. Wang, S.Q. Qin, J.B. Li, Phys.
Lett. A 376 (2012) 1166.
[21] A. Kumar, P.K. Ahluwalia, Mater. Chem. Phys. 135 (2012) 755.
[22] N.A. Dhas, K.S. Suslick, J. Am. Chem. Soc. 127 (2005) 2368.
[23] D. Kurbatov, V. Kosyak, A. Opanasyuk, V. Melnik, Physica B 404 (2009) 5002.
[24] X.J. He, T. He, Z.H. Wang, M.W. Zhao, Physica E 42 (2010) 2451.
[25] A.G.M. Das, C. Nyamhere, F.D. Auret, M. Hayes, Surf. Coat. Technol. 203 (2009)
2628.
[26] I.A. Buyanova, B. Monemat, J.L. Lindstrom, T. Hallberg, L.I. Murin, V.P.
Markevich, Mater. Sci. Eng. B 72 (2000) 146.
[27] D.P. Zhang, J.D. Shao, H.J. Qi, M. Fang, K. Yi, Z.X. Fan, Opt. Laser Technol. 38
(2006) 654.
[28] J.J. Thevaril, S.K. Oleary, Solid State Commun. 150 (2010) 1851.
[29] M. Makarova, Y. Okawa, M. Aono, J. Phys. Chem. C 116 (2012) 22411.
[30] C. Ataca, S. Ciraci, J. Phys. Chem. C 115 (2011) 13303.
[31] C. Ataca, H. Sahin, E. Akturk, S. Ciraci, J. Phys. Chem. C 115 (2011) 3934.
[32] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C.
Fiolhais, Phys. Rev. B 46 (1992) 6671.
[33] M.D. Segall, P.J.D. Lindan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J. Clark, J.
Phys.: Condens. Matter 14 (2002) 2717.
[34] T.S. Li, G. Galli, J. Phys. Chem. C 111 (2007) 16192.
[35] T. Boker, R. Severin, A. Muller, C. Janowitz, R. Manzke, Phys. Rev. B 64 (2001)
235305.
[36] K.F. Mak, C.G. Lee, J. Hone, J. Shan, T.F. Heinz, Phys. Rev. Lett. 105 (2010)
136805.
[37] Q. Yue, S. Chang, S. Qin, J. Li, Phys. Lett. A 377 (2013) 162.
[38] L.N. Kantorovich, Phys. Rev. B 60 (1999) 15476.
[39] D. Liu, Y. Guo, L. Fang, J. Robertson, Appl. Phys. Lett. 103 (2013) 183113.
[40] K. Matsunaga, T. Tanaka, T. Yamamoto, Phys. Rev. B 68 (2003) 085110.
[41] T. Tanaka, K. Matsunaga, Y. Ikuhara, T. Yamamoto, Phys. Rev. B 68 (2003)
205213.
[42] F.F. Ge, W.D. Wu, X.M. Wang, H.P. Wang, Y. Dai, H.B. Wang, J. Shen, Physica B
404 (2009) 3814.
[43] X. Luo, B. Wang, Y. Zheng, Phys. Rev. B 80 (2009) 104115.
[44] J.V. Lauritsen, J. Kibsgaard, H. Topsoe, B.S. Clausen, E. Laegsgaard, F.
Besenbacher, Nat. Nanotechnol. 2 (2007) 53.
[45] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.Y. Chim, G. Galli, F. Wang, Nano
Lett. 10 (2010) 1271.
[46] A. Kumar, P.K. Ahluwalia, Eur. Phys. J. B 85 (2012) 186.
[47] N. Singh, G. Jabbour, U. Schwingenschlgl, Eur. Phys. J. B 85 (2012) 392.
[48] J.R. Hahn, H. Kang, Phys. Rev. B 60 (1999) 6007.
[49] J.R. Hahn, H. Kang, Phys. Rev. B 53 (1996) 1725.
[50] R.S. Mulliken, J. Chem. Phys. 23 (1955) 1833.
[51] T. Toforova, V. Alexiev, R. Prins, T. Weber, Phys. Chem. Chem. Phys. 6 (2004)
3023.
[52] J.H. Yao, Y.W. Li, N. Li, S.R. Le, Physica B 407 (2012) 3888.
[53] H. Shi, H. Pan, Y.W. Zhang, B.I. Yakobson, Phys. Rev. B 87 (2013) 155304.
[54] A.H. Reshak, S. Auluck, Phys. Rev. B 68 (2003) 125101.
[55] J.V. Acrivos, W.Y. Liang, J.A. Wilson, J. Phys. C: Solid State Phys. 4 (1971) L18.
[56] H. Qiu, T. Xu, Z. Wang, W. Ren, H. Nan, Z. Ni, et al., Nat. Commun. 4 (2013)
2642.
[57] A. Kumar, P.K. Ahluwalia, Physica B 407 (2012) 4627.
[58] D.R. Penn, Phys. Rev. B 128 (1962) 2091.

Vous aimerez peut-être aussi