Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Letter

pubs.acs.org/NanoLett

Deterministic Two-Dimensional Polymorphism Growth of Hexagonal


n‑Type SnS2 and Orthorhombic p‑Type SnS Crystals
Ji-Hoon Ahn,†,∥ Myoung-Jae Lee,†,∥ Hoseok Heo,†,‡ Ji Ho Sung,†,‡ Kyungwook Kim,†,§ Hyein Hwang,†,§
and Moon-Ho Jo*,†,‡,§

Center for Artificial Low-Dimensional Electronic Systems, Institute for Basic Science (IBS), ‡Division of Advanced Materials Science,
and §Department of Materials Science and Engineering, Pohang University of Science and Technology (POSTECH), 77
Cheongam-Ro, Pohang 790-784, Korea
*
S Supporting Information
Downloaded from pubs.acs.org by UNIV OF WINNIPEG on 01/24/19. For personal use only.

ABSTRACT: van der Waals layered materials have large crystal


anisotropy and crystallize spontaneously into two-dimensional
(2D) morphologies. Two-dimensional materials with hexagonal
lattices are emerging 2D confined electronic systems at the limit
of one or three atom thickness. Often these 2D lattices also
form orthorhombic symmetries, but these materials have not
been extensively investigated, mainly due to thermodynamic
instability during crystal growth. Here, we show controlled poly-
morphic growth of 2D tin-sulfide crystals of either hexagonal
Nano Lett. 2015.15:3703-3708.

SnS2 or orthorhombic SnS. Addition of H2 during the growth


reaction enables selective determination of either n-type SnS2
or p-type SnS 2D crystal of dissimilar energy band gap of
2.77 eV (SnS2) or 1.26 eV (SnS) as a final product. Based on
this synthetic 2D polymorphism of p−n crystals, we also demonstrate p−n heterojunctions for rectifiers and photovoltaic cells,
and complementary inverters.
KEYWORDS: van der Waals layered materials, two-dimensional materials, tin disulfides, tin monosulfides, vapor transport synthesis,
polymorphism

I n hexagonal van der Waals layered crystals, such as


graphene,1,2 h-BN,3,4 and hexagonal transition-metal dichal-
cogenides,5−9 the constituent atoms within the monolayer plane
comprises zigzag double planes of the Sn and chalcogen atoms
separated by a van der Waals gap (Figure 1b). In bulk crystal
forms, hexagonal SnS2 and orthorhombic SnS exhibit n-type
are covalently bonded with a large bonding energy of 200− and p-type semiconductor characteristics, respectively.17,18
6000 meV, whereas the individual monolayer are vertically Growth of SnS2 crystals is thermodynamically stable in ambient
joined by weak van der Waals interactions with energies of conditions, and its 2D crystal absorbs visible light effectively.19 In
40−70 meV.10,11 Typically, these substances spontaneously form contrast, 2D SnS crystals have a narrow Eg and are therefore
two-dimensional (2D) crystals, and thereby establish unit-cell expected to be optically active in the near-infrared spectral
confined electronic systems in a hexagonal momentum space. In range.19 The fact that the SnS2 and SnS states exhibit n-type and
this regard, Sn-sulfides are particularly interesting class of the 2D p-type characters and dissimilar Eg and are thus optically active in
semiconductors with layered crystal structures because these complementary broad spectral ranges suggests that a synthetic
sulfides exist in diverse crystal phases, such as hexagonal and polymorphism of 2D p−n components may have various
orthorhombic, due to the versatile oxidation characteristics electronic and optical applications. Here, we report deterministic
of Sn and chalcogen elements. Notably, Sn-dichalcogenides, polymorphism growth of 2D hexagonal SnS2 and orthorhombic
such as SnS2 and SnSe2, crystallize two-dimensionally into SnS crystals by tuning the amount of H2 added during gas-phase
hexagonal unit cells with the Sn oxidation state of +4, to form synthesis.
semiconductors with a large band gap Eg12−14 in which the Sn In our study, the 2D Sn-sulfide crystals were synthesized on
ions are coordinated to six chalcogen ions in the octahedral sites SiO2/Si substrates by a vapor transport method from pure SnO2
with space group P3m ̅ 1 within a monolayer, which is stacked on and S powder precursors in a 12-in. hot-wall quartz-tube.20 Before
top of another monolayer by van der Waals interaction without synthesis, we performed simple thermodynamic calculations
translational displacements (Figure 1a). However, Sn-chalcogenides of the gas-phase reactions from SnO2 and S precursors. In the
can also crystallize in orthorhombic unit cells to form 2D
Sn-monochalcogenides15,16 in which Sn ions with oxidation Received: January 8, 2015
state of +2 are coordinated to three chalcogen ions to form an Revised: April 11, 2015
orthorhombic unit cell with the space group of Pnma, which Published: May 1, 2015

© 2015 American Chemical Society 3703 DOI: 10.1021/acs.nanolett.5b00079


Nano Lett. 2015, 15, 3703−3708
Nano Letters Letter

Figure 1. Crystal structures of (a) hexagonal SnS2 and (b) orthorhombic SnS crystals in different sectional views. The Sn and sulfur atoms are colored in
blue and orange, respectively. (c) Change in Gibbs free energy ΔG° during formation of SnS2 and SnS compounds from SnO2 and S in N2 atmosphere
and in N2−H2 atmosphere. Gray line, SnS2 and SnS in N2; blue line, SnS in N2; red line, SnS in N2−H2.

standard condition, the reaction can be predicted by the change Detailed growth procedures are depicted in the Figure S1 of
ΔG°rxn in the standard Gibbs free energy, which is a function of Supporting Information. Under pure N2 gas at 620−680 °C,
temperature as follows: large-area 2D crystals (more than several tens of microns in
width) with hexagonal or triangular facets were obtained
ΔG°rxn = ΔH °f,rxn − T ΔS°f,rxn (Figure 2a); i.e., reaction of SnO2 with S vapor under the inert
⎛ T ⎞ condition yields hexagonal SnS2 crystals with release of SO2 gas
= ⎜ΔH °298,rxn +

∫298 ΔCp,rxn dT ⎟⎠ as a byproduct. The thickness of the 2D crystals decreased
as the gas-flow rate decreased and could be as thin (∼1.1 nm) as a
⎛ T ΔCp ,rxn ⎞ unit lattice of hexagonal SnS2 (Figure 2c, inset). Atomic force
− Trxn⎜⎜ΔS°298,rxn + ∫298 dT ⎟ microscopy (AFM) line profiles confirmed that the facets were of
⎝ T ⎠ (1) uniform thickness except at the crystal centers which may be
where ΔXrxn = Σ(# of moles)Xproducts − Σ(# of moles)Xreactants, nuclei for the 2D growth.7
ΔH°f and S°f are the standard enthalpy and entropy of formation, Raman spectra (Figure 2c) were obtained for crystals of
ΔH°298 and S°298 are the standard enthalpy and entropy of various thickness. The A1g phonon mode at 317 cm−1 is assigned
formation at 298 K, T is the reaction temperature [K], and Cp is to the 2D SnS2 crystals,22−24 and the Eg phonon mode at 208 cm−1
the specific isobar heat capacity. When the SnO2 powders are corresponds to thick SnS2 crystals.25 As the thickness decreased
heated to react with S2 gas in an inert atmosphere, reaction forms to the nanometer scale, the Eg peak disappeared, presumably due
SnS or SnS2 by to the reduction in the scattering centers for in-plane scattering,26
and these match well with previously reported spectra of
SnO2 (s) + S2 (g) → SnS(s) + SO2 (g) (Reaction 1) nanostructured SnS2.23,24 When the carrier gas included H2, the
growth products were typically transformed to rectangular 2D
SnO2 (s) + 1.5S2 (g) → SnS2 (s) + SO2 (g) (Reaction 2) facets (Figure 2b, inset; Figure S2b). Addition of H2 to the gas
Using eq 1 and the thermodynamic data (Table S1),21 encouraged growth of orthorhombic SnS 2D crystals. Typically,
calculated values of ΔG°rxn of each reaction in the pertinent these crystals were favored at H2/N2 > 0.4. In this case, SnO2
temperature range 500−700 °C were positive for SnS growth but reacts with S vapor by H2 addition, then orthorhombic SnS crys-
negative for SnS2 growth (the red curve in Figure 1c); i.e., SnS tals grow with release of SO2 and H2S gas byproducts (Figure 2b).
growth from SnO2 and S precursors is not spontaneous in an The minimum thickness was ∼12.1 nm, which corresponds to
inert atmosphere. Nonetheless, because the Sn oxidation state in 10 or 11 unit cells. Below the H2/N2 ratio of 0.4, the final
SnS is +2, which is less than the +4 in SnS2, SnS growth can be products start to form irregular facets, and near the H2/N2 ratio
promoted in a reducing atmosphere. Therefore, we considered a ∼0, the hexagonal facet is stabilized (see Figure S2 for more
reaction with the addition of H2: details). The fact that the minimum achievable thickness of SnS is
thicker than that of the SnS2 may be due to the relatively larger
SnO2 (s) + 1.5S2(g) + H 2(g) van der Waals energy in the orthorhombic cell with its zigzag
arrangement of Sn and S atoms compared to the hexagonal cell.
→ SnS(s) + SO2 (g) + H 2S(g) (Reaction 3)
The Raman spectra (Figure 2d) typically showed four major
ΔG°rxn for SnS growth with H2 addition is negative and is also peaks; those at 94, 188, and 217 cm−1 can be assigned to the Ag
lower than that of SnS2 in the temperature range of interest phonon modes, and that at 160 cm−1 corresponds to the B3g
(Figure 1c). This result indicates that we can control the final mode of orthorhombic SnS.27,28
product by adding H2 to influence ΔG°rxn during growth. High-resolution transmission electron microscopy investiga-
Inspired by the result of this thermodynamic calculation, we tions revealed the polymorphic phases of the 2D SnS2 and SnS
designed a series of reactions to grow 2D SnS2 and SnS crystals crystals. Diffraction patterns constructed by fast Fourier
(Figure S1). To emulate the standard conditions of thermo- transform of each image (insets in Figure 2e,f) corroborate the
dynamics, we established an inert (N2) or a reducing (N2/H2) phase index in each growth condition. The measured interplanar
ambient (total pressure 700−800 Torr) during crystal growth. distances were 0.317 (Figure 2e) and 0.293 nm (Figure 2f),
3704 DOI: 10.1021/acs.nanolett.5b00079
Nano Lett. 2015, 15, 3703−3708
Nano Letters Letter

Figure 2. Growth schematics and representative optical microscope images of (a) hexagonal 2D SnS2 in N2 and (b) orthorhombic 2D SnS in N2−H2.
Raman spectra of (c) SnS2 2D crystals and (d) SnS 2D crystals of various thickness. Insets: atomic force microscope images. High resolution
transmission electron microscope images of (e) a SnS2 crystal and (f) a SnS crystal. The corresponding FFT-diffraction patterns of the insets clearly
show the hexagonal and orthorhombic lattices.

which are consistent with the (100) plane of the hexagonal SnS2 thickness dependence of spectral photocurrent (Figure 3a). The
and the (101) plane of the orthorhombic SnS, respectively. The range of photoresponse of SnS2 crystals was expanded to the
interaxial angles of 120° and 94.9°/85.1° are also consistently lower energy of ∼2.1 eV, and the corresponding absorption edge
assigned to SnS2 and SnS, respectively. shows a red-shift with increasing thickness. The optical band gap
To characterize the specific semiconductor properties of 2D notably increased as t decreased toward the monolayer regime
SnS2 and SnS crystals, we fabricated simple back-gated transistors (Figure 3b), and the extracted Eg of 2.77 eV in our 2D SnS2 is
that incorporate individual crystals on SiO2/degenerate Si sub- significantly higher than those of the bulk SnS2 of 1.82−2.2 eV
strates. Metal contacts were fabricated with Ti/Au and Ni/Au across the indirect Eg,12,32,33 which values match well with our
as the source/drain electrodes for SnS2 (3 nm thick) and SnS SnS2 above 10 nm in thickness. Thereby, this observation sug-
(30 nm thick) crystals, respectively. First, the spectral photo- gests that our 2D crystals exhibit optical confinement effects, by
current Iph responses were measured while illuminating the which the absorption edge progressively increases as crystal
devices with a supercontinuum laser equipped with a thickness decreases.34 Our observations are consistent with
monochromator. We confirmed that the major photoresponses predictions by density-functional tight-bonding calculation that
are spatially from the channels, not from the contact barriers, by indirect band gap size increases ∼2.81 eV for SnS2 monolayers.35
scanning Iph mapping (Figure S3a,b), thus confirming the In contrast, the 2D SnS crystals had an absorption edge of
effectively intrinsic photoresponses of the crystals. The 1.26 eV (Figure 3c), which is within the range of values reported
absorption edge can be determined by converting from measured previously (0.9−1.27 eV), suggesting absence of the size-effect in
spectral Iph to effective absorption coefficient α by the relation29 this thickness regime.36,37
α = −1/t(1 −(Iph/(1 − R))(hυ/eηP)), where hυ is the photon The dark electrical conductivity was 2.86 × 103 S/m for the 2D
energy, t is the thickness of the 2D crystal, P is the incident optical SnS crystal and 2.17 S/m for the 2D SnS2, and the photo-
power, R is the reflectance, and η is the photon-to-carrier con- conductivity was 3.85 × 103 and 85.23 S/m at 2.33 and 3.06 eV,
version efficiency. We assume that η = 1, and that the voltage respectively (Figure S3c,d). The higher conductivities of 2D SnS
application has negligible effect on R.30 The relation between hυ crystals can be attributed to the narrower Eg compared to 2D
and α is expressed by (αhυ)m = B(hυ − Eg), where B is a constant, SnS2.16,32 The gate voltage (Vg)-dependent transport character-
and with m = 2 for direct band gap transition and m = 1/2 for istics at Vds = 1.0 V show the typical n-type and p-type characters
indirect band gap transition.31 Because both SnS2 and SnS for the SnS2 and SnS field-effect transistors (FETs), respectively
semiconductors in the bulk forms to show indirect band gap (Figure 4a). The SnS2 n-FET had the on/off current ratio of
transitions, the optical band gap can be extracted by extrapolating 2 × 104, and the field effective electron mobility μe = 2.16 cm2 V−1 s−1
the linear region of a (αhυ)1/2 vs hυ plot. We investigated the at room temperature; this is larger than μe ≈ 1 cm2 V−1 s−1 reported
3705 DOI: 10.1021/acs.nanolett.5b00079
Nano Lett. 2015, 15, 3703−3708
Nano Letters Letter

Figure 3. Thickness dependence of (a) spectral responsivity (Iph/hυ) and (b) optical bandgap of SnS2 crystal extracted by extrapolating the linear region
of inset of (αhυ)1/2 vs hυ plot. (c) Spectral responsivity of 30 nm-thick SnS field-effect transistors. Insets: plots of (αhυ)1/2 vs hυ for determination of
band gap.

Figure 4. (a) Transfer characteristics of SnS2 and SnS back-gate transistors. Blue line, SnS2 transistor showing typical n-type characteristics; red line, SnS
transistor with p-type characteristics. (b) Gate tunable output characteristics of SnS2−SnS vertical heterojunctions; inset, device schematics. (c) Dark
I−V curve and Iph−V curve under 3.06 eV light excitation with a power of 3.2 μW; (left inset, photovoltaic I−V curves, showing the open-circuit voltage
of 0.21 V and the short-circuit current of 1.69 nA; right inset, corresponding band diagram of the SnS2−SnS vertical heterojunctions at Vb = 0 V,
illustrating the photovoltaic effect).

from an exfoliated 2D SnS2.17 The SnS p-FET had on/off ratio approximated by interlayer recombination processes between
of only ∼1.5, but the field effective hole mobility was μh = two majority carriers across the abrupt potential discontinuity,
10.55 cm2 V−1 s−1. Having established the synthetic Sn-sulfide such as by Langevin recombination or Schokley−Read−Hall
polymorphism for the 2D p−n components, we built vertical and recombination mediated by the interlayer defect states, as
lateral devices by incorporating individual 2D SnS2 and SnS suggested by the work of an MoS2/WSe2 monolayer p−n stack.39
crystals. We first stacked the two 2D crystals by manual transfer Investigations of these unique 2D phenomena will be a focus of
to construct the vertical p−n heterojunctions for rectifiers and our future work. Our polymorphic 2D p−n stack can operate
photovoltaic cells. The Vg-dependent output characteristics of as an ultrathin photovoltaic cell.40 Under 405 nm illumination
the 2D SnS2−SnS vertical heterojunction showed a rectifying with a power of 3.2 μW, the photoresponsivity for forward bias
diode behavior (Figure 4b), which is effectively modulated by the as a photoconductor was 4.56 mA/W and reverse bias as a
applied electric field Vg.38 The output current of the diode is photodiode was 27.09 mA/W (Figure 4c), which are moderate
largely governed by the higher resistivity of the n-type SnS2 than values with other 2D-based photodetectors.41 Our p−n junction
of the p-type SnS in the p−n series resistor and thus increases as shows a photovoltaic effect with an open-circuit voltage of
Vg increases. The rectification ratio = (forward current)/(reverse ∼0.21 V and a short-circuit current of ∼1.69 nA (inset, Figure 4c).
current) at bias voltage Vb = ±2 V increased from 9.4 to 33.7 as Vg The corresponding external quantum efficiency was calculated
increased from −40 to 40 V. The diode parameters from the to be ∼0.13%, which is comparable to those of other high-
Shockley diode equation with a series resistance Rs, which is performance monolayer semiconductor 2D p−n junctions.42,43
related to the metal/SnSx contacts were deduced; extracted On the basis of the extracted band gap and electron affinity
values were saturation current Is = 0.04 nA, Rs = 0.52 GΩ, and available in the literature of 4.2 eV for SnS2 and 3.14 eV for
ideality factor n = 6.7. These values of Rs and n differ greatly from SnS,44,45 the band diagram of the heterojunction at Vb = 0 V can
the ideal values. However, different from a conventional p−n be illustrated (right inset, Figure 4c). We assume a type-II flat
diode, the 2D p−n junction diodes do not allow a depletion band alignment for simplicity, as discussed above. Our photo-
region across the two adjacent layers, so the classical exponential voltaic cell is reminiscent of organic heterojunction cells in that
characteristics may not be representative. They can be better charge separation arises from discontinuous energy alignments at
3706 DOI: 10.1021/acs.nanolett.5b00079
Nano Lett. 2015, 15, 3703−3708
Nano Letters Letter

the heterointerfaces, in this case the band offsets of ∼0.2 eV V.; Grigorenko, A. N.; Geim, A. K.; Casriaghi, C.; Castro Neto, A. H.;
between the SnS2 conduction (EC,SnS2) and SnS valence (EV,SnS) Novoselov, K. S. Science 2013, 340, 1311−1314.
band. The maximum open-circuit voltage of our photovoltaic (10) Wilson, J. A.; Yoffe, A. D. Adv. Phys. 1969, 18, 193−335.
cells of ∼0.2 V under 405 nm illumination (inset, Figure 4c) is (11) Butler, S. Z.; Hollen, S. M.; Cao, L.; Cui, Y.; Gupta, J. A.;
Gutiérrez, H. R.; Heinz, T. F.; Hong, S. S.; Huang, J.; Ismach, A. F.;
consistent with this band-offset approximation.41,44 We ensure
Johnston-Halperin, E.; Kuno, M.; Plashnitsa, V. V.; Robinson, R. D.;
that the observed photovoltaic responses only pertain to the Ruoff, R. S.; Salahuddin, S.; Shan, J.; Shi, L.; Spencer, M. G.; Terrones,
junction from scanning Iph mapping, where the short circuit M.; Windl, W.; Goldberger, J. E. ACS Nano 2013, 7, 2898−2926.
current is localized at the junction between n-type SnS2 and (12) Sanchez-Juarez, A.; Ortiz, A. Semicond. Sci. Technol. 2002, 17,
p-type SnS (Figure S5). As another demonstration of 2D p−n 931−937.
polymorphism, we constructed a complementary metal−oxide− (13) Huang, Y.; Sutter, E.; Sadowski, J. T.; Cotlet, M.; Monti, O. L.;
semiconductor (CMOS) inverter (Figure S6); the observed general Racke, D. A.; Neupane, M. R.; Wickramaratne, D.; Lake, R. K.;
CMOS inverter features qualitatively suggest a possibility of 2D Parkinson, B. A.; Sutter, P. ACS Nano 2014, 8, 10743−10755.
logic operations based on the synthetic 2D p−n polymorphism. (14) Huang, L.; Yu, Y.; Li, C.; Cao, L. J. Phys. Chem. C 2013, 117,
In summary, we successfully synthesized 2D tin sulfide crystals 6469−6475.
of either hexagonal SnS2 or orthorhombic SnS and determined (15) Liu, X.; Li, Y.; Zhou, B.; Wang, X.; Cartwright, A. N.; Swihart, M.
that the type of crystal formed can be controlled by adding H2 to T. Chem. Mater. 2014, 26, 3515−3521.
(16) Deng, Z.; Cao, D.; He, J.; Lin, S.; Lindsay, S. M.; Liu, Y. ACS Nano
the feed gas to control thermodynamics during growth. Our 2D 2012, 6, 6197−6207.
polymorphic crystals show n-type (SnS2) and p-type (SnS) (17) De, D.; Manongdo, J.; See, S.; Zhang, V.; Guloy, A.; Peng, H.
semiconductor characteristics, and we demonstrated the Nanotechnology 2013, 24, 025202.
feasibility of using the crystals as polymorphic 2D heterostruc- (18) Hegde, S. S.; Kunjomana, A. G.; Chandrasekharan, K. A.; Ramesh,
ture device for rectifiers, photovoltaic cells, and complementary K.; Prashantha, M. Phys. B 2011, 406, 1143−1148.
inverters. Our methods may guide development of synthetic (19) Xu, M.; Liang, T.; Shi, M.; Chen, H. Chem. Rev. 2013, 113, 3766−
polymorphism of other 2D materials for the 2D electronic and 3798.
optoelectronic heterostructures. (20) Sung, J. H.; Heo, H.; Hwang, I.; Lim, M.; Lee, D.; Kang, K.; Choi,


H. C.; Park, J.-H.; Jhi, S.-H.; Jo, M. H. Nano Lett. 2014, 14, 4030−4035.
ASSOCIATED CONTENT (21) Kubaschewski, O.; et al. Materials Thermodynamics; Pergamon
Press: Oxford, U.K., 1993; pp 258−322.
*
S Supporting Information
(22) Wang, C.; Tang, K.; Yang, Q.; Qian, Y. Chem. Phys. Lett. 2002,
Experimental details of polymorphism growth, photocurrent 357, 371−375.
mapping and Raman characterization, and demonstration of (23) Qu, B.; Ma, C.; Ji, G.; Xu, C.; Xu, J.; Meng, Y. S.; Wang, T.; Lee, J.
CMOS inverter. The Supporting Information is available free of Y. Adv. Mater. 2014, 26, 3854−3859.
charge on the ACS Publications website at DOI: 10.1021/ (24) Du, Y.; Yin, Z.; Rui, X.; Zeng, Z.; Wu, X. J.; Liu, J.; Zhu, Y.; Zhu, J.;
acs.nanolett.5b00079. Huang, X.; Yan, Q.; Zhang, H. Nanoscale. 2013, 5, 1456−1459.

■ AUTHOR INFORMATION
Corresponding Author
(25) Price, L. P.; Parkin, I. P.; Hardy, A. M. E.; Clark, R. J. H.; Hibbert,
T. G.; Molloy, K. C. Chem. Mater. 1999, 11, 1792−1799.
(26) Ley, S.; Ge, L.; Liu, Z.; Najmaei, S.; Shi, G.; You, G.; Lou, J.; Vajtai,
R.; Ajayan, P. M. Nano Lett. 2013, 13, 2777−2781.
*E-mail: mhjo@postech.ac.kr.
(27) Zhang, X.; Yang, L.; Jiang, Y.; Yu, B. B.; Zou, Y. G.; Fang, Y.; Hu, J.
Author Contributions S.; Wan, L. J. Chem.Asian J. 2013, 8, 2483−2488.

These authors contributed equally to this work. (28) Chao, J.; Wang, Z.; Xu, X.; Xiang, Q.; Song, W.; Chen, G.; Hu, J.;
Notes Chen, D. RCS Adv. 2013, 3, 2746−2753.
The authors declare no competing financial interest. (29) Pankove, J. I. Optical Properties of Solid; Oxford University Press:


New York, 2001; Chapter 3.
ACKNOWLEDGMENTS (30) Jongthammanurak, S.; Liu, J.; Wada, K.; Cannon, D. D.;
Danielson, D. T.; Pan, D.; Kimerling, L. C.; Michel, J. Appl. Phys. Lett.
This work was supported by Institute for Basic Science (IBS), 2006, 89, 161115.
Korea, under the Project Code (IBS-R014-G1).


(31) Koffyberg, F. P.; Dwight, K.; Wold, A. Solid State Commun. 1979,
30, 433−437.
REFERENCES (32) Amalraj, L.; Sanjeeviraja, C.; Jayachandran, M. J. Cryst. Growth.
(1) Geim, A. K.; Novoselov, K. S. Nat. Mater. 2007, 6, 183−191. 2002, 234, 683−689.
(2) Geim, A. K. Science 2009, 324, 1530−1534. (33) Domingo, G.; Itoga, R. S.; Kannewurf, C. R. Phys. Rev. 1966, 143,
(3) Watanabe, K.; Taniguchi, T.; Kanda, H. Nat. Mater. 2004, 3, 404− 536−541.
409. (34) Yue, G. H.; Wang, L. S.; Wang, S.; Chen, Y. Z.; Peng, D. L.
(4) Dean, C. R.; Young, A. F.; Meric, I.; Lee, C.; Wang, L.; Sorgenfrei, Nanoscale Res. Lett. 2009, 4, 359−363.
S.; Watanabe, K.; Taniguchi, T.; Kim, P.; Shepard, K. L.; Hone, J. Nat. (35) Lorentz, T.; Joswig, J.; Seifert, G. Semicond. Sci. Technol. 2014, 29,
Nanotechnol. 2010, 5, 722−726. 064006.
(5) Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, (36) Zainal, Z.; Hussein, M. Z.; Ghazali, A. Sol. Energy Mater. Sol. Cells.
M. S. Nat. Nanotechnol. 2012, 7, 699−712. 1996, 40, 347−357.
(6) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang, H. (37) Lopez, S.; Ortiz, A. Semicond. Sci. Technol. 1994, 9, 2130−2133.
Nat. Chem. 2013, 5, 263−275. (38) Fang, H.; Battaglia, C.; Carraro, C.; Nemsak, S.; Ozdol, B.; Kang,
(7) Najmaei, S.; Liu, Z.; Zhou, W.; Zou, X.; Shi, G.; Lei, S.; Yakobson, J. S.; Bechtel, H. A.; Desai, S. B.; Kronast, F.; Unal, A. A.; Conti, G.;
B. I.; Idrobo, J.; Ajayan, P. M.; Lou, J. Nat. Mater. 2013, 12, 754−759. Colon, C.; Palsson, G. K.; Martin, M. C.; Minor, A. M.; Fadley, C. S.;
(8) van der Zande, A. M.; Huang, P. Y.; Chenet, D. A.; Berkelbach, T. Yablonovitch, E.; Maboudian, R.; Javey, A. Proc. Natl. Acad. Sci. U.S.A.
C.; You, Y.; Lee, G.-H.; Heinz, T. F.; Reichman, D. R.; Muller, D. A.; 2014, 111, 6198−6202.
Hone, J. C. Nat. Mater. 2013, 12, 554−561. (39) Lee, C.-H.; Lee, G.-H.; van der Zande, A. M.; Chen, W.; Li, Y.;
(9) Britnell, L.; Ribeiro, R. M.; Eckmann, A.; Jalil, R.; Belle, B. D.; Han, M.; Cui, X.; Arefe, G.; Nuckolls, C.; Heinz, T. F.; Guo, J.; Hone, J.;
Mishchenko, A.; Kim, Y.-J.; Gorbachev, R. V.; Georgiou, T.; Morozov, S. Kim, P. Nat. Nanotechnol. 2014, 9, 676−681.

3707 DOI: 10.1021/acs.nanolett.5b00079


Nano Lett. 2015, 15, 3703−3708
Nano Letters Letter

(40) Britnell, L.; Ribeiro, R. M.; Eckmann, A.; Jalil, R.; Belle, B. D.;
Mishchenko, A.; Kim, Y.-J.; Gorbachev, R. V.; Georgiou, T.; Morozov, S.
V.; Grigorenko, A. N.; Geim, A. K.; Casiraghi, C.; Castro Neto, A. H.;
Novoselov, K. S. Science 2013, 340, 1311−1314.
(41) Buscema, M.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; van
der Zant, H. S. J.; Castellanos-Gomez, A. Nano Lett. 2014, 14, 3347−
3352.
(42) Furchi, M. M.; Pospischil, A.; Libisch, F.; Burgdörfer, J.; Mueller,
T. Nano Lett. 2014, 14, 4785−4791.
(43) Deng, Y.; Luo, Z.; Conrad, N. J.; Liu, H.; Gong, Y.; Najmaei, S.;
Ajayan, P. M.; Lou, J.; Xu, X.; Ye, P. D. ACS Nano 2014, 8, 8292−8299.
(44) Williams, R. H.; Murray, R. B.; Govan, D. W.; Thomas, J. M.;
Evans, E. L. J. Phys. C: Solid State Phys. 1973, 6, 3631−3642.
(45) Devika, M.; Reddy, N. K.; Patolsky, F.; Gunasekhar, K. R. J. Appl.
Phys. 2008, 104, 124503.

3708 DOI: 10.1021/acs.nanolett.5b00079


Nano Lett. 2015, 15, 3703−3708

You might also like