Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/332016040

Electronic structures and transport properties of SnS-SnSe nanoribbon lateral


heterostructure

Article in Physical Chemistry Chemical Physics · March 2019


DOI: 10.1039/C9CP00427K

CITATIONS READS

5 273

6 authors, including:

Guo Yandong Dewei Rao


Macau University of Science and Technology Jiangsu University
50 PUBLICATIONS 223 CITATIONS 120 PUBLICATIONS 3,923 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

the cathode designing of Li-S battery View project

membrane separation View project

All content following this page was uploaded by Dewei Rao on 06 June 2019.

The user has requested enhancement of the downloaded file.


PCCP
PAPER

Electronic structures and transport properties of


SnS–SnSe nanoribbon lateral heterostructures†
Cite this: Phys. Chem. Chem. Phys.,
2019, 21, 9296
Yang Yang, ab Yuhao Zhou,ab Zhuang Luo,a Yandong Guo, b
Dewei Rao *a
and Xiaohong Yan*ab

The electronic structures of phosphorene-like SnS/SnSe nanoribbons and the transport properties of a
SnS–SnSe nanoribbon lateral heterostructure are investigated by using first-principles calculations
combined with nonequilibrium Green’s function (NEGF) theory. It is demonstrated that SnS and SnSe
nanoribbons with armchair edges (A-SnSNRs and A-SnSeNRs) are semiconductors, independent of the
width of the ribbon. Their bandgaps have an indirect-to-direct transition, which varies with the ribbon width.
In contrast, Z-SnSNRs and Z-SnSeNRs are metals. The transmission gap of armchair SnSNR–SnSeNR is
Received 23rd January 2019, different from the potential barrier of SnSNR and SnSeNR. The I–V curves of zigzag SnSNR–SnSeNR exhibit
Accepted 26th March 2019 a negative differential resistive (NDR) effect due to the bias-dependent transmission in the voltage window
DOI: 10.1039/c9cp00427k and are independent of the ribbon width. However, for armchair SnSNR–SnSeNR, which has a low
current under low biases, it is only about 10 6 mA. All the results suggest that phosphorene-like MX
rsc.li/pccp (M = Sn/Ge, X = S/Se) materials are promising candidates for next-generation nanodevices.

1 Introduction puckered structures, the electronic transport properties of MX


monolayers have been widely investigated in recent years.
Two-dimensional materials, such as graphene, hexagonal Zhang et al. showed that monolayer b-GeSe exhibits anisotropic
boron nitride (h-BN), phosphorene, and transition metal electron/hole transport with a high electron mobility of 2.93 
dichalcogenides (TMDs), have attracted increasing attention 104 cm2 V 1 along the armchair direction.29 An et al. studied
due to their unique properties and potential applications in the electronic transport properties of zigzag MX nanoribbons.30
nanodevices.1–10 Monolayer and few-layer materials exhibit Their calculations revealed that all these nanoribbons show a
novel physical properties, different from their bulk materials, negative differential resistive (NDR) effect. Li et al. explored
which inspired the design of new 2D materials with unique theoretically the electronic properties of monolayer GeS.31 They
properties. found that monolayer GeS has a high electron mobility, nearly
Recently, monolayers of group-IV monochalcogenides, MX 3680 cm2 V 1, which is much higher than that of monolayer
(M = Sn/Ge, X = S/Se), have triggered considerable interest.11–20 MoS2 (B200 cm2 V 1).
Theoretical studies have revealed that MX materials have Currently, two-dimensional heterostructures based on atom-
indirect bandgaps, except for monolayer GeSe.21,22 Several ically thin 2D materials have attracted considerable interest
studies found that their bandgaps can be tuned by strain and and play a crucial role in next generation high-performance
electric fields.23–27 It has been demonstrated that in-plane nanodevices.32–39 Two-dimensional heterostructures can be
compressive strain applied to a two-dimensional MX can effec- created by stacking 2D materials layer-by-layer vertically (vertical
tively reduce bandgaps and eventually lead to a semiconductor- heterostructures) or by aligning their in-plane interfaces (lateral or
to-metal transition that shows unique properties. Moreover, in-plane heterostructures). Compared to the vertical heterostruc-
Huang et al. reported that the difference in the bandgaps is tures, the lateral heterostructures show a smaller interface area, a
induced by structural anisotropy.28 Due to their fascinating higher epitaxial quality, simpler band alignment, and distinct
electronic structure, such as that shown by phosphorene-like phase separation.40 These properties are extremely beneficial for
the design of nanodevices. As a direct consequence, the lateral
heterostructures based on group-IV monochalcogenides have also
a
School of Materials Science and Engineering, Jiangsu University, been extensively studied. Zhao et al. investigated the band align-
Zhenjiang 212013, China. E-mail: dewei@ujs.edu.cn
b
ment and the electronic properties of lateral heterostructures made
College of Electronic and Optical Engineering, Nanjing University of Posts and
Telecommunications, Nanjing 210046, China. E-mail: yanxh@ujs.edu.cn
of monolayer group-IV monochalcogenides.41 They demonstrated
† Electronic supplementary information (ESI) available: Geometric and electronic that several lateral heterostructures show type-II band alignment,
structures of nanoribbons. See DOI: 10.1039/c9cp00427k which makes their use in photovoltaic applications possible.

9296 | Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 This journal is © the Owner Societies 2019
Paper PCCP

Because of their fascinating geometrical and electronic structures, of armchair nanoribbons of SnS and SnSe with a width from
the lateral heterostructures built with group-IV monochalcogenides 6 to 12 atomic chains and their zigzag nanoribbons were built.
may have potential applications in many fields, especially in These structures are denoted as nA-SnSNR or nA-SnSeNR, and
nanoelectronics. However, only a few works have reported their nZ-SnSNR or nZ-SnSeNR, with n = 6–12, respectively, and they
application in transport properties. are displayed in Fig. 1a and Fig. S1–S4 in the ESI.† Initially, the
In this work, the electronic structures and transport proper- electronic structures, including band structures and density of
ties of SnS–SnSe nanoribbon lateral heterostructures are inves- states (DOS), were studied. They are shown in Fig. S5–S8 (ESI†).
tigated. The results show that all the armchair nanoribbons are Clearly, all the zigzag ribbons are metallic since two bands
semiconductors and have an indirect-to-direct transition with cross the Fermi level, which are highlighted by red and blue
the expanding width of the ribbon. In contrast, all the zigzag lines, respectively, and their states are distributed both above
nanoribbons are metallic. The transport properties of the and below the Fermi level. The electronic structures of the
lateral SnSNR–SnSeNR heterostructures are discussed in detail armchair nanoribbons should be analysed in detail for their
in the following section. The zigzag SnS–SnSe nanoribbons tuneable band structure, which changes with their width.
exhibit the NDR effect, which is independent of the nanoribbon Fig. 1b shows that the band gaps of the armchair SnSNRs
width. For the armchair SnS–SnSe nanoribbons, a low current and SnSeNRs with increasing width fluctuate. This means
through the nanoribbon under low biases is observed. These that the nanoribbons of the group-IV monochalcogenides
data indicate that the SnSNR–SnSeNR lateral heterostructures can be regulated by varying their widths. Similar phenomena
have potential applications in the next-generation of nanodevices. have been reported for MoS2 nanoribbons and graphene nano-
This work may provide more guidance in the design of nanodevices ribbons. The band gaps of the armchair SnSNRs are in the
based on other two-dimensional materials, such as group-6 TMDs range of 1.3–1.45 eV and the band gaps of the SnSeNRs have an
and group-10 TMDs.42–45 energy range of 0.8–1.2 eV, which is smaller than that of the
SnSNRs. The bandgaps are significantly transformed between
indirect and direct gaps, as colored in Fig. 1b. For the SnSNRs,
2 Computational methods ribbons with a width of 7A and 11A are indirect-gap semicon-
ductors while 6A-, 8A-, and 10A-SnSeNRs are indirect-gap
The atomistix toolkit (ATK) code package based on density
semiconductors. Details of these band structures are displayed
functional theory combined with the nonequilibrium Green’s
in Fig. S5 and S6 (ESI†). These interesting phenomena can be
function and Vienna ab initio simulation package (VASP) were
explained by the quantum confinement effect.
employed in this work to perform the calculations.46–48 For ATK,
the generalized gradient approximation (GGA) of Perdew–Burke–
3.2 Transport properties of SnSNR–SnSeNR heterostructure
Ernzerhof (PBE) was used to describe the exchange–correlation
nanoribbons
potential. The SG15 pseudopotentials, which are considered to
be the state of the art for norm-conserving pseudopotentials, and As has been mentioned above, two-dimensional heterostructures
medium basis sets were employed in the transport calculation. could be potential candidates for the design of nanodevices.
The Kohn–Sham Hamiltonian was evaluated in a real space grid Previous works demonstrated that good electronic transport can
with a density mesh cut-off of 500 eV. The Brillouin zone be obtained from pristine MX nanoribbons. Despite this, their
integration was performed using a 1  1  11 Monkhorst–Pack heterostructure nanoribbons have not been studied systemati-
k-point sampling for the electronic structures and a 1  1  100 cally. In this section, the transport properties of both armchair
sampling for the transport calculations. The convergence criter- and zigzag SnSNR–SnSeNR heterostructure nanoribbons are
ion for the total energy was set at a value of 10 5 eV. All the explored. The two-probe device configuration of a 6A-SnSNR–
geometries were fully relaxed until the force on each atom SnSeNR is presented in Fig. 2a. It consists of three regions: a left
reached less than 0.01 eV Å 1. For VASP, ion–electron inter- electrode, a right electrode, and a central region. In the ATK, the
actions were described by a projected augmented wave, in which left and the right electrodes are periodic and perpendicular to
the exchange–correlation functional was Perdew–Burke–Ernzer-
hof within the generalized gradient approximation. Numerical
convergence was achieved with thresholds of 10 5 eV (10 6 eV
for DOS calculations) in energy and 10 2 eV Å 1 in force with a
cutoff energy of 500 eV. Furthermore, a vacuum spacing of at
least 10 Å was added in the direction perpendicular to the z axis
to avoid the interaction between adjacent layers.

3 Results and discussion


3.1 Electronic structures of SnS and SnSe nanoribbons Fig. 1 (a) The geometric structure of monolayer MX (M = Sn, Ge; X = S,
Se), in which the black arrows indicate armchair and zigzag directions of
The width is an important factor for the properties of nano- nanoribbons. (b) The band gaps of armchair SnS and SnSe nanoribbons as
ribbons. Therefore, to study the effect of the width, seven types a function of nanoribbon width.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 | 9297
PCCP Paper

Fig. 3 The I–V curves of 6A-SnSNR–SnSeNR (blue line) and 6Z-SnSNR–


SnSeNR (red line).

Fig. 2 (a) Two-probe device structure of SnSNR–SnSeNR. Left/right


electrodes are confined within the rectangles. The transmission spectra
of (b) 6A-SnSNR–SnSeNR and (c) 6Z-SnSNR–SnSeNR. The projected local (or close to zero), leading to the zero-transmission region in
densities of states (PLDOS) of (d) 6A-SnSNR–SnSeNR and (e) 6Z-SnSNR–
SnSeNR.
the transmission spectrum in Fig. 2b. For 6Z-SnSNR–SnSeNR
(Fig. 2e), the black area is located above the Fermi level and this
is consistent with the data shown in Fig. 2c.
the transport direction. The central region is finite and it To deeply understand the electronic transport properties,
consists of SnSNR and SnSeNR in the form of heterostructure the current–voltage (I–V) curves of the nanomaterials were
nanoribbons. Before exploring the properties of these devices, simulated. Fig. 3 shows the self-consistently calculated I–V
the heterostructure nanoribbons of 6A-SnSNR–SnSeNR and characteristic curves of 6A-SnSNR–SnSeNR and 6Z-SnSNR–
6Z-SnSNR–SnSeNR were optimized through the VASP code, as SnSeNR. The bias voltages are in the range of 0–1 V with an
well as the ATK code, which are both displayed in Fig. S9 (ESI†). interval of 0.1 V. The blue line in Fig. 3 shows the I–V
Clearly, the heterostructures calculated by the two software characteristic curve of 6A-SnSNR–SnSeNR. One can observe
programs are similar. Moreover, two nanoribbons are connected that the current is very low when the bias voltage is in the
tightly without obvious deformation and this indicates that these range of 0–0.8 V, which is only about 10 5–10 3 nA when the
structures are reasonably well designed. bias voltage is below 0.8 V. Surprisingly, the current increases
In general, the transmission spectrum is one of the most rapidly when the voltage increases to 1.0 V and reaches 58 nA: it
intuitive properties in the study of the quantum transport shows a four orders of magnitude increment in a narrow bias
behavior of a device. The transmission spectra of 6A-SnSNR– voltage range of 0.2 V. These interesting phenomena will be
SnSeNR and 6Z-SnSNR–SnSeNR were calculated and are illu- explained later in detail. The I–V curve of 6Z-SnSNR–SnSeNR is
strated in Fig. 2b and c, respectively. For 6Z-SnSNR–SnSeNR, a illustrated in Fig. 3 by a red line and it is different from the one
transmission peak appears near the Fermi level and this of 6A-SnSNR–SnSeNR. Obviously, upon the increase of the
indicates a high transmission coefficient. In contrast, the voltage bias from 0 to 1.0 V, the current decreased in the region
transmission spectrum of 6A-SnSNR–SnSeNR (Fig. 2b) exhibits of 0.3–0.6 V. This phenomenon is known as the NDR effect.49–53
a wide transmission gap near the Fermi level, which means that In the range of 0–0.3 V, the current initially increases and
the transmission coefficient for 6A-SnSNR–SnSeNR is zero reaches the maximum value (about 11155 nA, denoted as Ipeak)
(or close to zero) in this region. It should be noted that the at 0.3 V. The current then decreases slightly in the range of
potential barriers for electrons (Fe) and holes (Fh) are 0.85 eV 0.3–0.4 V. Evidently, the current decreases dramatically with a
and 0.48 eV, respectively. The energy difference between the bias voltage in the range 0.4–0.6 V from 10558 nA to 2173 nA.
Fermi level and the CBM (or VBM) for 6A-SnSNR is 0.82 eV It has been demonstrated that the current peak-to-valley ratio
(or 0.51 eV), while it is 0.44 eV (or 0.43 eV) for 6A-SnSeNR. (PVR) is an important characteristic to describe the NDR effect
The differences among 6A-SnSNR–SnSeNR, 6A-SnSNR, and and that it is defined as PVR = Ipeak/Ivalley. In most applications
6A-SnSeNR demonstrate that the heterostructures can effi- for NDR devices, a high current peak-to-valley ratio is expected.
ciently regulate the transport properties. In order to clearly For 6Z-SnSNR–SnSeNR, the corresponding PVR is 5.1, which
display the transport properties, the projected local densities of is higher than in the cases of the pristine zigzag SnS and
states (PLDOS), in which the local density of states of the device SnSe nanoribbons.30 In conclusion, the NDR effect of the
is projected along the z axis, of 6A-SnSNR–SnSeNR was calcu- 6Z-SnSNR–SnSeNR heterostructure has potential applications
lated. As shown in Fig. 2d, the black region of the PLDOS in logic gates, high-frequency oscillators, mixers, and electronic
indicates that the density of states in this region is zero amplifiers.

9298 | Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 This journal is © the Owner Societies 2019
Paper PCCP

Fig. 4 The bias-dependent transmission spectra of 6A-SnSNR–SnSeNR.


(a) 0.7 V, (b) 0.8 V, (c) 0.9 V and (d) 1.0 V. The blue dashed line indicates the
Fig. 5 The bias-dependent transmission spectra of 6Z-SnSNR–SnSeNR
bias window.
at the bias voltage of (a) 0.3 V, (b) 0.5 V, (c) 0.6 V, and (d) 1.0 V. The blue
dashed line indicates the bias window.

In order to explain the aforementioned phenomena in the


then disappears and the current increases again with a bias
I–V curves, the corresponding bias-dependent transmission
voltage up to 1.0 V.
spectra were calculated.54,55 The transmission spectra for
The transport properties of 7Z- and 8Z-SnSNR–SnSeNR, and
6A-SnSNR–SnSeNR with bias voltages of 0.7 V, 0.8 V, 0.9 V,
7A- and 8A-SnSNR–SnSeNR were also investigated here in
and 1.0 V are presented in Fig. 4. It was observed that the zero-
comparison to previous works on other nanoribbons,56,57
transmission crosses a broad range of energy near the Fermi
which exhibit the odd–even effect in the electronic transport
level with a bias voltage of 0.7 V and 0.8 V. In other words, there
of the nanoribbons (nanoribbons with odd chains exhibit
is a wide transmission gap around the Fermi level under the
different electronic transport properties compared to that of
applied voltage. It is difficult for the incident electrons to pass
even chains). The goal was to examine whether the SnSNR–
through and this results in a low current in the bias voltage
SnSeNR lateral heterostructure displays the odd–even effect. In
range of 0–0.8 V. In contrast, a transmission spectrum appears
Fig. 6, the I–V curves of 7Z- and 8Z-SnSNR–SnSeNR, and 7A- and
in the energy windows where the bias voltage reaches 1.0 V.
8A-SnSNR–SnSeNR are presented. One can notice that 7Z- and
A small transmission peak appears in the bias window and the
8Z-SnSNR–SnSeNR, and 7A- and 8A-SnSNR–SnSeNR exhibit I–V
corresponding current increases dramatically when the bias
curves similar to those of 6Z-SnSNR–SnSeNR and 6A-SnSNR–
voltage increases from 0.8 V to 1.0 V.
SnSeNR, respectively.
For 6Z-SnSNR–SnSeNR, the presence of the NDR effect is
The calculation results demonstrate that the ribbon widths
beneficial for its application in nanodevices. In order to explain
have little influence on the transport properties. All the zigzag
the NDR effect, the corresponding bias-dependent trans-
SnSNR–SnSeNR lateral heterostructures exhibit similar electronic
mission spectra are illustrated in Fig. 5. The transmission spectra
transport properties, including the NDR effect. Interestingly,
at the applied bias voltages of 0.3 V, 0.5 V, 0.6V, and 1.0 V were
for 7Z- and 8Z-SnSNR–SnSeNR, the NDR effect appears in two
studied. As is well known, the current through the nanoribbon is
regions with a maximum peak-to-valley ratio of 3.3 and 3.0 for
determined by the corresponding bias-dependent transmission
coefficient T(E). In the bias-dependent transmission spectrum,
the chemical potentials of the left and the right electrodes are
separated by the bias and distributed evenly around the Fermi
levels. Therefore, the current is determined by the integral of
the transmission spectrum within the bias window. One can
see that when the bias voltage increases from 0.3 V to 0.5 V, the
integral decreases dramatically and this indicates a decrease in
the current. This NDR effect is shown in Fig. 3. In addition,
dramatic decreases of the integrated areas are also observed in
the range of 0.5–0.6 V. With the increase of the bias voltage, the
transmission spectrum becomes narrower and smaller than Fig. 6 The I–V curves of (a) 6A-, 7A-, and 8A-SnSNR–SnSeNR and (b)
0.3 V when a bias voltage of 0.6 V is applied. The NDR effect 6Z-, 7Z-, and 8Z-SnSNR–SnSeNR.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 | 9299
PCCP Paper

7Z-SnSNR–SnSeNR and 8Z-SnSNR–SnSeNR, respectively. Similarly, Research & Practice Innovation Program of Jiangsu Province
6A-SnSNR–SnSeNR, 7A-SnSNR–SnSeNR, and 8A-SnSNR–SnSeNR (SJCX18_0727), and the project of Zhenjiang key laboratory of
exhibit similar transport properties. The current through the advanced sensing materials and devices (No. SS2018001).
nanoribbon is very low (only about 10 5–10 2 nA) under a low
bias. Then, the current increases linearly with the voltage and
reaches the maximum current. Specifically, the current reaches a References
maximum current of 497 nA for 7A-SnSNR–SnSeNR, which is
much higher than that of 6A-SnSNR–SnSeNR. The current of 1 A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6,
8A-SnSNR–SnSeNR decreases in the range of 0.9–1.0 V, which is 183–191.
mainly due to the decrease of the transmission spectrum. This is 2 B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and
consistent with the above conclusion that the decrease in the A. Kis, Nat. Nanotechnol., 2011, 6, 147–150.
current is induced by the decrease of the transmission coefficients 3 Z. Y. Yin, H. Li, H. Li, L. Jiang, Y. M. Shi, Y. H. Sun, G. Lu,
near the Fermi level. The bias-dependent transmission spectra of Q. Zhang, X. D. Chen and H. Zhang, ACS Nano, 2012, 6,
7A- and 8A-SnSNR–SnSeNR, and 7Z- and 8Z-SnSNR–SnSeNR were 74–80.
also calculated and are presented in Fig. S10–S13 (ESI†). They 4 F. Schwierz, Nat. Nanotechnol., 2010, 5, 487–496.
reasonably explain the I–V curves of 7Z- and 8Z-SnSNR–SnSeNR, 5 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,
and 7A- and 8A-SnSNR–SnSeNR. Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov,
Science, 2004, 306, 666–669.
6 L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H.
3 Conclusions Chen and Y. Zhang, Nat. Nanotechnol., 2014, 9, 372–377.
In summary, the electronic structures and the transport pro- 7 Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi,
perties of SnSNR–SnSeNR were investigated by using first- E. Kaxiras and P. Jarillo-Herrero, Nature, 2018, 556, 43–50.
principles calculations combined with nonequilibrium Green’s 8 O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic and
function theory. The results demonstrate that armchair SnSNRs A. Kis, Nat. Nanotechnol., 2013, 8, 497–501.
and SnSeNRs are semiconductors, independent of their ribbon 9 R. Krishna Kumar, D. A. Bandurin, F. M. D. Pellegrino,
width. Their band gaps fluctuate with the ribbon width and Y. Cao, A. Principi, H. Guo, G. H. Auton, M. Ben Shalom,
have an indirect-to-direct transition. In contrast, SnS and SnSe L. A. Ponomarenko, G. Falkovich, K. Watanabe, T. Taniguchi,
nanoribbons with zigzag edges are metals and have similar I. V. Grigorieva, L. S. Levitov, M. Polini and A. K. Geim,
band structures. The calculations on electronic transport reveal Nat. Phys., 2017, 13, 1182–1185.
that 6A-SnSNR–SnSeNR exhibits a wide transmission gap near 10 Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y.
the Fermi level and low current through the lateral hetero- Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi,
structure in the I–V curve. However, the I–V curve of 6Z-SnSNR– E. Kaxiras, R. C. Ashoori and P. Jarillo-Herrero, Nature, 2018,
SnSeNR exhibits an interesting negative differential resistive 556, 80–84.
effect with bias voltage in the range of 0.3–0.6 V. This is 11 L. C. Gomes and A. Carvalho, Phys. Rev. B: Condens. Matter
induced by the decrease of the bias-dependent transmission Mater. Phys., 2015, 92, 085406.
coefficient within the voltage window. In addition, the 12 A. K. Singh and R. G. Hennig, Appl. Phys. Lett., 2014,
transport properties of 7Z- and 8Z-SnSNR–SnSeNR, and 105, 042103.
7A- and 8A-SnSNR–SnSeNR were also investigated in order to 13 Z. Ma, B. Wang, L. Ou, Y. Zhang, X. Zhang and Z. Zhou,
study the odd–even effect in the electronic transport. It was Nanotechnology, 2016, 27, 415203.
demonstrated that 7A- and 8A-SnSNR–SnSeNR, and 7Z- and 14 S. D. Guo and Y. H. Wang, J. Appl. Phys., 2017, 121, 034302.
8Z-SnSNR–SnSeNR exhibit similar transport properties to those 15 P. Z. Hanakata, A. Carvalho, D. K. Campbell and H. S. Park,
of 6A-SnSNR–SnSeNR and 6Z-SnSNR–SnSeNR, respectively, Phys. Rev. B, 2016, 94, 035304.
independent of the nanoribbon width. The results demonstrate 16 Q. Wang, W. Yu, X. Fu, C. Qiao, C. Xia and Y. Jia, Phys.
that the SnSNR–SnSeNR lateral heterostructures have potential Chem. Chem. Phys., 2016, 18, 8158–8164.
applications in next-generation nanodevices and may provide 17 Z. Tian, C. Guo, M. Zhao, R. Li and J. Xue, ACS Nano, 2017,
more guidance in the design of nanodevices. 11, 2219–2226.
18 R. Li, H. Cao and J. Dong, Phys. Lett. A, 2017, 381,
3747–3753.
Conflicts of interest 19 M. Kar, B. Rajbanshi, S. Pal and P. Sarkar, J. Phys. Chem. C,
2018, 122, 5731–5741.
There are no conflicts to declare. 20 T. Hu and J. Dong, Phys. Chem. Chem. Phys., 2016, 18,
32514–32520.
Acknowledgements 21 R. Ibragimova, M. Ganchenkova, S. Karazhanov and
E. S. Marstein, Philos. Mag., 2017, 98, 710–726.
We acknowledge the support from the National Natural Science 22 Y. Hu, S. Zhang, S. Sun, M. Xie, B. Cai and H. Zeng, Appl.
Foundation of China (No. 51801075, 91750112), Postgraduate Phys. Lett., 2015, 107, 122107.

9300 | Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 This journal is © the Owner Societies 2019
Paper PCCP

23 Y. Huang, C. Ling, H. Liu and S. Wang, RSC Adv., 2014, 4, W. H. Chang, K. Suenaga and L. J. Li, Science, 2015, 349,
6933–6938. 524–528.
24 Y. Zhang, B. Shang, L. Li and J. Lei, RSC Adv., 2017, 7, 41 K. Cheng, Y. Guo, N. N. Han, Y. Su, J. F. Zhang and
30327–30333. J. J. Zhao, J. Mater. Chem. C, 2017, 5, 3788–3795.
25 A. K. Deb and V. Kumar, Phys. Status Solidi B, 2017, 42 C. J. Zhou, X. S. Wang, S. Raju, Z. Y. Lin, D. Villaroman,
254, 1600379. B. L. Huang, H. L. W. Chan, M. S. Chan and Y. Chai,
26 Z. Y. Li, M. Y. Liu, Y. Huang, Q. Y. Chen, C. Cao and Y. He, Nanoscale, 2015, 7, 8695–8700.
Phys. Chem. Chem. Phys., 2017, 20, 214–220. 43 C. J. Zhou, Y. D. Zhao, S. Raju, Y. Wang, Z. Y. Lin, M. S. Chan
27 H. I. Sirikumara and T. Jayasekera, J. Phys.: Condens. Matter, and Y. Chai, Adv. Funct. Mater., 2016, 26, 4223–4230.
2017, 29, 425501. 44 Y. D. Zhao, J. S. Qiao, P. Yu, Z. X. Hu, Z. Y. Lin, S. P. Lau,
28 L. Huang, F. Wu and J. Li, J. Chem. Phys., 2016, 144, 114708. Z. Liu, W. Ji and Y. Chai, Adv. Mater., 2016, 28, 2399–2407.
29 Y. Xu, H. Zhang, H. Shao, G. Ni, J. Li, H. Lu, R. Zhang, 45 Y. D. Zhao, J. S. Qiao, Z. H. Yu, P. Yu, K. Xu, S. P. Lau,
B. Peng, Y. Zhu, H. Zhu and C. M. Soukoulis, Phys. Rev. B, W. Zhou, Z. Liu, X. R. Wang, W. Ji and Y. Chai, Adv. Mater.,
2017, 96, 245421. 2017, 29, 1604230.
30 M. J. Zhang, Y. P. An, Y. Q. Sun, D. P. Wu, X. N. Chen, 46 Atomistix ToolKit version 2017.1, QuantumWise A/S.
T. X. Wang, G. L. Xu and K. Wang, Phys. Chem. Chem. Phys., 47 M. Brandbyge, J. L. Mozos, P. Ordejón, J. Taylor and
2017, 19, 17210–17215. K. Stokbro, Phys. Rev. B: Condens. Matter Mater. Phys.,
31 F. Li, X. Liu, Y. Wang and Y. Li, J. Mater. Chem. C, 2016, 4, 2002, 65, 165401.
2155–2159. 48 G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter
32 K. S. Novoselov, A. Mishchenko, A. Carvalho and A. H. Mater. Phys., 1996, 54, 11169.
Castro Neto, Science, 2016, 353, aac9439. 49 Y. P. An, M. J. Zhang, D. P. Wu, Z. M. Fu and K. Wang,
33 M. P. Levendorf, C. J. Kim, L. Brown, P. Y. Huang, R. W. J. Mater. Chem. C, 2016, 4, 10962–10966.
Havener, D. A. Muller and J. Park, Nature, 2012, 488, 627–632. 50 Y. P. An, M. J. Zhang, H. X. Da, Z. M. Fu, Z. Y. Jiao and
34 Z. Liu, L. Ma, G. Shi, W. Zhou, Y. Gong, S. Lei, X. Yang, Z. Y. Liu, J. Phys. D: Appl. Phys., 2016, 49, 245304.
J. Zhang, J. Yu, K. P. Hackenberg, A. Babakhani, J. C. Idrobo, 51 Y. P. An, J. T. Jiao, Y. S. Hou, H. Wang, R. D. Wu, C. Y. Liu,
R. Vajtai, J. Lou and P. M. Ajayan, Nat. Nanotechnol., 2013, 8, X. N. Chen, T. X. Wang and K. Wang, J. Phys.: Condens.
119–124. Matter, 2019, 31, 065301.
35 X. Duan, C. Wang, J. C. Shaw, R. Cheng, Y. Chen, H. Li, 52 Y. P. An, W. Ji and Z. Q. Yang, J. Phys. Chem. C, 2012, 116,
X. Wu, Y. Tang, Q. Zhang, A. Pan, J. Jiang, R. Yu, Y. Huang 5915–5919.
and X. Duan, Nat. Nanotechnol., 2014, 9, 1024–1030. 53 Y. P. An, Y. Q. Sun, M. J. Zhang, J. T. Jiao, D. P. Wu,
36 A. K. Geim and I. V. Grigorieva, Nature, 2013, 499, 419–425. T. X. Wang and K. Wang, IEEE Trans. Electron Devices,
37 Y. Gong, J. Lin, X. Wang, G. Shi, S. Lei, Z. Lin, X. Zou, G. Ye, 2018, 65, 4646–4651.
R. Vajtai, B. I. Yakobson, H. Terrones, M. Terrones, 54 Y. P. An, Y. Q. Sun, J. T. Jiao, M. J. Zhang, K. Wang,
B. K. Tay, J. Lou, S. T. Pantelides, Z. Liu, W. Zhou and X. N. Chen, D. P. Wu, T. X. Wang, Z. M. Fu and Z. Y. Jiao,
P. M. Ajayan, Nat. Mater., 2014, 13, 1135–1142. Org. Electron., 2017, 50, 43–47.
38 J. H. Yu, H. R. Lee, S. S. Hong, D. Kong, H. W. Lee, H. Wang, 55 Y. P. An, J. T. Jiao, Y. S. Hou, H. Wang, D. P. Wu, T. X. Wang,
F. Xiong, S. Wang and Y. Cui, Nano Lett., 2015, 15, 1031–1035. Z. M. Fu, G. L. Xu and R. Q. Wu, Phys. Chem. Chem. Phys.,
39 J. E. Padilha, A. Fazzio and A. J. da Silva, Phys. Rev. Lett., 2018, 20, 21552–21556.
2015, 114, 066803. 56 W. H. Duan, Z. Y. Li, H. Y. Qian, J. Wu and B. L. Gu, Phys.
40 M. Y. Li, Y. M. Shi, C. C. Cheng, L. S. Lu, Y. C. Lin, Rev. Lett., 2008, 100, 206802.
H. L. Tang, M. L. Tsai, C. W. Chu, K. H. Wei, J. H. He, 57 Y. P. An and Z. Q. Yang, Appl. Phys. Lett., 2011, 99, 192102.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 9296--9301 | 9301

View publication stats

You might also like