Band Gap Engineering and Layer-by-Layer Mapping of Selenium-Doped Molybdenum Disul Fide

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Letter

pubs.acs.org/NanoLett

Band Gap Engineering and Layer-by-Layer Mapping of Selenium-


Doped Molybdenum Disulfide
Yongji Gong,†,¶ Zheng Liu,‡,§,¶ Andrew R. Lupini,∥,¶ Gang Shi,‡ Junhao Lin,∥,⊥ Sina Najmaei,‡
Zhong Lin,# Ana Laura Elías,# Ayse Berkdemir,# Ge You,‡ Humberto Terrones,# Mauricio Terrones,#,○,∇
Robert Vajtai,‡ Sokrates T. Pantelides,∥,⊥ Stephen J. Pennycook,∥ Jun Lou,‡ Wu Zhou,*,∥
and Pulickel M. Ajayan*,†,‡

Department of Chemistry, Rice University, Houston, Texas 77005, United States

Department of Mechanical Engineering & Materials Science, Rice University, Houston, Texas 77005, United States
§
School of Materials Science and Engineering, School of Electrical and Electronic Engineering Nanyang Technological University,
639798, Singapore

Materials Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States

Department of Physics and Astronomy, Vanderbilt University, Nashville, Tennessee 37235, United States
#
Department of Physics and Center for 2-Dimensional and Layered Materials, The Pennsylvania State University, University Park,
Pennsylvania 16802, United States

Department of Chemistry, Department of Materials Science and Engineering & Materials Research Institute, The Pennsylvania State
University, University Park, Pennsylvania 16802, United States

Research Center for Exotic Nanocarbons (JST), Shinshu University, Wakasato 4-17-1, Nagano 380-8553, Japan
*
S Supporting Information

ABSTRACT: Ternary two-dimensional dichalcogenide alloys


exhibit compositionally modulated electronic structure, and
hence, control of dopant concentration within each individual
layer of these compounds provides a powerful tool to
efficiently modify their physical and chemical properties. The
main challenge arises when quantifying and locating the
dopant atoms within each layer in order to better understand
and fine-tune the desired properties. Here we report the
synthesis of molybdenum disulfide substitutionally doped with
a broad range of selenium concentrations, resulting in over
10% optical band gap modulations in atomic layers. Chemical
analysis using Z-contrast imaging provides direct maps of the
dopant atom distribution in individual MoS2 layers and hence a measure of the local optical band gaps. Furthermore, in a bilayer
structure, the dopant distribution is imaged layer-by-layer. This work demonstrates that each layer in the bilayer system contains
similar local Se concentrations, randomly distributed, providing new insights into the growth mechanism and alloying behavior in
two-dimensional dichalcogenide atomic layers. The results show that growth of uniform, ternary, two-dimensional dichalcogenide
alloy films with tunable electronic properties is feasible.
KEYWORDS: Monolayer molybdenum disulfide (MoS2), ternary alloy, Se doping, dopant distribution, ADF imaging,
band gap engineering

T wo-dimensional van der Waals (2D-vdW) materials have


attracted considerable attention in recent years due to
their fascinating properties,1−3 which has led to the develop-
the atomic scale within layers and understanding their influence
on the material properties still remain as key challenges.
The isomorphism of the transition metal dichalcogenide
ment of novel synthetic approaches for large scale preparation (TMD) families (MX2: M = Mo, W; X = S, Se, Te) makes
of layered materials.4−6 However, in order to make these them great candidates to form ternary vdW alloys that will not
materials more versatile, further modification of their physical suffer from phase separation. Molybdenum disulfide (MoS2) is
and chemical properties is required. A possible way to fine-tune
these properties is via chemical doping. However, only a few Received: August 28, 2013
ternary vdW alloys have so far been realized by chemical Revised: December 11, 2013
doping.7,8 Efficiently quantifying the distribution of dopants at

© XXXX American Chemical Society A dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 1. Synthesis and morphologies of MoS2(1−x)Se2x atomic layers. (a) Controlled synthesis of MoS2(1‑x)Se2x layers via CVD with S/Se powder
positioned in the low temperature zone as the source of chalcogens and MoO3 located at the center of the tube as the Mo source. (b−d) Typical
optical images of monolayer and single-crystal domains, merged domains, and continuous MoS2(1−x)Se2x (x = 0.1) films, respectively. Additional
layers are found at triangle centers (b) and at the interface between merged triangles (d). (e) AFM height topography of a MoS2(1−x)Se2x (x = 0.1)
monolayer single-crystal triangle. (f,g) Height and phase topographies of the edges in triangular domains. Inset in panel f is the height profile
showing the thickness of ∼0.7 nm, measured along the white line.

a representative member of the TMD family, consisting of two the single-atom level and hence have been able to assign the
layers of sulfur atoms and a layer of Mo sandwiched in between. local optical band gap, in this 2D-vdW alloyed system. Our
Recently, this material has been revisited and highlighted layer-by-layer quantitative dopant analysis reveals that the two
because the unique band structure of monolayered MoS2, with Se-doped MoS2 layers at AB stacking bilayer regions probably
its breakdown of inversion symmetry, provides excellent grow simultaneously under our synthesis conditions, as
opportunities for applications in optical and electronic opposed to the sequential growth observed for bilayer
devices.9−12 Synthesis of large-area and high-quality MoS2 graphene.17
atomic layers has recently been reported.13−15 Theoretical In order to synthesize atomic layered MoS2(1−x)Se2x, sulfur
investigations suggest that the band structure of MoS2 could be and selenium fine powder are mixed as the chalcogen sources
tuned by substituting sulfur atoms with other chalcogen (S and Se), while molybdenum oxide powder is used as the Mo
elements to form TMD alloys.16 This enables a tunable source. Monolayer and bilayer MoS2(1−x)Se2x are directly grown
mechanism for varying the physical and chemical properties of on SiO2 via CVD at ∼800 °C (Figure 1a). The as-prepared
MoS2. samples with a variety of morphologies are depicted in Figure
Here, we report a one-step direct synthesis of MoS2(1−x)Se2x 1b−d. By prolonging the growing time, nucleation of a high
atomic mono- and bilayers with tunable compositions on SiO2 density of domains results in merged triangles and continuous
by CVD. Both isolated triangular single-crystal islands as well as films (Figure 1c,d). The MoS2(1−x)Se2x nucleation sites are
continuous films could be grown by controlling the growth noticeable in some triangles, and some bilayer regions are also
parameters. We demonstrate that, through controlled selenium found (dark purple in Figure 1d), similar to what has been
doping, the band structure of MoS2 could be modified and the observed in the growth of pristine MoS2.14,15 The size of the as-
optical band gap could be continuously tuned by over 200 meV. prepared single-crystal domains and continuous films is of the
Using atomic resolution Z-contrast imaging, we demonstrate order of a few tens (Figure 1b,c) to hundreds of micrometers
direct atomic identification of Se dopants within the MoS2 (Figure 1d), respectively. Atomic force microscopy (AFM)
lattice with almost 100% detection efficiency. Furthermore, by height topography confirms that most of the MoS2(1−x)Se2x
combining a novel site separation method and quantitative films are single-layered (Figure 1e,g).
image intensity analysis, we have achieved layer-by-layer By changing the ratio of the S and Se powders during
visualization and quantification of the dopant distribution at synthesis, the concentration of Se in the MoS2 lattice can be
B dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 2. Band gap engineering of MoS2(1−x)Se2x atomic layers. (a−c) XPS spectra of S 2p, Se 3d, and Mo 3d core levels in the monolayer
MoS2(1−x)Se2x with different Se concentration (x = 0.1, 0.3, 0.50, and 0.75). (d) PL spectra of pristine MoS2 (blue), MoS1.4Se0.6 (green), MoS1Se1
(purple), MoS0.5Se1.5 (orange), and MoSe2 (red), respectively, measured with a 488 nm laser. Inset: red dots, the Se concentration dependency of
the PL optical gap of MoS2(1−x)Se2x showing a linear decrease from ∼1.85 eV (pure MoS2) to ∼1.54 eV (pure MoSe2). We multiplied the PL
intensity of MoSe2 by 0.1 so that all the spectra can be seen clearly. Green dots, calculated LDA band gap shows the same trend as the experimental
optical gap (the similarity of values is coincidental as the experimental gap is excitonic; see text). (e−g) Optical image of MoS2(1−x)Se2x triangles and
the corresponding PL intensity (f) and peak position (g) mappings.

well controlled from 0 to ∼75%, as confirmed by X-ray the Bethe−Salpeter equation (BSE). In Figure 2d, we also show
photoelectron spectroscopy (XPS) analysis. Figure 2a−c plots the LDA bandgap dependency on the Se concentrations (Se%),
the S 2p, Se 1s, and Mo 3d spectra for MoS2(1−x)Se2x samples revealing the same trend as the optical (excitonic) gap. It is a
with 10%, 30%, 50%, and 75% Se doping. With increasing Se mere coincidence, however, that the LDA band gap has roughly
concentration, the intensity of S 2p peaks (∼162.3 and 161.2 the same value as the excitonic gap. The similarity of the trends
eV) decrease, while peaks from Se 3p3/2 (∼159.8 eV) and Se 3d reveals that the GW + BSE correction is roughly independent
(∼54 eV, Figure 2b) appear and become dominant. The peaks of Se%. Figure 2d further shows that, for high concentrations of
of Mo 3d3/2 and 3d5/2 core levels are located at ∼231.5 and Se in MoS2, the PL intensity decreases, possibly due to the
228.3 eV, respectively, in agreement with values from MoS2 and increased local electronic states within the samples, as predicted
MoSe2 systems.18,19 by density functional theory (DFT) calculations,16,23 as
We next use photoluminescence (PL) spectroscopy20 to evidenced by the minimum intensity in 50% Se-doped MoS2
examine the modification of the optical (excitonic) gap in MoS2 (PL spectrum in purple in Figure 2d), where the local
as a function of Se doping. Figure 2d shows PL spectra of electronic states are expected to be maximum. An optical image
monolayer pristine MoS 2 (blue), 30% Se-doped MoS 2 and the corresponding PL intensity and peak position mapping
(MoS1.4Se0.6, green), 50% Se-doped MoS2 (MoS1Se1, purple), of a typical MoS2(1−x)Se2x triangle are shown in Figure 2e−g,
75% Se-doped MoS2 (MoS0.5Se1.5, orange), and MoSe2 (red). respectively. It can be clearly observed from Figure 2g that, at a
Only one dominant peak is observed in the spectra from Se- given Se concentration, the optical gap of MoS2(1−x)Se2x varies
doped samples, suggesting that there is no separation of doped- slightly (typically ∼30 meV) at the micrometer scale within the
and undoped-MoS2 domains in the sample within the size of triangular crystal, thus suggesting the presence of some
the optical probe (∼ 500 nm). As the Se concentration compositional fluctuations within the MoS2(1−x)Se2x sample
increases, a noticeable red-shift in the peak position from 670 (some other effects may also contribute to the PL peak shift at
to 805 nm is observed, corresponding to an optical gap change the edge of the triangles). The MoS2(1−x)Se2x alloy is also
from 1.85 to 1.54 eV. This correlation is plotted as an inset in characterized by Raman spectroscopy. The vibration modes
Figure 2d, where the Se concentration dependence of the from both MoS2 and MoSe2 can be identified (Supplementary
optical gap (red dots) can be linearly extrapolated to pure Figure S1). In addition, our measurements suggest that
MoSe2 (100% Se) with an optical gap of ∼1.60 eV. The optical MoS2(1−x)Se2x based field-effect transistors (FETs) have a
gap values at the MoS2 and MoSe2 cases, around 1.9 and 1.5 comparable mobility (3.8 to 15.3 cm2/(V s)) and ON/OFF
eV, respectively, are in excellent agreement with published ratio (106) to mechanically exfoliated MoS2 FET devices
theoretical values.21,22 These calculations used the GW scheme (Supplementary Figure S2).24
to correct the band gap obtained in the local-density In order to quantify the local Se concentration at the nm-
approximation (LDA) and then included excitonic effects via scale and map out the atomic distribution of Se dopants, we
C dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 3. Atom-by-atom dopant analysis in monolayer MoS2. (a) ADF image of pristine MoS2. (b) ADF image of Se-doped MoS2 with ∼12% local
Se concentration. (c) Higher magnification ADF image showing Se dopants. The green and cyan circles highlight single- and double-Se substituted
S2 sites, respectively. (d) Comparison of experimental and simulated image intensity line profiles from single- and double-Se substitution at S2 sites.
The experimental profiles were obtained along the green and cyan dashed lines in panel c. (e) Site-separated image intensity histogram analysis of
pristine and Se-doped MoS2 monolayers. The dashed lines are Gaussian fits to the intensity peaks. The analysis for Se-doped MoS2 was based on the
image shown in panel b; while the analysis for pristine MoS2 was obtained from an ADF image with larger field of view (Figure S3 and Table S1,
Supporting Information). (f) Structure model obtained from histogram analysis showing the distribution of single- and double-Se substituted S2 sites.
The local Se concentration is ∼12%, corresponding to an optical band gap of ∼1.79 eV. Red, Mo sites; dark green, S2 sites; bright green, Se + S sites;
white, Se2 sites.

performed atomic annular dark field (ADF) imaging, also in Figure 3e) shows only one intensity peak for the Mo sites
known as Z-contrast imaging, on an aberration-corrected and one main peak for the S2 sites with a small tail at half-
scanning transmission electron microscope (STEM),25 operat- intensity arising from S2 sites with monosulfur vacancies.28 The
ing at 60 kV (see Supporting Information for experimental histogram from Se-doped MoS2 (right panel in Figure 3e),
details). ADF imaging has been demonstrated to be an efficient however, shows an additional peak for the S2 sites with single
way for atom-by-atom chemical analysis in monolayer Se substitution, at about the same brightness as the Mo
materials7,10,26−28 since the image intensity is directly related intensity peak, with a few double Se substitution sites displaying
to the atomic number of the imaged species.26,29 Figure 3a,b higher intensity than Mo. No change is observed for the image
compares ADF images from monolayered pristine MoS2 and intensity distribution at the Mo sites, suggesting that Se
Se-doped MoS2 with about 12% local Se concentration (with substitution only happens at S sites.
respect to the total amount of chalcogen atoms). The ADF Since the image intensities for S2, Se + S, and Se2 sites are
image from pristine MoS2 (Figure 3a) shows an alternating well separated in the histogram after site-separation, the
atomic arrangement of Mo and S2 sites in the hexagonal lattice, distribution of these three types of sites can be easily mapped
with Mo atoms being brighter than the S2 sites. In contrast, out by selecting different ranges of image intensity in the
with Se dopants, some S2 sites become much brighter and histogram (see Methods). This provides a quick and reliable
display intensities close to or even higher than the Mo sites way to perform Se dopant mapping at a relatively large scale
(Figures 3b,c), while the intensities of the Mo sites remain (64−1000 nm2 field of view per image) at the single-atom level
unchanged. Quantitative image intensity analysis and image in a largely automatic fashion. The output of such an analysis
simulations (Figure 3d) demonstrate that the higher image based on the ADF image shown in Figure 3b is provided in
intensity at the S2 sites arises from Se substitution. S2 sites with Figure 3f, where the single Se substitution and double
single Se substitution (i.e., Se + S, highlighted in green in substitution sites are presented in bright green and white,
Figure 3c) have image intensity slightly lower than that of a respectively. The local Se concentration and thus the local
single Mo atom; while double substitution sites (i.e., Se2, optical band gap based on the linear dependency shown in
highlighted in cyan in Figure 3c) are brighter than Mo. Figure 2d, at the nm scale, could then be calculated directly
Statistics of the image intensity from Mo and S2 sites are from the number of substitution sites.
provided in Figure 3e. In order to examine the doping at the In order to understand the growth mechanism and
different sites, here the Mo and S2 sites are analyzed separately properties, it is of particular importance to obtain the Se
(see Methods). The histogram from pristine MoS2 (left panel distributions in each MoS2 layer in the bilayer regions in our
D dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Figure 4. Layer-by-layer dopant analysis in AB stacked bilayer MoS2 at the single atom level. (a) Filtered ADF image of Se-doped bilayer MoS2 with
AB stacking. (b) Schematic of the three types of sites in AB stacked bilayer MoS2. (c) A zoom-in view of the AB stacked Se-doped bilayer MoS2
exhibiting a wide range of image contrast levels. (d) Site-separated image intensity histogram analysis of the image shown in panel a. The dashed
lines are Gaussian fits to the intensity peaks. (e,f) Structure model obtained from the histogram analysis showing the distribution of single- and
double-Se substituted S2 sites in the A layer with 32% local in-layer Se concentration and in the B layer with 28% local Se concentration, respectively.
Red, Mo sites; dark green, S2 sites; bright green, Se + S sites; white, Se2 sites. (g) Interlayer and inter-regional compositional variations obtained from
site-separated ADF image analysis. The seven analyses, each containing between four and eleven thousand atoms, were obtained from random
regions in the same sample. The interlayer compositional variation has a mean value of ∼2%, which is much smaller than the inter-regional variations.
(h,i) The alloying degree versus Se concentration. Mo, S, and Se atoms are shown in purple, yellow, and green, respectively, in panels b and h.

CVD-grown sample. For example, a material in which dichalcogen sites in each layer. By selectively considering each
subsequent atomic layers have different dopant concentrations intensity peak for site 1 and site 3, the distribution of these
might have very different properties from an alloy of equivalent three types of dichalcogen sites in the two MoS2 layers can be
layers. Atom-by-atom chemical analysis via quantitative ADF mapped out individually. The outputs of this layer-by-layer site-
imaging, however, is very challenging for samples thicker than a separated dopant analysis are shown in Figure 4e,f, where the
monolayer, due to the overlap of atoms in different layers.26,27 atomic positions of site 3 represent Mo sites in the A layer, and
An example of a Se-doped MoS2 bilayer is shown in Figure 4a, the atomic positions of site 2 are Mo sites in the B layer. The
where diverse levels of image contrast (see Supplementary in-layer local Se concentration can then be quantified from
Figures S4 and S5) can be observed at the atomic sites, thus these two outputs. Similar site-separated image analysis method
making the image difficult to interpret visually. We notice that can also be applied to bilayers with 2H stacking as shown in
the bilayer regions in our CVD-grown MoS2(1−x)Se2x are Supplementary Figure S6, although the 2H stacking is rarely
predominantly AB stacking (Figure 4b ,c), which can be observed in our sample.
separated into three distinct types of atomic sites: S2 sites in the Figure 4g summarizes the local Se concentration and the
A layer, Mo sites in the B layer, and Mo in the A layer plus S2 corresponding optical band gap in the A and B layers in seven
sites in the B layer (labeled as sites 1, 2, and 3 in Figure 4b), random bilayer regions (with no direct spatial correlation with
where we use “A” and “B” as relative notations. With this site the regions examined by optical methods as presented in Figure
separation, the S2 sites in the A and B layers can be 2), usually with an area of ∼10 × 10 nm2 and ∼50−200 nm
distinguished and analyzed individually. The site-separated apart, in the same film. It is important to note that, at random
histogram obtained from Figure 4a is shown in Figure 4d. bilayer regions, the difference of the local Se concentration
Three intensity peaks were observed for both site 1 and site 3, between layers is relatively small (green and blue columns in
corresponding to S2, Se + S, and Se2, respectively, at the Figure 4g), with a mean variation of ∼2%. In contrast, the local
E dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

Se concentration can change as much as 19% between different shown in Figure 3f and Figure 4e−f, after counting the numbers
regions within the same sample. This compositional variation is of S and Se occupying the six nearest neighbor sites for all
in general consistent with observations from PL mapping dichalcogen sites in each MoS2 layer (Figure 4h). Following a
(Figure 2g) and Raman mapping (Supplementary Figure S1); modified method based on ref 7, the degree of alloying (JA_B) is
however, the spatial resolution obtained from direct ADF quantified as
imaging is on the 10 nm level, much higher than with the PL
mapping. The fluctuating flow rate of Se vapor evaporated from JA_B = Pobserved /Prandom × 100%
the Se powder may be responsible for the variation of the local
Se concentration. Furthermore, our calculations show that Se where Pobserved is the averaged percentage of B-type nearest
diffusion between MoS2 layers is orders of magnitude lower neighbors for the A-type sites, and Prandom is the total
than diffusion within the same layer (Supplementary Figure percentage of B-type sites in the examined layer (see
S7). Therefore, the small interlayer compositional variation, as Supporting Information). The results are summarized in Figure
compared to the large inter-regional variations, suggests that 4i for local Se concentrations ranging from 10% to 33%. All the
the two Se-doped MoS2 layers at the bilayer regions most likely J values are close to 100%, i.e., within statistical error the
grow simultaneously, instead of sequentially, i.e., under similar distributions of Se-doped sites are totally random, and there is
transient growth conditions, indicating the possibilities of layer- no preference for Se dopants to cluster in the MoS2 layer.
controlled growth of ternary two-dimensional alloys. In summary, we have demonstrated a novel approach for
Using this site separation and quantitative image intensity preparing Se-doped MoS2 atomic layers with controllable
analysis method, layer-by-layer quantification of dopant composition using chemical vapor deposition. We show that
distribution can be achieved in bilayer 2D materials with single the optical band gap of the material can be fine-tuned between
atom sensitivity and sub-nm depth resolution based on the 1.85 and 1.60 eV by changing the Se concentration. The
known symmetry of the crystal structure. This is a significant excellent transport properties and control over the band
improvement as compared with methods such as depth- structure of the material pave the way for future investigation of
sectioning30,31 or tilt-tomography,32 which would be difficult to physical properties and applications of MoS2(1−x)Se2x ternary
apply to large-size 2D materials. The accuracy of this analytical systems. A novel and robust ADF-imaging-based histogram
method is estimated to be about 1−2% (see Supporting analysis method has been developed for single layer and bilayer
Information for details), which compares favorably to conven- materials to map the distribution of dopants with single-atom
tional chemical analysis methods such as X-ray energy resolution and single layer discrimination. Our quantification
dispersive spectroscopy (XEDS), and electron energy-loss shows that the substitutional Se doping of MoS2 is random and
spectroscopy (EELS), at this scale.33 Moreover, by analyzing homogeneous at the nanoscale. This method could also be
the site-separated image histogram, several thousand atoms can applied to other two-dimensional systems.
be identified in a single analysis from a single ADF image, Methods. Synthesis of Se-Doped MoS2. In order to
which makes it a very dose-efficient way of performing atom- synthesize MoS2(1−x)Se2x, we adapted a technique developed
by-atom chemical analysis compared to those techniques. Thus, for the synthesis of MoS2 and WS2 and used a mixture of
it is particularly suitable for high-spatial-resolution chemical sublimated sulfur and selenium powders.6,14,15 We positioned
analysis of electron beam sensitive 2D materials. the chalcogen precursors upstream in the low temperature zone
In addition, our method provides new insights into the of the furnace. Molybdenum oxide (MoO3) powder was used
alloying behavior at the single atom level. Table 1 summarizes as a precursor for molybdenum and was positioned on a SiO2/
Si wafer close to the designated growth substrate (also SiO2/Si)
Table 1. Comparison of the Percentages of S2 Sites with and in the center of the reactor, i.e., the hot zone (Figure 1a).
Single- and Double-Se Substitution Obtained from Using a slow heating ramp of 40 °C/min, the system was
Experimental Data and Calculated Based on a Binomial brought to the reaction temperature of ca. 800 °C. The S and
Distribution at the Same Local Se Concentration Se sources start evaporating and reacting with MoO3. These
Se
experiments were performed in an inert quartz environment
image S2 sites Se + S Se2 sites concentration facilitated by a steady 100 sccm flow of Ar. The ratio between
index (%) sites (%) (%) (%) Se and S was adjusted by changing the weight of Se and S.
Figure 3f experiment 77.5 20.7 1.8 12.2 Bandgap and Diffusion Barrier Calculations. Our calcu-
binomial 77.1 21.4 1.5 lations were performed using the plane wave code CASTEP,34
Figure 4e experiment 45.8 44.0 10.2 32.2 as implemented in the Materials Studio, on 3 × 3 hexagonal
binomial 46.0 43.7 10.4 cells under the Local Density Approximation (LDA),
Figure 4f experiment 51.8 40.6 7.6 27.9 considering the Ceperly−Alder−Perdew−Zunger (CA-
binomial 52.0 40.2 7.8 PZ)35,36 functional with 6 × 6 × 3 Monkhorst−Pack k-points
and a plane waves cut off of 600 eV. All the structures were
the number of S2 sites with single- and double-Se substitution relaxed, including the cells, until the forces became smaller than
obtained from the histogram analysis, compared with a 0.01 eV/Å and the energy tolerances were less than 5 × 10−6
binomial distribution. The numbers of single- and double-Se eV/atom. It is well-known that DFT-LDA usually under-
substitution sites observed in the sample are close to what estimates the true band gap; however, in semiconducting
could be expected from a binomial distribution, suggesting that transition metal dichalcogenides, the electronic gap obtained by
there is no strong preference for forming any particular Se and DFT-LDA is very close to the optical band gap obtained
S configuration in nearest neighbor chalcogen sites. This experimentally and thus can be used as an approximation.37
observation is consistent with theoretical calculations shown in The intralayer and interlayer diffusion barrier of the Se atoms in
Supplementary Figure S8 and previous results.16 The degree of the MoS2 lattice were calculated using the nudged elastic band
alloying could also be determined directly from the outputs (NEB) method38 implemented in the VASP package39 with
F dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX
Nano Letters Letter

dispersion-corrected (van der Waals) functional, on 12 × 12 Computing and Cyberinfrastructure unit of Information
hexagonal cells and 6 × 6 hexagonal bilayer cells. Technology Services, and Penn-State Center for Nanoscale
Analysis of Dopant Distribution from STEM-ADF Images. Science. This research was supported in part by a Wigner
In order to determine the positions and types of the atoms, we Fellowship through the Laboratory Directed Research and
used a method based on an initial indication of the number of Development Program of Oak Ridge National Laboratory
unit cells, their approximate positions and atom locations (ORNL), managed by UT-Battelle, LLC, for the U.S. DOE (to
within a unit cell, which are then refined in an iterative manner. W.Z.); by the U.S. Department of Energy (DOE), Basic Energy
For each iteration the center (xj,yj) of the intensity (I) for the Sciences, Materials Sciences and Engineering Division (to
jth atomic column was calculated as (xj,yj) = (ΣiIijxij,ΣiIijyij)/ΣiIij A.R.L., S.T.P., and S.J.P.), U.S. DOE grant DE-FG02-
where the denominator is the summed intensity, and the sums 09ER46554 (to J.L. and S.T.P.), and through a user project
over i include all pixels within a user-defined radius of the supported by ORNL’s Center for Nanophase Materials
center from the previous step. This process gives a subpixel Sciences (CNMS), which is sponsored by the Scientific User
estimate of the column position and, since the intensity is Facilities Division, Office of Basic Energy Sciences, U.S. DOE.
averaged over several pixels, it is less sensitive to noise than M.T. thanks JST-Japan for funding the Research Center for
taking a single pixel maximum. This procedure was typically Exotic NanoCarbons, under the Japanese regional Innovation
iterated for about 20 steps and the convergence monitored. Strategy Program by the Excellence. M.T. acknowledges
After the process, atomic columns that were too far from the support from the Penn State Center for Nanoscale Science
initial estimate were treated as potentially unreliable. The for seed grant on 2-D Layered Materials (DMR-0820404). The
histograms were plotted, and the intensity values at which the authors also acknowledge the Center for 2-Dimensional and
columns are assigned as S2, S + Se, or Se2 for site 1 (or Mo plus Layered Materials. The authors are grateful to Lázaro Calderiń
S2, S + Se, or Se2 for site 3) can be specified. Although the for technical assistance and to Drs. Kazu Suenaga and Ryo
peaks in the histogram are distinctly separated, there could be Ishikawa for helpful discussions.
some overlap between them, and the area of this overlap can be
used to give an indication of the error in the assignment of the
dopants (see Supporting Information for details).
■ REFERENCES
(1) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L. J.; Loh, K. P.; Zhang,


*
ASSOCIATED CONTENT
S Supporting Information
H. The chemistry of two-dimensional layered transition metal
dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263−275.
(2) Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-like two-
dimensional materials. Chem. Rev. 2013, 113, 3766−98.
Additional experimental details and figures. This material is (3) Huang, X.; Zeng, Z. Y.; Zhang, H. Metal dichalcogenide
available free of charge via the Internet at http://pubs.acs.org.


nanosheets: preparation, properties and applications. Chem. Soc. Rev.
2013, 42, 1934−1946.
AUTHOR INFORMATION (4) Li, X.; et al. Large-area synthesis of high-quality and uniform
Corresponding Authors graphene films on copper foils. Science 2009, 324, 1312−1314.
*(W.Z.) E-mail: wu.zhou.stem@gmail.com. (5) Song, L.; et al. Large scale growth and characterization of atomic
*(P.M.A.) E-mail: ajayan@rice.edu. hexagonal boron nitride layers. Nano Lett. 2010, 10, 3209−3215.
(6) Elias, A. L.; et al. Controlled synthesis and transfer of large-area
Author Contributions WS2 sheets: from single layer to few layers. ACS Nano 2013, 7, 5235−

Y.G., Z.L., and A.R.L. contributed equally to this work. Y.G., 42.
Z.L., and W.Z. designed the project. Y.G. synthesized the (7) Dumcenco, D. O.; Kobayashi, H.; Liu, Z.; Huang, Y.-S.; Suenaga,
MoS2(1−x)Se2x alloy and performed the Raman and XPS K. Visualization and quantification of transition metal atomic mixing in
characterization. W.Z. performed the STEM experiments and Mo1−xWxS2 single layers. Nat. Commun. 2013, 4, 1351.
analyzed the results with ARL. G.S. made MoS2(1−x)Se2x FET (8) Ci, L.; et al. Atomic layers of hybridized boron nitride and
devices and carried out the electric measaurement. Z.L. graphene domains. Nat. Mater. 2010, 9, 430−435.
(9) Li, H.; et al. Fabrication of single- and multilayer MoS2 film-based
performed the PL characitziation of the material. J.L. calculated field-effect transistors for sensing NO at room temperature. Small
the Se dopant formation energy and diffusion barrier under 2012, 8, 63−67.
supervision of S.T.P. A.B. calculated the bandgap of (10) Yin, Z.; et al. Single-layer MoS2 phototransistors. ACS Nano
MoS2(1−x)Se2x. G.Y. performed AFM characterization. Y.G., 2012, 6, 74−80.
Z.L., A.R.L.,and W.Z. wrote the paper with advice from J.L., (11) Mak, K. F.; He, K.; Shan, J.; Heinz, T. F. Control of valley
S.J.P., and P.M.A. All authors participated in discussions. polarization in monolayer MoS2 by optical helicity. Nat. Nanotechnol.
Notes 2012, 7, 494−498.
(12) Eda, G.; Fujita, T.; Yamaguchi, H.; Voiry, D.; Chen, M.;
The authors declare no competing financial interest.


Chhowalla, M. Coherent atomic and electronic heterostructures of
single-layer MoS2. ACS Nano 2012, 6, 7311−7317.
ACKNOWLEDGMENTS (13) Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P. M.; Lou, J. Large-area
This work was supported by the Army Research Office MURI vapor-phase growth and characterization of MoS2 atomic layers on a
grant W911NF-11-1-0362, the FAME Center, one of six SiO2 substrate. Small 2012, 8, 966−971.
centers of STARnet, a Semiconductor Research Corporation (14) Najmaei, S.; et al. Vapour phase growth and grain boundary
program sponsored by MARCO and DARPA, the U.S. Office structure of molybdenum disulphide atomic layers. Nat. Mater. 2013,
12, 754−759.
of Naval Research MURI grant N000014-09-1-1066, Welch
(15) van der Zande, A. M.; et al. Grains and grain boundaries in
Foundation grant C-1716, the NSF grant DMR-0928297, the highly crystalline monolayer molybdenum disulphide. Nat. Mater.
National Research Foundation Singapore under NRF RF 2013, 12, 554−561.
Award No. NRF-RF2013-08, the Nanoelectronics Research (16) Komsa, H. P.; Krasheninnikov, A. V. Two-dimensional
Corporation contract S201006, the Materials Simulation transition metal dichalcogenide alloys: stability and electronic
Center of the Materials Research Institute, the Research properties. J. Phys. Chem. Lett. 2012, 3, 3652−3656.

G dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX


Nano Letters Letter

(17) Fang, W.; et al. Rapid identification of stacking orientation in


isotopically labeled chemical-vapor grown bilayer graphene by Raman
spectroscopy. Nano Lett. 2013, 13, 1541−1548.
(18) Bougouma, M.; Batan, A.; Guel, B.; Segato, T.; Legma, J. B.;
Reniers, F.; Delplancke-Ogletree, M.-P.; Buess-Herman, C.; Doneux,
T. Growth and characterization of large, high quality MoSe2 single
crystals. J. Cryst. Growth 2013, 363, 122−127.
(19) Abdallah, W. E.; Nelson, A. E. Characterization of MoSe2(0001)
and ion-sputtered MoSe2 by XPS. J. Mater. Sci. 2005, 40, 2679−2681.
(20) Sundaram, R.; Engel, M.; Lombardo, A.; Krupke, R.; Ferrari, A.
C.; Avouris, P.; Steiner, M. Electroluminescence in single layer MoS2.
Nano Lett. 2013, 13, 1416−1421.
(21) Conley, H. J.; Wang, B.; Ziegler, J. I.; Haglund, R. F.; Pantelides,
S. T.; Bolotin, K. I. Bandgap engineering of strained monolayer and
bilayer MoS2. Nano Lett. 2013, 13, 3626−3630.
(22) Ramasubramaniam, A. Large excitonic effects in monolayers of
molybdenum and tungsten dichalcogenides. Phys. Rev. B 2012, 86,
115409.
(23) Chen, Y.; Xi, J.; Dumcenco, D. O.; Liu, Z.; Suenaga, K.; Wang,
D.; Shuai, Z.; Huang, Y.-S.; Xie, L. Tunable band gap photo-
luminescence from atomically thin transition-metal dichalcogenide
alloys. ACS Nano 2013, 7, 4610−4616.
(24) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A.
Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147−150.
(25) Krivanek, O. L.; Corbin, G. J.; Dellby, N.; Elston, B. F.; Keyse,
R. J.; Murfitt, M. F.; Own, C. S.; Szilagyi, Z. S.; Woodruff, J. W. An
electron microscope for the aberration-corrected era. Ultramicroscopy
2008, 108, 179−195.
(26) Krivanek, O. L.; et al. Atom-by-atom structural and chemical
analysis by annular dark-field electron microscopy. Nature 2010, 464,
571−574.
(27) Zhou, W.; Oxley, M. P.; Lupini, A. R.; Krivanek, O. L.;
Pennycook, S. J.; Idrobo, J. C. Single atom microscopy. Microsc.
Microanal. 2012, 18, 1342−1354.
(28) Zhou, W.; et al. Intrinsic structural defects in monolayer
molybdenum disulfide. Nano Lett. 2013, 13, 2615−22.
(29) Krivanek, O. L.; Zhou, W.; Chisholm, M. F.; Idrobo, J. C.;
Lovejoy, T. C.; Ramasse, Q. M.; Dellby, N. Gentle STEM of Single
Atoms: Low keV Imaging and Analysis at Ultimate Detection Limits.
In Low Voltage Electron Microscopy: Principles and Applications; Bell, D.,
Erdman, N., Eds.; Wiley-Blackwell: London, U.K., 2013.
(30) van Benthem, K.; et al. Three-dimensional imaging of individual
hafnium atoms inside a semiconductor device. Appl. Phys. Lett. 2005,
87, 034104.
(31) Allen, J. E.; et al. High-resolution detection of Au catalyst atoms
in Si nanowires. Nat. Nanotechnol. 2008, 3, 168−173.
(32) Midgley, P. A.; Dunin-Borkowski, R. E. Electron tomography
and holography in materials science. Nat. Mater. 2009, 8, 271−280.
(33) Williams, D. B.; Carter, C. B. Transmission Electron Microscopy:
A Textbook for Materials Science, 2nd ed.; Springer: New York, 2009.
(34) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert,
M. I. J.; Refson, K.; Payne, M. C. First principles methods using
CASTEP. Z. Kristallogr. 2005, 220, 567−570.
(35) Ceperley, D. M.; Alder, B. J. Ground state of the electron gas by
a stochastic method. Phys. Rev. Lett. 1980, 45, 566−569.
(36) Perdew, J. P.; Zunger, A. Self-interaction correction to density-
functional approximations for many-electron systems. Phys. Rev. B
1981, 23, 5048−5079.
(37) Ataca, C.; Şahin, H.; Ciraci, S. Stable, single-layer MX2
transition-metal oxides and dichalcogenides in a honeycomb-like
structure. J. Phys. Chem. C 2012, 116, 8983.
(38) Henkelman, G.; Uberuaga, B. P.; Jonsson, H. A climbing image
nudged elastic band method for finding saddle points and minimum
energy paths. J. Chem. Phys. 2000, 113, 9901−9904.
(39) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid
metals. Phys. Rev. B 1993, 47, 558−561.

H dx.doi.org/10.1021/nl4032296 | Nano Lett. XXXX, XXX, XXX−XXX

You might also like