Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Grain boundary-mediated nanopores in molybdenum disulfide grown by chemical vapor deposition

Kenan Elibol a, Toma Susi a, Maria O′Brien b, Bernhard C. Bayer a, Timothy J. Pennycook a, Niall McEvoy b, Georg S. Duesberg b, Jannik C. Meyer a and Jani Kotakoski *a
aFaculty of Physics, University of Vienna, Boltzmanngasse 5, 1090 Wien, Austria. E-mail: jani.kotakoski@univie.ac.at
bSchool of Chemistry, Centre for the Research on Adaptive Nanostructures and Nanodevices (CRANN) and Advanced Materials and BioEngineering Research (AMBER), Trinity College Dublin, Dublin 2, Ireland

Received 17th November 2016 , Accepted 22nd December 2016

First published on 26th December 2016


Abstract

Molybdenum disulfide (MoS2) is a particularly interesting member of the family of two-dimensional (2D) materials due to its semiconducting and tunable electronic properties. Currently, the most reliable method for obtaining high-quality industrial scale amounts of 2D materials is chemical vapor deposition (CVD), which results in polycrystalline samples. As grain boundaries (GBs) are intrinsic defect lines within CVD-grown 2D materials, their atomic structure is of paramount importance. Here, through atomic-scale analysis of micrometer-long GBs, we show that covalently bound boundaries in 2D MoS2 tend to be decorated by nanopores. Such boundaries occur when differently oriented MoS2 grains merge during growth, whereas the overlap of grains leads to boundaries with bilayer areas. Our results suggest that the nanopore formation is related to stress release in areas with a high concentration of dislocation cores at the grain boundaries, and that the interlayer interaction leads to intrinsic rippling at the overlap regions. This provides insights for the controlled fabrication of large-scale MoS2 samples with desired structural properties for applications.


Introduction

Two-dimensional (2D) materials such as graphene, hexagonal boron nitride (hBN), and transition metal dichalcogenides like molybdenum disulfide (MoS2) show great potential for applications, especially in their monolayer (ML) forms (for examples, see ref. 1). A monolayer of MoS2 consists of two atomic layers of sulfide (S) and one atomic layer of molybdenum (Mo) between the sulfide layers. Within each layer, the S–Mo–S bonds are covalent, whereas the interlayer interaction is dominated by weak van der Waals forces.2,3 MoS2 exhibits semiconducting character due to its moderate band gap4 in contrast to semi-metallic graphene and insulating hBN. The attractive properties of MoS2 originate from its unsaturated d-electron interactions and quantum confinement2,5,6 and can be tuned by applying strain, through defect engineering, or by controlling its rippling.7–9 Currently, chemical vapor deposition (CVD) is the only way to produce macroscopic samples of ML MoS2. Since CVD leads to the growth of polycrystalline samples, understanding the atomic structure of grain boundaries is important for the growth of high-quality large area MoS2 films suitable for applications.

Here we present an atomic-scale scanning transmission electron microscopy (STEM) analysis of the longest grain boundary (GB) areas reported to date in CVD-grown MoS2. Covalently bound (stitched) GBs in CVD-MoS2 occur via the merging of triangular grains during the growth process.10,11 When two grains grow together but do not merge, an overlap region with a bilayer (BL) structure is formed instead. Surprisingly, we observed a chain of nanometer-sized pores (nanopores, NP) in all studied stitched GBs with size distributions that appear to be associated with the angle between the two grains: a decreasing misorientation angle leads to an increasing density of dislocation cores (DCs)12–14 and to larger NPs.

Despite being detrimental to some properties, NPs in 2D materials can also be beneficial. For example, they show promise for DNA sequencing, single molecule detection, filtration and power generation.15,16 In fact, NPs in MoS2 are thought to be particularly suited for, e.g., water desalination.15,17 The spontaneous formation of pores during film manufacture may provide an alternative to hitherto used dedicated post-growth processing via electron or ion beam irradiation or by electrochemical reactions.18–21 In addition to reducing the number of required processing steps, this also avoids drawbacks of those other methods such as material degradation during irradiation and contamination during chemical processes.

Our analysis of the BL areas at overlap regions reveals intrinsic rippling resulting from the Moiré pattern of just a slight misorientation between the grains (ca. 0.8°). Due to the sensitivity of the electronic system of MoS2, this is expected to lead to local variation of the optical and electronic properties due to varying interlayer separation and stacking as well as intralayer strain, revealing another way to affect the properties of CVD-MoS2 directly through the growth process.

Results and discussion

Our MoS2 samples were synthesized on SiO2/Si substrates by CVD in a micro-cavity setup that allowed the solid state precursor to be supplied in close proximity.22 They were transferred onto carbon-supported transmission electron microscopy (TEM) grids (for STEM experiments) and SiN/Si chips with multiple holes (for other experiments) following the method reported in ref. 23. Formation of different grain boundary areas is schematically illustrated in Fig. 1. CVD growth of MoS2 on SiO2 occurs through the growth of randomly placed and oriented triangle-shaped grains (see the atomic force microscopy image in ESI Fig. 1) that merge together.24 The colored frames in Fig. 1 show high angle annular dark field (HAADF)-STEM images of MoS2 including a stitched GB between two ML grains (red), ML without a grain boundary (yellow) and two overlapping regions corresponding to BL stacking with a 0.8° (blue) and a 10° (green) rotation between the layers. While a clear Moiré pattern occurs for the larger rotation angle, it is not apparent for the structure with the smaller rotation due to its large Moiré period. This schematic also shows how the grains create triangular empty spaces when the growth finishes prematurely (see also ESI Fig. 2).
image file: c6nr08958e-f1.tif
Fig. 1 A schematic representation of the formation of stitched grain boundary (GB) structures between two monolayer (ML) grains and bilayer (BL) stacking in overlap areas. HAADF-STEM images of MoS2 GB and BL stacking of MoS2 are shown in the colored frames.

In Fig. 2, we present large-scale images of stitched GBs in CVD-MoS2 along with the corresponding NP size distributions. The insets of Fig. 2a–c show HAADF-STEM images of three different samples with different but relatively similar misorientation angles between the grains. We observed NPs at all analyzed stitched GBs, but not within the grains themselves. We point out that although electron irradiation in a TEM experiment can cause nanopore formation in MoS2 even in near ultra-high vacuum due to sputtering,25 such pores are easy to distinguish from the intrinsic ones while imaging (see also the ESI for an irradiation-induced nanopore and two examples of the evolution of a GB area during imaging). For the GBs shown in Fig. 2, the NP areas varied from 0.12 nm2 to 25.45 nm2 with the larger pores being comparatively rare (less than ∼5% of pores had areas larger than 10 nm2). Although we could not directly ascertain this, the rare larger pores may have been formed during the sample transfer process. If this is the case, these might be avoided by improving the sample preparation method. The geometric mean sizes (±standard error) of the pores are 0.82 ± 0.17, 0.74 ± 0.20 and 0.60 ± 0.13 nm2 for misorientations of 7.9°, 8.9° and 9.6°, respectively. These misorientation angles were measured directly from the STEM images for atomic rows next to the boundaries with a field of view of 20 nm. Misorientations measured from Fourier transforms of the same images (dominated by atomic rows away from the GB) lead to values of 8.7°, 9.2° and 10.5°, respectively. The difference stems from local adjustment of the grain orientations toward each other at the boundary. Notice that the first bins in the histograms (smallest pores in inset of Fig. 2a–c) correspond to single and double Mo vacancies along with their neighboring S atoms. The longest atomic-scale images of MoS2-GBs, created by merging several HAADF-STEM images, are shown in Fig. 2d–f (see ESI for full-resolution atomically resolved composite images of the GBs). The bright contrast along the GBs is caused by hydrocarbon contamination attracted to the chemically more reactive GB structures (as compared to pristine MoS2). We point out that this contamination is also present in most of the NPs observed during STEM imaging, although its contrast in the images is lower than that of MoS2.


image file: c6nr08958e-f2.tif
Fig. 2 Large-scale images of stitched grain boundaries. (a–c) Histograms showing the area distribution of nanopores (NP) for 7.9°, 8.9° and 9.6° angle boundaries. Insets are HAADF-STEM images showing overviews of each of the analyzed structures. The red arrows show the location of the boundaries (they are also easy to recognize from the brighter contrast resulting from hydrocarbon contamination). (d–f) Atomic-resolution HAADF-STEM images of the areas marked with yellow rectangles in the overview images of panels (a–c). The green dashed lines indicate the boundary between the grains.

To further understand the formation of GB-NPs, we analyzed the dislocation cores along the stitched GBs, as shown in Fig. 3. The original HAADF-STEM image (in false color) is presented in Fig. 3a. In Fig. 3b, the same image is treated with a Wiener filter to enhance the contrast of the atomic structure. On top of this image, we have indicated the locations of five easily visible DCs as well as the NP by drawing Burger circuits around them. The location of each DC is confirmed by Fig. 3c, which shows a strain map obtained from geometric phase analysis (GPA)26–28 of the same image (DCs are marked with squares): the highest amounts of tensile and compressive strains appear at the DCs. Interestingly, it can be seen that one DC (highlighted by the white square) was hidden under the bright contamination directly left of the NP. A more detailed analysis of the strain around one of the DC structures is presented in Fig. 3d–f. As can be seen from the plots, the εxx strain fits both to the Peierls-Nabarro (PN) and the Foreman (FM) dislocation models.29 Indeed, for the FM model, the experimental strain profile is best fitted with a parameter a = 1 (see ESI), at which point the two models converge. We point out that although it is difficult to measure absolute strain values with a scanning device due to the possibility of sample drift, its (possible) outcome would be a uniform strain field superimposed on the actual strain, which would not alter any of our conclusions.


image file: c6nr08958e-f3.tif
Fig. 3 Stress analysis of a nanopore and dislocation cores at a stitched grain boundary. (a) HAADF-STEM image of a short section of a boundary with a nanopore. (b) The same image with a Wiener filter applied. White lines indicate Burger circuits around dislocation cores. The red and blue arrows show the lattice unit and Burger vectors, respectively. (c) εxx strain map of the image. Dislocation cores are indicated by squares (the white square shows a location of a dislocation core hidden under contamination in panels (a) and (b)). (d) Strain map around one of the dislocation cores. (e) εxx strain profile taken at a radius of 0.08 nm for image in panel (d) and fits to two dislocation models (PN and FM) with different fitting parameters. (f) εxx strain profiles taken at different radii along with fits to the FM dislocation model with parameter a = 1.

We hypothesize that the NP formation is associated with a buildup of stress at GBs with a high concentration of DCs. As can be seen in Fig. 3, the NP covers an area in which there would otherwise need to be four DCs. While the surrounding area shows that such a DC density is indeed possible, it is likely that the two grains would not be able to sustain the resulting strain over long distances. Analysis of the DC density of the three GBs presented in Fig. 2 is shown in Fig. 4. We start again with atomic-resolution images of example areas of the GBs (Fig. 4a–c). Fig. 4d–f show the corresponding mapping of εxx, εyy and shear (εxy) strain fields and rotation (γxy) obtained from GPA. Again, the points indicating the highest levels of tensile and compressive strain fields are DCs in the GBs. The insets in Fig. 4a–c show the distributions of distances between DCs in each GB. The average distance (±standard error) between the dislocation cores is measured as 2.12 ± 0.04, 2.01 ± 0.07 and 1.89 ± 0.03 nm for the 7.9°, 8.9° and 9.6° GBs, respectively. Even with the admittedly relatively low statistics, this clearly indicates that the DC separation decreases for increasing misorientation. To quantify this result, we compare it against the Frank-Bilby equation, given as δ = b/[2sin(θ/2)] in ref. 30 (where b is magnitude of the Burger vector and δ is the spacing of GB DCs for the misorientation angle θ). According to the equation, a large spacing between DCs should be observed for a small misorientation angle, consistent with our results. The values are shown in Table 1 along with our experimental estimates.


image file: c6nr08958e-f4.tif
Fig. 4 Dislocation density analysis at stitched grain boundaries. (a–c) Atomic-resolution HAADF-STEM images of the three grain boundaries. Insets show the distributions for distances between the dislocation cores for each boundary. (d–f) Strain maps for each boundary, calculated with geometric phase analysis: shear strain (εxy), rotation (γxy), εxx and εyy strain field. The features at the edges of the images are boundary artifacts.
Table 1 The average distances (δexp ± standard error) between dislocations and the geometric mean of the sizes (A ± standard error) of nanopores for each misorientation angle. δFBE are the expected dislocation distances based on the Frank-Bilby equation
θ (°) δ exp (nm) δ FBE (nm) A (nm2)
9.6 1.89 ± 0.03 1.86 0.60 ± 0.13
8.9 2.01 ± 0.07 2.01 0.74 ± 0.20
7.9 2.12 ± 0.04 2.26 0.82 ± 0.17


Unfortunately, due to the hydrocarbon contamination, the exact atomic structure at the DCs remain in most cases elusive even in our atomic-resolution images. However, in ESI Fig. 3, we present an example of a clean DC with its complete atomic structure resolved and compared to a model simulated with density functional theory (DFT).

Finally, we turn to the analysis of the overlap areas. It has already been shown that triangular MoS2 grains can grow on top of each other without creating chemical bonds.10 In our example case, the rotation between the two MoS2 layers is 0.8° with a Moiré periodicity of ∼14 nm, as shown in Fig. 5. Our DFT calculations (see Methods) confirm that AB stacking is energetically favored over AA (ref. 31) and the AB–AA transition region, all of which are observed in the experimental image. Since these different stackings also have different equilibrium distances (by less than 0.6 Å), it is reasonable to expect that one or both of the overlapping layers have an undulating structure. As a simple model of this situation, we created two model grains with the same misorientation as in the experiment, and artificially introduced a small ripple to one of them so that the different stackings each have an interlayer distance consistent with DFT calculations, and with a smooth transition between them with the same periodicity as in the experimental image. The simulated HAADF-STEM image based on this model (Fig. 5b) agrees well with the experiment, indicating that even such a small rotation leads to intrinsic rippling in BL MoS2. This result is also consistent with our analysis of the electron diffraction patterns of BL areas within the samples (see the ESI). Due to the sensitivity of the electronic properties of MoS2 to the stacking order (see ESI) and external influences,32 this allows at least in principle tuning those properties by controlling the misorientation of the overlapping grains.


image file: c6nr08958e-f5.tif
Fig. 5 Atomic-resolution analysis of a bilayer overlap area with a misorientation angle of 0.8°. (a) Double Gaussian filtered33 HAADF-STEM image of the area (unprocessed image is included as ESI Fig. 4). Squares with different colors on the image show different stacking orders. (b) Simulated HAADF image corresponding to (a). (c) The model used in the simulations. Note that the undulation, albeit small, is visible in the side view image. The graph under the model shows the calculated interlayer distances for different stacking orders and the equations describe the smooth transition between them as implemented for our model structure (L denotes distance between the different stackings). (d) Experimental and simulated images from the highlighted frames of (a) and (b).

Conclusions

Through atomic-resolution analysis of the longest grain boundary structures in MoS2 samples reported to date, we showed that stitched boundaries tend to be decorated by a chain of nanopores. The boundaries that we found formed between grains with misorientation angles ranging from 7.9° to 9.6°, and the observed pores had geometric mean sizes between 0.82 and 0.60 nm2 with a tendency for decreasing pore sizes with increasing angle. Through geometric phase analysis, we show that the average distances between the dislocation cores at MoS2 grain boundaries is greater for larger misorientation angles following the Frank-Bilby equation, which can help estimating the expected pore size for arbitrary misorientation angles. Further, we also presented an atomic-resolution analysis of an overlap region where two MoS2 grains have joined without covalent bonding, resulting instead in the formation of a bilayer with a misorientation of approximately 0.8°. Our analysis shows that the Moiré structure of the two overlapping lattices leads to a rippling of the material due to the different interlayer distances for different MoS2 stackings, hinting towards a possibility to create bilayer structures with locally varying electronic properties. Although defective grain boundaries have certainly a negative impact on properties important for some potential applications, the large number of possible applications for nanoporous MoS2 in sensing, DNA sequencing and water desalination gives promise that careful control over the grain orientation and the grain boundary structures during CVD-growth may lead to significant advances in large scale production of MoS2-based devices.

Methods

Sample growth

The MoS2 samples were synthesized on SiO2/Si using the CVD microreactor method described in ref. 22 and transferred onto carbon-supported TEM grids or SiN/Si chips with multiple holes following the method reported in ref. 23.

Transmission electron microscopy

STEM images were obtained using a Nion UltraSTEM 100 electron microscope with an accelerating voltage of 60 kV in near ultra-high vacuum (2 × 10−7 Pa) using the high angle annular dark field (HAADF) detector with a collection angle of 80–200 mrad. Diffraction experiments (ESI) were carried out using a Delong instruments LVEM5 table-top transmission electron microscope operated at 5 kV.

Modeling

Our atomistic simulations were carried out with density functional theory as implemented in the GPAW software package.34,35 To study the energetics of MoS2 bilayers, we created models with two six-atom unit cells of MoS2 in AA, AB and intermediate stacking. Using a plane-wave basis (cutoff energy 700 eV, 16 × 16 × 1 k-point mesh) we relaxed the atomic positions using the C09 van der Waals functional36 (we tested other exchange correlation functionals but this one seemed to give the best agreement with ref. 31 and 37) so that maximum forces were below 0.01 eV Å−1. Comparing the total energies (E) for each stacking order, we found EAB, EAA, and EAA–AB to be 22.926 eV, 22.971 eV and 22.969 eV, respectively. The interlayer distances for the AA, AB–AA and AB stackings were 6.936 Å, 7.077 Å and 6.521 Å, respectively. These results were used in the model structure (Fig. 5). To model the dislocation core structure (ESI Fig. 3), we created a 273-atom model and relaxed its atomic positions with periodic boundary conditions and a 3 × 3 × 1 k-point mesh so that maximum forces were <0.05 eV Å−1. Since this large structure has 2352 electrons, we used a double-zeta linear combination of atomic orbitals basis to speed up the calculations.38 The STEM image simulations were carried out with the QSTEM software39 with parameters matching our experimental conditions.

Atomic force microscopy and Raman spectroscopy

Atomic force microscopy (AFM) and Raman spectroscopy results (ESI Fig. 1) were recorded using a NT-MDT NTEGRA Spectra combined AFM-Raman spectrometer with a 473 nm wavelength. A Pt coated Si AFM cantilever with visual access (NT-MDT VIT_P_Pt) was used in the AFM experiment.

Acknowledgements

K. E. and J. C. M. acknowledge support from the Austrian Science Fund (FWF) through project P25721-N20 and T. S. through P28322-N36. B. C. B. and T. J. P. acknowledge funding from the European Union's Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 656214-2DInterFOX (B. C. B.) and 655760-DIGIPHASE (T. J. P.). M. O. B. acknowledges an Irish Research Council scholarship via the Enterprise Partnership Scheme, Project 201517, Award 12508. N. M. acknowledges support from SFI through 15/SIRG/3329. G. S. D. acknowledges the support of SFI under Contract No. 12/RC/2278 and PI_10/IN.1/I3030. J. K. acknowledges support from the Vienna Science and Technology Fund (WWTF) through project MA14-009. Computational resources through the Vienna Scientific Cluster are also gratefully acknowledged.

References

  1. Y. P. Venkata Subbaiah, K. J. Saji and A. Tiwari, Atomically Thin MoS2: A Versatile Nongraphene 2D Material, Adv. Funct. Mater., 2016, 26, 2046–2069 CrossRef CAS.
  2. K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Atomically Thin MoS2: A New Direct-Gap Semiconductor, Phys. Rev. Lett., 2010, 105, 136805 CrossRef PubMed.
  3. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis, Single-layer MoS2 transistors, Nat. Nanotechnol., 2011, 6, 147–150 CrossRef CAS PubMed.
  4. Q. Liu, et al., Tuning Electronic Structure of Bilayer MoS2 by Vertical Electric Field: A First-Principles Investigation, J. Phys. Chem. C, 2012, 116, 21556–21562 CAS.
  5. R. Ganatra and Q. Zhang, Few-Layer MoS2: A Promising Layered Semiconductor, ACS Nano, 2014, 8, 4074–4099 CrossRef CAS PubMed.
  6. P. Waduge, et al., Direct and Scalable Deposition of Atomically Thin Low-Noise MoS2 Membranes on Apertures, ACS Nano, 2015, 9, 7352–7359 CrossRef CAS PubMed.
  7. K. He, C. Poole, K. F. Mak and J. Shan, Experimental Demonstration of Continuous Electronic Structure Tuning via Strain in Atomically Thin MoS2, Nano Lett., 2013, 13, 2931–2936 CrossRef CAS PubMed.
  8. A. M. van der Zande, et al., Grains and grain boundaries in highly crystalline monolayer molybdenum disulphide, Nat. Mater., 2013, 12, 554–561 CrossRef CAS PubMed.
  9. Z. Lin, et al., Defect engineering of two-dimensional transition metal dichalcogenides, 2D Mater., 2016, 3, 022002 CrossRef.
  10. S. Najmaei, et al., Vapour phase growth and grain boundary structure of molybdenum disulphide atomic layers, Nat. Mater., 2013, 12, 754–759 CrossRef CAS PubMed.
  11. V. I. Artyukhov, Z. Hu, Z. Zhang and B. I. Yakobson, Topochemistry of Bowtie- and Star-Shaped Metal Dichalcogenide Nanoisland Formation, Nano Lett., 2016, 16, 3696–3702 CrossRef CAS PubMed.
  12. Y. Liu, F. Xu, Z. Zhang, E. S. Penev and B. I. Yakobson, Two-Dimensional Mono-Elemental Semiconductor with Electronically Inactive Defects: The Case of Phosphorus, Nano Lett., 2014, 14, 6782–6786 CrossRef CAS PubMed.
  13. X. Zou, Y. Liu and B. I. Yakobson, Predicting Dislocations and Grain Boundaries in Two-Dimensional Metal-Disulfides from the First Principles, Nano Lett., 2013, 13, 253–258 CrossRef CAS PubMed.
  14. O. V. Yazyev and S. G. Louie, Topological defects in graphene: Dislocations and grain boundaries, Phys. Rev. B: Condens. Matter, 2010, 81, 195420 CrossRef.
  15. M. Heiranian, A. B. Farimani and N. R. Aluru, Water desalination with a single-layer MoS2 nanopore, Nat. Commun., 2015, 6, 8616 CrossRef CAS PubMed.
  16. J. Feng, et al., Single-layer MoS2 nanopores as nanopower generators, Nature, 2016, 536, 197–200 CrossRef CAS PubMed.
  17. W. Li, Y. Yang, J. K. Weber, G. Zhang and R. Zhou, Tunable, Strain-Controlled Nanoporous MoS2 Filter for Water Desalination, ACS Nano, 2016, 10, 1829–1835 CrossRef CAS PubMed.
  18. J. Feng, et al., Electrochemical Reaction in Single Layer MoS2: Nanopores Opened Atom by Atom, Nano Lett., 2015, 15, 3431–3438 CrossRef CAS PubMed.
  19. X. Liu, et al., Top–down fabrication of sub-nanometre semiconducting nanoribbons derived from molybdenum disulfide sheets, Nat. Commun., 2013, 4, 1776 CrossRef PubMed.
  20. D. S. Fox, et al., Nanopatterning and Electrical Tuning of MoS2 Layers with a Subnanometer Helium Ion Beam, Nano Lett., 2015, 15, 5307–5313 CrossRef CAS PubMed.
  21. D. Emmrich, et al., Nanopore fabrication and characterization by helium ion microscopy, Appl. Phys. Lett., 2016, 108, 163103 CrossRef.
  22. M. O'Brien, et al., Transition Metal Dichalcogenide Growth via Close Proximity Precursor Supply, Sci. Rep., 2014, 4, 7374 CrossRef PubMed.
  23. K. Elibol, et al., Visualising the strain distribution in suspended two-dimensional materials under local deformation, Sci. Rep., 2016, 6, 28485 CrossRef CAS PubMed.
  24. X. Yin, et al., Edge Nonlinear Optics on a MoS2 Atomic Monolayer, Science, 2014, 344, 488–490 CrossRef CAS PubMed.
  25. H.-P. Komsa, et al., Two-dimensional transition metal dichalcogenides under electron irradiation: defect production and doping, Phys. Rev. Lett., 2012, 109, 035503 CrossRef PubMed.
  26. V. Grillo and E. Rotunno, STEM_CELL: A software tool for electron microscopy: Part I—simulations, Ultramicroscopy, 2013, 125, 97–111 CrossRef CAS PubMed.
  27. V. Grillo and F. Rossi, STEM_CELL: A software tool for electron microscopy. Part 2 analysis of crystalline materials, Ultramicroscopy, 2013, 125, 112–129 CrossRef CAS PubMed.
  28. M. J. Hÿtch, E. Snoeck and R. Kilaas, Quantitative measurement of displacement and strain fields from HREM micrographs, Ultramicroscopy, 1998, 74, 131–146 CrossRef.
  29. C. W. Zhao, Y. M. Xing, C. E. Zhou and P. C. Bai, Experimental examination of displacement and strain fields in an edge dislocation core, Acta Mater., 2008, 56, 2570–2575 CrossRef CAS.
  30. L. Wang, et al., Grain rotation mediated by grain boundary dislocations in nanocrystalline platinum, Nat. Commun., 2014, 5, 4402 CAS.
  31. J. T. Jiang, et al., Indirect–direct bandgap transition and gap width tuning in bilayer MoS2 superlattices, Chem. Phys. Lett., 2014, 613, 74–79 CrossRef CAS.
  32. B. Cao and T. Li, Interlayer Electronic Coupling in Arbitrarily Stacked MoS2 Bilayers Controlled by Interlayer S–S Interaction, J. Phys. Chem. C, 2015, 119, 1247–1252 CAS.
  33. O. L. Krivanek, et al., Atom-by-atom structural and chemical analysis by annular dark-field electron microscopy, Nature, 2010, 464, 571–574 CrossRef CAS PubMed.
  34. J. J. Mortensen, L. B. Hansen and K. W. Jacobsen, Real-space grid implementation of the projector augmented wave method, Phys. Rev. B: Condens. Matter, 2005, 71, 035109 CrossRef.
  35. J. Enkovaara, et al., Electronic structure calculations with GPAW: a real-space implementation of the projector augmented-wave method, J. Phys.: Condens. Matter, 2010, 22, 253202 CrossRef CAS PubMed.
  36. V. R. Cooper, van der Waals density functional: An appropriate exchange functional, Phys. Rev. B: Condens. Matter, 2010, 81, 161104 CrossRef.
  37. S. Yang, J. Kang, Q. Yue and K. Yao, Vapor Phase Growth and Imaging Stacking Order of Bilayer Molybdenum Disulfide, J. Phys. Chem. C, 2014, 118, 9203–9208 CAS.
  38. A. H. Larsen, M. Vanin, J. J. Mortensen, K. S. Thygesen and K. W. Jacobsen, Localized atomic basis set in the projector augmented wave method, Phys. Rev. B: Condens. Matter, 2009, 80, 195112 CrossRef.
  39. C. Koch, Determination of Core Structure Periodicity and Point Defect Density along Dislocations, Arizona State University, 2002 Search PubMed.
  40. H. Li, et al., From Bulk to Monolayer MoS2: Evolution of Raman Scattering, Adv. Funct. Mater., 2012, 22, 1385–1390 CrossRef CAS.
  41. H. J. Conley, et al., Bandgap Engineering of Strained Monolayer and Bilayer MoS2, Nano Lett., 2013, 13, 3626–3630 CrossRef CAS PubMed.
  42. X. Zhang, et al., Raman spectroscopy of shear and layer breathing modes in multilayer MoS2, Phys. Rev. B: Condens. Matter, 2013, 87, 115413 CrossRef.
  43. S. Mignuzzi, et al., Effect of disorder on Raman scattering of single-layer MoS2, Phys. Rev. B: Condens. Matter, 2015, 91, 195411 CrossRef.
  44. S. Mathew, et al., Magnetism in MoS2 induced by proton irradiation, Appl. Phys. Lett., 2012, 101, 102103 CrossRef.
  45. J. Brivio, D. T. L. Alexander and A. Kis, Ripples and Layers in Ultrathin MoS2 Membranes, Nano Lett., 2011, 11, 5148–5153 CrossRef CAS PubMed.
  46. K. Liu, et al., Elastic Properties of Chemical-Vapor-Deposited Monolayer MoS2, WS2, and Their Bilayer Heterostructures, Nano Lett., 2014, 14, 5097–5103 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: AFM and Raman spectroscopy results, STEM-HAADF images of triangular holes between grains, an atomic-resolution image of a dislocation core and the corresponding simulated model, image of a transition between a monolayer and bilayer area, TEM diffraction analysis, strain analysis of an electron-beam-induced nanopore, theoretical equations describing the strain around dislocations, untreated image of the structure shown in Fig. 5a and a table of energies, interlayer distances and band gaps calculated for bilayers with different stacking orders. Ref. 40–46 are cited in the ESI. Supplementary files with atomic-scale images of the three studied covalently bound grain boundaries are additionally provided as full-resolution composite figures. See DOI: 10.1039/c6nr08958e

This journal is © The Royal Society of Chemistry 2017