Chemistry:Molybdenum disulfide

From HandWiki
Molybdenum disulfide
Molybdenum disulfide
Molybdenite-3D-balls.png
Names
IUPAC name
Molybdenum disulfide
Other names
Molybdenum(IV) sulfide
Identifiers
3D model (JSmol)
ChEBI
ChemSpider
RTECS number
  • QA4697000
UNII
Properties
MoS2
Molar mass 160.07 g/mol[1]
Appearance black/lead-gray solid
Density 5.06 g/cm3[1]
Melting point 2,375 °C (4,307 °F; 2,648 K)[4]
insoluble[1]
Solubility decomposed by aqua regia, hot sulfuric acid, nitric acid
insoluble in dilute acids
Band gap 1.23 eV (indirect, 3R or 2H bulk)[2]
~1.8 eV (direct, monolayer)[3]
Structure
hP6, P63/mmc, No. 194 (2H)

hR9, R3m, No 160 (3R)[5]

a = 0.3161 nm (2H), 0.3163 nm (3R), c = 1.2295 nm (2H), 1.837 (3R)
Trigonal prismatic (MoIV)
Pyramidal (S2−)
Thermochemistry
62.63 J/(mol K)
-235.10 kJ/mol
-225.89 kJ/mol
Hazards
Safety data sheet External MSDS
Related compounds
Other anions
Molybdenum(IV) oxide
Molybdenum diselenide
Molybdenum ditelluride
Other cations
Tungsten disulfide
Related lubricants
Graphite
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N verify (what is ☑Y☒N ?)
Infobox references

Molybdenum disulfide (or moly) is an inorganic compound composed of molybdenum and sulfur. Its chemical formula is MoS2.

The compound is classified as a transition metal dichalcogenide. It is a silvery black solid that occurs as the mineral molybdenite, the principal ore for molybdenum.[6] MoS2 is relatively unreactive. It is unaffected by dilute acids and oxygen. In appearance and feel, molybdenum disulfide is similar to graphite. It is widely used as a dry lubricant because of its low friction and robustness. Bulk MoS2 is a diamagnetic, indirect bandgap semiconductor similar to silicon, with a bandgap of 1.23 eV.[2]

Production

Molybdenite

MoS2 is naturally found as either molybdenite, a crystalline mineral, or jordisite, a rare low temperature form of molybdenite.[7] Molybdenite ore is processed by flotation to give relatively pure MoS2. The main contaminant is carbon. MoS2 also arises by thermal treatment of virtually all molybdenum compounds with hydrogen sulfide or elemental sulfur and can be produced by metathesis reactions from molybdenum pentachloride.[8]

Structure and physical properties

Electron microscopy of antisites (a, Mo substitutes for S) and vacancies (b, missing S atoms) in a monolayer of molybdenum disulfide. Scale bar: 1 nm.[9]

Crystalline phases

All forms of MoS2 have a layered structure, in which a plane of molybdenum atoms is sandwiched by planes of sulfide ions. These three strata form a monolayer of MoS2. Bulk MoS2 consists of stacked monolayers, which are held together by weak van der Waals interactions.

Crystalline MoS2 exists in one of two phases, 2H-MoS2 and 3R-MoS2, where the "H" and the "R" indicate hexagonal and rhombohedral symmetry, respectively. In both of these structures, each molybdenum atom exists at the center of a trigonal prismatic coordination sphere and is covalently bonded to six sulfide ions. Each sulfur atom has pyramidal coordination and is bonded to three molybdenum atoms. Both the 2H- and 3R-phases are semiconducting.[10]

A third, metastable crystalline phase known as 1T-MoS2 was discovered by intercalating 2H-MoS2 with alkali metals.[11] This phase has trigonal symmetry and is metallic. The 1T-phase can be stabilized through doping with electron donors such as rhenium,[12] or converted back to the 2H-phase by microwave radiation.[13] The 2H/1T-phase transition can be controlled via the incorporation of S vacancies.[14]

Allotropes

Nanotube-like and buckyball-like molecules composed of MoS2 are known.[15]

Exfoliated MoS2 flakes

While bulk MoS2 in the 2H-phase is known to be an indirect-band gap semiconductor, monolayer MoS2 has a direct band gap. The layer-dependent optoelectronic properties of MoS2 have promoted much research in 2-dimensional MoS2-based devices. 2D MoS2 can be produced by exfoliating bulk crystals to produce single-layer to few-layer flakes either through a dry, micromechanical process or through solution processing.

Micromechanical exfoliation, also pragmatically called "Scotch-tape exfoliation", involves using an adhesive material to repeatedly peel apart a layered crystal by overcoming the van der Waals forces. The crystal flakes can then be transferred from the adhesive film to a substrate. This facile method was first used by Konstantin Novoselov and Andre Geim to obtain graphene from graphite crystals. However, it can not be employed for a uniform 1-D layers because of weaker adhesion of MoS2 to the substrate (either Si, glass or quartz); the aforementioned scheme is good for graphene only.[16] While Scotch tape is generally used as the adhesive tape, PDMS stamps can also satisfactorily cleave MoS2 if it is important to avoid contaminating the flakes with residual adhesive.[17]

Liquid-phase exfoliation can also be used to produce monolayer to multi-layer MoS2 in solution. A few methods include lithium intercalation[18] to delaminate the layers and sonication in a high-surface tension solvent.[19][20]

Mechanical properties

MoS2 excels as a lubricating material (see below) due to its layered structure and low coefficient of friction. Interlayer sliding dissipates energy when a shear stress is applied to the material. Extensive work has been performed to characterize the coefficient of friction and shear strength of MoS2 in various atmospheres.[21] The shear strength of MoS2 increases as the coefficient of friction increases. This property is called superlubricity. At ambient conditions, the coefficient of friction for MoS2 was determined to be 0.150, with a corresponding estimated shear strength of 56.0 MPa (megapascals).[21] Direct methods of measuring the shear strength indicate that the value is closer to 25.3 MPa.[22]

The wear resistance of MoS2 in lubricating applications can be increased by doping MoS2 with Cr. Microindentation experiments on nanopillars of Cr-doped MoS2 found that the yield strength increased from an average of 821 MPa for pure MoS2 (at 0% Cr) to 1017 MPa at 50% Cr.[23] The increase in yield strength is accompanied by a change in the failure mode of the material. While the pure MoS2 nanopillar fails through a plastic bending mechanism, brittle fracture modes become apparent as the material is loaded with increasing amounts of dopant.[23]

The widely used method of micromechanical exfoliation has been carefully studied in MoS2 to understand the mechanism of delamination in few-layer to multi-layer flakes. The exact mechanism of cleavage was found to be layer dependent. Flakes thinner than 5 layers undergo homogenous bending and rippling, while flakes around 10 layers thick delaminated through interlayer sliding. Flakes with more than 20 layers exhibited a kinking mechanism during micromechanical cleavage. The cleavage of these flakes was also determined to be reversible due to the nature of van der Waals bonding.[24]

In recent years, MoS2 has been utilized in flexible electronic applications, promoting more investigation into the elastic properties of this material. Nanoscopic bending tests using AFM cantilever tips were performed on micromechanically exfoliated MoS2 flakes that were deposited on a holey substrate.[17][25] The yield strength of monolayer flakes was 270 GPa,[25] while the thicker flakes were also stiffer, with a yield strength of 330 GPa.[17] Molecular dynamic simulations found the in-plane yield strength of MoS2 to be 229 GPa, which matches the experimental results within error.[26]

Bertolazzi and coworkers also characterized the failure modes of the suspended monolayer flakes. The strain at failure ranges from 6 to 11%. The average yield strength of monolayer MoS2 is 23 GPa, which is close to the theoretical fracture strength for defect-free MoS2.[25]

The band structure of MoS2 is sensitive to strain.[27][28][29]

Chemical reactions

Molybdenum disulfide is stable in air and attacked only by aggressive reagents. It reacts with oxygen upon heating forming molybdenum trioxide:

2 MoS2 + 7 O2 → 2 MoO3 + 4 SO2

Chlorine attacks molybdenum disulfide at elevated temperatures to form molybdenum pentachloride:

2 MoS2 + 7 Cl2 → 2 MoCl5 + 2 S2Cl2

Intercalation reactions

Molybdenum disulfide is a host for formation of intercalation compounds. This behavior is relevant to its use as a cathode material in batteries.[30][31] One example is a lithiated material, LixMoS2.[32] With butyl lithium, the product is LiMoS2.[6]

Applications

Lubricant

A tube of commercial graphite powder lubricant with molybdenum disulfide additive (called "molybdenum")[33]

Due to weak van der Waals interactions between the sheets of sulfide atoms, MoS2 has a low coefficient of friction. MoS2 in particle sizes in the range of 1–100 µm is a common dry lubricant.[34] Few alternatives exist that confer high lubricity and stability at up to 350 °C in oxidizing environments. Sliding friction tests of MoS2 using a pin on disc tester at low loads (0.1–2 N) give friction coefficient values of <0.1.[35][36]

MoS2 is often a component of blends and composites that require low friction. For example, it is added to graphite to improve sticking.[33] A variety of oils and greases are used, because they retain their lubricity even in cases of almost complete oil loss, thus finding a use in critical applications such as aircraft engines. When added to plastics, MoS2 forms a composite with improved strength as well as reduced friction. Polymers that may be filled with MoS2 include nylon (trade name Nylatron), Teflon and Vespel. Self-lubricating composite coatings for high-temperature applications consist of molybdenum disulfide and titanium nitride, using chemical vapor deposition.

Examples of applications of MoS2-based lubricants include two-stroke engines (such as motorcycle engines), bicycle coaster brakes, automotive CV and universal joints, ski waxes[37] and bullets.[38]

Other layered inorganic materials that exhibit lubricating properties (collectively known as solid lubricants (or dry lubricants)) includes graphite, which requires volatile additives and hexagonal boron nitride.[39]

Catalysis

Fingerprint revealed by molybdenum disulfide

MoS2 is employed as a cocatalyst for desulfurization in petrochemistry, for example, hydrodesulfurization. The effectiveness of the MoS2 catalysts is enhanced by doping with small amounts of cobalt or nickel. The intimate mixture of these sulfides is supported on alumina. Such catalysts are generated in situ by treating molybdate/cobalt or nickel-impregnated alumina with H2S or an equivalent reagent. Catalysis does not occur at the regular sheet-like regions of the crystallites, but instead at the edge of these planes.[40]

MoS2 finds use as a hydrogenation catalyst for organic synthesis.[41] It is derived from a common transition metal, rather than group 10 metal as are many alternatives, MoS2 is chosen when catalyst price or resistance to sulfur poisoning are of primary concern. MoS2 is effective for the hydrogenation of nitro compounds to amines and can be used to produce secondary amines via reductive amination.[42] The catalyst can also can effect hydrogenolysis of organosulfur compounds, aldehydes, ketones, phenols and carboxylic acids to their respective alkanes.[41] The catalyst suffers from rather low activity however, often requiring hydrogen pressures above 95 atm and temperatures above 185 °C.

Research

MoS2 plays an important role in condensed matter physics research.[43]

Hydrogen evolution

MoS2 and related molybdenum sulfides are efficient catalysts for hydrogen evolution, including the electrolysis of water;[44][45] thus, are possibly useful to produce hydrogen for use in fuel cells.[46]

Oxygen reduction and evolution

MoS2@Fe-N-C core/shell[47] nanosphere with atomic Fe-doped surface and interface (MoS2/Fe-N-C) can be used as a used an electrocatalyst for oxygen reduction and evolution reactions (ORR and OER) bifunctionally because of reduced energy barrier due to Fe-N4 dopants and unique nature of MoS2/Fe-N-C interface.

Microelectronics

As in graphene, the layered structures of MoS2 and other transition metal dichalcogenides exhibit electronic and optical properties[48] that can differ from those in bulk.[49] Bulk MoS2 has an indirect band gap of 1.2 eV,[50][51] while MoS2 monolayers have a direct 1.8 eV electronic bandgap,[52] supporting switchable transistors[53] and photodetectors.[54][49][55]

MoS2 nanoflakes can be used for solution-processed fabrication of layered memristive and memcapacitive devices through engineering a MoOx/MoS2 heterostructure sandwiched between silver electrodes.[56] MoS2-based memristors are mechanically flexible, optically transparent and can be produced at low cost.

The sensitivity of a graphene field-effect transistor (FET) biosensor is fundamentally restricted by the zero band gap of graphene, which results in increased leakage and reduced sensitivity. In digital electronics, transistors control current flow throughout an integrated circuit and allow for amplification and switching. In biosensing, the physical gate is removed and the binding between embedded receptor molecules and the charged target biomolecules to which they are exposed modulates the current.[57]

MoS2 has been investigated as a component of flexible circuits.[58][59]

In 2017, a 115-transistor, 1-bit microprocessor implementation was fabricated using two-dimensional MoS2.[60]

MoS2 has been used to create 2D 2-terminal memristors and 3-terminal memtransistors.[61]

Valleytronics

Due to the lack of spatial inversion symmetry, odd-layer MoS2 is a promising material for valleytronics because both the CBM and VBM have two energy-degenerate valleys at the corners of the first Brillouin zone, providing an exciting opportunity to store the information of 0s and 1s at different discrete values of the crystal momentum. The Berry curvature is even under spatial inversion (P) and odd under time reversal (T), the valley Hall effect cannot survive when both P and T symmetries are present. To excite valley Hall effect in specific valleys, circularly polarized lights were used for breaking the T symmetry in atomically thin transition-metal dichalcogenides.[62] In monolayer MoS2, the T and mirror symmetries lock the spin and valley indices of the sub-bands split by the spin-orbit couplings, both of which are flipped under T; the spin conservation suppresses the inter-valley scattering. Therefore, monolayer MoS2 have been deemed an ideal platform for realizing intrinsic valley Hall effect without extrinsic symmetry breaking.[63]

Photonics and photovoltaics

MoS2 also possesses mechanical strength, electrical conductivity, and can emit light, opening possible applications such as photodetectors.[64] MoS2 has been investigated as a component of photoelectrochemical (e.g. for photocatalytic hydrogen production) applications and for microelectronics applications.[53]

Superconductivity of monolayers

Under an electric field MoS2 monolayers have been found to superconduct at temperatures below 9.4 K.[65]

See also

References

  1. 1.0 1.1 1.2 Haynes, William M., ed (2011). CRC Handbook of Chemistry and Physics (92nd ed.). Boca Raton, FL: CRC Press. p. 4.76. ISBN 1439855110. 
  2. 2.0 2.1 Kobayashi, K.; Yamauchi, J. (1995). "Electronic structure and scanning-tunneling-microscopy image of molybdenum dichalcogenide surfaces". Physical Review B 51 (23): 17085–17095. doi:10.1103/PhysRevB.51.17085. PMID 9978722. Bibcode1995PhRvB..5117085K. 
  3. Yun, Won Seok; Han, S. W.; Hong, Soon Cheol; Kim, In Gee; Lee, J. D. (2012). "Thickness and strain effects on electronic structures of transition metal dichalcogenides: 2H-MX2 semiconductors (M = Mo, W; X = S, Se, Te)". Physical Review B 85 (3): 033305. doi:10.1103/PhysRevB.85.033305. Bibcode2012PhRvB..85c3305Y. 
  4. "Molybdenum Disulfide". PubChem. https://pubchem.ncbi.nlm.nih.gov/compound/14823#section=Color. 
  5. Schönfeld, B.; Huang, J. J.; Moss, S. C. (1983). "Anisotropic mean-square displacements (MSD) in single-crystals of 2H- and 3R-MoS2". Acta Crystallographica Section B 39 (4): 404–407. doi:10.1107/S0108768183002645. 
  6. 6.0 6.1 Sebenik, Roger F. et al. (2005) "Molybdenum and Molybdenum Compounds", Ullmann's Encyclopedia of Chemical Technology. Wiley-VCH, Weinheim. doi: 10.1002/14356007.a16_655
  7. "Jordisite". https://www.mindat.org/min-2114.html. 
  8. Murphy, Donald W.; Interrante, Leonard V.; Kaner; Mansuktto (1995). Metathetical Precursor Route to Molybdenum Disulfide. Inorganic Syntheses. 30. pp. 33–37. doi:10.1002/9780470132616.ch8. ISBN 9780470132616. 
  9. Hong, J.; Hu, Z.; Probert, M.; Li, K.; Lv, D.; Yang, X.; Gu, L.; Mao, N. et al. (2015). "Exploring atomic defects in molybdenum disulphide monolayers". Nature Communications 6: 6293. doi:10.1038/ncomms7293. PMID 25695374. Bibcode2015NatCo...6.6293H. 
  10. (in de) Gmelin Handbook of Inorganic and Organometallic Chemistry - 8th edition. https://www.springer.com/series/562. 
  11. Wypych, Fernando; Schöllhorn, Robert (1992-01-01). "1T-MoS2, a new metallic modification of molybdenum disulfide" (in en). Journal of the Chemical Society, Chemical Communications (19): 1386–1388. doi:10.1039/C39920001386. ISSN 0022-4936. http://pubs.rsc.org/is/content/articlehtml/1992/c3/c39920001386. 
  12. Enyashin, Andrey N.; Yadgarov, Lena; Houben, Lothar; Popov, Igor; Weidenbach, Marc; Tenne, Reshef; Bar-Sadan, Maya; Seifert, Gotthard (2011-12-22). "New Route for Stabilization of 1T-WS2 and MoS2 Phases". The Journal of Physical Chemistry C 115 (50): 24586–24591. doi:10.1021/jp2076325. ISSN 1932-7447. 
  13. Xu, Danyun; Zhu, Yuanzhi; Liu, Jiapeng; Li, Yang; Peng, Wenchao; Zhang, Guoliang; Zhang, Fengbao; Fan, Xiaobin (2016). "Microwave-assisted 1T to 2H phase reversion of MoS 2 in solution: a fast route to processable dispersions of 2H-MoS 2 nanosheets and nanocomposites" (in en). Nanotechnology 27 (38): 385604. doi:10.1088/0957-4484/27/38/385604. ISSN 0957-4484. PMID 27528593. Bibcode2016Nanot..27L5604X. 
  14. Gan, Xiaorong; Lee, Lawrence Yoon Suk; Wong, Kwok-yin; Lo, Tsz Wing; Ho, Kwun Hei; Lei, Dang Yuan; Zhao, Huimin (2018-09-24). "2H/1T Phase Transition of Multilayer MoS 2 by Electrochemical Incorporation of S Vacancies" (in en). ACS Applied Energy Materials 1 (9): 4754–4765. doi:10.1021/acsaem.8b00875. ISSN 2574-0962. https://pubs.acs.org/doi/10.1021/acsaem.8b00875. 
  15. Tenne, R.; Redlich, M. (2010). "Recent progress in the research of inorganic fullerene-like nanoparticles and inorganic nanotubes". Chemical Society Reviews 39 (5): 1423–34. doi:10.1039/B901466G. PMID 20419198. 
  16. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. (2004-10-22). "Electric Field Effect in Atomically Thin Carbon Films" (in en). Science 306 (5696): 666–669. doi:10.1126/science.1102896. ISSN 0036-8075. PMID 15499015. Bibcode2004Sci...306..666N. 
  17. 17.0 17.1 17.2 Castellanos-Gomez, Andres; Poot, Menno; Steele, Gary A.; van der Zant, Herre S. J.; Agraït, Nicolás; Rubio-Bollinger, Gabino (2012-02-07). "Elastic Properties of Freely Suspended MoS2 Nanosheets" (in en). Advanced Materials 24 (6): 772–775. doi:10.1002/adma.201103965. ISSN 1521-4095. PMID 22231284. Bibcode2012AdM....24..772C. 
  18. Wan, Jiayu; Lacey, Steven D.; Dai, Jiaqi; Bao, Wenzhong; Fuhrer, Michael S.; Hu, Liangbing (2016-12-05). "Tuning two-dimensional nanomaterials by intercalation: materials, properties and applications" (in en). Chemical Society Reviews 45 (24): 6742–6765. doi:10.1039/C5CS00758E. ISSN 1460-4744. PMID 27704060. 
  19. Coleman, Jonathan N.; Lotya, Mustafa; O’Neill, Arlene; Bergin, Shane D.; King, Paul J.; Khan, Umar; Young, Karen; Gaucher, Alexandre et al. (2011-02-04). "Two-Dimensional Nanosheets Produced by Liquid Exfoliation of Layered Materials" (in en). Science 331 (6017): 568–571. doi:10.1126/science.1194975. ISSN 0036-8075. PMID 21292974. Bibcode2011Sci...331..568C. 
  20. Zhou, Kai-Ge; Mao, Nan-Nan; Wang, Hang-Xing; Peng, Yong; Zhang, Hao-Li (2011-11-11). "A Mixed-Solvent Strategy for Efficient Exfoliation of Inorganic Graphene Analogues" (in en). Angewandte Chemie 123 (46): 11031–11034. doi:10.1002/ange.201105364. ISSN 1521-3757. Bibcode2011AngCh.12311031Z. 
  21. 21.0 21.1 Donnet, C.; Martin, J. M.; Le Mogne, Th.; Belin, M. (1996-02-01). "Super-low friction of MoS2 coatings in various environments". Tribology International 29 (2): 123–128. doi:10.1016/0301-679X(95)00094-K. 
  22. Oviedo, Juan Pablo; KC, Santosh; Lu, Ning; Wang, Jinguo; Cho, Kyeongjae; Wallace, Robert M.; Kim, Moon J. (2015-02-24). "In Situ TEM Characterization of Shear-Stress-Induced Interlayer Sliding in the Cross Section View of Molybdenum Disulfide". ACS Nano 9 (2): 1543–1551. doi:10.1021/nn506052d. ISSN 1936-0851. PMID 25494557. 
  23. 23.0 23.1 Tedstone, Aleksander A.; Lewis, David J.; Hao, Rui; Mao, Shi-Min; Bellon, Pascal; Averback, Robert S.; Warrens, Christopher P.; West, Kevin R. et al. (2015-09-23). "Mechanical Properties of Molybdenum Disulfide and the Effect of Doping: An in Situ TEM Study". ACS Applied Materials & Interfaces 7 (37): 20829–20834. doi:10.1021/acsami.5b06055. ISSN 1944-8244. PMID 26322958. 
  24. Tang, Dai-Ming; Kvashnin, Dmitry G.; Najmaei, Sina; Bando, Yoshio; Kimoto, Koji; Koskinen, Pekka; Ajayan, Pulickel M.; Yakobson, Boris I. et al. (2014-04-03). "Nanomechanical cleavage of molybdenum disulphide atomic layers" (in en). Nature Communications 5: 3631. doi:10.1038/ncomms4631. PMID 24698887. Bibcode2014NatCo...5.3631T. 
  25. 25.0 25.1 25.2 Bertolazzi, Simone; Brivio, Jacopo; Kis, Andras (2011). "Stretching and Breaking of Ultrathin MoS2" (in en). ACS Nano 5 (12): 9703–9709. doi:10.1021/nn203879f. PMID 22087740. http://infoscience.epfl.ch/record/170263. 
  26. Jiang, Jin-Wu; Park, Harold S.; Rabczuk, Timon (2013-08-12). "Molecular dynamics simulations of single-layer molybdenum disulphide (MoS2): Stillinger-Weber parametrization, mechanical properties, and thermal conductivity". Journal of Applied Physics 114 (6): 064307–064307–10. doi:10.1063/1.4818414. ISSN 0021-8979. Bibcode2013JAP...114f4307J. 
  27. Li, H.; Wu, J.; Yin, Z.; Zhang, H. (2014). "Preparation and Applications of Mechanically Exfoliated Single-Layer and Multilayer MoS2 and WSe2 Nanosheets". Acc. Chem. Res. 47 (4): 1067–75. doi:10.1021/ar4002312. PMID 24697842. 
  28. Amorim, B.; Cortijo, A.; De Juan, F.; Grushin, A.G.; Guinea, F.; Gutiérrez-Rubio, A.; Ochoa, H.; Parente, V. et al. (2016). "Novel effects of strains in graphene and other two dimensional materials". Physics Reports 1503: 1–54. doi:10.1016/j.physrep.2015.12.006. Bibcode2016PhR...617....1A. 
  29. Zhang, X.; Lai, Z.; Tan, C.; Zhang, H. (2016). "Solution-Processed Two-Dimensional MoS2 Nanosheets: Preparation, Hybridization, and Applications". Angew. Chem. Int. Ed. 55 (31): 8816–8838. doi:10.1002/anie.201509933. PMID 27329783. 
  30. Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. (2014). "Lithium Ion Battery Applications of Molybdenum Disulfide (MoS2) Nanocomposites". Energy Environ. Sci. 7: 209–31. doi:10.1039/C3EE42591F. 
  31. Benavente, E.; Santa Ana, M. A.; Mendizabal, F.; Gonzalez, G. (2002). "Intercalation chemistry of molybdenum disulfide". Coordination Chemistry Reviews 224 (1–2): 87–109. doi:10.1016/S0010-8545(01)00392-7. 
  32. Müller-Warmuth, W.; Schöllhorn, R. (1994). Progress in intercalation research. Springer. ISBN 978-0-7923-2357-0. https://books.google.com/books?id=IyB_rPo3osUC&pg=PA50. 
  33. 33.0 33.1 High Performance, Dry Powdered Graphite with sub-micron molybdenum disulfide. pinewoodpro.com
  34. Claus, F. L. (1972), "Solid Lubricants and Self-Lubricating Solids", New York: Academic Press, Bibcode1972slsl.book.....C 
  35. Miessler, Gary L.; Tarr, Donald Arthur (2004). Inorganic Chemistry. Pearson Education. ISBN 978-0-13-035471-6. https://books.google.com/books?id=oLQPAQAAMAAJ. 
  36. Shriver, Duward; Atkins, Peter; Overton, T. L.; Rourke, J. P.; Weller, M. T.; Armstrong, F. A. (17 February 2006). Inorganic Chemistry. W. H. Freeman. ISBN 978-0-7167-4878-6. https://books.google.com/books?id=so8oAQAAMAAJ. 
  37. "On dry lubricants in ski waxes". Swix Sport AX. http://www.swixsport.com/dav/8dde5f4784.pdf. 
  38. "Barrels retain accuracy longer with Diamond Line". Norma. http://www.norma.cc/en/Ammunition-Academy/Barrel-wear/. 
  39. Bartels, Thorsten (2002). "Lubricants and Lubrication". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley VCH. doi:10.1002/14356007.a15_423. ISBN 978-3527306732. 
  40. Topsøe, H.; Clausen, B. S.; Massoth, F. E. (1996). Hydrotreating Catalysis, Science and Technology. Berlin: Springer-Verlag. 
  41. 41.0 41.1 Nishimura, Shigeo (2001). Handbook of Heterogeneous Catalytic Hydrogenation for Organic Synthesis (1st ed.). New York: Wiley-Interscience. pp. 43–44 & 240–241. ISBN 9780471396987. https://books.google.com/books?id=RjZRAAAAMAAJ&pg=PA43. 
  42. Dovell, Frederick S.; Greenfield, Harold (1964). "Base-Metal Sulfides as Reductive Alkylation Catalysts". The Journal of Organic Chemistry 29 (5): 1265–1267. doi:10.1021/jo01028a511. 
  43. Wood, Charlie (2022-08-16). "Physics Duo Finds Magic in Two Dimensions" (in en). https://www.quantamagazine.org/physics-duo-finds-magic-in-two-dimensions-20220816/. 
  44. Kibsgaard, Jakob; Jaramillo, Thomas F.; Besenbacher, Flemming (2014). "Building an appropriate active-site motif into a hydrogen-evolution catalyst with thiomolybdate [Mo3S132− clusters"]. Nature Chemistry 6 (3): 248–253. doi:10.1038/nchem.1853. PMID 24557141. Bibcode2014NatCh...6..248K. https://zenodo.org/record/889641. 
  45. Laursen, A. B.; Kegnaes, S.; Dahl, S.; Chorkendorff, I. (2012). "Molybdenum Sulfides – Efficient and Viable Materials for Electro- and Photoelectrocatalytic Hydrogen Evolution". Energy Environ. Sci. 5 (2): 5577–91. doi:10.1039/c2ee02618j. 
  46. "Superior hydrogen catalyst just grows that way" (news release). Sandia Labs. https://share-ng.sandia.gov/news/resources/news_releases/superior-catalyst/#.Wia84bbMw5s. "a spray-printing process that uses molybdenum disulfide to create a “flowering” hydrogen catalyst far cheaper than platinum and reasonably close in efficiency." 
  47. Yan, Yan; Liang, Shuang; Wang, Xiang; Zhang, Mingyue; Hao, Shu-Meng; Cui, Xun; Li, Zhiwei; Lin, Zhiqun (2021-10-05). "Robust wrinkled MoS 2 /N-C bifunctional electrocatalysts interfaced with single Fe atoms for wearable zinc-air batteries" (in en). Proceedings of the National Academy of Sciences 118 (40): e2110036118. doi:10.1073/pnas.2110036118. ISSN 0027-8424. PMID 34588309. Bibcode2021PNAS..11810036Y. 
  48. Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S. (2012). "Electronics and optoelectronics of two-dimensional transition metal dichalcogenides". Nature Nanotechnology 7 (11): 699–712. doi:10.1038/nnano.2012.193. PMID 23132225. Bibcode2012NatNa...7..699W. http://infoscience.epfl.ch/record/182177. 
  49. 49.0 49.1 Ganatra, R.; Zhang, Q. (2014). "Few-Layer MoS2: A Promising Layered Semiconductor". ACS Nano 8 (5): 4074–99. doi:10.1021/nn405938z. PMID 24660756. 
  50. Zhu, Wenjuan; Low, Tony; Lee, Yi-Hsien; Wang, Han; Farmer, Damon B.; Kong, Jing; Xia, Fengnian; Avouris, Phaedon (2014). "Electronic transport and device prospects of monolayer molybdenum disulphide grown by chemical vapour deposition". Nature Communications 5: 3087. doi:10.1038/ncomms4087. PMID 24435154. Bibcode2014NatCo...5.3087Z. 
  51. Hong, Jinhua; Hu, Zhixin; Probert, Matt; Li, Kun; Lv, Danhui; Yang, Xinan; Gu, Lin; Mao, Nannan et al. (2015). "Exploring atomic defects in molybdenum disulphide monolayers". Nature Communications 6: 6293. doi:10.1038/ncomms7293. PMID 25695374. Bibcode2015NatCo...6.6293H. 
  52. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, J.; F.; Wang, Feng (2010). "Emerging Photoluminescence in Monolayer MoS2". Nano Letters 10 (4): 1271–1275. doi:10.1021/nl903868w. PMID 20229981. Bibcode2010NanoL..10.1271S. 
  53. 53.0 53.1 Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. (2011). "Single-layer MoS2 transistors". Nature Nanotechnology 6 (3): 147–150. doi:10.1038/nnano.2010.279. PMID 21278752. Bibcode2011NatNa...6..147R. http://infoscience.epfl.ch/record/164049. 
  54. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. (2013). "Ultrasensitive photodetectors based on monolayer MoS2". Nature Nanotechnology 8 (7): 497–501. doi:10.1038/nnano.2013.100. PMID 23748194. Bibcode2013NatNa...8..497L. http://infoscience.epfl.ch/record/183895. 
  55. Rao, C. N. R.; Ramakrishna Matte, H. S. S.; Maitra, U. (2013). "Graphene Analogues of Inorganic Layered Materials". Angew. Chem. 52 (50): 13162–85. doi:10.1002/anie.201301548. PMID 24127325. 
  56. Bessonov, A. A.; Kirikova, M. N.; Petukhov, D. I.; Allen, M.; Ryhänen, T.; Bailey, M. J. A. (2014). "Layered memristive and memcapacitive switches for printable electronics". Nature Materials 14 (2): 199–204. doi:10.1038/nmat4135. PMID 25384168. Bibcode2015NatMa..14..199B. 
  57. "Ultrasensitive biosensor from molybdenite semiconductor outshines graphene". R&D Magazine. 4 September 2014. http://www.rdmag.com/news/2014/09/ultrasensitive-biosensor-molybdenite-semiconductor-outshines-graphene?et_cid=4135513&et_rid=677699018&location=top. 
  58. Akinwande, Deji; Petrone, Nicholas; Hone, James (2014-12-17). "Two-dimensional flexible nanoelectronics". Nature Communications 5: 5678. doi:10.1038/ncomms6678. PMID 25517105. Bibcode2014NatCo...5.5678A. 
  59. Chang, Hsiao-Yu; Yogeesh, Maruthi Nagavalli; Ghosh, Rudresh; Rai, Amritesh; Sanne, Atresh; Yang, Shixuan; Lu, Nanshu; Banerjee, Sanjay Kumar et al. (2015-12-01). "Large-Area Monolayer MoS2 for Flexible Low-Power RF Nanoelectronics in the GHz Regime". Advanced Materials 28 (9): 1818–1823. doi:10.1002/adma.201504309. PMID 26707841. 
  60. Wachter, Stefan; Polyushkin, Dmitry K.; Bethge, Ole; Mueller, Thomas (2017-04-11). "A microprocessor based on a two-dimensional semiconductor" (in en). Nature Communications 8: 14948. doi:10.1038/ncomms14948. ISSN 2041-1723. PMID 28398336. Bibcode2017NatCo...814948W. 
  61. "Memtransistors advance neuromorphic computing | NextBigFuture.com" (in en-US). NextBigFuture.com. 2018-02-24. https://www.nextbigfuture.com/2018/02/memtransistors-advance-neuromorphic-computing.html. 
  62. Mak, Kin Fai; He, Keliang; Shan, Jie; Heinz, Tony F. (2012). "Control of valley polarization in monolayer MoS2 by optical helicity" (in en). Nature Nanotechnology 7 (8): 494–498. doi:10.1038/nnano.2012.96. PMID 22706698. Bibcode2012NatNa...7..494M. https://www.nature.com/articles/nnano.2012.96. 
  63. Wu, Zefei; Zhou, Benjamin T.; Cai, Xiangbin; Cheung, Patrick; Liu, Gui-Bin; Huang, Meizhen; Lin, Jiangxiazi; Han, Tianyi et al. (2019-02-05). "Intrinsic valley Hall transport in atomically thin MoS2". Nature Communications 10 (1): 611. doi:10.1038/s41467-019-08629-9. PMID 30723283. Bibcode2019NatCo..10..611W. 
  64. Coxworth, Ben (September 25, 2014). "Metal-based graphene alternative "shines" with promise". Gizmag. http://www.gizmag.com/molybdenum-di-sulphide-metal-graphene/33980. 
  65. Taniguchi, Kouji; Matsumoto, Akiyo; Shimotani, Hidekazu; Takagi, Hidenori (July 23, 2012). "Electric-field-induced superconductivity at 9.4 K in a layered transition metal disulphide MoS2". Applied Physics Letters 101 (4): 042603. doi:10.1063/1.4740268. Bibcode2012ApPhL.101d2603T. https://aip.scitation.org/doi/abs/10.1063/1.4740268. 

External links