Molybdenum disulfide

Molybdenum disulfide
Names
IUPAC name
Molybdenum disulfide
Identifiers
1317-33-5 Yes
ChEBI CHEBI:30704 Yes
ChemSpider 14138 Yes
Jmol-3D images Image
PubChem 14823
RTECS number QA4697000
Properties
MoS
2
Molar mass 160.07 g/mol[1]
Appearance black/lead-gray solid
Density 5.06 g/cm3[1]
Melting point 1,185 °C (2,165 °F; 1,458 K) decomposes
insoluble[1]
Solubility decomposed by aqua regia, hot sulfuric acid, nitric acid
insoluble in dilute acids
Band gap 1.23 eV (2H)[2]
Structure
Crystal structure hP6, space group P6
3
/mmc, No 194 (2H)

hR9, space group R3m, No 160 (3R)[3]

Lattice constant a = 0.3161 nm (2H), 0.3163 nm (3R), c = 1.2295 nm (2H), 1.837 (3R)
Trigonal prismatic (MoIV)
Pyramidal (S2−)
Hazards
MSDS External MSDS
EU Index not listed
Related compounds
Other anions
Molybdenum(IV) oxide
Other cations
Tungsten disulfide
Related lubricants
Graphite
Except where noted otherwise, data is given for materials in their standard state (at 25 °C (77 °F), 100 kPa)
  verify (what is: Yes/?)
Infobox references

Molybdenum disulfide is the inorganic compound with the formula MoS
2
.

The compound is classified as a metal dichalcogenide. It is a silvery black solid that occurs as the mineral molybdenite, the principal ore for molybdenum.[4] MoS
2
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 solid lubricant because of its low friction properties and robustness.

Production

Molybdenite

Molybdenite ore is processed by flotation to give relatively pure MoS
2
, the main contaminant being carbon. MoS
2
also arises by the thermal treatment of virtually all molybdenum compounds with hydrogen sulfide or elemental sulfur. It can also be produced by metathesis reactions from molybdenum pentachloride.[5]

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.[6]

MoS
2
usually consists of a mixture of two major polytypes of similar structure, 2H and 3R, with the former being most abundant.[3] In 2H-MoS
2
, each Mo(IV) center occupies a trigonal prismatic coordination sphere that is bound to six sulfide ligands. Each sulfur centre is pyramidal and is connected to three Mo centres. In this way, the trigonal prisms are interconnected to give a layered structure, wherein molybdenum atoms are sandwiched between layers of sulfur atoms.[7] Because of the weak van der Waals interactions between the sheets of sulfide atoms, MoS
2
has a low coefficient of friction, producing its lubricating properties. Other layered inorganic materials exhibit lubricating properties (collectively known as solid lubricants (or dry lubricants)) including graphite, which requires volatile additives and hexagonal boron nitride.[8]

While bulk material forms a layered structure, nanoparticulate MoS
2
forms fullerene and nanotubular microstructures.[9]

Bulk MoS
2
is a diamagnetic, indirect bandgap semiconductor similar to silicon, with a bandgap of 1.23 eV.[2] The natural amorphous form is known as the rarer mineral jordisite.

Two-dimensional, single- or few-layer MoS
2
, is a two-dimensional semiconductor, with the band structure very sensitive to strain.[10]

Chemical reactions

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

2 MoS
2
+ 9 O
2
→ 2 MoO
3
+ 4 SO
3

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

2 MoS
2
+ 7 Cl
2
→ 2 MoCl
5
+ 2 S
2
Cl
2

Molybdenum disulfide is a host for formation of intercalation compounds.[11] One example is lithiated material, Li
x
MoS
2
.[12] With butyl lithium, the product is LiMoS
2
.[4]

Applications

Lubricant

MoS
2
with particle sizes in the range of 1–100 µm is a common dry lubricant.[13] Few alternatives exist that confer high lubricity and stability at up to 350 °C in oxidizing environments. Sliding friction tests of MoS
2
using a pin on disc tester at low loads (0.1–2 N) give friction coefficient values of <0.1.[14][15]

MoS
2
is often a component of blends and composites that require low friction. 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, MoS
2
forms a composite with improved strength as well as reduced friction. Polymers filled with MoS
2
include nylon (with the 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 MoS
2
-based lubricants include two-stroke engines (e.g., motorcycle engines), bicycle coaster brakes, automotive CV and universal joints, ski waxes,[16] and even bullets.[17]

Petroleum refining

MoS
2
is employed as a cocatalyst for desulfurization in petrochemistry; e.g., hydrodesulfurization.[18] The effectiveness of the MoS
2
catalysts is enhanced by doping with small amounts of cobalt or nickel, supporting the intimate mixture on alumina. Such catalysts are generated in situ by treating molybdate/cobalt or nickel-impregnated alumina with H
2
S
or an equivalent reagent.

Water splitting

In 2014, a 30-year-old recipe was found where MoS
2
was used as a catalyst in the electrolysis of water.[19]

Research

Much research is focused on unusual morphologies of MoS2. Multilayer sheets are produced by liquid phase exfoliation.[20][21] Nanotubes and buckyball-like molecules composed of MoS
2
exhibit unusual tribological and electronic properties.[22] MoS
2
has been investigated as a component of photoelectrochemical (e.g. for photocatalytic hydrogen production) applications and for microelectronics applications.[23]

MoS
2
and other transition metal dichalcogenides form bulk crystals composed of two-dimensional layers stacked in the vertical direction. Such two-dimensional layers are similar in form to graphene and express diverse electronic and optical properties[24] that can differ from those in bulk. Whereas bulk MoS
2
has an indirect band gap of 1.2 eV, MoS
2
monolayers have a direct 1.8 eV electronic bandgap,[25] allowing the production of switchable transistors[23] and sensitive photodetectors.[26] As a transition metal di-chalcogenide, MoS
2
possesses some of graphene's desirable qualities (such as mechanical strength and electrical conductivity), and can emit light, opening possible applications such as photodetectors.[27]

The sulfur group on MoS
2
surface interacts with noble metals, including gold. The bond between MoS
2
and gold nanostructures was found to act as a highly coupled gate capacitor with a reduced carrier-transport thermal-barrier and increased thermal conductivity.[28][29]

Calculations indicate that MoS
2
transistors would consume on the order of 100,000 times less energy than silicon transistors while in the "off" state. For molybdenum disulfide sheets, Raman spectra cannot be fully explained as phenomena involving certain characteristic vibrations of the crystal network.[30]

MoS
2
nanoflakes can be used for solution-processed fabrication of layered memristive and memcapacitive devices through engineering a MoO
x
/MoS
2
heterostructure sandwiched between silver electrodes.[31] The MoS
2
-based memristors are mechanically flexible, optically transparent and can be produced at low cost.

In 2014 a two-dimensional, MoS
2
-based semiconductor material for biosensing was announced. Compared to graphene, it offers higher sensitivity, better scalability and lends itself to high-volume manufacturing. The wide bandgap pf MoS
2
prevents leakage and results in more sensitive and accurate readings.[32]

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.[32]

The demonstration biosensors provide protein sensing with a sensitivity of 196 even at 100 femtomolar, similar to one drop of milk dissolved in one hundred tons of water. A MoS2-based pH sensor achieved sensitivity of 713 for a pH change by one unit over a pH range of 3–9.[32]

See also

Wikimedia Commons has media related to Molybdenum disulfide.

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. doi:10.1103/PhysRevB.51.17085.
  3. 3.0 3.1 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 Structural Science 39 (4): 404. doi:10.1107/S0108768183002645.
  4. 4.0 4.1 Sebenik, Roger F. et al. (2005) "Molybdenum and Molybdenum Compounds" in Ullmann's Encyclopedia of Chemical Technology. Wiley-VCH, Weinheim. doi: 10.1002/14356007.a16_655
  5. Murphy, Donald W.; Interrante, Leonard V.; Kaner; Mansuktto (1995). "Metathetical Precursor Route to Molybdenum Disulfide". Inorganic Syntheses. Inorganic Syntheses 30: 33–37. doi:10.1002/9780470132616.ch8. ISBN 9780470132616.
  6. Hong, J.; Hu, Z.; Probert, M.; Li, K.; Lv, D.; Yang, X.; Gu, L.; Mao, N.; Feng, Q.; Xie, L.; Zhang, J.; Wu, D.; Zhang, Z.; Jin, C.; Ji, W.; Zhang, X.; Yuan, J.; Zhang, Z. (2015). "Exploring atomic defects in molybdenum disulphide monolayers". Nature Communications 6: 6293. Bibcode:2015NatCo...6E6293H. doi:10.1038/ncomms7293. PMC 4346634. PMID 25695374.
  7. Wells, A.F. (1984). Structural Inorganic Chemistry. Oxford: Clarendon Press. ISBN 0-19-855370-6.
  8. Thorsten Bartels et al. (2002). "Lubricants and Lubrication". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley VCH. doi:10.1002/14356007.a15_423.
  9. Tenne, R.; Redlich, M. (2010). "Recent progress in the research of inorganic fullerene-like nanoparticles and inorganic nanotubes". Chemical Society Reviews 39 (5): 1423. doi:10.1039/B901466G. PMID 20419198.
  10. Amorim, B.; De Juan, F.; Grushin, A. G.; Guinea, F.; Gutiérrez-Rubio, A.; Ochoa, H.; Parente, V.; Roldán, R.; San-José, P.; Schiefele, J.; Sturla, M.; Vozmediano, M. A. H. et al. (2 March 2015). "Novel effects of strains in graphene and other two dimensional materials" 1503. arxiv.org. p. 747. arXiv:1503.00747. Bibcode:2015arXiv150300747A. Retrieved 12 March 2015.
  11. Benavente, E.; Santa Ana, M. A.; Mendizabal, F.; Gonzalez, G. (2002). "Intercalation chemistry of molybdenum disulfide". Coordination Chemistry Reviews 224: 87–109. doi:10.1016/S0010-8545(01)00392-7.
  12. Müller-Warmuth, W. and Schöllhorn, R. (1994). Progress in intercalation research. Springer. ISBN 0-7923-2357-2.
  13. Claus, F. L. (1972) Solid Lubricants and Self-Lubricating Solids, Academic Press, New York.
  14. Miessler, G. L. and Tarr, D. A. (2004). Inorganic Chemistry, 3rd Ed. Pearson/Prentice Hall publisher. ISBN 0-13-035471-6.
  15. Shriver, D. F.; Atkins, P. W.; Overton, T. L.; Rourke, J. P.; Weller, M. T.; Armstrong, F. A. (2006). Inorganic Chemistry. New York: W. H. Freeman. ISBN 0-7167-4878-9.
  16. "On dry lubricants in ski waxes" (PDF). Swix Sport AX. Retrieved 2011-01-06.
  17. "Barrels retain accuracy longer with Diamond Line". Norma. Retrieved 2009-06-06.
  18. Topsøe, H.; Clausen, B. S.; Massoth, F. E. (1996). Hydrotreating Catalysis, Science and Technology. Berlin: Springer-Verlag.
  19. Kibsgaard, Jakob; Jaramillo, Thomas F.; Besenbacher, Flemming (2014). "Building an appropriate active-site motif into a hydrogen-evolution catalyst with thiomolybdate [Mo3S13]2− clusters". Nature Chemistry 6 (3): 248–253. Bibcode:2014NatCh...6..248K. doi:10.1038/nchem.1853. PMID 24557141.
  20. Novel exfoliation method for molybdenum disulfide. Materials Today. 6 January 2014
  21. A new twist to sodium ion battery technology. Materials Today. 30 January 2014
  22. Feldman, Y.; Wasserman, E.; Srolovitz, D. J.; Tenne, R. (1995). "High-Rate, Gas-Phase Growth of MoS2 Nested Inorganic Fullerenes and Nanotubes". Science 267 (5195): 222–225. Bibcode:1995Sci...267..222F. doi:10.1126/science.267.5195.222. PMID 17791343.
  23. 23.0 23.1 Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. (2011). "Single-layer MoS2 transistors". Nature Nanotechnology 6 (3): 147–150. Bibcode:2011NatNa...6..147R. doi:10.1038/nnano.2010.279. PMID 21278752.
  24. 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.
  25. 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. Bibcode:2010NanoL..10.1271S. doi:10.1021/nl903868w. PMID 20229981.
  26. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. (2013). "Ultrasensitive photodetectors based on monolayer MoS2". Nature Nanotechnology 8 (7): 497–501. Bibcode:2013NatNa...8..497L. doi:10.1038/nnano.2013.100. PMID 23748194.
  27. Coxworth, Ben (September 25, 2014). "Metal-based graphene alternative "shines" with promise". Gizmag. Retrieved September 30, 2014.
  28. Sreeprasad, T. S.; Nguyen, P.; Kim, N.; Berry, V. (2013). "Controlled, Defect-Guided, Metal-Nanoparticle Incorporation onto MoS2via Chemical and Microwave Routes: Electrical, Thermal, and Structural Properties". Nano Letters: 130813110201000. doi:10.1021/nl402278y.
  29. "Another breakthrough in replacing silicon in transistors". KurzweilAI. Retrieved 2013-09-11.
  30. "Material rivaling graphene may one day be mined from rocks". R&D Magazine. 12 March 2014.
  31. 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. Bibcode:2015NatMa..14..199B. doi:10.1038/nmat4135. PMID 25384168.
  32. 32.0 32.1 32.2 "Ultrasensitive biosensor from molybdenite semiconductor outshines graphene". R&D Magazine. 4 September 2014.