Inverse vulcanization

From Wikipedia, the free encyclopedia
Preparation of poly(sulfur-co-1,3-diisopropylbenzene)

Inverse vulcanization is a solvent-free copolymerization process that produces polymers containing chains of sulfur atoms.[1] These materials exploit the wide availability of high purity sulfur, a by-product from the crude oil and natural gas refining processes. The inverse vulcanization promises to produce a low cost and chemically stable sulfur-rich material, which has diverse potential applications.

Synthesis[]

Like thiokols and inverse vulcanization exploits the tendency of sulfur catenate. The polymers produced by inverse vulcanization consist of long sulfur linear chains interspersed with organic linkers. In contrast, traditional sulfur vulcanization produces a cross-linked material with short sulfur bridges, even made by one or two sulfur atoms.

The polymerization process begins with heating of elemental sulfur above its melting point (115.21 °C), in order to favor the ring-opening polymerization process (ROP) of the S8 monomer, occurring at 159°C. As a result, the liquid sulfur is constituted by linear polysulfide chains with diradical ends, which can be easily bridged together with a modest amount of small dienes, such as (DIB),[1] ,[2] limonene,[3] divinylbenzene (DVB),[4] dicyclopentadiene,[5] styrene,[6] 4-vinylpyridine,[7] cycloalkene[8] and ethylidene norbornene,[9] or longer organic molecules as polybenzoxazines,[10] squalene[11] and triglyceride.[12] Chemically, the diene carbon-carbon double bond (C=C) of the substitutional group disappears, forming the carbon-sulfur single bond (C-S) which binds together the sulfur linear chains. The huge advantage of such a polymerization is the absence of a solvent (solvent-free): sulphur acts as comonomer and solvent. This makes the process highly scalable at the industrial scale. As evidence, the kilogram-scale synthesis of the poly(S-r-DIB) has been already correctly accomplished.[13]

Inverse vulcanization process of sulfur through .

Products. Characterization and properties[]

Physical appearance of poly(sulfur-random-

Vibrational spectroscopy was performed to investigate the chemical structure of the copolymers: the presence of the C-S bonds was detected through Infrared or Raman spectroscopies.[14] The high amount of S-S bonds makes the copolymer highly IR inactive in the near and mid-infrared spectrum. As a consequence, sulfur-rich materials made via inverse vulcanization are characterized by a high refractive index (n~1.8), whose value depends again upon the composition and crosslinking species.[15] As shown by the thermogravimetric analysis (TGA), the copolymer thermal stability increases with the amount of the added crosslinker; in any case, all the tested compositions degrade above 222 °C.[2][4]

Focusing on the mechanical features, the copolymer behavior, included the glass-transition temperature, depends upon composition and crosslinking species. For given comonomers, the behavior of the copolymers as a function of the temperature depends on the chemical composition, for example, the poly(sulfur-random-divinylbenzene) behaves as a plastomer for a diene content between 15-25%wt, and as a viscous resin with the 30–35%wt of DVB. On the other hand, the poly(sulfur-random-) acts as thermoplastic at 15–25%wt of DIB, while it becomes a thermoplastic-thermosetting polymer for a diene concentration of 30-35%wt.[16] The possibility to break and reform the chemical bonds along the polysulfides chains (S-S) allows to repair the copolymer by simply heating above 100 °C. This feature increases the reforming and recyclability of the high molecular weight copolymer.[17] The high amount of S-S bonds makes the copolymer highly IR inactive in the near and mid-infrared spectrum. As a consequence, sulfur-rich materials made via inverse vulcanization are characterized by a high refractive index (n~1.8), whose value depends again upon the composition and crosslinking species.[18]

Potential applications[]

The sulfur-rich copolymers made via inverse vulcanization could in principle find diverse applications thanks to the simple synthesis process and their thermoplasticity.

Lithium-sulfur batteries[]

This new way of sulfur processing has been exploited for the cathode preparation of long-cycling lithium-sulfur batteries. Such electrochemical systems are characterized by a greater energy density than commercial Li-ion batteries, but they are not stable for a long service life. Simmonds et al.[19] first demonstrated an improved capacity retention for over 500 cycles with an inverse vulcanization copolymer, suppressing the typical capacity fading of sulfur-polymer composites. Indeed, the poly(sulfur-random-1,3-diisopropenylbenzene), briefly defined as poly(S-r-DIB), showed a higher composition homogeneity compared with other cathodic materials, together with a greater sulfur retention and an enhanced adjustment of the polysulfides volume variations. These advantages made it possible to assembly a stable and durable Li-S cell. After that, other copolymers via inverse vulcanization were synthetized and tested inside these electrochemical devices, again providing exceptional stability over cycles.

Battery performances
Cathode Date Source Specific Capacity after cycling
Poly(sulfur-random-1,3-diisopropylbenzene) 2014 University of Arizona[19] 1005 mA⋅h/g after 100 cycles (at 0.1 C)
Poly(sulfur-random-) 2015 University of Arizona[2] 800 mA⋅h/g after 300 cycles (at 0.2 C)
Poly(sulfur-random-divinylbenzene) 2016 University of the Basque Country[20] 700 mA⋅h/g after 500 cycles (at 0.25 C)
Poly(sulfur-random-diallyl disulfide) 2016 University of the Basque Country[21] 616 mA⋅h/g after 200cycles (at 0.2 C)
Poly(sulfur-random-bismaleimide-divinylbenzene) 2016 Istanbul Technical University[22] 400 mA⋅h/g after 50 cycles (at 0.1 C)
Poly(sulfur-random-styrene) 2017 University of Arizona[6] 485 mA⋅h/g after 1000 cycles (at 0.2 C)

In order to overcome the great disadvantage related to the materials' low electrical conductivity (1015–1016 Ω·cm),[16] researchers started to add special carbon-based particles, to increase the electron transport inside the copolymer. Furthermore, such carbonaceous additives improve the polysulfides retention at the cathode through the polysulfides-capturing effect, increasing the battery performances. Examples of employed nanostructures are long carbon nanotubes,[23] graphene[11] and .[24]

Mercury capture[]

The new materials could be used to remove toxic metals from soil or water. However, pure sulfur cannot be employed to manufacture a functional filter, because of its low mechanical properties. Therefore, inverse vulcanization was investigated to produce porous materials, in particular for the mercury capturing process. The liquid metal binds together with the sulfur-rich copolymer, remaining mostly inside the filter. Mercury is dangerous for the environment and highly toxic for humans, making its removal fundamental.[25][26][27]

Infrared transmission[]

Polymers are used for IR optical applications because of their low refractive index (n~1.5-1.6); their poor transparency towards the infrared radiation limits their exploitation in this sector. On the other hand, inorganic materials (n~2-5) are characterized by high-cost and complex processability, detrimental factors for the large-scale production.

Sulfur-rich copolymers, made via inverse vulcanization, represent a great alternative thanks to the simple manufacturing process, low cost reagents and high refractive index. As mentioned before, the latter depends upon the S-S bonds concentration, leading to the possibility of tuning the optical properties of the material by simply modifying the chemical formulation. Such possibility of changing the material refractive index to fulfil the specific application requirements, makes these copolymers applicable in the military, civil or medical fields.[28][29][30][31]

Others[]

The inverse vulcanization process can also be employed for the synthesis of activated carbon with narrow pore-size distributions. The sulfur-rich copolymer acts here as a template where the carbons are produced. The final material is doped with sulfur and exhibits a micro-porous network and high gas selectivity. Therefore, inverse vulcanization could be also applied in the gas separation sector.[32]

See also[]

  • Sulfur
  • Free-radical polymerization
  • Lithium-sulfur batteries

References[]

  1. ^ a b Chung, Woo Jin; Griebel, Jared J.; Kim, Eui Tae; Yoon, Hyunsik; Simmonds, Adam G.; Ji, Hyun Jun; Dirlam, Philip T.; Glass, Richard S.; Wie, Jeong Jae; Nguyen, Ngoc A.; Guralnick, Brett W.; Park, Jungjin; Somogyi, Árpád; Theato, Patrick; Mackay, Michael E.; Sung, Yung-Eun; Char, Kookheon; (14 April 2013). "The use of elemental sulfur as an alternative feedstock for polymeric materials". Nature Chemistry. 5 (6): 518–524. doi:10.1038/NCHEM.1624. PMID 23695634.
  2. ^ a b c Dirlam, Philip T.; Simmonds, Adam G.; Kleine, Tristan S.; Nguyen, Ngoc A.; Anderson, Laura E.; Klever, Adam O.; Florian, Alexander; Costanzo, Philip J.; Theato, Patrick; Mackay, Michael E.; Glass, Richard S.; Char, Kookheon; Pyun, Jeffrey (2015). "Inverse vulcanization of elemental sulfur with 1,4-diphenylbutadiyne for cathode materials in Li–S batteries". RSC Advances. 5 (31): 24718–24722. doi:10.1039/c5ra01188d.
  3. ^ Crockett, Michael P.; Evans, Austin M.; Worthington, Max J. H.; Albuquerque, Inês S.; Slattery, Ashley D.; Gibson, Christopher T.; Campbell, Jonathan A.; Lewis, David A.; Bernardes, Gonçalo J. L.; Chalker, Justin M. (26 January 2016). "Sulfur-Limonene Polysulfide: A Material Synthesized Entirely from Industrial By-Products and Its Use in Removing Toxic Metals from Water and Soil". Angewandte Chemie International Edition. 55 (5): 1714–1718. doi:10.1002/anie.201508708. PMC 4755153. PMID 26481099.
  4. ^ a b Salman, Mohamed Khalifa; Karabay, Baris; Karabay, Lutfiye Canan; Cihaner, Atilla (20 July 2016). "Elemental sulfur-based polymeric materials: Synthesis and characterization". Journal of Applied Polymer Science. 133 (28). doi:10.1002/app.43655.
  5. ^ Parker, D. J.; Jones, H. A.; Petcher, S.; Cervini, L.; Griffin, J. M.; Akhtar, R.; Hasell, T. (2017). "Low cost and renewable sulfur-polymers by inverse vulcanisation, and their potential for mercury capture" (PDF). Journal of Materials Chemistry A. 5 (23): 11682–11692. doi:10.1039/C6TA09862B.
  6. ^ a b Zhang, Yueyan; Griebel, Jared J.; Dirlam, Philip T.; Nguyen, Ngoc A.; Glass, Richard S.; Mackay, Michael E.; Char, Kookheon; Pyun, Jeffrey (1 January 2017). "Inverse vulcanization of elemental sulfur and styrene for polymeric cathodes in Li-S batteries". Journal of Polymer Science Part A: Polymer Chemistry. 55 (1): 107–116. doi:10.1002/pola.28266.
  7. ^ Berk, Hasan; Balci, Burcu; Ertan, Salih; Kaya, Murat; Cihaner, Atilla (June 2019). "Functionalized polysulfide copolymers with 4-vinylpyridine via inverse vulcanization". Materials Today Communications. 19: 336–341. doi:10.1016/j.mtcomm.2019.02.014.
  8. ^ Omeir, Meera Y.; Wadi, Vijay S.; Alhassan, Saeed M. (January 2020). "Inverse vulcanized sulfur–cycloalkene copolymers: Effect of ring size and unsaturation on thermal properties". Materials Letters. 259: 126887. doi:10.1016/j.matlet.2019.126887.
  9. ^ Smith, Jessica A.; Wu, Xiaofeng; Berry, Neil G.; Hasell, Tom (15 August 2018). "High sulfur content polymers: The effect of crosslinker structure on inverse vulcanization". Journal of Polymer Science Part A: Polymer Chemistry. 56 (16): 1777–1781. doi:10.1002/pola.29067. PMC 6175008. PMID 30333680.
  10. ^ Arslan, Mustafa; Kiskan, Baris; Yagci, Yusuf (22 January 2016). "Combining Elemental Sulfur with Polybenzoxazines via Inverse Vulcanization". Macromolecules. 49 (3): 767–773. doi:10.1021/acs.macromol.5b02791.
  11. ^ a b Sahu, Tuhin Subhra; Choi, Sinho; Jaumaux, Pauline; Zhang, Jinqiang; Wang, Chengyin; Zhou, Dong; Wang, Guoxiu (April 2019). "Squalene-derived sulfur-rich copolymer@ 3D graphene-carbon nanotube network cathode for high-performance lithium-sulfur batteries". Polyhedron. 162: 147–154. doi:10.1016/j.poly.2019.01.068.
  12. ^ Tikoalu, Alfrets D.; Lundquist, Nicholas A.; Chalker, Justin M. (13 February 2020). "Mercury Sorbents Made By Inverse Vulcanization of Sustainable Triglycerides: The Plant Oil Structure Influences the Rate of Mercury Removal from Water". Advanced Sustainable Systems. 4 (3): 1900111. doi:10.1002/adsu.201900111.
  13. ^ Griebel, Jared J.; Li, Guoxing; Glass, Richard S.; Char, Kookheon; Pyun, Jeffrey (15 January 2015). "Kilogram scale inverse vulcanization of elemental sulfur to prepare high capacity polymer electrodes for Li-S batteries". Journal of Polymer Science Part A: Polymer Chemistry. 53 (2): 173–177. doi:10.1002/pola.27314.
  14. ^ Bastian, Ernest J.; Martin, R. Bruce (April 1973). "Disulfide vibrational spectra in the sulfur-sulfur and carbon-sulfur stretching region". The Journal of Physical Chemistry. 77 (9): 1129–1133. doi:10.1021/j100628a010.
  15. ^ Griebel, Jared J.; Namnabat, Soha; Kim, Eui Tae; Himmelhuber, Roland; Moronta, Dominic H.; Chung, Woo Jin; Simmonds, Adam G.; Kim, Kyung-Jo; van der Laan, John; Nguyen, Ngoc A.; Dereniak, Eustace L.; Mackay, Michael E.; Char, Kookheon; Glass, Richard S.; Norwood, Robert A.; Pyun, Jeffrey (May 2014). "New Infrared Transmitting Material via Inverse Vulcanization of Elemental Sulfur to Prepare High Refractive Index Polymers". Advanced Materials. 26 (19): 3014–3018. doi:10.1002/adma.201305607. PMID 24659231.
  16. ^ a b Diez, Sergej; Hoefling, Alexander; Theato, Patrick; Pauer, Werner (15 February 2017). "Mechanical and Electrical Properties of Sulfur-Containing Polymeric Materials Prepared via Inverse Vulcanization". Polymers. 9 (12): 59. doi:10.3390/polym9020059. PMC 6432436. PMID 30970741.
  17. ^ Chalker, Justin M.; Worthington, Max J. H.; Lundquist, Nicholas A.; Esdaile, Louisa J. (20 May 2019). "Synthesis and Applications of Polymers Made by Inverse Vulcanization". Topics in Current Chemistry. 377 (3): 16. doi:10.1007/s41061-019-0242-7. PMID 31111247. S2CID 160013607.
  18. ^ Griebel, Jared J.; Namnabat, Soha; Kim, Eui Tae; Himmelhuber, Roland; Moronta, Dominic H.; Chung, Woo Jin; Simmonds, Adam G.; Kim, Kyung-Jo; van der Laan, John; Nguyen, Ngoc A.; Dereniak, Eustace L.; Mackay, Michael E.; Char, Kookheon; Glass, Richard S.; Norwood, Robert A.; Pyun, Jeffrey (May 2014). "New Infrared Transmitting Material via Inverse Vulcanization of Elemental Sulfur to Prepare High Refractive Index Polymers". Advanced Materials. 26 (19): 3014–3018. doi:10.1002/adma.201305607. PMID 24659231.
  19. ^ a b Simmonds, Adam G.; Griebel, Jared J.; Park, Jungjin; Kim, Kwi Ryong; Chung, Woo Jin; Oleshko, Vladimir P.; Kim, Jenny; Kim, Eui Tae; Glass, Richard S.; Soles, Christopher L.; Sung, Yung-Eun; Char, Kookheon; Pyun, Jeffrey (20 February 2014). "Inverse Vulcanization of Elemental Sulfur to Prepare Polymeric Electrode Materials for Li–S Batteries". ACS Macro Letters. 3 (3): 229–232. doi:10.1021/mz400649w.
  20. ^ Gomez, Iñaki; Mecerreyes, David; Blazquez, J. Alberto; Leonet, Olatz; Ben Youcef, Hicham; Li, Chunmei; Gómez-Cámer, Juan Luis; Bondarchuk, Oleksandr; Rodriguez-Martinez, Lide (October 2016). "Inverse vulcanization of sulfur with divinylbenzene: Stable and easy processable cathode material for lithium-sulfur batteries". Journal of Power Sources. 329: 72–78. doi:10.1016/j.jpowsour.2016.08.046.
  21. ^ Gomez, Iñaki; Leonet, Olatz; Blazquez, J. Alberto; Mecerreyes, David (20 December 2016). "Inverse Vulcanization of Sulfur using Natural Dienes as Sustainable Materials for Lithium-Sulfur Batteries". ChemSusChem. 9 (24): 3419–3425. doi:10.1002/cssc.201601474. PMID 27910220.
  22. ^ Arslan, Mustafa; Kiskan, Baris; Cengiz, Elif Ceylan; Demir-Cakan, Rezan; Yagci, Yusuf (July 2016). "Inverse vulcanization of bismaleimide and divinylbenzene by elemental sulfur for lithium sulfur batteries". European Polymer Journal. 80: 70–77. doi:10.1016/j.eurpolymj.2016.05.007.
  23. ^ Tiwari, Vimal K.; Song, Hyeonjun; Oh, Yeonjae; Jeong, Youngjin (March 2020). "Synthesis of sulfur-co-polymer/porous long carbon nanotubes composite cathode by chemical and physical binding for high performance lithium-sulfur batteries". Energy. 195: 117034. doi:10.1016/j.energy.2020.117034.
  24. ^ Choudhury, Soumyadip; Srimuk, Pattarachai; Raju, Kumar; Tolosa, Aura; Fleischmann, Simon; Zeiger, Marco; Ozoemena, Kenneth I.; Borchardt, Lars; Presser, Volker (2018). "Carbon onion/sulfur hybrid cathodes inverse vulcanization for lithium–sulfur batteries". Sustainable Energy & Fuels. 2 (1): 133–146. doi:10.1039/c7se00452d.
  25. ^ Crockett, Michael P.; Evans, Austin M.; Worthington, Max J. H.; Albuquerque, Inês S.; Slattery, Ashley D.; Gibson, Christopher T.; Campbell, Jonathan A.; Lewis, David A.; Bernardes, Gonçalo J. L.; Chalker, Justin M. (26 January 2016). "Sulfur-Limonene Polysulfide: A Material Synthesized Entirely from Industrial By-Products and Its Use in Removing Toxic Metals from Water and Soil". Angewandte Chemie International Edition. 55 (5): 1714–1718. doi:10.1002/anie.201508708. PMC 4755153. PMID 26481099.
  26. ^ Hasell, T.; Parker, D. J.; Jones, H. A.; McAllister, T.; Howdle, S. M. (2016). "Porous inverse vulcanised polymers for mercury capture". Chemical Communications. 52 (31): 5383–5386. doi:10.1039/c6cc00938g. PMID 26931278.
  27. ^ Parker, D. J.; Jones, H. A.; Petcher, S.; Cervini, L.; Griffin, J. M.; Akhtar, R.; Hasell, T. (2017). "Low cost and renewable sulfur-polymers by inverse vulcanisation, and their potential for mercury capture" (PDF). Journal of Materials Chemistry A. 5 (23): 11682–11692. doi:10.1039/c6ta09862b.
  28. ^ Baumgartner, Thomas; Jäkle, Frieder (19 December 2017). Main group strategies towards functional hybrid materials. Wiley. ISBN 9781119235972.
  29. ^ Griebel, Jared J.; Namnabat, Soha; Kim, Eui Tae; Himmelhuber, Roland; Moronta, Dominic H.; Chung, Woo Jin; Simmonds, Adam G.; Kim, Kyung-Jo; van der Laan, John; Nguyen, Ngoc A.; Dereniak, Eustace L.; Mackay, Michael E.; Char, Kookheon; Glass, Richard S.; Norwood, Robert A.; Pyun, Jeffrey (May 2014). "New Infrared Transmitting Material via Inverse Vulcanization of Elemental Sulfur to Prepare High Refractive Index Polymers". Advanced Materials. 26 (19): 3014–3018. doi:10.1002/adma.201305607. PMID 24659231.
  30. ^ Griebel, Jared J.; Nguyen, Ngoc A.; Namnabat, Soha; Anderson, Laura E.; Glass, Richard S.; Norwood, Robert A.; Mackay, Michael E.; Char, Kookheon; Pyun, Jeffrey (16 August 2015). "Dynamic Covalent Polymers via Inverse Vulcanization of Elemental Sulfur for Healable Infrared Optical Materials". ACS Macro Letters. 4 (9): 862–866. doi:10.1021/acsmacrolett.5b00502.
  31. ^ Kleine, Tristan S.; Nguyen, Ngoc A.; Anderson, Laura E.; Namnabat, Soha; LaVilla, Edward A.; Showghi, Sasaan A.; Dirlam, Philip T.; Arrington, Clay B.; Manchester, Michael S.; Schwiegerling, Jim; Glass, Richard S.; Char, Kookheon; Norwood, Robert A.; Mackay, Michael E.; Pyun, Jeffrey (23 September 2016). "High Refractive Index Copolymers with Improved Thermomechanical Properties via the Inverse Vulcanization of Sulfur and 1,3,5-Triisopropenylbenzene". ACS Macro Letters. 5 (10): 1152–1156. doi:10.1021/acsmacrolett.6b00602.
  32. ^ Bear, Joseph C.; McGettrick, James D.; Parkin, Ivan P.; Dunnill, Charles W.; Hasell, Tom (September 2016). "Porous carbons from inverse vulcanised polymers". Microporous and Mesoporous Materials. 232: 189–195. doi:10.1016/j.micromeso.2016.06.021.

External links[]

Retrieved from ""