Polyphosphazene

Polyphosphazene general structure
General structure of polyphosphazenes. Gray spheres represent any organic or inorganic group.

Polyphosphazenes include a wide range of hybrid inorganic-organic polymers with a number of different skeletal architectures that contain alternating phosphorus and nitrogen atoms.[1] Nearly all of these molecules contain two organic or organometallic side groups attached to each phosphorus atom. These include linear polymers with the formula (N=PR1R2)n, where R1 and R2 are organic or organometallic side groups. The linear polymers are the largest group, with the general structure shown schematically in the picture. Other known architectures are cyclolinear and cyclomatrix polymers in which small phosphazene rings are connected together by organic chain units. Other architectures are available, such as block copolymer, star, dendritic, or comb-type structures. More than 700 different polyphosphazenes are known, with different side groups (R) and different molecular architectures. Many of these polymers were first synthesized and studied in the research group of Harry R. Allcock at The Pennsylvania State University.[1][2][3][4][5]

Synthesis

The method of synthesis depends on the type of polyphosphazene. The most widely used method for linear polymers is based on a two-step process.[1][2][3][4] In the first step a cyclic small molecule phosphazene, known as hexachlorocyclotriphosphazene, with the formula (NPCl2)3, is heated in a sealed system at 250 °C to convert it to a long chain linear polymer with typically 15,000 or more repeating units. In the second step the chlorine atoms linked to phosphorus in the polymer are replaced by organic groups through reactions with alkoxides, aryloxides, amines or organometallic reagents. Because many different reagents can participate in this macromolecular substitution reaction, and because two or more different reagents may be used, a large number of different polymers can be produced, each with a different combination of properties. Variations to this process are possible using poly(dichlorophosphazene) made by condensation reactions.[6]

Polyphosphazene synthesis

Another synthetic process uses a living cationic polymerization that allows the formation of block copolymers or comb, star, or dendritic architectures.[7][8] Other synthetic methods include the condensation reactions of organic-substituted phosphoranimines.[9][10][11][12]

Cyclomatrix type polymers made by linking small molecule phosphazene rings together employ difunctional organic reagents to replace the chlorine atoms in (NPCl2)3, or the introduction of allyl or vinyl substituents, which are then polymerized by free-radical methods.[13] Such polymers may be useful as coatings or thermosetting resins, often prized for their thermal stability.

Properties and uses

The linear high polymers have the geometry shown in the picture. More than 700 different macromolecules that correspond to this structure are known with different side groups or combinations of different side groups. In these polymers the properties are controlled partly by the high flexibility of the backbone, its radiation resistance, high refractive index, ultraviolet and visible transparency, and its fire resistance. However, the side groups exert an equal or even greater influence on the properties since they impart properties such as hydrophobicity, hydrophilicity, color, useful biological properties such as bioerodibility, or ion transport properties to the polymers. Representative examples of these polymers are shown below. Polyphosphazene examples

Thermoplastics

The first stable thermoplastic poly(organophosphazenes), isolated in the mid 1960s by Allcock, Kugel, and Valan, were macromolecules with trifluoroethoxy, phenoxy, methoxy, ethoxy, or various amino side groups.[2][3][4] Of these early species, poly[bis(trifluoroethoxyphosphazene], [NP(OCH2CF3)2]n, has proved to be the subject of intense research due to its crystallinity, high hydrophobicity, biological compatibility, fire resistance, general radiation stability, and ease of fabrication into films, microfibers and nanofibers. It has also been a substrate for various surface reactions to immobilize biological agents. The polymers with phenoxy or amino side groups have also been studied in detail.

Phosphazene elastomers

The first large-scale commercial uses for linear polyphosphazenes were in the field of high technology elastomers, with a typical example containing a combination of trifluoroethoxy and longer chain fluoroalkoxy groups.[14][15][16][17] The mixture of two different side groups eliminates the crystallinity found in single-substituent polymers and allows the inherent flexibility and elasticity to become manifest. Glass transition temperatures as low as -60 °C are attainable, and properties such as oil-resistance and hydrophobicity are responsible for their utility in land vehicles and aerospace components. They have also been used in biostable biomedical devices.[18]

Other side groups, such as non-fluorinated alkoxy or oligo-alkyl ether units, yield hydrophilic or hydrophobic elastomers with glass transitions over a broad range from -100 °C to + 100 °C.[19] Polymers with two different aryloxy side groups have also been developed as elastomers for fire-resistance as well as thermal and sound insulation applications.

Polymer electrolytes

Linear polyphosphazenes with oligo-ethyleneoxy side chains are gums that are good solvents for salts such as lithium triflate. These solutions function as electrolytes for lithium ion transport, and they have been the focus of much research designed to incorporate them into fire-resistant rechargeable lithium-ion polymer battery.[20][21][22] The same polymers are also of interest as the electrolyte in experimental dye-sensitized solar cells.[23] Other polyphosphazenes with sulfonated aryloxy side groups are proton conductors of interest for use in the membranes of proton exchange membrane fuel cells.[24]

Hydrogels

Water-soluble poly(organophosphazenes) with oligo-ethyleneoxy side chains can be cross-linked by gamma-radiation techniques. The cross-linked polymers absorb water to form hydrogels which are responsive to temperature changes, expanding to a limit defined by the cross-link density below a critical solution temperature, but contracting above that temperature. This is the basis of controlled permeability membranes. Other polymers with both oligo-ethyleneoxy and carboxyphenoxy side groups expand in the presence of monovalent cations but contract in the presence of di- or tri-valent cations, which form ionic cross-links.[25][26][27][28][29] Phosphazene hydrogels have been utilized for controlled drug release and other medical applications.[26]

Bioerodible polyphosphazenes

The ease with which properties can be controlled and fine-tuned by the linkage of different side groups to polyphosphazene chains has prompted major efforts to address biomedical materials challenges using these polymers. Different polymers have been studied as macromolecular drug carriers, as membranes for the controlled delivery of drugs, as biostable elastomers, and especially as tailored bioerodible materials for the regeneration of living bone.[30][31][32][33] An advantage for this last application is that poly(dichlorophosphazene) reacts with amino acid ethyl esters (such as ethyl glycinate or the corresponding ethyl esters of numerous other amino acids) through the amino terminus to form polyphosphazenes with amino acid ester side groups. These polymers hydrolyze slowly to a near-neutral, pH-buffered solution of the amino acid, ethanol, phosphate, and ammonium ion. The speed of hydrolysis depends on the amino acid ester, with half-lives that vary from weeks to months depending on the structure of the amino acid ester. Nanofibers and porous constructs of these polymers assist osteoblast replication and accelerate the repair of bone in animal model studies.

Commercial aspects

The cyclic trimer, (NPCl2)3, is commercially available and has formed the starting point for most commercial developments. Prominent among these developments has been the high performance elastomers known as PN-F or Eypel-F, which have been manufactured for seals, O-rings, and dental devices. An aryloxy-substituted polymer has also been developed as a fire resistant expanded foam for thermal and sound insulation. The patent literature contains many references to cyclomatrix polymers derived from cyclic trimeric phosphazenes incorporated into cross-linked resins for fire resistant circuit boards and related applications.

References

  1. 1 2 3 Allcock, H. R., Kugel, R. L. (1965). "Synthesis of High Polymeric Alkoxy and Aryloxyphosphonitriles". J. Am. Chem. Soc. 87: 4216 4217. doi:10.1021/ja01096a056.
  2. 1 2 3 Allcock, H. R., Kugel, R. L., Valan, K. J. (1966). "High Molecular Weight Poly(alkoxy and aryloxy-phosphazenes)". Inorg. Chem. 5: 1709 1715. doi:10.1021/ic50044a016.
  3. 1 2 3 Allcock, H. R., Kugel, R. L. (1966). "High Molecular Weight Poly(diaminophosphazenes)". Inorg. Chem. 5: 1716 1718. doi:10.1021/ic50044a017.
  4. 1 2 3 "Allcock Research Group Web Site".
  5. Allcock, Harry R. (2003). Chemistry and Applications of Polyphosphazenes. Wiley-Interscience.
  6. Gleria, M., De Jaeger, R. and Potin, P. (2004). Synthesis and Characterization of Poly(organophosphazenes). New York: Nova Science Publishers.
  7. Honeyman, C. H., Manners, I., Morrissey, C. T., Allcock, H. R. (1995). "Ambient Temperature Synthesis of Poly(dichlorophosphazene) with Molecular Weight Control". J. Am. Chem. Soc. 117: 7035–7036. doi:10.1021/ja00131a040.
  8. Allcock, H. R., Crane, C. A., Morrissey, C. T., Nelson, J. M., Reeves, S. D., Honeyman, C. H., Manners, I. (1996). ""Living" Cationic Polymerization of Phosphoranimines as an Ambient Temperature Route to Polyphosphazenes with Controlled Molecular Weights". Macromolecules. 29: 7740–7747. doi:10.1021/ma960876j.
  9. Wisian-Neilson, P.; Neilson, R. H. (1980). "Poly(dimethylphosphazene)". J. Am. Chem. Soc. 102: 2848. doi:10.1021/ja00528a060.
  10. Neilson, R. H., Wisian Neilson, P. (1988). "Poly(alkyl/arylphosphazenes) and their precursors". Chem. Rev.: 541–562. doi:10.1021/cr00085a005.
  11. Montague, R. A., Matyjaszewski, K. (1990). "Synthesis of Poly[bis(trifluoroethoxy)phosphazene] under Mild Conditions using a Fluoride Initiator". J. Am. Chem. Soc. 112: 6721. doi:10.1021/ja00174a047.
  12. Matyjaszewski, K., Moore, M. M., White (1993). "Synthesis of Polyphosphazene Block Copolymers bearing Alkoxyethoxy and Trifluoroethoxy Groups". Macromolecules. 26: 6741. doi:10.1021/ma00077a008.
  13. Allen, C. W., Shaw, J. C., Brown, D. E. (1988). "Copolymerization of ((alpha-methylethenyl)phenyl)pentafluorocyclotriphosphazenes with Styrene and Methyl Methacrylate". Macromolecules. 21: 2653–2657. doi:10.1021/ma00187a001.
  14. Rose, S. H. (1968). "Synthesis of Phosphonitrilic Fluoroelastomers". J. Polymer Sci. 6: 837–839. doi:10.1002/pol.1968.110061203.
  15. Singler, R. E., Schneider, N. S., Hagnauer, G. L. (1975). "Polyphosphazenes: Synthesis—properties—Applications". Polymer Engineering Sci. 15: 321–338. doi:10.1002/pen.760150502.
  16. Kolich, C. H., Klobucar, W. D., Books, J. T. (1990). "Process for Surface Treating Phosphonitrilic Fluoroelastomers". U.S. Patent , 4,945,139.
  17. Tate, D. P. (1974). "Polyphosphazene Elastomers". J. Polymer Sci. (Symposia). 48: 33–45. doi:10.1002/polc.5070480106.
  18. Gettleman, L.; Farris, C. L.; Rawls, H. R. & LeBouef, R. J. (1984). "Soft and Firm Denture Liner for a Composite Denture and Method of Fabricating". U.S. Patent No. 4432730.
  19. Weikel, A. L., Lee, D., Krogman, N. R., Allcock, H. R. (2010). Polymer Engineering Sci. 92A: 114–125. Missing or empty |title= (help)
  20. Blonsky, P. M.; Shriver, D. F.; Austin, P. E.; Allcock, H. R. (1984). "Polyphosphazene solid electrolytes". J. Am. Chem. Soc. 106: 6854–6855. doi:10.1021/ja00334a071.
  21. , H. R.; O’Connor, S. J. M.; Olmeijer, D. L.; Napierala, M. E.; Cameron, C. G. (1996). "Cation Complexation and Conductivity in Crown Ether Bearing Polyphosphazenes". Macromolecules. 29: 7544–7552. doi:10.1021/ma960592z.
  22. Fei, S.-T.; Allcock, H. R. (2010). "Methoxyethoxyethoxyphosphazenes as Ionic Conductive Fire Retardant /additives for Lithium Battery Systems". J. Power Sources. 195 (7): 2082–2088. doi:10.1016/j.jpowsour.2009.09.043.
  23. Fei, S.-T; Lee, S.-H. A; Pursel, S. M.; Basham, J.; Hess, A.; Grimes, C. A.; Horn, M. W.; Mallouk, T. E.; Allcock, H. R. (2011). "Electrolyte Infiltration in Phosphazene-Based Dye-Sensitized Solar Cells". J. Power Sources. 21: 2641–2651.
  24. Tang, H.; Pintauro, P. N. (2001). "Polyphosphazene membranes. IV. Polymer morphology and proton conductivity in sulfonated poly[bis(3-methylphenoxy)phosphazene] films". J. Applied Polymer Sci. 79: 49–59. doi:10.1002/1097-4628(20010103)79:1<49::aid-app60>3.0.co;2-j.
  25. H. R. Allcock; S. Kwon; G. H. Riding; R. J. Fitzpatrick; J. L. Bennett (1988). "Hydrophilic Polyphosphazenes as Hydrogels: Radiation Cr osslinking and Hydrogel Characteristics of Poly[bis(methoxyethoxyethoxy)phosphazene]". Biomaterials. 9: 509–513. doi:10.1016/0142-9612(88)90046-4.
  26. 1 2 Kim, J.; Chun, C.; Kim, B.; Hong, J. M.; Cho J.–K; Lee. S. H. & Song, S.–C. (2012). "Thermosensitive/magnetic poly(organophosphazene) hydrogel as a long-term magnetic resonance contrast platform". Biomaterials. 33: 218–224. doi:10.1016/j.biomaterials.2011.09.033.
  27. H. R. Allcock, S. R. Pucher, M. L. Turner and R. J. Fitzpatrick (1992). "Poly(organophosphazenes) with Poly(alkyl ether) Side Groups: A Study of Their Water Solubility and the Swelling Characteristics of Their Hydrogels". Macromolecules. 25: 5573–5577. doi:10.1021/ma00047a002.
  28. . R. Allcock; R. J. Fitzpatrick; K. B. Visscher (1992). "Thin Layer Grafts of Poly[bis(methoxyethoxy¬¬ethoxy)¬-phosphazene] on Organic Polymer Surfaces". Chemistry of Materials. 4: 775–780. doi:10.1021/cm00022a007.
  29. H. R. Allcock; A. M. A. Ambrosio (1996). "Synthesis and Characterization of pH-Senstitive Poly(organophosphazene) Hydrogels". Biomaterials. 17: 2295–2302. doi:10.1016/0142-9612(96)00073-7.
  30. Allcock, H. R.; Pucher, S. R.; Scopelianos, A. G. (1994). "Poly[amino acid ester)phosphazenes] as Substrates for the Controlled Release of Small Molecules". Biomaterials. 15: 563–569. doi:10.1016/0142-9612(94)90205-4.
  31. Deng, M., Kumbar, S. G., Wan, Y. Toti, U. S. Allcock, H. R., Laurencin, C. T. (2010). "Polyphosphazene Polymers for Tissue Engineering: An Analysis of Material Synthesis, Characterization, and Applications". Soft Matter. 6: 3119–3132. doi:10.1039/b926402g.
  32. Deng, M., Kumbar, S. G., Nair, L. S. Arlin L. Weikel, A. L, Allcock, H. R., Laurencin, C. T. (2011). "Biomimetic Structures: Biological Implications of Dipeptide-Substituted Polyphosphazene–Polyester Blend Nanofiber Matrices for Load-Bearing Bone Regeneration". Adv. Functional Mater. 21: 2641–2651. doi:10.1002/adfm.201100275.
  33. Allcock, H. R.; Morozowich, N. (2012). "Bioerodible Polyphosphazenes and their Medical Potential". Polymer Chemistry. 3: 578–590. doi:10.1039/c1py00468a.

Further information

"H. R. Allcock Research Group". 2007-01-22. Retrieved 2007-10-26. 

This article is issued from Wikipedia - version of the 6/5/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.