Chain walking

Last updated

In polymer chemistry, chain walking (CW) or chain running or chain migration is a mechanism that operates during some alkene polymerization reactions. CW can be also considered as a specific case of intermolecular chain transfer (analogous to radical ethene polymerization). This reaction gives rise to branched and hyperbranched/dendritic hydrocarbon polymers. This process is also characterized by accurate control of polymer architecture and topology. [1] The extent of CW, displayed in the number of branches formed and positions of branches on the polymers are controlled by the choice of a catalyst. The potential applications of polymers formed by this reaction are diverse, from drug delivery to phase transfer agents, nanomaterials, and catalysis. [2]

Contents

Catalysts

Catalysts that promote chain walking were discovered in the 1980-1990s. Nickel(II) and palladium(II) complexes of α-diimine ligands were known to efficiently catalyze polymerization of alkenes. They are also referred as Brookhart's catalysts after being used for making of high molar mass polyolefins for the first time at University of North Carolina at Chapel Hill in 1995. [3] Currently nickel and palladium complexes bearing α-diimine ligands, such as the two examples shown, are the most thoroughly described chain walking catalysts in scientific literature. [4] Ligand design influences not only CW extent but also regio- and stereoselectivity [5] [6] and also the sensitivity of the catalyst to undergo chain-breaking reactions, mainly β-H elimination, influencing achievable molar mass and also the possibility to achieve living polymerization behaviour. [7] [8] [9] [10] [11] Thus stereo block copolymers could be made by combination of living and stereospecific CW polymerization catalysts. [6] [12] Continuous research effort led to design of other ligands which provide CW polymerization catalysts upon complexation to late transition metals. Examples are β-diimine, α-keto-β-diimine, [13] amine-imine [14] and most recently diamine ligands. [15] As the vast majority of CW polymerization catalysts is based on late-transition metal complexes, having generally lower oxophilicity, these complexes were demonstrated also to provide copolymerisation of olefins with polar monomers like acrylates, alkylvinylketones, ω-alken-1-ols, ω-alken-1-carboxylic acids etc., which was the main initial intention of development of this class of catalysts. [16] These random copolymers could further be utilized in the construction of sophisticated amphiphilic grafted copolymers with hydrophobic polyolefin core and shell based on hydrophilic arms, in some cases made of stimuli-responsive polymers. [17]

Example of a nickel precatalyst that, upon activation with suitable cocatalyst (MAO, organoaluminiums), promotes chain walking. Ni Catalyst.png
Example of a nickel precatalyst that, upon activation with suitable cocatalyst (MAO, organoaluminiums), promotes chain walking.
Example of a palladium catalyst that promotes chain walking. Pd Catalyst.png
Example of a palladium catalyst that promotes chain walking.

Mechanism

Structure of polymers obtained in CW polymerization of ethene vs. higher olefins CW C2 vs C3+.png
Structure of polymers obtained in CW polymerization of ethene vs. higher olefins

CW occurs after the polymer chain has grown somewhat on the metal catalyst. The precursor is a 16 e complex with the general formula [ML2(C2H4)(chain)]+. The ethylene ligand (the monomer) dissociates to produce a highly unsaturated 14 e cation. This cation is stabilized by an agostic interaction. β-Hydride elimination then occurs to give a hydride-alkene complex. Subsequent reinsertion of the M-H into the C=C bond, but in the opposite sense gives a metal-alkyl complex. [18]

Mechanism of CW ethene polymerization leading to the formation of short chain branches Chain walkingC2.gif
Mechanism of CW ethene polymerization leading to the formation of short chain branches

This process, a step in the chain walk, moves the metal from the end of a chain to a secondary carbon center. At this stage, two options are available: (1) chain walking can continue or (2) a molecule of ethylene can bind to reform the 16e complex. At this second resting state, the ethylene molecule can insert to grow the polymer or dissociate inducing further chain walking. If many branches can form, a hyperbranched topology results. Therefore, ethene only homopolymerization can provide branched polymer whereas the same mechanism leads to chain straightening in α-olefin polymerization. [16] The variation of CW by changing T, monomer concentration, or catalyst switch [19] [20] [21] can be used to produce block copolymer with amorphous and semi-crystalline blocks or with blocks of different topology.

ChainWalk.png

Related Research Articles

A Ziegler–Natta catalyst, named after Karl Ziegler and Giulio Natta, is a catalyst used in the synthesis of polymers of 1-alkenes (alpha-olefins). Two broad classes of Ziegler–Natta catalysts are employed, distinguished by their solubility:

In polymer chemistry, living polymerization is a form of chain growth polymerization where the ability of a growing polymer chain to terminate has been removed. This can be accomplished in a variety of ways. Chain termination and chain transfer reactions are absent and the rate of chain initiation is also much larger than the rate of chain propagation. The result is that the polymer chains grow at a more constant rate than seen in traditional chain polymerization and their lengths remain very similar. Living polymerization is a popular method for synthesizing block copolymers since the polymer can be synthesized in stages, each stage containing a different monomer. Additional advantages are predetermined molar mass and control over end-groups.

<span class="mw-page-title-main">Kaminsky catalyst</span> Ethylene polymerization catalyst

A Kaminsky catalyst is a catalytic system for alkene polymerization. Kaminsky catalysts are based on metallocenes of group 4 transition metals activated with methylaluminoxane (MAO). These and other innovations have inspired development of new classes of catalysts that in turn led to commercialization of novel engineering polyolefins.

A post-metallocene catalyst is a kind of catalyst for the polymerization of olefins, i.e., the industrial production of some of the most common plastics. "Post-metallocene" refers to a class of homogeneous catalysts that are not metallocenes. This area has attracted much attention because the market for polyethylene, polypropylene, and related copolymers is large. There is a corresponding intense market for new processes as indicated by the fact that, in the US alone, 50,000 patents were issued between 1991-2007 on polyethylene and polypropylene.

<span class="mw-page-title-main">Wacker process</span>

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Olefin metathesis</span> Organic reaction involving the breakup and reassembly of alkene double bonds

In organic chemistry, olefin metathesis is an organic reaction that entails the redistribution of fragments of alkenes (olefins) by the scission and regeneration of carbon-carbon double bonds. Because of the relative simplicity of olefin metathesis, it often creates fewer undesired by-products and hazardous wastes than alternative organic reactions. For their elucidation of the reaction mechanism and their discovery of a variety of highly active catalysts, Yves Chauvin, Robert H. Grubbs, and Richard R. Schrock were collectively awarded the 2005 Nobel Prize in Chemistry.

Polyketones are a family of high-performance thermoplastic polymers. The polar ketone groups in the polymer backbone of these materials gives rise to a strong attraction between polymer chains, which increases the material's melting point (255 °C for copolymer, 220 °C for terpolymer. Trade names include Poketone, Carilon, Karilon, Akrotek, and Schulaketon. Such materials also tend to resist solvents and have good mechanical properties. Unlike many other engineering plastics, aliphatic polyketones such as Shell Chemicals' Carilon are relatively easy to synthesize and can be derived from inexpensive monomers. Carilon is made with a palladium catalyst from ethylene and carbon monoxide. A small fraction of the ethylene is generally replaced with propylene to reduce the melting point somewhat. Shell Chemical commercially launched Carilon thermoplastic polymer in the U.S. in 1996, but discontinued it in 2000. Hyosung announced that they would launch production in 2015.

Coordination polymerisation is a form of polymerization that is catalyzed by transition metal salts and complexes.

<span class="mw-page-title-main">Constrained geometry complex</span>

In organometallic chemistry, a "constrained geometry complex" (CGC) is a kind of catalyst used for the production of polyolefins such as polyethylene and polypropylene. The catalyst was one of the first major departures from metallocene-based catalysts and ushered in much innovation in the development of new plastics.

<span class="mw-page-title-main">Maurice Brookhart</span>

Maurice S. Brookhart is an American chemist, and professor of chemistry at the University of Houston since 2015.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

<span class="mw-page-title-main">Concurrent tandem catalysis</span>

Concurrent tandem catalysis (CTC) is a technique in chemistry where multiple catalysts produce a product otherwise not accessible by a single catalyst. It is usually practiced as homogeneous catalysis. Scheme 1 illustrates this process. Molecule A enters this catalytic system to produce the comonomer, B, which along with A enters the next catalytic process to produce the final product, P. This one-pot approach can decrease product loss from isolation or purification of intermediates. Reactions with relatively unstable products can be generated as intermediates because they are only transient species and are immediately used in a consecutive reaction.

<span class="mw-page-title-main">Thiol-yne reaction</span>

The thiol-yne reaction is an organic reaction between a thiol and an alkyne. The reaction product is an alkenyl sulfide. The reaction was first reported in 1949 with thioacetic acid as reagent and rediscovered in 2009. It is used in click chemistry and in polymerization, especially with dendrimers.

Catalytic chain transfer (CCT) is a process that can be incorporated into radical polymerization to obtain greater control over the resulting products.

<span class="mw-page-title-main">Brookhart's acid</span> Chemical compound

Brookhart's acid is the salt of the diethyl ether oxonium ion and tetrakis[3,5-bis(trifluoromethyl)phenyl]borate (BAr′4). It is a colorless solid, used as a strong acid. The compound was first reported by Volpe, Grant, and Brookhart in 1992.

Diiminopyridines are a class of diimine ligands. They featuring a pyridine nucleus with imine sidearms appended to the 2,6–positions. The three nitrogen centres bind metals in a tridentate fashion, forming pincer complexes. Diiminopyridines are notable as non-innocent ligand that can assume more than one oxidation state. Complexes of DIPs participate in a range of chemical reactions, including ethylene polymerization, hydrosilylation, and hydrogenation.

Diimines are organic compounds containing two imine (RCH=NR') groups. Common derivatives are 1,2-diketones and 1,3-diimines. These compounds are used as ligands and as precursors to heterocycles. Diimines are prepared by condensation reactions where a dialdehyde or diketone is treated with amine and water is eliminated. Similar methods are used to prepare Schiff bases and oximes.

Functionalized polyolefins are olefin polymers with polar and nonpolar functionalities attached onto the polymer backbone. There has been an increased interest in functionalizing polyolefins due to their increased usage in everyday life. Polyolefins are virtually ubiquitous in everyday life, from consumer food packaging to biomedical applications; therefore, efforts must be made to study catalytic pathways towards the attachment of various functional groups onto polyolefins in order to affect the material's physical properties.

The Mukaiyama hydration is an organic reaction involving formal addition of an equivalent of water across an olefin by the action of catalytic bis(acetylacetonato)cobalt(II) complex, phenylsilane and atmospheric oxygen to produce an alcohol with Markovnikov selectivity.

β-Carbon elimination is a type of reaction in organometallic chemistry wherein an allyl ligand bonded to a metal center is broken into the corresponding metal-bonded alkyl (aryl) ligand and an alkene. It is a subgroup of elimination reactions. Though less common and less understood than β-hydride elimination, it is an important step involved in some olefin polymerization processes and transition-metal-catalyzed organic reactions.

References

  1. Guan, Z.; Cotts, PM; McCord, EF; McLain, SJ (1999). "Chain Walking: A New Strategy to Control Polymer Topology". Science. 283 (5410): 2059–2062. Bibcode:1999Sci...283.2059G. doi:10.1126/science.283.5410.2059. PMID   10092223.
  2. Guan, Z. (2010). "Recent Progress of Catalytic Polymerization for Controlling Polymer Topology". Chem. Asian J. 5 (5): 1058–1070. doi:10.1002/asia.200900749. PMID   20391469.
  3. Johnson, Lynda K.; Killian, Christopher M.; Brookhart, Maurice (June 1995). "New Pd(II)- and Ni(II)-Based Catalysts for Polymerization of Ethylene and .alpha.-Olefins". Journal of the American Chemical Society. 117 (23): 6414–6415. doi:10.1021/ja00128a054. ISSN   0002-7863.
  4. Domski, G. J.; Rose, J. M.; Coates, G. W.; Bolig, A. D.; Brookhart, M. (2007). "Living alkene polymerization: New methods for the precision synthesis of polyolefins". Prog. Polym. Sci. 32: 30–92. doi:10.1016/j.progpolymsci.2006.11.001.
  5. Cherian, Anna E.; Rose, Jeffrey M.; Lobkovsky, Emil B.; Coates, Geoffrey W. (October 2005). "A C 2 -Symmetric, Living α-Diimine Ni(II) Catalyst: Regioblock Copolymers from Propylene". Journal of the American Chemical Society. 127 (40): 13770–13771. doi:10.1021/ja0540021. ISSN   0002-7863. PMID   16201780.
  6. 1 2 Rose, Jeffrey M.; Deplace, Fanny; Lynd, Nathaniel A.; Wang, Zhigang; Hotta, Atsushi; Lobkovsky, Emil B.; Kramer, Edward J.; Coates, Geoffrey W. (2008-12-23). "C 2 -Symmetric Ni(II) α-Diimines Featuring Cumyl-Derived Ligands: Synthesis of Improved Elastomeric Regioblock Polypropylenes". Macromolecules. 41 (24): 9548–9555. doi:10.1021/ma8019943. ISSN   0024-9297.
  7. Killian, Christopher M.; Tempel, Daniel J.; Johnson, Lynda K.; Brookhart, Maurice (January 1996). "Living Polymerization of α-Olefins Using Ni II −α-Diimine Catalysts. Synthesis of New Block Polymers Based on α-Olefins". Journal of the American Chemical Society. 118 (46): 11664–11665. doi:10.1021/ja962516h. ISSN   0002-7863.
  8. Gottfried, Amy C.; Brookhart, Maurice (February 2001). "Living Polymerization of Ethylene Using Pd(II) α-Diimine Catalysts". Macromolecules. 34 (5): 1140–1142. doi:10.1021/ma001595l. ISSN   0024-9297.
  9. Gottfried, Amy C.; Brookhart, M. (May 2003). "Living and Block Copolymerization of Ethylene and α-Olefins Using Palladium(II)−α-Diimine Catalysts". Macromolecules. 36 (9): 3085–3100. doi:10.1021/ma025902u. ISSN   0024-9297.
  10. Leone, Giuseppe; Mauri, Massimiliano; Bertini, Fabio; Canetti, Maurizio; Piovani, Daniele; Ricci, Giovanni (2015-03-10). "Ni(II) α-Diimine-Catalyzed α-Olefins Polymerization: Thermoplastic Elastomers of Block Copolymers". Macromolecules. 48 (5): 1304–1312. doi:10.1021/ma502427u. ISSN   0024-9297.
  11. Peleška, Jan; Hoštálek, Zdeněk; Hasalíková, Darja; Merna, Jan (January 2011). "Living/controlled hex-1-ene polymerization initiated by nickel diimine complexes activated by non-MAO cocatalysts: Kinetic and UV–vis study". Polymer. 52 (2): 275–281. doi:10.1016/j.polymer.2010.11.029.
  12. Sokolohorskyj, Anatolij; Železník, Ondřej; Císařová, Ivana; Lenz, Johannes; Lederer, Albena; Merna, Jan (2017-08-01). "α-keto-β-diimine nickel-catalyzed olefin polymerization: Effect of ortho-aryl substituents and preparation of stereoblock copolymers". Journal of Polymer Science Part A: Polymer Chemistry. 55 (15): 2440–2449. doi:10.1002/pola.28631.
  13. Azoulay, Jason D.; Rojas, Rene S.; Serrano, Abigail V.; Ohtaki, Hisashi; Galland, Griselda B.; Wu, Guang; Bazan, Guillermo C. (2009-01-26). "Nickel α-Keto-β-Diimine Initiators for Olefin Polymerization". Angewandte Chemie International Edition. 48 (6): 1089–1092. doi:10.1002/anie.200804661. PMID   19065683.
  14. Gao, Haiyang; Hu, Haibin; Zhu, Fangming; Wu, Qing (2012). "A thermally robust amine–imine nickel catalyst precursor for living polymerization of ethylene above room temperature". Chemical Communications. 48 (27): 3312–3314. doi:10.1039/c2cc17154f. ISSN   1359-7345. PMID   22361652.
  15. Liao, Heng; Zhong, Liu; Xiao, Zefan; Zheng, Ting; Gao, Haiyang; Wu, Qing (2016-09-19). "α-Diamine Nickel Catalysts with Nonplanar Chelate Rings for Ethylene Polymerization". Chemistry: A European Journal. 22 (39): 14048–14055. doi:10.1002/chem.201602467. PMID   27514844.
  16. 1 2 Ittel, Steven D.; Johnson, Lynda K.; Brookhart, Maurice (April 2000). "Late-Metal Catalysts for Ethylene Homo- and Copolymerization". Chemical Reviews. 100 (4): 1169–1204. doi:10.1021/cr9804644. ISSN   0009-2665. PMID   11749263.
  17. Chen, Yongsheng; Wang, Li; Yu, Haojie; Zhao, Yulai; Sun, Ruoli; Jing, Guanghui; Huang, Jin; Khalid, Hamad; Abbasi, Nasir M. (June 2015). "Synthesis and application of polyethylene-based functionalized hyperbranched polymers". Progress in Polymer Science. 45: 23–43. doi:10.1016/j.progpolymsci.2015.01.004.
  18. Atkins, P.; Overton, T.; Rourke, J.; Weller, M.; Armstrong, F.; Hagerman, M. (2010). Inorganic Chemistry (5th ed.). New York: W. H. Freeman and Company. p. 574. ISBN   978-0-19-923617-6.
  19. Mundil, Robert; Wilson, Lucy E.; Schaarschmidt, Dieter; Císařová, Ivana; Merna, Jan; Long, Nicholas J. (September 2019). "Redox-switchable α-diimine palladium catalysts for control of polyethylene topology". Polymer. 179: 121619. doi:10.1016/j.polymer.2019.121619. hdl: 10044/1/73501 .
  20. Anderson, W. Curtis; Rhinehart, Jennifer L.; Tennyson, Andrew G.; Long, Brian K. (2016-01-27). "Redox-Active Ligands: An Advanced Tool To Modulate Polyethylene Microstructure". Journal of the American Chemical Society. 138 (3): 774–777. doi:10.1021/jacs.5b12322. ISSN   0002-7863. PMID   26722675.
  21. Zhao, Minhui; Chen, Changle (2017-11-03). "Accessing Multiple Catalytically Active States in Redox-Controlled Olefin Polymerization". ACS Catalysis. 7 (11): 7490–7494. doi: 10.1021/acscatal.7b02564 . ISSN   2155-5435.