Ring-opening polymerization

Last updated
IUPAC definition

A polymerization in which a cyclic monomer yields a monomeric unit which is acyclic or contains fewer cycles than the monomer. Note: If monomer is polycyclic, the opening of a single ring is sufficient to classify the reaction as ring-opening polymerization.

Contents

Modified from the earlier definition. [1] [2]

Penczek S.; Moad, G. Pure Appl. Chem., 2008, 80(10), 2163-2193

General scheme ionic propagation. Propagating center can be radical, cationic or anionic. General scheme ionic prop.png
General scheme ionic propagation. Propagating center can be radical, cationic or anionic.

In polymer chemistry, ring-opening polymerization (ROP) is a form of chain-growth polymerization in which the terminus of a polymer chain attacks cyclic monomers to form a longer polymer (see figure). The reactive center can be radical, anionic or cationic. Some cyclic monomers such as norbornene or cyclooctadiene can be polymerized to high molecular weight polymers by using metal catalysts. ROP is a versatile method for the synthesis of biopolymers.

Ring-opening of cyclic monomers is often driven by the relief of bond-angle strain. Thus, as is the case for other types of polymerization, the enthalpy change in ring-opening is negative. [3]

Monomers

Cyclic monomers that are amenable to ROP include epoxides, [4] [5] cyclic trisiloxanes,[ citation needed ] some lactones [4] [6] and lactides, [6] cyclic anhydrides, [5] cyclic carbonates, [7] and amino acid N-carboxyanhydrides. [8] [9] Many strained cycloalkenes, e.g norbornene, are suitable monomers via ring-opening metathesis polymerization. Even highly strained cycloalkane rings, such as cyclopropane [10] and cyclobutane [11] derivatives, can undergo ROP.

History

Ring-opening polymerization has been used since the beginning of the 1900s to produce polymers. Synthesis of polypeptides which has the oldest history of ROP, dates back to the work in 1906 by Leuchs. [12] Subsequently, the ROP of anhydro sugars provided polysaccharides, including synthetic dextran, xanthan gum, welan gum, gellan gum, diutan gum, and pullulan. Mechanisms and thermodynamics of ring-opening polymerization were established in the 1950s. [13] [14] The first high-molecular weight polymers (Mn up to 105) with a repeating unit were prepared by ROP as early as in 1976. [15] [16]

An industrial application is the production of nylon-6 from caprolactam.

Mechanisms

Ring-opening polymerization can proceed via radical, anionic, or cationic polymerization as described below. [17] Additionally, radical ROP is useful in producing polymers with functional groups incorporated in the backbone chain that cannot otherwise be synthesized via conventional chain-growth polymerization of vinyl monomers. For instance, radical ROP can produce polymers with ethers, esters, amides, and carbonates as functional groups along the main chain. [17] [18]

Anionic ring-opening polymerization (AROP)

The general mechanism for anionic ring-opening polymerization. Polarized functional group is represented by X-Y, where the atom X (usually a carbon atom) becomes electron deficient due to the highly electron-withdrawing nature of Y (usually an oxygen, nitrogen, sulfur, etc.). The nucleophile will attack atom X, thus releasing Y . The newly formed nucleophile will then attack the atom X in another monomer molecule, and the sequence would repeat until the polymer is formed. Wiki566665.tif
The general mechanism for anionic ring-opening polymerization. Polarized functional group is represented by X-Y, where the atom X (usually a carbon atom) becomes electron deficient due to the highly electron-withdrawing nature of Y (usually an oxygen, nitrogen, sulfur, etc.). The nucleophile will attack atom X, thus releasing Y . The newly formed nucleophile will then attack the atom X in another monomer molecule, and the sequence would repeat until the polymer is formed.

Anionic ring-opening polymerizations (AROP) involve nucleophilic reagents as initiators. Monomers with a three-member ring structure - such as epoxides, aziridines, and episulfides - undergo anionic ROP. [18]

A typical example of anionic ROP is that of ε-caprolactone, initiated by an alkoxide. [18]

Cationic ring-opening polymerization

Cationic initiators and intermediates characterize cationic ring-opening polymerization (CROP). Examples of cyclic monomers that polymerize through this mechanism include lactones, lactams, amines, and ethers. [19] CROP proceeds through an SN1 or SN2 propagation, chain-growth process. [17] The mechanism is affected by the stability of the resulting cationic species. For example, if the atom bearing the positive charge is stabilized by electron-donating groups, polymerization will proceed by the SN1 mechanism. [18] The cationic species is a heteroatom and the chain grows by the addition of cyclic monomers thereby opening the ring system.

Synthesis of Spandex. PTMEG synthesis.svg
Synthesis of Spandex.

The monomers can be activated by Bronsted acids, carbenium ions, onium ions, and metal cations. [17]

CROP can be a living polymerization and can be terminated by nucleophilic reagents such as phenoxy anions, phosphines, or polyanions. [17] When the amount of monomers becomes depleted, termination can occur intra or intermolecularly. The active end can "backbite" the chain, forming a macrocycle. Alkyl chain transfer is also possible, where the active end is quenched by transferring an alkyl chain to another polymer.

Ring-opening metathesis polymerization

Ring-opening metathesis polymerisation (ROMP) produces unsaturated polymers from cycloalkenes or bicycloalkenes. It requires organometallic catalysts. [17]

The mechanism for ROMP follows similar pathways as olefin metathesis. The initiation process involves the coordination of the cycloalkene monomer to the metal alkylidene complex, followed by a [2+2] type cycloaddition to form the metallacyclobutane intermediate that cycloreverts to form a new alkylidene species. [21] [22]

General scheme of the mechanism for ROMP. Romp mechanism.png
General scheme of the mechanism for ROMP.

Commercially relevant unsaturated polymers synthesized by ROMP include polynorbornene, polycyclooctene, and polycyclopentadiene. [23]

Thermodynamics

The formal thermodynamic criterion of a given monomer polymerizability is related to a sign of the free enthalpy (Gibbs free energy) of polymerization:

where:

x and y indicate monomer and polymer states, respectively (x and/or y = l (liquid), g (gaseous), c (amorphous solid), c' (crystalline solid), s (solution));
ΔHp(xy) is the enthalpy of polymerization (SI unit: joule per kelvin);
ΔSp(xy) is the entropy of polymerization (SI unit: joule);
T is the absolute temperature (SI unit: kelvin).

The free enthalpy of polymerization (ΔGp) may be expressed as a sum of standard enthalpy of polymerization (ΔGp°) and a term related to instantaneous monomer molecules and growing macromolecules concentrations:

where:

R is the gas constant;
M is the monomer;
(m)i is the monomer in an initial state;
m* is the active monomer.

Following Flory–Huggins solution theory that the reactivity of an active center, located at a macromolecule of a sufficiently long macromolecular chain, does not depend on its degree of polymerization (DPi), and taking in to account that ΔGp° = ΔHp° TΔSp° (where ΔHp° and ΔSp° indicate a standard polymerization enthalpy and entropy, respectively), we obtain:

At equilibrium (ΔGp = 0), when polymerization is complete the monomer concentration ([M]eq) assumes a value determined by standard polymerization parameters (ΔHp° and ΔSp°) and polymerization temperature:

Polymerization is possible only when [M]0 > [M]eq. Eventually, at or above the so-called ceiling temperature (Tc), at which [M]eq = [M]0, formation of the high polymer does not occur.

For example, tetrahydrofuran (THF) cannot be polymerized above Tc = 84 °C, nor cyclo-octasulfur (S8) below Tf = 159 °C. [24] [25] [26] [27] However, for many monomers, Tc and Tf, for polymerization in the bulk, are well above or below the operable polymerization temperatures, respectively. The polymerization of a majority of monomers is accompanied by an entropy decrease, due mostly to the loss in the translational degrees of freedom. In this situation, polymerization is thermodynamically allowed only when the enthalpic contribution into ΔGp prevails (thus, when ΔHp° < 0 and ΔSp° < 0, the inequality |ΔHp| > TΔSp is required). Therefore, the higher the ring strain, the lower the resulting monomer concentration at equilibrium.

Additional reading

Related Research Articles

In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

<span class="mw-page-title-main">Eutectic system</span> Mixture with a lower melting point than its constituents

A eutectic system or eutectic mixture is a homogeneous mixture that has a melting point lower than those of the constituents. The lowest possible melting point over all of the mixing ratios of the constituents is called the eutectic temperature. On a phase diagram, the eutectic temperature is seen as the eutectic point.

<span class="mw-page-title-main">Gibbs free energy</span> Type of thermodynamic potential

In thermodynamics, the Gibbs free energy is a thermodynamic potential that can be used to calculate the maximum amount of work, other than pressure-volume work, that may be performed by a thermodynamically closed system at constant temperature and pressure. It also provides a necessary condition for processes such as chemical reactions that may occur under these conditions. The Gibbs free energy is expressed as

In physical chemistry, Henry's law is a gas law that states that the amount of dissolved gas in a liquid is directly proportional to its partial pressure above the liquid. The proportionality factor is called Henry's law constant. It was formulated by the English chemist William Henry, who studied the topic in the early 19th century.

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">Host–guest chemistry</span> Supramolecular structures held together other than by covalent bonds

In supramolecular chemistry, host–guest chemistry describes complexes that are composed of two or more molecules or ions that are held together in unique structural relationships by forces other than those of full covalent bonds. Host–guest chemistry encompasses the idea of molecular recognition and interactions through non-covalent bonding. Non-covalent bonding is critical in maintaining the 3D structure of large molecules, such as proteins and is involved in many biological processes in which large molecules bind specifically but transiently to one another.

<span class="mw-page-title-main">Radical polymerization</span> Polymerization process involving free radicals as repeating units

In polymer chemistry, free-radical polymerization (FRP) is a method of polymerization by which a polymer forms by the successive addition of free-radical building blocks. Free radicals can be formed by a number of different mechanisms, usually involving separate initiator molecules. Following its generation, the initiating free radical adds (nonradical) monomer units, thereby growing the polymer chain.

<span class="mw-page-title-main">Flory–Huggins solution theory</span> Lattice model of polymer solutions

Flory–Huggins solution theory is a lattice model of the thermodynamics of polymer solutions which takes account of the great dissimilarity in molecular sizes in adapting the usual expression for the entropy of mixing. The result is an equation for the Gibbs free energy change for mixing a polymer with a solvent. Although it makes simplifying assumptions, it generates useful results for interpreting experiments.

The Eyring equation is an equation used in chemical kinetics to describe changes in the rate of a chemical reaction against temperature. It was developed almost simultaneously in 1935 by Henry Eyring, Meredith Gwynne Evans and Michael Polanyi. The equation follows from the transition state theory, also known as activated-complex theory. If one assumes a constant enthalpy of activation and constant entropy of activation, the Eyring equation is similar to the empirical Arrhenius equation, despite the Arrhenius equation being empirical and the Eyring equation based on statistical mechanical justification.

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

A photopolymer or light-activated resin is a polymer that changes its properties when exposed to light, often in the ultraviolet or visible region of the electromagnetic spectrum. These changes are often manifested structurally, for example hardening of the material occurs as a result of cross-linking when exposed to light. An example is shown below depicting a mixture of monomers, oligomers, and photoinitiators that conform into a hardened polymeric material through a process called curing.

The Van 't Hoff equation relates the change in the equilibrium constant, Keq, of a chemical reaction to the change in temperature, T, given the standard enthalpy change, ΔrH, for the process. The subscript means "reaction" and the superscript means "standard". It was proposed by Dutch chemist Jacobus Henricus van 't Hoff in 1884 in his book Études de Dynamique chimique.

In biochemistry, equilibrium unfolding is the process of unfolding a protein or RNA molecule by gradually changing its environment, such as by changing the temperature or pressure, pH, adding chemical denaturants, or applying force as with an atomic force microscope tip. If the equilibrium was maintained at all steps, the process theoretically should be reversible during equilibrium folding. Equilibrium unfolding can be used to determine the thermodynamic stability of the protein or RNA structure, i.e. free energy difference between the folded and unfolded states.

<span class="mw-page-title-main">Thermodynamic databases for pure substances</span> Thermodynamic properties list

Thermodynamic databases contain information about thermodynamic properties for substances, the most important being enthalpy, entropy, and Gibbs free energy. Numerical values of these thermodynamic properties are collected as tables or are calculated from thermodynamic datafiles. Data is expressed as temperature-dependent values for one mole of substance at the standard pressure of 101.325 kPa, or 100 kPa. Both of these definitions for the standard condition for pressure are in use.

In polymer chemistry, chain transfer is a polymerization reaction by which the activity of a growing polymer chain is transferred to another molecule:

<span class="mw-page-title-main">Transition state theory</span> Theory describing the reaction rates of elementary chemical reactions

In chemistry, transition state theory (TST) explains the reaction rates of elementary chemical reactions. The theory assumes a special type of chemical equilibrium (quasi-equilibrium) between reactants and activated transition state complexes.

<span class="mw-page-title-main">Enthalpy of fusion</span> Enthalpy change when a substance melts

In thermodynamics, the enthalpy of fusion of a substance, also known as (latent) heat of fusion, is the change in its enthalpy resulting from providing energy, typically heat, to a specific quantity of the substance to change its state from a solid to a liquid, at constant pressure.

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

In polymer chemistry, cationic polymerization is a type of chain growth polymerization in which a cationic initiator transfers charge to a monomer, which then becomes reactive. This reactive monomer goes on to react similarly with other monomers to form a polymer. The types of monomers necessary for cationic polymerization are limited to alkenes with electron-donating substituents and heterocycles. Similar to anionic polymerization reactions, cationic polymerization reactions are very sensitive to the type of solvent used. Specifically, the ability of a solvent to form free ions will dictate the reactivity of the propagating cationic chain. Cationic polymerization is used in the production of polyisobutylene and poly(N-vinylcarbazole) (PVK).

Ceiling temperature is a measure of the tendency of a polymer to revert to its constituent monomers. When a polymer is at its ceiling temperature, the rate of polymerization and depolymerization of the polymer are equal. Generally, the ceiling temperature of a given polymer is correlated to the steric hindrance of the polymer’s monomers. Polymers with high ceiling temperatures are often commercially useful. Polymers with low ceiling temperatures are more readily depolymerizable.

In colloidal chemistry, the critical micelle concentration (CMC) of a surfactant is one of the parameters in the Gibbs free energy of micellization. The concentration at which the monomeric surfactants self-assemble into thermodynamically stable aggregates is the CMC. The Krafft temperature of a surfactant is the lowest temperature required for micellization to take place. There are many parameters that affect the CMC. The interaction between the hydrophilic heads and the hydrophobic tails play a part, as well as the concentration of salt within the solution and surfactants.

References

  1. IUPAC, Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) "Ring-opening polymerization". doi:10.1351/goldbook.R05396
  2. Jenkins, A. D.; Kratochvíl, P.; Stepto, R. F. T.; Suter, U. W. (1996). "Glossary of basic terms in polymer science (IUPAC Recommendations 1996)". Pure and Applied Chemistry. 68 (12): 2287–2311. doi: 10.1351/pac199668122287 .
  3. Young, Robert J. (2011). Introduction to Polymers. Boca Raton: CRC Press. ISBN   978-0-8493-3929-5.
  4. 1 2 Yann Sarazin; Jean-François Carpentier (2015). "Discrete Cationic Complexes for Ring-Opening Polymerization Catalysis of Cyclic Esters and Epoxides". Chemical Reviews. 115 (9): 3564–3614. doi:10.1021/acs.chemrev.5b00033. PMID   25897976.
  5. 1 2 Longo, Julie M.; Sanford, Maria J.; Coates, Geoffrey W. (2016). "Ring-Opening Copolymerization of Epoxides and Cyclic Anhydrides with Discrete Metal Complexes: Structure–Property Relationships". Chemical Reviews. 116 (24): 15167–15197. doi:10.1021/acs.chemrev.6b00553. PMID   27936619.
  6. 1 2 JEROME, C; LECOMTE, P (2008-06-10). "Recent advances in the synthesis of aliphatic polyesters by ring-opening polymerization☆". Advanced Drug Delivery Reviews. 60 (9): 1056–1076. doi:10.1016/j.addr.2008.02.008. hdl: 2268/3723 . ISSN   0169-409X. PMID   18403043.
  7. Matsumura, Shuichi; Tsukada, Keisuke; Toshima, Kazunobu (May 1997). "Enzyme-Catalyzed Ring-Opening Polymerization of 1,3-Dioxan-2-one to Poly(trimethylene carbonate)". Macromolecules. 30 (10): 3122–3124. Bibcode:1997MaMol..30.3122M. doi:10.1021/ma961862g.
  8. Kricheldorf, H. R. (2006). "Polypeptides and 100 Years of Chemistry of α-Amino Acid N-Carboxyanhydrides". Angewandte Chemie International Edition. 45 (35): 5752–5784. doi:10.1002/anie.200600693. PMID   16948174.
  9. Nikos Hadjichristidis; Hermis Iatrou; Marinos Pitsikalis; Georgios Sakellariou (2009). "Synthesis of Well-Defined Polypeptide-Based Materials via the Ring-Opening Polymerization of α-Amino Acid N-Carboxyanhydrides". Chemical Reviews. 109 (11): 5528–5578. doi:10.1021/cr900049t. PMID   19691359.
  10. Scott, R. J.; Gunning, H. E. (1952). "The Polymerization of Cyclopropane". J. Phys. Chem. 56 (1): 156–160. doi:10.1021/j150493a031.
  11. Yokozawa, Tsutomu; Tsuruta, Ei-ichi (1996). "Ring-Opening Polymerization of the Cyclobutane Adduct of Methyl Tricyanoethylenecarboxylate and Ethyl Vinyl Ether". Macromolecules. 29 (25): 8053–8056. doi:10.1021/ma9608535.
  12. Leuchs, H. (1906). "Glycine-carbonic acid". Berichte der Deutschen Chemischen Gesellschaft. 39: 857. doi:10.1002/cber.190603901133.
  13. Dainton, F. S.; Devlin, T. R. E.; Small, P. A. (1955). "The thermodynamics of polymerization of cyclic compounds by ring opening". Transactions of the Faraday Society. 51: 1710. doi:10.1039/TF9555101710.
  14. Conix, André; Smets, G. (January 1955). "Ring opening in lactam polymers". Journal of Polymer Science. 15 (79): 221–229. Bibcode:1955JPoSc..15..221C. doi:10.1002/pol.1955.120157918.
  15. Kałuz̀ynski, Krzysztof; Libiszowski, Jan; Penczek, Stanisław (1977). "Poly(2-hydro-2-oxo-1,3,2-dioxaphosphorinane). Preparation and NMR spectra". Die Makromolekulare Chemie. 178 (10): 2943–2947. doi:10.1002/macp.1977.021781017. ISSN   0025-116X.
  16. Libiszowski, Jan; Kałużynski, Krzysztof; Penczek, Stanisław (June 1978). "Polymerization of cyclic esters of phosphoric acid. VI. Poly(alkyl ethylene phosphates). Polymerization of 2-alkoxy-2-oxo-1,3,2-dioxaphospholans and structure of polymers". Journal of Polymer Science: Polymer Chemistry Edition. 16 (6): 1275–1283. Bibcode:1978JPoSA..16.1275L. doi:10.1002/pol.1978.170160610.
  17. 1 2 3 4 5 6 Nuyken, Oskar; Stephen D. Pask (25 April 2013). "Ring-Opening Polymerization—An Introductory Review". Polymers. 5 (2): 361–403. doi: 10.3390/polym5020361 .
  18. 1 2 3 4 5 Dubois, Philippe (2008). Handbook of ring-opening polymerization (1. Aufl. ed.). Weinheim: Wiley-VCH. ISBN   978-3-527-31953-4.
  19. Cowie, John McKenzie Grant (2008). Polymers: Chemistry and Physics of Modern Materials. Boca Raton, Florida: CRC Press. pp. 105–107. ISBN   978-0-8493-9813-1.
  20. Pruckmayr, Gerfried; Dreyfuss, P.; Dreyfuss, M. P. (1996). "Polyethers, Tetrahydrofuran and Oxetane Polymers". Kirk‑Othmer Encyclopedia of Chemical Technology. John Wiley & Sons.
  21. Sutthasupa, Sutthira; Shiotsuki, Masashi; Sanda, Fumio (13 October 2010). "Recent advances in ring-opening metathesis polymerization, and application to synthesis of functional materials". Polymer Journal. 42 (12): 905–915. doi: 10.1038/pj.2010.94 .
  22. Hartwig, John F. (2010). Organotransition metal chemistry: from bonding to catalysis. Sausalito, California: University Science Books. ISBN   978-1-891389-53-5.
  23. Walsh, Dylan J.; Lau, Sii Hong; Hyatt, Michael G.; Guironnet, Damien (2017-09-25). "Kinetic Study of Living Ring-Opening Metathesis Polymerization with Third-Generation Grubbs Catalysts". Journal of the American Chemical Society. 139 (39): 13644–13647. doi:10.1021/jacs.7b08010. ISSN   0002-7863. PMID   28944665.
  24. Tobolsky, A. V. (July 1957). "Equilibrium polymerization in the presence of an ionic initiator". Journal of Polymer Science. 25 (109): 220–221. Bibcode:1957JPoSc..25..220T. doi:10.1002/pol.1957.1202510909.
  25. Tobolsky, A. V. (August 1958). "Equilibrium polymerization in the presence of an ionic initiator". Journal of Polymer Science. 31 (122): 126. Bibcode:1958JPoSc..31..126T. doi: 10.1002/pol.1958.1203112214 .
  26. Tobolsky, Arthur V.; Eisenberg, Adi (May 1959). "Equilibrium Polymerization of Sulfur". Journal of the American Chemical Society. 81 (4): 780–782. doi:10.1021/ja01513a004.
  27. Tobolsky, A. V.; Eisenberg, A. (January 1960). "A General Treatment of Equilibrium Polymerization". Journal of the American Chemical Society. 82 (2): 289–293. doi:10.1021/ja01487a009.