Fact-checked by Grok 2 weeks ago

Ring-opening polymerization

Ring-opening polymerization (ROP) is a technique in which cyclic monomers undergo ring scission to form linear or branched polymers, typically driven by the release of energy or an increase in . This method contrasts with other polymerizations by avoiding the loss of small molecules during chain growth, enabling the synthesis of high-molecular-weight polymers with controlled architectures. ROP is industrially significant for producing a wide array of materials, including biodegradable polyesters, plastics, and specialty elastomers. The mechanisms of ROP vary depending on the monomer and catalyst, encompassing anionic ROP (AROP), cationic ROP (CROP), coordination-insertion ROP, radical ROP (RROP), and ring-opening metathesis polymerization (ROMP). In AROP and CROP, nucleophilic or electrophilic initiators attack the strained ring, propagating the chain through repeated openings, while coordination-insertion involves metal catalysts facilitating monomer insertion into a metal-alkoxide bond. ROMP, often catalyzed by transition metal complexes like ruthenium or molybdenum, targets cyclic olefins and is noted for its tolerance to functional groups. These mechanisms allow for living or controlled polymerizations, yielding polymers with narrow molecular weight distributions and block copolymer capabilities. Common monomers for ROP include cyclic esters (e.g., lactide and ε-caprolactone for polylactides and polycaprolactones), cyclic ethers (e.g., epoxides and for polyethers), cyclic carbonates (e.g., trimethylene carbonate), lactams (e.g., ε-caprolactam for ), and cyclic siloxanes (for polysiloxanes like silicones). Other classes encompass cyclic anhydrides, phosphazenes, and disulfides, enabling diverse polymer properties from rigid thermoplastics to flexible elastomers. in three- to seven-membered rings, such as 116 kJ/mol for oxiranes, provides the thermodynamic driving force, though larger rings rely on changes. ROP's applications span biomedical devices, sustainable packaging, and electronics, with notable examples including biodegradable polylactic acid (PLA) for medical implants and food containers, and for high-performance engineering parts. The process supports the production of approximately 500,000 tons of PLA annually as of 2024, addressing plastic waste through hydrolyzable linkages. Recent advances emphasize metal-free organocatalytic ROP and enantioselective variants using chiral catalysts to produce stereoregular polymers from racemic monomers, enhancing mechanical strength via control (e.g., selectivity factors up to 53 for ROP). These developments underscore ROP's role in creating recyclable, high-value materials for a .

Overview

Definition and basic principles

Ring-opening polymerization (ROP) is a method in which cyclic undergo ring scission to form polymers composed of acyclic repeating units. This process typically produces high-molecular-weight polymers with controlled microstructure, leveraging the inherent reactivity of the cyclic structure to achieve precise chain lengths and narrow molecular weight distributions. Unlike traditional monomer additions, ROP relies on the transformation of strained or activated rings into linear or extended chains without the elimination of small molecules in most cases. The basic principles of ROP are governed by the thermodynamic favorability of ring opening, primarily driven by the relief of ring strain in the monomer. For instance, three- and four-membered rings such as epoxides and lactones exhibit significant strain energies (e.g., approximately 116 kJ/mol for oxiranes), providing an enthalpic driving force that offsets the entropic penalty of polymerization. The reaction proceeds through the addition of the cyclic monomer to a propagating species, initiated by nucleophiles or electrophiles, leading to sequential ring openings that extend the polymer chain. This process occurs below the ceiling temperature, where the equilibrium favors polymer formation over depolymerization. A general reaction scheme for ROP can be depicted as follows, where a cyclic (M) reacts with an to form a propagating chain with opened rings: n \ \ce{M(cyclic)} + \ce{I} \rightarrow \ce{I-[M(acyclic)]_n-X} Here, I represents the and X the terminating group, illustrating the conversion of n cyclic units into a linear segment. ROP distinguishes itself from step-growth polymerizations, which involve reactions between bifunctional monomers, and from conventional chain-growth methods like , which rely on double-bond additions without ring involvement. The resulting polymers can exhibit linear topologies for most standard ROP processes, though branched or even cyclic structures may arise depending on conditions and monomer functionality.

Advantages and limitations

Ring-opening polymerization (ROP) offers several advantages over traditional polymerization methods, primarily stemming from the exothermic relief of in cyclic monomers, which drives the reaction toward high molecular weights without the need for harsh conditions. This process enables the synthesis of polymers with molecular weights often exceeding 100,000 g/ and narrow polydispersity indices (typically PDI < 1.5), particularly in living ROP systems where chain-end fidelity is maintained. Unlike addition polymerization of vinyl monomers, ROP does not introduce unsaturation or initiator-derived end groups into the polymer backbone, resulting in more uniform structures suitable for biomedical applications. Additionally, ROP produces minimal volume shrinkage or even expansion during polymerization, which is beneficial for applications requiring dimensional stability, such as in composites or coatings. A key benefit is access to polymers with unique properties, including biodegradability; for instance, ROP of lactones like ε-caprolactone yields poly(ε-caprolactone) (PCL), a hydrolytically degradable polyester used in drug delivery and tissue engineering. The living nature of many ROP processes, such as anionic ROP of cyclic esters, allows for the sequential addition of different monomers to form well-defined block copolymers with precise architectures, enhancing material versatility for advanced applications like self-assembling nanostructures. Despite these strengths, ROP has notable limitations that can complicate its implementation. The process is highly sensitive to impurities, particularly moisture and protic contaminants, which can initiate unwanted side reactions like hydrolysis or chain transfer, reducing molecular weight control and yield. In some systems, such as the polymerization of larger ring lactones or cyclic siloxanes, thermodynamic equilibrium favors ring-chain interconversion, leading to cyclization and lower linear polymer yields unless conditions are optimized to shift the equilibrium. Controlling tacticity and end-group functionality often requires specialized catalysts, as racemization or transesterification can occur in chiral monomers without stereoselective initiators. Furthermore, certain catalysts, like tin(II) octoate commonly used for lactone ROP, raise toxicity concerns for biomedical uses, necessitating additional purification steps.
AspectROPAddition Polymerization (e.g., Radical)Condensation Polymerization (e.g., Step-Growth)
By-productsNoneNone, but potential initiator fragmentsSmall molecules eliminated (e.g., water)
Backbone StructureNo unsaturation; heteroatom-richSaturated carbon chain; possible branchesVariable; often requires functional groups
Volume ChangeMinimal shrinkage or expansionSignificant shrinkage (~20-30%)Variable, often shrinkage
Molecular Weight ControlExcellent in living systems (PDI < 1.5)Moderate (PDI ~1.5-3)Requires high conversion for high MW

Monomers

Cyclic esters and lactones

Cyclic esters, commonly referred to as lactones, are key monomers in ring-opening polymerization (ROP) due to their ability to form biodegradable polyesters. These compounds feature an intramolecular ester linkage, where a hydroxy acid cyclizes, resulting in a general formula represented as a cyclic -O-C(=O)-(CH₂)ₙ- structure, with n determining the ring size. Lactones are categorized by ring size: γ-lactones form 5-membered rings, δ-lactones 6-membered rings, and ε-lactones 7-membered rings. Prominent examples include ε-caprolactone, a 7-membered ε-lactone derived from 6-hydroxyhexanoic acid, and the 6-membered cyclic diesters glycolide (from glycolic acid) and lactide (from lactic acid), which serve as dilactones for producing polyesters with tailored properties. The ring strain in these monomers decreases with increasing ring size—highest in γ-lactones and lowest in ε-lactones—driving their propensity for ROP, as strain relief provides a thermodynamic favorability to polymerization. Additionally, the inherent polarity of the ester functionality in lactones enhances their solubility in polar aprotic solvents like dichloromethane or toluene, which is crucial for controlling ROP conditions and achieving high molecular weights. In terms of reactivity, lactones exhibit a strong preference for coordination-insertion mechanisms involving metal alkoxides or organocatalysts, as well as anionic pathways with strong nucleophiles, due to the electrophilic carbonyl carbon susceptible to nucleophilic attack. This leads to the formation of linear polyesters with ester linkages. A representative case is the anionic ROP of , initiated by an alkoxide, which opens the ring to propagate the chain: \ce{RO^- + \overset{|}{C}O-(CH2)5-O -> RO-(CH2)5-C(=O)O^-} Subsequent propagation involves the end-group adding to additional monomers, yielding poly(ε-caprolactone) (PCL). PCL, characterized by its semicrystalline nature and slow biodegradability, finds extensive use in applications, where it enables sustained release of therapeutics over months via hydrolytic .

Cyclic ethers and other heterocycles

Cyclic ethers, particularly and , serve as key monomers in ring-opening polymerization (ROP) due to their strained structures that facilitate nucleophilic attack and chain propagation. (EO), a symmetrical three-membered with the formula \ce{(CH2)2O}, and (PO), its unsymmetrical counterpart \ce{(CH2)CH(CH3)O}, are the most common epoxide monomers, while represent four-membered cyclic ethers such as unsubstituted \ce{(CH2)3O} or 3-methyloxetane. Nitrogen-containing heterocycles like , the three-membered analogs of epoxides (e.g., unsubstituted \ce{(CH2)2NH} or 2-methylaziridine), extend this class to produce amine-functional polymers. These monomers are selected for ROP based on their inherent , which mirrors that in cyclic esters but yields polyether or backbones instead of polyesters. The high ring strain in these small heterocycles—approximately 110–116 kJ/mol (26–28 kcal/mol) for three-membered epoxides and , and about 25 kcal/mol for four-membered oxetanes—provides the thermodynamic driving force for by relieving angular and torsional distortions upon ring opening. This enables facile ROP under mild conditions, but it also contributes to challenges in monomer handling; for instance, is a volatile gas ( 10.7°C) that requires pressurized storage and careful manipulation to prevent explosive or toxicity risks. Similarly, are highly reactive and toxic, often stabilized with bases like hydroxides for safe industrial handling at scales up to 9000 tons/year as of 2006. In contrast, larger-ring ethers like exhibit lower (~15 kJ/mol) and are less prone to unintended reactions but polymerize more slowly. Reactivity of these monomers is highly dependent on and substituents, with three- and four-membered rings favoring both cationic and anionic pathways. Epoxides like undergo anionic ROP efficiently with initiators to yield linear poly() (PEO), a hydrophilic polymer used in due to its linkages and solubility in . PO, however, shows propensity for in anionic conditions, often requiring coordination catalysts for higher molecular weights, while cationic with Lewis acids (e.g., BF₃) produces atactic poly() (). Oxetanes polymerize primarily via cationic mechanisms with solid acids or Lewis acids, forming polyoxetanes with pendant groups for functional materials. , as nitrogen analogs, are polymerized anionically using N-activated derivatives (e.g., N-sulfonylaziridines) to access linear polyamines, or cationically to generate hyperbranched polyethylenimines (PEI) with primary:secondary:tertiary amine ratios of 1:2:1. These processes highlight the versatility of heterocycle ROP for tailoring polymer architecture and functionality. Other important heterocycles for ROP include cyclic carbonates and lactams. Cyclic carbonates, such as the six-membered trimethylene carbonate (TMC, \ce{(CH2)3(OCO)}), undergo ROP to form aliphatic polycarbonates with carbonate linkages in the backbone, offering biocompatibility and tunable degradation rates. TMC ROP is typically conducted via coordination-insertion or organocatalytic mechanisms, yielding poly(trimethylene carbonate) (PTMC) for biomedical applications like tissue engineering scaffolds due to its flexibility and low toxicity. Lactams, cyclic amides like the seven-membered ε-caprolactam (\ce{(CH2)5C(O)NH}), are polymerized anionically or hydrolytic-activation methods to produce polyamides. The ROP of ε-caprolactam yields Nylon 6, a high-strength engineering plastic used in textiles and automotive parts, with the mechanism involving nucleophilic attack at the carbonyl to open the ring and form amide bonds.

Cyclic siloxanes

Cyclic siloxanes, such as (D4, a four-membered of alternating and O atoms with methyl substituents), are essential monomers for ROP to produce polysiloxanes (silicones). These monomers feature low (around 6-10 kJ/mol for D4) but polymerize due to gain upon ring opening. The ROP is typically anionic, catalyzed by strong bases like KOH, proceeding via nucleophilic attack on to propagate chains with Si-O-Si linkages. This process yields (PDMS), known for its thermal stability, low surface energy, and elasticity, widely used in sealants, lubricants, and medical devices. Recent advances include metal-free organocatalytic ROP to achieve controlled molecular weights and reduce cyclic byproducts.

Historical Development

Early discoveries

The foundations of ring-opening polymerization (ROP) were laid in the late and early through exploratory work on cyclic monomers, particularly epoxides and lactones. In 1929, and H. A. Schweitzer reported the polymerization of using various catalysts to produce poly() (), demonstrating for the first time the potential of ROP to generate high-molecular-weight polyethers from strained three-membered rings. This seminal study involved heating with alkaline catalysts, yielding polymers with molecular weights up to several thousand, though with limited control over chain length and polydispersity. Staudinger's efforts highlighted the in epoxides as a driving force for , influencing subsequent investigations into chain-growth mechanisms. A pivotal advancement came in the early 1930s from Wallace H. Carothers and his collaborators at , who synthesized ε-caprolactone and explored its ROP to form poly(ε-caprolactone) (PCL), one of the first synthetic biodegradable polyesters. Their 1934 publication detailed the thermal and catalytic of ε-caprolactone, revealing an equilibrium between the cyclic and linear chains, with favored at high temperatures due to gains from ring opening. This work, conducted without modern coordination catalysts, produced low-molecular-weight PCL (typically below 10,000 Da) but established ROP of seven-membered lactones as a route to polyesters with potential industrial applications, such as in coatings and adhesives. The research underscored the versatility of ROP for cyclic esters, distinct from the condensation methods Carothers had previously championed for polyamides like . Another key early development was the anionic ROP of ε-caprolactam to produce , invented in 1938 by Paul Schlack at (now ). This process involved hydrolytic or anionic initiation of the seven-membered ring, leading to high-molecular-weight polyamides suitable for fibers and textiles. Commercial production began in 1946 in , marking one of the first large-scale applications of ROP and demonstrating its potential for commodity polymers. The mechanism features nucleophilic attack on the lactam carbonyl, propagating via amide bond formation without byproduct elimination, contrasting with step-growth polyamides. Initial anionic ROP attempts emerged in the 1940s, with Paul J. Flory investigating the base-initiated polymerization of , observing that propagation could proceed with minimal termination or transfer under controlled conditions, foreshadowing living polymerization paradigms. These early experiments, building on Staudinger's foundation, often suffered from uncontrolled side reactions—such as and —resulting in broad molecular weight distributions and low yields (frequently below 50%). Without effective catalysts, reactions required harsh conditions like high temperatures or strong bases, limiting reproducibility and polymer quality in the pre-coordination chemistry era. Despite these hurdles, the pioneering contributions of Staudinger, Carothers, Schlack, and Flory provided critical insights into the and of ROP, setting the stage for more refined techniques in later decades.

Key advancements post-1950s

In the 1950s, significant progress in enabled the stereoselective ring-opening polymerization (ROP) of epoxides, with demonstrating the synthesis of isotactic poly() using aluminum alkyl-based initiators, marking an early application of coordination mechanisms to heterocyclic monomers. This work built on broader advances in organometallic , establishing foundational mechanisms and for ROP processes. The 1963 awarded to and for their development of Ziegler-Natta catalysts profoundly influenced coordination-insertion ROP, inspiring the design of metal-based systems that facilitated controlled propagation in cyclic ester and ether polymerizations by mimicking olefin coordination pathways. Their catalysts, which enabled stereoregular synthesis, provided a template for subsequent ROP innovations, shifting focus toward precise monomer insertion and reduced side reactions. By the 1970s, the concept of living polymerization, pioneered by Michael Szwarc in the 1950s for monomers, was extended to ROP, enabling the synthesis of well-defined polyethers and polyesters with narrow molecular weight distributions and copolymer architectures through controlled , including anionic for oxiranes and oxetanes, and cationic for , as well as coordination-insertion for ε-caprolactone. Researchers such as Yamashita and Boileau applied Szwarc's principles to achieve living anionic ROP of oxiranes and oxetanes, and living cationic ROP of , minimizing termination and transfer reactions to produce polymers with predictable chain lengths. The 1980s saw breakthroughs in ring-opening metathesis polymerization (ROMP) with the introduction of well-defined ruthenium-based catalysts by , who developed alkylidene complexes that tolerated functional groups and enabled living ROMP of strained cycloolefins like , yielding polymers with precise microstructures for . These catalysts, evolving from earlier ill-defined systems, revolutionized ROMP by providing high activity under mild conditions and facilitating aqueous and stereoselective variants. In the 1990s, the commercialization of poly() () via ROP of represented a major industrial milestone, with Dow launching large-scale commercial production in 2002 using tin octoate-catalyzed ROP to produce biodegradable polyesters from renewable resources, achieving molecular weights exceeding 100,000 g/mol for packaging and biomedical applications. This process scaled ROP from to annual capacities over 100,000 tons, emphasizing solvent-free conditions and high purity to meet commercial viability. The introduced enzymatic ROP as a alternative, leveraging s such as Candida antarctica lipase B (immobilized as Novozym 435) to catalyze the of lactones and lactides under mild, metal-free conditions, yielding polyesters with controlled and minimal for biomedical uses. Key advancements included bulk and solvent-free enzymatic processes achieving molecular weights up to 10,000 g/mol, with exponential growth in applications driven by and . Yoshiki Chujo and collaborators advanced ROP through the development of "immortal polymerization" in the late 1980s and 1990s, a variant of living ROP using aluminum catalysts that allowed reversible chain exchange, enabling the incorporation of thousands of units per initiator and the synthesis of high-molecular-weight without termination. This technique, distinct from traditional by permitting initiator regeneration, facilitated multifunctional architectures and influenced subsequent organocatalytic strategies.

Polymerization Mechanisms

Anionic ring-opening polymerization

Anionic ring-opening polymerization (AROP) is a chain-growth process characterized by nucleophilic and , where an active chain-end anion attacks heterocyclic , leading to ring scission and chain extension. typically occurs through the reaction of strong nucleophiles, such as or carbanions, with the to generate the initial active species, often an for ester-based cycles. proceeds via repeated nucleophilic attacks by this anion on subsequent units; for cyclic esters like lactones, the attack targets the carbonyl carbon, resulting in acyl-oxygen cleavage and formation of an extended ester chain, while for cyclic like epoxides, the nucleophile attacks the least substituted carbon, yielding linkages. The propagation step can be represented as: \text{RO}^- + \text{cyclic monomer} \rightarrow \text{RO-(\text{monomer})}_n^- where RO⁻ denotes the alkoxide chain end, and the reaction maintains the anionic character for further additions. Termination is optional and can be induced by proton sources like water or alcohols to yield hydroxyl end-groups, though in living AROP, chains remain active indefinitely, enabling precise control over molecular weight. Common initiators include alkoxides, such as tert-butoxide, which are effective for lactones like ε-caprolactone under mild conditions (e.g., room temperature in THF), producing poly(ε-caprolactone) with predictable molecular weights. For epoxides like , however, transfer reactions to limit control, necessitating stronger, non-coordinating bases like superbases (e.g., t-BuP₄) to achieve high efficiency and suppress side reactions at temperatures around 80–100°C. These catalysts enable metal-free AROP, enhancing and in biomedical applications. AROP exhibits living characteristics, evidenced by linear molecular weight increase with conversion and narrow polydispersity indices (typically PDI < 1.2), facilitating the synthesis of block copolymers by sequential addition. For instance, sequential AROP of ε-caprolactone followed by propylene oxide yields poly(ε-caprolactone)-b-poly(propylene oxide) with defined segments. Additionally, careful selection of initiators and conditions allows stereocontrol, achieving high syndiotacticity in polymers from chiral s like β-butyrolactone.

Cationic ring-opening polymerization

Cationic ring-opening polymerization (CROP) involves the electrophilic initiation of strained cyclic monomers, such as epoxides and cyclic vinyl ethers, using Brønsted or Lewis acids to generate positively charged propagating species. This process contrasts with anionic mechanisms by relying on acid catalysis, which can lead to faster but less controlled polymerization rates. It is particularly suited for monomers with high ring strain, enabling the formation of polyethers and related polymers through nucleophilic attack on activated rings. The mechanism proceeds via initiation, where a proton or Lewis acid coordinates to the monomer's oxygen, facilitating ring opening to form an oxonium or carbocation intermediate. Propagation occurs through the active chain-end mechanism, in which the cationic species attacks another monomer molecule, or via the activated monomer mechanism, where the chain-end nucleophile (e.g., a hydroxyl group) attacks a protonated monomer. A key propagation step involves the oxonium ion at the chain end reacting with an epoxide: \text{R-O}^{+} + \ce{CH2-CH2-O} \rightarrow \text{R-O-CH2-CH2-O}^{+} This extends the chain while regenerating the cationic end, though carbocation rearrangements can occur, complicating stereochemistry. Common catalysts include Lewis acids like BF₃·OEt₂ and Brønsted acids such as triflic acid (CF₃SO₃H), often paired with co-initiators like water or alcohols to form onium ions (e.g., sulfonium or oxonium salts). Onium salts, such as diaryliodonium or triarylsulfonium hexafluorophosphates, serve as photoinitiators for UV-triggered CROP. These systems enable control over molecular weight but face challenges in stereochemistry due to the planar carbocation intermediates, leading to atactic polymers from chiral monomers like propylene oxide. CROP is characterized by rapid propagation rates, making it industrially viable for producing polymers like , but it is prone to side reactions such as back-biting, where the chain-end attacks intra-molecularly to form cyclic oligomers, and via hydride abstraction or collapse. These issues can limit molecular weight control, though "living" CROP is achievable under anhydrous conditions with weakly coordinating anions like SbF₆⁻. The method excels in synthesizing telechelic polymers with functional end-groups, useful for block formation.

Coordination-insertion mechanisms

In coordination-insertion ring-opening polymerization (ROP), a metal coordinates to the , typically a cyclic or , facilitating its insertion into a metal- bond of the growing . This proceeds without the formation of free ions, distinguishing it from cationic or anionic pathways by providing enhanced control over molecular weight and polydispersity. The process begins with the coordination of the 's carbonyl oxygen to the acidic metal center, activating the acyl-oxygen bond for nucleophilic attack by the . Subsequent migratory insertion cleaves the acyl-oxygen bond, elongating the while regenerating the active metal- for . The key propagation step can be represented as: \ce{M-OR + \overset{O}{C}-O-R' -> M-O-C(=O)-R'-OR} where M denotes the metal center, OR is the initiator or propagating group, and the cyclic (e.g., ) is depicted in simplified form. This insertion maintains the metal's , enabling rapid turnover and high catalytic efficiency. Common catalysts include aluminum complexes, such as those with ligands, which offer stereocontrol in the ROP of to produce isotactic polylactide (). For instance, chiral Al(salicylaldiminato) complexes achieve high activity and selectivity by stabilizing the during insertion. Tin(II) compounds, particularly stannous octoate (Sn(Oct)2), are widely employed due to their robustness and commercial availability. These operate via coordination-insertion, with the metal formed from co-initiators, yielding polymers with predictable end groups. Zinc-based catalysts, such as Zn(II) or β-diiminato complexes, provide milder conditions and faster rates, often exhibiting turnover frequencies exceeding 1000 h-1 for ε-caprolactone ROP. A notable feature of this mechanism is its tolerance to functional groups, such as hydroxyl or moieties, which might interfere in ionic polymerizations; the metal center selectively activates the without side reactions. This enables copolymerizations and the of functionalized polyesters. In industrial PLA production, Sn(Oct)2-catalyzed ROP of achieves high molecular weights (up to 150,000 g/mol) under bulk conditions at 180–200°C, contributing to global PLA production of approximately 0.87 million metric tons annually as of 2025. Variants include living ROP, where chain transfer is minimized to afford narrow polydispersity (Đ < 1.1), and immortal ROP, which incorporates exogenous alcohols for controlled chain shuttling, allowing high monomer conversion with multiple active chains per catalyst. Aluminum and zinc systems often support immortal modes, enhancing scalability for biomedical applications.

Ring-opening metathesis polymerization

Ring-opening metathesis polymerization (ROMP) is a specialized variant of that employs olefin metathesis catalysis to convert cyclic alkenes into linear or cyclic polymers containing carbon-carbon double bonds in the backbone. This process is thermodynamically favored by the relief of ring strain in the monomers, such as in bicyclic or medium-sized rings, and is initiated by transition metal carbene complexes that facilitate the rearrangement of C=C bonds. Unlike traditional ROP mechanisms involving heterolytic bond cleavage, ROMP proceeds via a metal-mediated carbene-alkene exchange, enabling precise control over polymer microstructure and molecular weight. The mechanism of ROMP follows the Chauvin pathway, beginning with the coordination of a cyclic olefin to a metal carbene initiator, forming a [2+2] cycloaddition product known as a metallacyclobutane intermediate. This intermediate then undergoes a retro-[2+2] cycloaddition, releasing the ring-opened alkylidene and propagating the chain while inserting the monomer unit. For example, the propagation step can be represented as: \ce{M=CH2 + \overset{\frown}{CH=CH-(CH2)_{n-2}} -> M=CH-(CH2)_{n-2}-CH=CH2} where M denotes the metal center and the cyclic olefin is depicted generically. This cycle continues, with each step driven by the formation and cleavage of metallacyclobutanes, leading to living polymerization when chain transfer and termination are minimized. Seminal work by Schrock and demonstrated this living character using well-defined catalysts, allowing narrow polydispersity indices (PDI ≈ 1.1–1.3) and block copolymer synthesis via sequential monomer addition. Key catalysts for ROMP include Schrock's high-oxidation-state and alkylidene complexes, such as Mo(NAr)(CHCMe₂Ph)(OR)₂ (where Ar = 2,6-i-Pr₂C₆H₃ and R = CMe(CF₃)₂), introduced in 1990, which provide exceptional activity for strained monomers but are highly sensitive to air and moisture due to their oxophilicity. In contrast, Grubbs' -based catalysts, exemplified by the second-generation variant RuCl₂(PCy₃)₂(CHPh) (Grubbs II), offer greater tolerance to air, moisture, and polar functional groups, enabling ROMP in protic media. These systems, developed from 1992 onward, complement Schrock catalysts by broadening substrate scope while maintaining living polymerization traits. Common monomers for ROMP are strained cyclic olefins like (bicyclic, with ≈18 kcal/ strain energy) and cyclooctene (medium ring, ≈2–3 kcal/ strain), where polymerization relieves angular and transannular strain to drive exothermicity (ΔH ≈ -20 to -30 kcal/ per monomer unit). Bicyclic systems such as exhibit rapid polymerization rates (k_p up to 10^4 M⁻¹ s⁻¹ with Schrock catalysts) due to enhanced reactivity from σ-donation in the metallacyclobutane formation. ROMP characteristics include tunable cis/trans ratios in the polymer backbone—achieving >95% cis with certain chiral Schrock catalysts via stereoselective metallacyclobutane formation—and improved functional group tolerance in systems, which handle ethers, alcohols, and esters without deactivation.

Radical ring-opening polymerization

Radical ring-opening polymerization (RROP) is a free- chain-growth method that combines the benefits of with ring scission, typically using monomers featuring a polymerizable unsaturated (e.g., or exomethylene) adjacent to a strained heterocycle. This allows the synthesis of polymers with ester or other heteroatom linkages in the backbone without metal catalysts, offering tolerance to functional groups and compatibility with other radical techniques like or ATRP for control. RROP is particularly useful for producing degradable polyesters from non-strained cyclic monomers. The involves by a species (e.g., from AIBN or peroxides) adding to the exomethylene or of the , forming a carbon-centered adjacent to the ring. This then undergoes rapid β-scission of the strained ring (e.g., in cyclic s), relocating the to end and incorporating the opened unit as an link. Propagation continues with the new adding to another , favoring ring-opening over direct addition due to the exothermicity of scission (ΔH ≈ -20 to -40 kcal/mol). A simplified propagation step for a cyclic (e.g., 2-methylene-1,3-dioxolane) is: \text{R}^\bullet + \ce{(CH2=)C-O-CH2-CH2-O} \rightarrow \text{R-CH2-C(=O)-O-CH2-CH2-O-CH2}^\bullet where the ring opens to form the propagating radical. Termination occurs via standard radical pathways, such as combination or disproportionation. Common monomers include five- or six-membered cyclic ketene acetals (CKAs) like 2,2-dimethyl-1,3-dioxol-4-ene (for acrylic-type polyesters) and 1,3-dioxolan-2-one derivatives, which yield poly(acrylate-co-ester)s with high ring-opening fidelity (>90%) under controlled conditions. Initiators are typical free-radical sources, such as azobisisobutyronitrile (AIBN) at 60–80°C in organic solvents, though controlled RROP using reversible-deactivation techniques achieves low polydispersity (PDI ≈ 1.2–1.5) and block copolymer capability. Challenges include competing vinyl polymerization, which reduces ester content, mitigated by monomer design with high ring-strain or electron-withdrawing groups. RROP enables access to hydrophilic or functional polyesters for biomedical uses, such as drug delivery systems.

Thermodynamics and Kinetics

Thermodynamic driving forces

Ring-opening polymerization (ROP) is primarily driven by the enthalpic relief of in the cyclic , which provides a significant negative contribution to the change (ΔG). For small-ring monomers such as epoxides, the ring energy is approximately 20-25 kcal/mol, leading to an exothermic polymerization (ΔH_p) on the order of -110 to -115 kJ/mol for . This arises from the deviation of bond angles in three-membered rings from the tetrahedral , favoring the opening to form linear chains with more stable bond angles. In contrast, larger rings exhibit minimal , reducing the enthalpic driving force; for example, seven-membered ε-caprolactone has a ΔH_p of -28.8 kJ/mol (-6.9 kcal/mol), while four-membered shows an endothermic ΔH_p of +5.1 kJ/mol, making its thermodynamically unfavorable under standard conditions. The entropic contribution (ΔS_p) to ROP is typically negative due to the loss of translational and rotational freedom when multiple cyclic condense into a single chain, with values ranging from -20 to -50 J/mol·K for most systems. This unfavorable is often outweighed by the enthalpic gain at lower temperatures, but for larger, strainless rings (e.g., macrolactones >12 members), ΔS_p can become positive owing to increased conformational flexibility in the , shifting the driving force toward . The overall thermodynamics are governed by the equation for the standard of per monomer unit: \Delta G_p^\circ = \Delta H_p^\circ - T \Delta S_p^\circ Polymerization proceeds spontaneously when ΔG_p < 0, which holds below the ceiling temperature (T_c). The ceiling temperature represents the equilibrium point where the forward polymerization rate equals the depolymerization rate, resulting in a monomer-to-polymer ratio determined by the initial monomer concentration ([M]0). At T_c, ΔG_p = 0, yielding the simplified derivation T_c ≈ ΔH_p^\circ / ΔS_p^\circ for [M] ≈ 1 M, though more precisely it incorporates concentration dependence as T_c = (ΔH_p^\circ + RT \ln [M]{eq}) / ΔS_p^\circ, where [M]_{eq} is the equilibrium monomer concentration. Factors such as temperature, pressure, and [M]_0 influence T_c; higher pressure favors polymerization by reducing the entropic penalty, while elevated temperatures promote depolymerization. For epoxides like ethylene oxide, T_c exceeds 300°C, enabling high conversions, whereas for γ-butyrolactone, T_c < 0°C limits practical polymerization unless under high pressure or in copolymerizations. Lactones illustrate this gradient: ε-caprolactone has T_c ≈ 261°C at 1 M, supporting efficient ROP, compared to larger rings like pentadecanolide (16-membered), where minimal ΔH_p (+3 kJ/mol) results in a lower T_c driven more by entropy.

Kinetic control and propagation steps

In ring-opening polymerization (ROP), kinetic control is governed by the rates of initiation, propagation, and any termination or transfer processes, which collectively determine the molecular weight, polydispersity, and overall polymer architecture. The propagation step, central to chain growth, follows a typical rate law R_p = k_p [M][P^*], where k_p is the propagation rate constant, [M] is the monomer concentration, and [P^*] is the concentration of active chain ends. This second-order dependence ensures that the reaction rate accelerates with increasing monomer concentration and active centers, enabling efficient polymerization under controlled conditions. Initiation efficiency is crucial for precise control, requiring the initiation rate constant k_i to be at least comparable to or greater than k_p (k_i \geq k_p) to generate active chains quantitatively without lag. In living ROP systems, high initiation efficiency (often near 100%) is achieved with nucleophilic or electrophilic initiators that rapidly activate the monomer, minimizing dormant species and ensuring all chains start growing simultaneously. For instance, in anionic ROP of ε-caprolactone, potassium-based initiators with crown ethers achieve near-complete initiation, leading to uniform chain lengths. Distinguishing living from non-living ROP hinges on the absence or presence of termination (k_t) and chain transfer (k_{tr}) rates. Living ROP proceeds without irreversible termination or transfer (k_t = 0, k_{tr} = 0), allowing chains to remain active indefinitely and enabling block copolymer synthesis through sequential monomer addition. Chain transfer constants, defined as C_{tr} = k_{tr}/k_p, must be low (typically <10^{-4}) for living character; higher values lead to non-living behavior with broader distributions due to premature chain stopping or redistribution. The polydispersity index (PDI or Đ) in living ROP approaches the Poisson limit, approximated as \Đ = 1 + 1/\overline{DP}_n, where \overline{DP}_n is the number-average degree of polymerization, often yielding Đ ≈ 1.1–1.2 for high molecular weights. In contrast, non-living systems exhibit Đ > 1.5 due to transfer-dominated . Propagation rates vary significantly across mechanisms, with anionic ROP typically exhibiting faster kinetics than coordination-insertion mechanisms. For example, in anionic ROP of ε-caprolactone, k_p for free anions can reach 350 L mol^{-1} s^{-1} in THF at 20°C, reflecting highly reactive alkoxide chain ends, whereas coordination ROP with tin(II) octoate catalysts yields k_p on the order of 10^{-4} to 10^{-2} L mol^{-1} s^{-1}, prioritizing selectivity over speed. Temperature profoundly influences these rates via Arrhenius dependence, k_p = A \exp(-E_a / RT), where activation energies E_a range from 30–60 kJ mol^{-1}; elevated temperatures accelerate propagation but can enhance transfer reactions, narrowing the window for living conditions (e.g., optimal at -20 to 50°C for many lactones). In ideal living ROP, molecular weight is directly controlled by the monomer-to-initiator ratio, expressed as M_n = \frac{[M]_0}{[I]_0} \times M_{mon}, assuming complete and no , where [M]_0 and [I]_0 are initial concentrations and M_{mon} is the molecular weight. This linear relationship allows predictable of polymers with targeted M_n from 10^3 to 10^5 g mol^{-1}, as demonstrated in anionic ROP of cyclic esters. Deviations arise from incomplete or minor , but high-fidelity systems maintain narrow distributions.

Applications and Industrial Uses

Biomedical and pharmaceutical applications

Ring-opening polymerization (ROP) of cyclic esters such as , glycolide, and ε-caprolactone yields polyesters like , , and poly(ε-caprolactone) (PCL), which are prized for their and biodegradability in biomedical contexts. These polymers degrade via into non-toxic byproducts, making them suitable for temporary implants and drug carriers. For instance, PLA and PGA are employed in resorbable sutures and orthopedic implants, where their mechanical strength supports tissue healing before complete degradation. A prominent example is Vicryl sutures, composed of a PLA-PGA copolymer synthesized via ROP, which provide tensile strength for 2-3 weeks post-implantation and fully degrade within 60-90 days, eliminating the need for surgical removal; these received FDA approval in the 1970s. In drug delivery, PCL's slow degradation rate—tunable from months to years through ROP control of molecular weight—enables sustained release from implants like the Capronor contraceptive device, which maintains efficacy for up to two years. PLA and PLGA nanoparticles, formed into micelles via ROP, encapsulate hydrophobic drugs such as β-lapachone for targeted cancer therapy, improving solubility and bioavailability while minimizing systemic toxicity. In , ROP-derived PLA/PCL scaffolds support and for and repair, with porous structures fabricated to mimic extracellular matrices. Degradation rates can be precisely adjusted by varying ratios or chain during ROP, allowing synchronization with tissue regrowth. Notably, stereoregular PLA produced through enantioselective ROP enables chiral particles that exhibit enantio-differentiating release of drugs like naproxen, preserving stereochemical integrity for enhanced therapeutic specificity. FDA approvals for PLGA-based systems date back to 1989 (e.g., Lupron Depot), underscoring their established safety profile.

Commodity and specialty polymers

Ring-opening polymerization (ROP) enables the production of several commodity polymers with widespread industrial applications. Poly(lactic acid) (), synthesized via ROP of derived from renewable resources such as , serves as a biodegradable alternative to petroleum-based plastics in , fibers, and films. Global PLA production exceeded 500,000 tons annually by , driven by its compostability and reduced compared to traditional plastics, which can lower carbon footprints by up to 25% through substitution. Polyethers derived from the ring-opening polymerization of , such as (PEG) and fatty alcohol ethoxylates, serve as key nonionic in detergents, enhancing cleaning efficiency in household and industrial formulations on a multimillion-ton scale. Specialty polymers from ROP offer tailored properties for niche uses. Polynorbornene, produced by ring-opening metathesis polymerization (ROMP) of , finds application in tires and optical components due to its high optical transparency, low , and rubber-like elasticity after . A prominent example is Norsorex, a hydrogenated polynorbornene rubber developed via ROMP and commercialized by for tire sidewalls and vibration-dampening parts, providing superior impact resistance and aging stability. Polyamides, such as -6 produced by ring-opening polymerization of ε-caprolactam (primarily hydrolytic), are vital in textiles and automotive components, with global production contributing over 50% of the nylon market for durable fibers and molded parts. Industrial processes highlight ROP's scalability for fibers and rubbers. ROP of ε-caprolactone yields poly(ε-caprolactone) (PCL), a biodegradable extruded into fibers for composites and textiles, with commercial production emphasizing melt processing for controlled degradation profiles. ROMP followed by produces specialty rubbers like hydrogenated polynorbornene, used in high-performance elastomers for and belts due to enhanced thermal and oxidative resistance. Scaling cationic ROP processes presents challenges, including thermodynamic hurdles and slow that result in prolonged reaction times and limited molecular weight control, hindering large-scale adoption for commodity production.

References

  1. [1]
    Ring-Opening Polymerization—An Introductory Review - MDPI
    This short, introductory review covers the still rapidly growing and industrially important field of ring opening polymerization (ROP).
  2. [2]
    Organocatalytic Ring-Opening Polymerization | Chemical Reviews
    Nov 8, 2007 · In this review, we highlight some of the important advances in organocatalytic polymerization reactions and the utility of organocatalytic methods.Cationic Ring-Opening... · Anionic Ring-Opening... · Enzymatic Ring-Opening...
  3. [3]
    Coordination Ring-Opening Polymerization of Cyclic Esters - NIH
    Ring-opening polymerization (ROP) of cyclic esters (lactones, lactides, cyclic carbonates and phosphates) is an effective tool to synthesize biocompatible ...
  4. [4]
    Recent advances in enantioselective ring-opening polymerization ...
    Sep 29, 2023 · This review focuses on recent advances in developing enantioselective catalysts in ROP and ROCOP to produce degradable polymers from racemic ...
  5. [5]
  6. [6]
  7. [7]
  8. [8]
  9. [9]
  10. [10]
    Poly-є-caprolactone based formulations for drug delivery and tissue ...
    Feb 28, 2012 · This review aims to provide an up to date of drugs incorporated in different PCL based formulations, their purpose and brief outcomes.Review · 3. Pcl Microspheres · 5. Pcl Scaffolds And...
  11. [11]
    Polymerization of Ethylene Oxide, Propylene Oxide, and Other ...
    Dec 29, 2015 · To avoid this issue, cationic ring-opening polymerization is often employed for polymerizing disubstituted epoxide monomers like cyclohexene ...
  12. [12]
    Aziridines and azetidines: building blocks for polyamines by anionic ...
    May 13, 2019 · This review highlights the recent advances on the polymerizations of aziridine and azetidine. It provides an overview of the different routes to produce ...
  13. [13]
    Polyethylene Oxides - an overview | ScienceDirect Topics
    In 1929 Staudinger and Schweitzer synthesized a series of PEG from ethylene oxide as the monomer and separated the polymers on the basis of their molecular ...
  14. [14]
    Living Polymerization—Emphasizing the Molecule in Macromolecules
    Sep 17, 2017 · Staudinger described polymerization as a chain reaction in 1935 (“Über die Polymerization als Kettenreaktion”), (8) but while typical ...
  15. [15]
    [PDF] Thermodynamics and Kinetics of Ring - Opening Polymerization 1
    Cyclic monomers that have been polymerized via ring - opening encompass a variety of structures, such as alkanes, alkenes, compounds containing heteroatoms.
  16. [16]
    Stereoselective Epoxide Polymerization and Copolymerization
    Significant advances have been made in stereocontrolled epoxide polymerization by well-defined catalysts that allow for the extraordinarily fast production of ...Missing: post- | Show results with:post-
  17. [17]
    The Influence of Ziegler-Natta and Metallocene Catalysts on ...
    Ziegler and Natta are both awarded the Nobel Prize for Chemistry 1963 in recognition of their work on the Ziegler-Natta catalyst. 1957, Commercial production ...
  18. [18]
    Ziegler-Natta catalysis: 50 years after the Nobel Prize | MRS Bulletin
    Mar 13, 2013 · This introductory article will provide a short historical review of the discovery of catalyzed olefin polymerization by Ziegler, Natta, and others<|separator|>
  19. [19]
    Living Ring-Opening Polymerization of Heterocyclic Monomers
    Aug 6, 2025 · These processes, conceptually, are based on the idea developed by Szwarc after discovering living conditions for anionic vinyl polymerization.
  20. [20]
    The Development of L2X2RuCHR Olefin Metathesis Catalysts
    Starting in the late 1980s, the potential of ruthenium catalysts for ROMP applications was reexamined. ... The Development of Functional Group Tolerant ROMP ...
  21. [21]
    [PDF] Robert H. Grubbs - National Academy of Sciences
    Grubbs. 1988. Catalytic Organometallic Chemistry in Water: The. Aqueous Ring-Opening Metathesis Polymerization of 7-Oxanorbornene Derivatives. J. Am. Chem ...
  22. [22]
    Cargill Rearrangement - an overview | ScienceDirect Topics
    Polylactic acid is a rigid thermoplastic polyester from renewable resources. It was commercialized by Cargill in 1994 (Table 2). The production of polylactic ...
  23. [23]
    About NatureWorks
    What began in 1989 as a Cargill research project looking for innovative uses of carbohydrates from plants as feedstock for more sustainable plastics, ...Missing: ROP 1990s
  24. [24]
    Recent developments in enzyme-catalyzed ring-opening ...
    An exponential growth has been observed in the last decade where enzymes were used as catalysts for polymerization of different monomers and due to enzyme's ...
  25. [25]
    Enzymatic Ring-Opening Polymerization (ROP) of Polylactones - NIH
    Some lactones are naturally occurring saturated or unsaturated γ- and δ-lactones, which are cyclic esters of the corresponding hydroxy fatty acids. But more ...
  26. [26]
    Synthesis of star block copolymers by controlled ring-opening ...
    Yoshiki Chujo, Akio Naka, Martina Krämer, Kazuki Sada, Takeo Saegusa. ... Living and immortal polymerizations. Progress in Polymer Science 1994, 19 (3) ...
  27. [27]
    Living ring-opening polymerizations of heterocyclic monomers
    Mechanisms of living ring-opening polymerizations (LROP), including anionic, coordination, cationic, activated monomer, and polymerization in dispersed media ...
  28. [28]
    Kinetics and mechanisms in anionic ring‐opening polymerization
    Recent advances in the anionic ring-opening polymerization (AROP), including covalent (pseudoanionic) polymerization, are reviewed. Thermodynamics, kinetics ...Missing: early | Show results with:early
  29. [29]
    From activated or metal-free anionic ring-opening polymerization of ...
    Flory observed in the 1940s that propagation may proceed without side reaction in anionic polymerization of ethylene oxide [1]. Alkali metal derivatives ...
  30. [30]
    Activation in anionic polymerization: Why phosphazene bases are ...
    Phosphazene bases, which are characterized by an exceptional basicity, have been used in the domain of anionic polymerization.
  31. [31]
    Cationic Ring-Opening Polymerization - ScienceDirect.com
    Cationic ring-opening polymerization is defined as a nucleophilic substitution reaction involving the polymerization of heterocyclic monomers, where growing ...Missing: paper | Show results with:paper
  32. [32]
    Ring Opening Polymerisation - an overview | ScienceDirect Topics
    ROP has obvious advantages in processing technologies, for example, reactive injection molding, over preformed high-molecular-weight polymers due to the low ...
  33. [33]
    Coordination Insertion Mechanism of Ring‐Opening Polymerization ...
    Feb 28, 2021 · In coordination-insertion ROP of lactide, complexes can interconvert very fast by the refresh resulting from countless intermolecular collisions ...
  34. [34]
    Review – recent development of ring-opening polymerization of ...
    In this review, the mechanism for ring-opening polymerizations of cyclic ester, synthesis of aluminum complexes containing tetradendete ligand, tridentate ...Missing: early history
  35. [35]
    Are Well Performing Catalysts for the Ring Opening Polymerization ...
    Jun 8, 2022 · In Scheme 2, we report the ROP cycle catalyzed by a Zn–OCH3 catalyst. The ROP operates with a coordination–insertion mechanism in which the ...
  36. [36]
    Multinuclear metal catalysts in ring-opening polymerization of ε ...
    Jan 15, 2023 · Coordinated CL shifted to external Al, and then the alkoxide initiator initiated to CL to open the CL ring and become a new initiator (Fig.Missing: equation | Show results with:equation
  37. [37]
    (PDF) Metal-Catalyzed Immortal Ring-Opening Polymerization of ...
    Aug 6, 2025 · This Perspective article summarizes efforts paid in our group to develop efficient metal-based catalysts for the immortal ring-opening ...
  38. [38]
    [PDF] Richard R. Schrock - Nobel Lecture
    The molybdenum bisalkoxide catalysts are also remarkably active for a wide range of metathesis reactions, again especially when the alkoxide is the high- ly ...
  39. [39]
    None
    ### Summary of Grubbs Catalysts for Ring-Opening Metathesis Polymerization (ROMP)
  40. [40]
  41. [41]
  42. [42]
    Thermodynamic Presynthetic Considerations for Ring-Opening ...
    Jan 21, 2016 · The increased order of the system upon polymerization will finally outweigh the ring strain, making the transition from monomer to polymer ...
  43. [43]
  44. [44]
    Medical applications and prospects of polylactic acid materials - PMC
    This review summarizes current applications of the PLA-based biomaterials in drug delivery systems, orthopedic treatment, tissue regenerative engineering, and ...
  45. [45]
    Poly(lactic acid)/poly(lactic-co-glycolic acid)-based microparticles
    Jun 11, 2019 · Vicryl®, a bioresorbable suture based on PLGA, was launched on the market by Ethicon in 1974 (Ratner et al. 2013) (Fig. 2). In the early 1970s, ...
  46. [46]
    Polycaprolactone: How a Well-Known and Futuristic Polymer Has ...
    Jan 20, 2020 · It was first synthesized in the early 1930s by ring-opening polymerization of the cyclic monomer of Ɛ-caprolactone. ... PCL and PCL-based ...
  47. [47]
    β-Lapachone-containing PEG-PLA Polymer Micelles as Novel ...
    The results indicate that the drug is encapsulated within the PLA micelle core and the micelle surface is stabilized with a mobile PEG corona.2.2 β-Lap Micelle... · 3. Results · 3.3 β-Lap Micelle...
  48. [48]
    Polylactide-based chiral particles with enantio-differentiating release ...
    Jul 15, 2018 · PLA-based chiral particles were successfully prepared and demonstrated enantioselectivity in releasing chiral drug, naproxen.
  49. [49]
    Polylactic Acid (PLA) Market Size, Price, Demand and Forecast, 2034
    Aug 6, 2024 · The polylactic acid (PLA) market was 512 thousand tons in 2023 and is projected to reach 2784 thousand tons by 2036, growing at a CAGR of ...
  50. [50]
    Bio-plastics VS Petroleum-based Plastics | Civic Issues Blog - SP 23
    Apr 27, 2023 · A 2017 study determined that switching from traditional plastic to corn-based PLA would cut U.S. greenhouse gas emissions by 25 percent. The ...
  51. [51]
    Polynorbornene - an overview | ScienceDirect Topics
    Applications for these polymers are flat panel displays89 and optical wave guides, both of which are accessible due to the high optical transmission and low ...
  52. [52]
    Synthesis of hydrogenated functional polynorbornene (HFPNB) and ...
    Norbornene and its derivatives have been polymerized in three different ways: (1) ring-opening metathesis polymerization (ROMP), (2) vinylic polymerization, and ...
  53. [53]
    Beyond nylon 6: polyamides via ring opening polymerization of ...
    Sep 1, 2022 · Ring opening polymerization (ROP) of lactams is a highly efficient and versatile method to synthesize polyamides.Missing: cyclic | Show results with:cyclic
  54. [54]
    Poly (ε-caprolactone) Fiber: An Overview - Sage Journals
    The present article presents a review on the chemistry and different properties of PCL, production of PCL fiber by various methods and correlations between.
  55. [55]
    S/O and vinyl isomerization enables ultrafast cationic ring-opening ...
    Oct 29, 2025 · However, their general ring-opening polymerizations face challenges with thermodynamics and kinetics, generally resulting in long reaction times ...Introduction · Results · Methods