Fact-checked by Grok 2 weeks ago

Step-growth polymerization

Step-growth polymerization is a fundamental method of synthesizing polymers in which bifunctional or multifunctional monomers react stepwise with one another through their reactive functional groups to form dimers, trimers, and progressively longer chains, ultimately yielding high-molecular-weight polymers, often but not always with the elimination of small byproduct molecules such as . This process typically involves the formation of carbon-heteroatom bonds (e.g., or linkages) via reactions between groups like carboxylic acids and alcohols or amines. Unlike chain-growth polymerization, which proceeds rapidly through sequential addition of monomers to an active chain end initiated by a reactive species and forms carbon-carbon bonds, step-growth polymerization builds polymers slowly through reactions between any two molecules bearing complementary functional groups, regardless of chain length, resulting in a broader distribution of chain lengths and typically lower initial molecular weights. The kinetics of step-growth reactions generally follow a second-order rate law, with the rate depending on the concentrations of the reacting functional groups, and the process may be catalyzed by the reacting functional groups (e.g., carboxylic acids in esterifications). Prominent examples of step-growth polymers include polyesters such as () formed from diols like and dicarboxylic acids like , and polyamides such as produced from and . The (DP) is described by the , DP = 1 / (1 - p), where p is the ; achieving high molecular weights requires p to approach 1 (typically >0.99), often facilitated by removing byproducts or using stoichiometric control of monomer ratios. These polymers exhibit strong intermolecular forces, such as hydrogen bonding, leading to desirable properties like high tensile strength and crystallinity, making them essential in applications ranging from textiles and fibers to engineering plastics.

Overview and History

Definition and Basic Principles

Step-growth polymerization is a polymerization mechanism in which monomers possessing two or more reactive functional groups undergo successive intermolecular reactions to form dimers, trimers, and larger oligomers, which continue to couple until high-molecular-weight polymers are obtained. This process relies on the reactivity between complementary functional groups, such as hydroxyl (-OH) and carboxyl (-COOH), without requiring a separate initiator or involving distinct phases of chain initiation and propagation. Instead, chain growth occurs gradually through step-wise coupling of any two species—monomers, oligomers, or polymers—that bear appropriate functional groups. The fundamental principle of step-growth polymerization emphasizes the progressive nature of bond formation, where the molecular weight increases incrementally with each reaction step. For bifunctional monomers, which carry exactly two reactive sites (often denoted as A and B groups in an AB-type monomer), the initial reaction forms a dimer that can further react to yield linear polymer chains. This can be illustrated by the general reaction scheme: \ce{A-B + A-B ->[catalyst] A-B-A-B + by-product} Subsequent steps extend the chain similarly, as the terminal functional groups on oligomers remain reactive throughout the process. When monomers with more than two functional groups are employed, the potential for branching arises, as additional reactions can occur at non-terminal sites, leading to more complex architectures. Representative examples of step-growth polymerization include the synthesis of polyesters via esterification between a diol and a diacid, where the hydroxyl groups of the diol react with the carboxyl groups of the diacid to form ester linkages, typically eliminating water as a byproduct. Another common case is polyamidation, as seen in the production of polyamides like Nylon 6,6, where a diamine reacts with a diacid to create amide bonds through similar condensation. These reactions highlight the versatility of step-growth mechanisms in forming covalent linkages central to many engineering polymers.

Historical Development

The study of natural polymers such as and in the laid the groundwork for synthetic analogs by highlighting their repetitive molecular structures and potential for chemical modification. Chemists like Anselme Payen analyzed as a distinct polymeric substance in , while 's protein was investigated for its linkages, inspiring efforts to replicate these properties synthetically through reactions. These observations, though empirical, demonstrated how natural macromolecules could serve as models for durable materials, prompting early experiments with derivatives like in the 1840s by Christian Schönbein. The first fully synthetic step-growth polymer emerged in 1907 with Leo Baekeland's invention of Bakelite, a phenol-formaldehyde resin produced via condensation polymerization, marking the transition from modified natural materials to true synthetics. Baekeland's controlled reaction between phenol and formaldehyde under heat and pressure yielded a thermosetting material, commercialized by 1910 and widely used in electrical insulators and consumer goods. This breakthrough validated step-growth mechanisms for creating high-molecular-weight substances, influencing subsequent research. Meanwhile, Hermann Staudinger's work in the 1920s established the macromolecular hypothesis, proposing that polymers like rubber and cellulose consist of long covalent chains rather than aggregates, a theory initially met with skepticism but pivotal for rational synthesis. Staudinger's 1920 paper on rubber and his 1922 lectures formalized this view, earning him the 1953 Nobel Prize in Chemistry. In the late 1920s, Wallace Carothers at DuPont conducted the first systematic studies of step-growth polymerization, synthesizing polyesters from diols and dicarboxylic acids in 1930 and demonstrating their linear chain growth. Carothers derived the foundational Carothers equation relating monomer conversion to degree of polymerization, providing a quantitative framework for the process. His team produced the first nylon polyamide, nylon 6,6, on February 28, 1935, from hexamethylenediamine and adipic acid, leading to commercial production at DuPont's Seaford plant on December 15, 1939, and public availability in 1940 for stockings and parachutes. Concurrently, Paul Flory, joining DuPont in 1934, advanced the theoretical foundations during the 1930s and World War II era, developing statistical models for molecular weight distributions and gelation in step-growth systems through papers in the early 1940s, as detailed in his 1953 book Principles of Polymer Chemistry. Flory's work on polycondensation kinetics, including the 1946 review in Chemical Reviews, shifted understanding from empirical trial-and-error to mechanistic predictions. The 1940s saw further milestones with the development of polyurethanes by Otto Bayer at IG Farben in 1937, using diisocyanates and polyols in a step-growth process; initial commercialization as coatings and foams occurred in Germany by 1942, expanding post-war for foams and elastomers. These innovations built on Carothers' and Flory's frameworks, enabling diverse applications. Post-1950, advances in spectroscopic techniques and computational modeling facilitated a deeper mechanistic understanding, moving beyond kinetic models to elucidate reaction pathways and side reactions in step-growth systems, as evidenced by Flory's ongoing contributions and the field's maturation into a cornerstone of materials science.

Mechanisms and Types

Condensation Polymerization

Condensation polymerization involves the stepwise reaction between monomers bearing complementary functional groups, resulting in the formation of covalent bonds and the simultaneous elimination of small byproduct molecules, typically water or other low-molecular-weight species. This process forms the backbone of many step-growth polymers, where each reaction step links oligomers or monomers of varying sizes, progressively building chain length. The mechanism relies on nucleophilic attack by one functional group on another, such as the hydroxyl group of an alcohol attacking the carbonyl carbon of a carboxylic acid to form an ester linkage, with proton transfers facilitating the departure of the byproduct. A in these reactions is the removal of the eliminated byproduct, which shifts the toward formation according to . For instance, in polyesterification, is distilled off under reduced or high to prevent reversal of the reaction and promote higher conversions. Without such removal, the often favors monomers or low-molecular-weight oligomers, limiting yield and chain length. Common reactions include polyesterification, where diols react with diacids to produce polyesters; polyamidation, involving diamines and diacids to form polyamides; and polyanhydride formation from diacids activated by anhydrides. In polyesterification, the balanced equation is: n \ce{HO-R-OH} + n \ce{HOOC-R'-COOH} \rightarrow \ce{[-O-R-OC(O)-R'-C(O)-]_n} + 2n \ce{H2O} This reaction, exemplified by the synthesis of polyethylene terephthalate from ethylene glycol and terephthalic acid, was foundational in Wallace Carothers' work at DuPont in the 1930s. For polyamidation, as in nylon 6,6 production, the equation is: n \ce{H2N-(CH2)6-NH2} + n \ce{HOOC-(CH2)4-COOH} \rightarrow \ce{[-HN-(CH2)6-NHC(O)-(CH2)4-C(O)-]_n} + 2n \ce{H2O} This yields durable materials like textiles and fibers. Polyanhydrides, used in biomedical applications for their biodegradability, form via melt condensation with acetic anhydride: n \ce{HOOC-R-COOH} + n \ce{(CH3CO)2O} \rightarrow \ce{[-OC(O)-R-C(O)-O-]_n} + 2n \ce{CH3COOH} Here, acetic acid is removed under vacuum to drive the process. Successful requires monomers with at least two complementary functional groups per molecule to enable linear or branched chain extension, such as bifunctional diols pairing with bifunctional diacids. These groups must be reactive toward each other without excessive side reactivity, and stoichiometric balance is critical to achieve high molecular weights. Monomers lacking such compatibility, like those with only one reactive site, cannot propagate the chain. Despite these principles, condensation polymerization has limitations, particularly its sensitivity to side reactions at low conversions. Cyclization, where reactive end groups on the same or intramolecularly react to form rings, competes with intermolecular chain growth, reducing the of high-molecular-weight linear polymers and favoring cyclic s. This issue is pronounced in dilute solutions or early stages, often necessitating high concentrations and precise control to minimize it.

Non-Condensation Step-Growth Processes

Non-condensation step-growth polymerization encompasses mechanisms where bifunctional monomers react through direct displacements or couplings without the loss of small molecules like water or alcohol, distinguishing it from classical condensation processes. These reactions typically involve nucleophilic substitutions or oxidative processes that form covalent bonds while incorporating all atoms from the monomers into the polymer chain, often with ionic or trapped byproducts that do not require extensive removal. This approach enables the synthesis of polymers such as polyethers, polysulfones, polyurethanes, and conjugated systems with high purity and controlled architectures. A prominent example is the formation of polyethers via the Williamson synthesis, where an alkyl halide reacts with an alkoxide ion in a stepwise nucleophilic substitution. The general reaction is represented as: \text{R-X} + \text{R'-O}^- \rightarrow \text{R-O-R'} + \text{X}^- This process proceeds under basic conditions, such as with sodium alkoxides in polar aprotic solvents, yielding linear polyethers like poly(alkylene oxides) without volatile byproducts beyond the halide salt. Model compounds, such as H[(CH₂)ₓO]ᵧ(CH₂)ₓH where x=4–10 and y=2–4, have been synthesized to demonstrate the step-growth nature, achieving high conversions through iterative dimerization and chain extension. Arylene/alkylene polyethers with pendant groups, formed from dihydric phenols and dihalides like 1,4-dichlorobutane, further illustrate the versatility, producing polymers with intrinsic viscosities up to 0.094 dL/g. Another important class is polyurethanes, formed by the step-growth reaction between diisocyanates and diols or polyols without elimination of small molecules. The general reaction is: n \ce{OCN-R-NCO} + n \ce{HO-R'-OH} \rightarrow \ce{[-OCN H-R-NHC(O)O-R'-O-]_n} This proceeds via nucleophilic attack of the alcohol on the carbon, forming linkages. Polyurethanes are widely used in foams, coatings, adhesives, and elastomers due to their versatility and properties like elasticity and toughness. Polysulfones are synthesized through nucleophilic aromatic substitution (SNAr), where an activated aryl fluoride or chloride reacts with a phenoxide. A typical mechanism involves the displacement at an electron-deficient aromatic ring, as in: \text{Ar-F} + \text{HO-Ar'} \rightarrow \text{Ar-O-Ar'} + \text{HF} The HF byproduct is often trapped by base, minimizing elimination issues and enabling high molecular weights. This step-growth method uses dialkali salts of dihydric phenols with dihalodiphenyl sulfones in dipolar aprotic solvents like dimethyl sulfoxide, yielding poly(aryl ether sulfones) with excellent thermal stability. Self-polycondensation of AB-type monomers, such as hydroxyphenyl sulfone halides, further exemplifies this, proceeding via iterative SNAr to form ether linkages without additional catalysts in some cases. Oxidative polymerization of dithiols provides another non-condensation route, forming poly(disulfide)s through the coupling of thiol groups. The mechanism involves oxidation to thiyl radicals or direct disulfide formation, as in 2 RS⁻ → RS-SR, often catalyzed by air, hydrogen peroxide, or bases like triethylamine under ambient conditions. This living polymerization of dithiols like 2-[2-(2-sulfanylethoxy)ethoxy]ethanethiol achieves number-average molecular weights up to 250,000 g/mol with polydispersity indices as low as 1.15, producing water as the only byproduct in oxygen-mediated reactions. Metal-catalyzed couplings, such as the Sonogashira reaction, enable the step-growth synthesis of polyynes and conjugated polymers. In this process, a terminal alkyne couples with an aryl or vinyl halide using Pd/Cu catalysis: \text{Ar-X} + \text{HC≡C-Ar'} \rightarrow \text{Ar-C≡C-Ar'} + \text{HX} The HX is neutralized by base, avoiding small-molecule loss, and the reaction proceeds via oxidative addition and transmetalation cycles to form carbon-carbon triple bonds. This method constructs poly(aryleneethynylene)s from dihalides and diynes, yielding materials for optoelectronics with tunable conjugation lengths. These non-condensation processes offer advantages including cleaner reaction profiles and reduced need for byproduct distillation, facilitating scalable production and end-group control. However, challenges arise from the requirement for functional group orthogonality, as competing reactions can lead to branching or gelation if halide or thiol reactivities are not balanced.

Comparison with Chain-Growth Polymerization

Fundamental Differences

Step-growth polymerization builds polymer chains through the random reaction of functional groups on any two species present in the system, including monomers, oligomers, and longer chains, leading to stepwise elongation without a dedicated propagating site. In contrast, chain-growth polymerization proceeds via the sequential addition of monomers exclusively to active centers at chain ends, such as radicals, carbanions, or cations, forming a chain reaction mechanism. A fundamental distinction lies in initiation: step-growth reactions do not require an external initiator, as they rely solely on the inherent reactivity of complementary functional groups on bifunctional (or higher) monomers to commence oligomer formation. Chain-growth polymerization, however, demands an initiator or catalyst to generate the initial active species, which then propagates by adding monomers until termination or transfer occurs. The growth patterns differ markedly in terms of chain length development and molecular weight distribution. In step-growth, polymerization yields a broad distribution of chain lengths from the outset, as any reactive species can couple with another, resulting in exponential increases in average chain length only at high conversions. Chain-growth, especially in without termination, features simultaneous and uniform growth of all chains from active centers, producing polymers with a narrow molecular weight distribution early in the process. Key kinetic contrasts highlight these differences in monomer consumption. Step-growth follows second-order kinetics with respect to functional group concentration, causing a rapid initial drop in monomer levels that decelerates later; the integrated rate law is \frac{1}{[M]} = kt + \frac{1}{[M]_0} where [M] is the functional group concentration, k is the rate constant, t is time, and [M]_0 is the initial concentration, with the extent of reaction given exactly by p = 1 - \frac{1}{1 + k t [M]_0} and for early stages p \approx k t [M]_0. In chain-growth, propagation is typically first-order in monomer, resulting in rapid exponential decay of monomer concentration: [M] = [M]_0 e^{-k t} with high molecular weights achieved even at low monomer conversions. Structurally, step-growth is particularly susceptible to branching when multifunctional monomers (functionality f > 2) are employed, as additional reactive groups enable random intermolecular linkages that can culminate in crosslinked networks, as quantified by Flory-Stockmayer theory. Chain-growth tends to favor linear architectures unless branching agents are deliberately introduced.

Practical Implications

In step-growth polymerization, achieving high molecular weights necessitates conversions exceeding 99%, as the scales inversely with the extent of unreacted functional groups, contrasting with chain-growth methods where high molecular weights are attainable at much lower conversions of around 10-20%. This requirement stems from the random coupling of oligomers, demanding precise stoichiometric control to minimize low-molecular-weight species. Consequently, synthesis strategies for step-growth polymers emphasize shifts through byproduct removal, such as in formation, whereas chain-growth relies on rapid propagation from active centers. Step-growth polymers typically retain reactive functional end-groups, like carboxylic acids or amines, enabling post-polymerization modifications such as cross-linking or , which enhances versatility in tailoring properties for specific applications. However, these systems are highly sensitive to impurities, including trace or monofunctional contaminants, which disrupt and cap chain growth prematurely, leading to brittle materials with inferior mechanical strength. In comparison, chain-growth polymers often feature inert or less reactive ends, reducing such vulnerabilities but limiting derivatization options. Industrially, step-growth processes are well-suited for melt-phase reactions, facilitating the production of polyesters like () for bottles and films, where high-temperature avoids solvents. Nylon-6,6, a step-growth , exemplifies durability through hydrogen bonding, offering superior tensile strength (up to 80 MPa) and for textiles and gears, though its recyclability is complicated by sensitivity. Conversely, chain-growth is processed via solution or , yielding highly recyclable films and containers with excellent chemical inertness but lower thermal stability (melting at 110-130°C). These differences influence scalability: step-growth demands slower rates and energy-intensive purification to achieve purity levels above 99.9%, increasing operational costs compared to the faster, solvent-based chain-growth routes.

Polymer Architectures

Linear Step-Growth Polymers

Linear step-growth polymers are synthesized exclusively using difunctional monomers, each containing precisely two reactive functional groups, ensuring the formation of unbranched, linear chain architectures. These monomers typically react in an A-A + B-B fashion, where one monomer bears two identical A functional groups (e.g., carboxylic acids) and the complementary monomer has two B groups (e.g., hydroxyls), or in an A-B configuration where each monomer carries one A and one B group. The stepwise reaction between these functional groups progressively builds linear chains denoted as -[A-B]-_n, with each step involving the coupling of any two reactive species—monomers, oligomers, or polymers—without the need for initiators or chain carriers. The structure of these polymers features sequences of repeating units that alternate in A-A + B-B systems, promoting regular arrangements, or appear more randomly in A-B systems due to self-condensation possibilities. ends terminate with unreacted functional groups, such as -COOH or -OH, which retain reactivity and influence further extension or end-capping reactions. This linear contrasts with more complex architectures and supports like processability in the melt state. Key characteristics of linear step-growth polymers include their potential for high crystallinity, driven by intermolecular interactions like hydrogen bonding, as seen in polyamides such as nylons, which enables superior mechanical performance. The average degree of polymerization, \bar{DP}_n, is described by the Carothers equation for stoichiometric systems: \bar{DP}_n = \frac{1}{1 - p} where p represents the extent of reaction; high molecular weights require p > 0.99, emphasizing the need for precise stoichiometry. Representative examples include polyethylene terephthalate (PET), formed via condensation of terephthalic acid and ethylene glycol, which exhibits a tensile strength of approximately 55–75 MPa and is widely used in fibers and bottles due to its clarity and strength. Similarly, nylon 6,6, produced from hexamethylenediamine and adipic acid, demonstrates high crystallinity from amide hydrogen bonding, yielding a tensile strength around 80 MPa and exceptional toughness for applications in textiles and engineering plastics. A significant advantage of linear step-growth polymerization lies in the facile control of chain-end functionality, enabling the production of telechelic polymers with predefined reactive groups at both termini for subsequent or modification, which enhances versatility in block and .

Branched and Network Polymers

In step-growth polymerization, the introduction of monomers with functionality greater than two leads to branched architectures, diverging from the linear chains formed by strictly bifunctional monomers. Trifunctional or higher monomers, such as alcohols or acids with three or more reactive groups, enable chain extension in multiple directions, resulting in dendritic or randomly branched structures. This multifunctionality promotes the formation of branch points during the stepwise reaction of functional groups, increasing the complexity of the polymer topology. A critical aspect of branching is the onset of gelation, where the polymer system transitions from soluble, finite molecules to an insoluble, infinite . According to the Flory-Stockmayer theory, the gel point occurs when the branching coefficient \alpha, defined as the probability that a on a branch unit has reacted to form another branch, reaches \alpha_c = \frac{1}{f_{\text{avg}} - 1}, with f_{\text{avg}} being the average functionality of the monomers. For trifunctional monomers (f = 3), this critical value is \alpha_c = 0.5, meaning gelation happens at 50% conversion of the branching units under ideal conditions. This theory assumes random reaction statistics and no cyclization, providing a foundational model for predicting gelation in systems like polyesters formed from trifunctional and difunctional diacids, where glycerol acts as the branching agent to yield hyperbranched or dendritic poly(glycerol adipate). Branched structures can evolve into crosslinked networks when higher-functionality monomers (e.g., tetrafunctional) are incorporated or through post-polymerization crosslinking reactions, leading to an infinite three-dimensional network at or beyond the critical conversion. In such networks, all chains interconnect, forming a gel that dominates the system's properties above the gel point. Representative examples include epoxy resins, where difunctional epoxy monomers undergo step-growth polyaddition with multifunctional amines to form highly crosslinked thermoset networks with exceptional mechanical integrity, and urea-formaldehyde resins, synthesized via stepwise condensation of urea (difunctional) and formaldehyde (multifunctional under reaction conditions) to produce rigid thermosets used in adhesives and composites. These networks exhibit enhanced mechanical strength and thermal stability due to the extensive interconnections, but at the cost of increased melt viscosity and complete loss of solubility, rendering them insoluble in solvents. To control branching and prevent premature gelation, the fraction of multifunctional monomers is often limited to below the critical threshold, allowing for branched polymers with manageable molecular weights without forming networks. For instance, in polyester synthesis, using less than 50 mol% trifunctional glycerol relative to diols keeps the system below the gel point, producing branched but soluble materials with tailored rheology. This approach balances the benefits of branching, such as improved processability in melts compared to linear analogs, against the risks of uncontrolled network formation.

Kinetics

General Kinetic Models

Step-growth polymerization follows second-order kinetics under the assumption of bimolecular reactions between functional groups. For an A-B type monomer, where each molecule contains one A and one B functional group that react to form a link, the rate law is expressed as -\frac{d[A]}{dt} = k [A]^2, where [A] represents the concentration of unreacted functional groups (equivalent to [B] due to stoichiometry), and k is the second-order rate constant. Integrating this differential equation yields \frac{1}{[A]} = \frac{1}{[A]_0} + kt, with [A]_0 as the initial concentration, providing a direct relationship between time t and the extent of reaction. The extent of reaction p, defined as the fraction of functional groups that have reacted, is given by p = 1 - \frac{[A]}{[A]_0} = 1 - \frac{1}{1 + [A]_0 kt}. This parameter quantifies polymerization progress, linking kinetics to molecular structure. The number-average degree of polymerization \overline{DP}_n, which indicates the average number of monomer units per chain, is related to p by the Carothers equation: \overline{DP}_n = \frac{1}{1 - p}. Thus, high molecular weight polymers require p approaching 1, such as p > 0.99 for \overline{DP}_n > 100. For systems involving A-A and B-B difunctional monomers with equal initial concentrations of A and B groups, the kinetics simplify similarly to the A-B case, as [A] = [B] throughout the reaction, yielding the same second-order form and integrated rate law. The Carothers equation applies identically, assuming stoichiometric balance. These models rest on key assumptions: all functional groups exhibit equal reactivity regardless of the chain to which they belong, and no side reactions or intramolecular cyclizations occur, ensuring purely intermolecular growth. In reality, deviations from ideal arise, particularly from intramolecular cyclization, where functional groups on the same react to form rings, reducing the concentration of reactive ends available for and thus lowering the effective rate compared to predictions. This effect becomes more pronounced at higher conversions or in dilute solutions, where chain flexibility favors loop formation over intermolecular links.

Catalyzed vs. Uncatalyzed Reactions

In step-growth polymerization, particularly for condensation reactions like polyesterification and polyamidation, the kinetics differ significantly between uncatalyzed (self-catalyzed) processes and those employing external catalysts. In uncatalyzed reactions, the functional groups on the monomers, such as carboxyl groups in dicarboxylic acids, serve as intrinsic acid catalysts, leading to a rate dependence on the carboxyl concentration. The reaction rate is expressed as k = k' [\ce{COOH}], where k' is the specific rate constant, resulting in autoacceleration as the concentration of carboxyl groups evolves during the reaction. This self-catalysis arises because the carboxyl groups protonate the hydroxyl or amine groups, facilitating nucleophilic attack, and the process exhibits third-order kinetics overall: -\frac{d[\ce{COOH}]}{dt} = k [\ce{COOH}]^2 [\ce{OH}] (or equivalently k [\ce{COOH}]^3 under stoichiometric conditions). In terms of the extent of reaction p (fraction of functional groups reacted), the differential rate equation is \frac{dp}{dt} = k' (1-p)^3, based on the concentrations of the reacting functional groups./03%3A_Kinetics_and_Thermodynamics_of_Polymerization/3.02%3A_Kinetics_of_Step-Growth_Polymerization) A classic example of self-catalyzed step-growth is the polycondensation of adipic acid and hexamethylenediamine to form nylon 6,6, where the carboxyl groups of adipic acid catalyze the amidation without added agents. Here, the reaction proceeds at elevated temperatures (around 250–280°C), with autoacceleration becoming pronounced after initial monomer consumption, as the local concentration of catalytic carboxyl ends increases relative to the diluting effect of water byproduct. The integrated form of the rate equation shows that the number-average degree of polymerization \overline{X}_n grows approximately as the square root of time, \overline{X}_n \approx (1 + 2 k' [\ce{M}]_0^2 t)^{1/2}, making high molecular weights achievable only at near-complete conversion (p > 0.99). This self-driven mechanism is common in polyamide and polyester syntheses but is inherently slower at lower temperatures due to the limited intrinsic catalysis./03%3A_Kinetics_and_Thermodynamics_of_Polymerization/3.02%3A_Kinetics_of_Step-Growth_Polymerization) In contrast, externally catalyzed reactions introduce added species to enhance the rate, typically following pseudo-second-order kinetics independent of chain length. Common catalysts include strong acids (e.g., p-toluenesulfonic acid), bases, or metal compounds like titanium(IV) alkoxides, Ti(OR)_4, which coordinate to carbonyl groups to activate them toward nucleophilic attack. The rate is first-order in catalyst concentration: \frac{dp}{dt} = k [\ce{cat}] (1-p)^2, where k incorporates the catalyst's efficiency, allowing linear growth in \overline{X}_n with time, \overline{X}_n = 1 + k [\ce{cat}] [\ce{M}]_0 t. For instance, in the industrial synthesis of poly(ethylene terephthalate) (PET), Ti(OR)_4 (e.g., titanium tetraisopropoxide) is employed at concentrations of 10–50 ppm, accelerating esterification and transesterification steps at 250–290°C while maintaining equilibrium by distilling ethylene glycol. This catalysis boosts the rate constant by factors of 10–100 compared to uncatalyzed conditions, enabling shorter reaction times. The use of external catalysis offers practical advantages over self-catalyzed processes, particularly in controlling reaction conditions and product quality. By increasing the effective rate, catalysts permit operation at lower temperatures (e.g., 200–250°C for Ti-catalyzed PET versus 280°C for uncatalyzed analogs), which reduces energy costs and minimizes side reactions such as thermal decarboxylation, ether formation from diols, or hydrolytic degradation. In PET production, Ti(OR)_4 specifically lowers the onset of side reactions like acetaldehyde formation or cyclization, yielding polymers with higher intrinsic viscosity and fewer defects. Overall, while self-catalysis suffices for laboratory-scale polyamide syntheses like nylon 6,6, external catalysis is essential for scalable polyester processes, balancing kinetics with polymer integrity.

Molecular Weight Distribution and Control

Distribution Functions

In linear step-growth polymerization, the Flory-Schulz model provides a statistical framework for the molecular weight distribution, assuming random reaction between functional groups of equal reactivity and neglecting cyclization or side reactions. This model treats the growth of polymer chains as a probabilistic process where the extent of reaction p represents the fraction of functional groups that have reacted, determining the probability of chain propagation versus termination. Specifically, the probability of propagation q equals p, while the probability of termination (or non-reaction at an end group) is $1 - p. The extent of reaction p arises from the underlying kinetics of the polymerization, as discussed in prior sections on kinetic models. The number-average molecular weight distribution, or number fraction x_n (the probability that a randomly selected molecule contains n monomer units), follows a geometric distribution derived from binomial probabilities. Consider a chain starting from a monomer; for it to have exactly n units, there must be n-1 successful propagations (each with probability p) followed by a termination (with probability $1-p). This yields the expression: x_n = (1 - p) p^{n-1} for n = 1, 2, 3, \dots. This distribution is normalized such that \sum_{n=1}^{\infty} x_n = 1, confirming its probabilistic validity. The weight-average molecular weight distribution, or weight fraction w_n (the fraction of total mass contributed by chains of n units), accounts for the fact that longer chains contribute more mass proportionally to n. It is obtained by weighting the number fraction by n and renormalizing: w_n = n (1 - p)^2 p^{n-1} for n = 1, 2, 3, \dots, with \sum_{n=1}^{\infty} w_n = 1. These distributions highlight the most probable chain length, which peaks at lower n for the number fraction and shifts to higher n for the weight fraction, reflecting the skew toward longer chains in mass terms. The polydispersity index (PDI), defined as the ratio of weight-average molecular weight \overline{M}_w to number-average molecular weight \overline{M}_n (PDI = \overline{M}_w / \overline{M}_n), quantifies the breadth of the distribution. Using the moments of the distribution—where the first moment relates to \overline{M}_n and the second to \overline{M}_w—the PDI simplifies to $1 + p. For high conversions (p \approx 1), typical in polymer synthesis to achieve substantial molecular weights, PDI approaches 2, indicating a relatively broad distribution compared to many chain-growth processes. Plots of these distributions versus n (or molecular weight, scaled by monomer mass) illustrate their behavior: the number fraction x_n decays exponentially from n=1, while the weight fraction w_n rises to a maximum near n \approx 1/(1-p) before decaying, becoming broader and shifting rightward as p increases toward 1. These visualizations underscore the model's prediction of a most probable distribution, where no single chain length dominates overwhelmingly, but the ensemble averages grow with p.

Stoichiometric and Impurity Effects

In step-growth polymerization involving bifunctional A-A and B-B monomers, the stoichiometric ratio r is defined as the initial concentration ratio of the limiting functional group to the excess one, typically r = \frac{[B]_0}{[A]_0} \leq 1. This ratio critically influences the achievable molecular weight, as deviations from r = 1 limit chain growth by leaving excess unreactive end groups on one type of chain. The number-average degree of polymerization \overline{DP}_n is approximated by the modified Carothers-Flory equation: \overline{DP}_n = \frac{1 + r}{1 + r - 2 r p}, where p is the extent of reaction of the limiting functional group; the maximum \overline{DP}_n occurs at r = 1, reducing to the stoichiometric case \overline{DP}_n = \frac{1}{1 - p}. Monofunctional impurities, such as acetic acid in polyester syntheses or monocarboxylic acids in polyamide production, introduce chain-terminating end groups that cap growing chains and disrupt stoichiometry. These impurities effectively reduce the maximum attainable extent of reaction to p_{\max} = 1 - \frac{[imp]}{[func]_0}, where [imp] is the impurity concentration and [func]_0 is the initial concentration of functional groups. Consequently, the number-average degree of polymerization is limited to approximately \overline{DP}_n = \frac{1}{\frac{[imp]}{[func]_0}}, preventing high molecular weights even at near-complete conversion of reactive groups. Achieving \overline{DP}_n > 1000 (corresponding to molecular weights suitable for many applications) requires stringent control: a stoichiometric ratio r > 0.99 and impurity levels below 0.1% relative to functional groups. For instance, at r = 0.99 and p = 0.9988, \overline{DP}_n drops to around 100-200, yielding molecular weights of approximately 17,000-20,000 g/mol for typical monomers, compared to over 80,000 g/mol under ideal stoichiometric conditions. Stoichiometric imbalances and impurities shift the molecular weight distribution away from the ideal Flory-Schulz form (most probable distribution with PDI = 2), resulting in a broader low-molecular-weight tail and an increase in polydispersity index (PDI) to values greater than 2 due to the presence of excess low-molecular-weight species.

Advanced Applications and Developments

High-Performance Step-Growth Polymers

High-performance step-growth polymers, such as aromatic polyethers, polyethersulfones, aromatic polysulfides, and aromatic polyimides, are engineered through rigid monomer designs that incorporate aromatic rings and heteroatoms to achieve exceptional thermal, mechanical, and chemical stability for demanding engineering applications. These materials are typically synthesized via nucleophilic aromatic substitution or condensation reactions, where the incorporation of stiff, conjugated backbones restricts chain mobility, elevating glass transition temperatures (Tg) and enabling operation at elevated temperatures. For instance, the use of activated aryl halides or dianhydrides with bisphenols or diamines promotes high molecular weights while maintaining processability, distinguishing them from commodity polymers. Aromatic polyethers, exemplified by polyetheretherketone (PEEK), are produced through step-growth nucleophilic aromatic substitution of hydroquinone with 4,4'-difluorobenzophenone in diphenyl sulfone solvent at temperatures ranging from 150°C to 340°C. The ether and ketone linkages, derived from rigid aromatic monomers, confer high thermal stability with a Tg of approximately 143°C and melting temperature (Tm) between 334°C and 350°C, allowing PEEK to withstand prolonged exposure to heat without degradation. This rigidity enhances mechanical strength and chemical resistance, making PEEK suitable for aerospace components and medical implants. Polyethersulfones (PES) are synthesized similarly via nucleophilic aromatic substitution between dihalogenated diphenyl sulfones and bisphenol A, yielding polymers where sulfonyl groups interrupt the aromatic backbone to improve toughness and flexibility. These groups contribute to outstanding thermal and oxidative stability, with PES exhibiting a high Tg around 225°C and resistance to hydrolysis, which is critical for long-term durability. PES is widely employed in membrane applications, such as ultrafiltration and gas separation, due to its mechanical robustness and ability to form porous structures via phase inversion. Aromatic polysulfides, particularly polyphenylene sulfide (PPS), are formed by the step-growth polycondensation of p-dichlorobenzene with sodium sulfide in a polar aprotic solvent like N-methylpyrrolidone at elevated temperatures and pressures, as developed in the Edmonds-Hill process. The all-aromatic structure provides semicrystalline character, enabling melt processability while offering superior chemical resistance to acids, bases, and solvents, alongside a continuous use temperature up to 240°C. PPS's rigidity from para-linked phenylene units ensures dimensional stability, positioning it for use in automotive and electronic components exposed to harsh environments. Aromatic polyimides, such as Kapton-type variants, are prepared through a two-step step-growth process involving the reaction of pyromellitic dianhydride (PMDA) with 4,4'-oxydianiline (ODA) to form a poly(amic acid) precursor, followed by thermal imidization at 240–330°C. The fully aromatic, heterocyclic imide rings impart extreme heat resistance, with materials sustaining short-term exposure up to 400°C and continuous use up to 240°C, and decomposition onset above 500°C in inert atmospheres. Monomer selection emphasizes rigid, planar structures to minimize free volume and enhance chain packing, resulting in high tensile strength and low dielectric constants for applications in flexible electronics and aerospace insulation.

Modern Synthetic Strategies

Modern synthetic strategies in step-growth polymerization emphasize precision, efficiency, and environmental sustainability, enabling the creation of tailored polymers with enhanced functionalities such as recyclability and self-healing. These approaches build on traditional condensation mechanisms but integrate orthogonal reactions and bio-derived components to overcome limitations in molecular weight control and byproduct formation. Innovations since the 2000s have focused on modular assembly techniques that allow for high yields under mild conditions, reducing energy demands and waste. Telechelic oligomers, which are short chains with functional end-groups, serve as versatile building blocks for constructing block copolymers through step-growth coupling. For instance, α,ω-dihydroxy-terminated oligomers can be coupled via condensation reactions to form well-defined multiblock architectures, enabling precise control over segment lengths and interfaces in materials like compatibilizers for polymer blends. Recent developments include the synthesis of telechelic dithiol copolymers with tunable molecular weights and compositions, achieved through controlled radical polymerization followed by step-growth thiol-ene linking, yielding networks with low polydispersity indices (PDI) approaching 1.2-1.5. Chain-stoppers, such as monofunctional thiols, are employed to regulate chain extension and PDI in these systems, preventing excessive branching and ensuring uniform distributions essential for advanced applications. Click chemistry has revolutionized step-growth polymerization by providing orthogonal, high-efficiency linkages that proceed without catalysts or byproducts in many cases. Copper-catalyzed azide-alkyne cycloaddition (CuAAC) and thiol-ene reactions exemplify this, allowing rapid coupling of telechelic precursors under ambient conditions with yields exceeding 95%. Thiol-ene click step-growth, in particular, leverages nucleophilic Michael additions to form E/Z stereocontrolled polymers, offering atom-economic routes to sustainable biomaterials with minimal purification needs. These methods enable the assembly of complex architectures, such as graft copolymers, where azide- or thiol-terminated chains are linked to alkyne- or ene-functionalized backbones, achieving PDIs below 1.5 and molecular weights up to 50 kDa. Sustainable advances prioritize bio-based monomers and enzymatic catalysis to align step-growth processes with green chemistry principles. Itaconic acid, derived from renewable fermentation of sugars, serves as a key diacid monomer for polyesters, where its vinyl functionality is preserved during polycondensation to yield copolymers with pendant reactivity for further modification. Lipase-catalyzed ring-opening addition condensation of itaconic anhydride with diols produces bio-based polyesters at temperatures below 100°C, achieving molecular weights over 10 kDa with PDIs around 2.0 and eliminating harsh metal catalysts. Enzymatic strategies extend to co-polymerization of succinate, itaconate, and butanediol, enabling tunable biodegradability and thermal properties in aliphatic polyesters. These bio-derived systems reduce reliance on petroleum feedstocks, with recent integrations of thiolactone-functionalized itaconic acid enabling dual step-growth and radical polymerizations for versatile networks. Developments in the 2020s have centered on recyclable networks incorporating dynamic covalent bonds, such as in vitrimers, which combine the permanence of cross-linked thermosets with reprocessability via bond exchange. Vitrimers formed by step-growth polycondensation of telechelic diols with boric acid exhibit associative dynamic networks, allowing reshaping at elevated temperatures without depolymerization, with topology freeze demonstrated up to 1.0 W/(m·K) thermal conductivity in crystalline variants. Self-healing step-growth polymers leverage dynamic bonds like disulfides for autonomous repair; for example, polyurethanes with exchangeable urea or disulfide linkages recover over 90% tensile strength after damage through thermal activation. These materials address end-of-life challenges, with biobased vitrimers from polyhydroxyalkanoates showing elastomeric recovery and closed-loop recyclability. Representative examples illustrate these strategies' impact. Poly(arylene ether sulfone)-poly(disulfide) multiblock copolymers, synthesized via step-growth coupling of telechelic segments with dynamic disulfide links, exhibit self-healing energies as low as 50 J/m² due to rapid bond exchange, enabling repair at 150°C. PDI control via chain-stoppers in thiol-ene systems, such as adding monothiol terminators during coupling, limits molecular weight distributions to PDI < 1.3, facilitating precise engineering of mechanical properties in recyclable networks. These innovations underscore step-growth's adaptability for high-performance, circular economy materials.

References

  1. [1]
    Polymers - MSU chemistry
    The polyester Dacron and the polyamide Nylon 66, shown here, are two examples of synthetic condensation polymers, also known as step-growth polymers.
  2. [2]
    [PDF] Kinetics of Step Growth Polymerization (Chapter 2)
    10.569 Synthesis of Polymers. Prof. Paula Hammond. Lecture 3: Step Growth Polymerization, Types of Monomers, Kinetics and.Missing: examples | Show results with:examples
  3. [3]
    Kinetics and Thermodynamics in P
    Condensation polymers grow through a a step-growth process. The reaction generally involves substitution at a carboxyloid. For example, polyester formation ...Missing: definition | Show results with:definition
  4. [4]
    [PDF] Step-growth polymerization
    May 1, 2005 · We shall sometimes use the terms step-growth and condensation polymers as synonyms, although step-growth polymerization encompasses a wider ...
  5. [5]
    Step-growth polymerization | Journal of Chemical Education
    In step-growth polymerization, the molecular weight of the polymer chain builds up slowly and there is only one reaction mechanism for the formation of polymer.
  6. [6]
    step-growth polymerization - Wiley Online Library
    In step-growth polymerization, only one kind of reaction is involved in the formation of a polymer and the reaction proceeds step by step. The main feature of ...
  7. [7]
    Step-Growth Polymerization - an overview | ScienceDirect Topics
    Step-growth polymerization, also known as polycondensation, is defined as a method for synthesizing polymers by exploiting the reactivity of functional ...
  8. [8]
    History and Future of Plastics
    The first synthetic polymer was invented in 1869 by John Wesley Hyatt, who was inspired by a New York firm's offer of $10,000 for anyone who could provide a ...Missing: 19th | Show results with:19th
  9. [9]
    Hermann Staudinger Foundation of Polymer Science - Landmark
    Industrial Significance of Polymer Science. During the late 19th century, cellulose derivatives were developed as artificial substitutes for silk and ivory.Missing: inspiring | Show results with:inspiring
  10. [10]
    Bakelite® First Synthetic Plastic - American Chemical Society
    Many 19th century manufacturers modified colloids and natural polymers to form new materials. In 1870, the American inventor John Wesley Hyatt used chemically ...
  11. [11]
    Hermann Staudinger | Science History Institute
    Hermann Staudinger's macromolecular theory of polymers eventually superseded the reigning aggregate theory. He was awarded the Nobel Prize in Chemistry in 1953.Missing: inspiration | Show results with:inspiration
  12. [12]
    Wallace Carothers and the Development of Nylon - Landmark
    To test his theory, he proposed to create polymers by using well-known chemical reactions to join together many small molecules.
  13. [13]
    Wallace Hume Carothers | Science History Institute
    Through his inventions of nylon and neoprene in 1930, Carothers helped demonstrate the macromolecular theory of polymers and establish the field of modern ...
  14. [14]
    Fundamental Principles of Condensation Polymerization.
    Fundamental Principles of Condensation Polymerization. ... Beyond Flory's principle: Cyclization and unequal reactivity in step-growth linear polymerization.
  15. [15]
    Polyurethanes: An Introduction | ACS Symposium Series
    Aug 8, 2025 · 1.1 Historical Development. Otto Bayer created the first PU around eighty-seven years ago, in 1937. To compete with nylon fibers, Bayer set ...
  16. [16]
  17. [17]
    Condensation Polymer - an overview | ScienceDirect Topics
    Condensation polymers are formed from bi- or polyfunctional monomers by reactions that involve elimination of some smaller molecule. Polyesters (e.g., 1-5) and ...
  18. [18]
    Polyanhydride Chemistry | Biomacromolecules - ACS Publications
    Nov 23, 2022 · The aromatic PAs, poly(p-CPH) (n = 6) and poly(p-CPP) (n = 3) and their copolymers have been synthesized by melt condensation polymerization of ...
  19. [19]
    Polycondensation Reaction - an overview | ScienceDirect Topics
    In order to generate polymers, two or more functional groups must be present in each monomer, e.g., bifunctional diamine and diacid monomers can generate a ...Missing: compatibility | Show results with:compatibility
  20. [20]
    What Does Conversion Mean in Polymer Science? - Kricheldorf - 2021
    Mar 26, 2021 · A quantitative conversion can only result from cyclization reactions, and recent experiments have confirmed that in polycondensations of “abn” ...
  21. [21]
    Model polyethers I—Synthesis by the Williamson reaction
    Abstract. Model polyalkylene ethers H[(CH2)xO]y (CH2)xH and trialkyleneglycols OH[(CH2)xO]3H, where x=4, 5, 6 and 10 and y=2, 3 and 4 have been synthesized by ...
  22. [22]
    Synthesis of an arylene/alkylene polyether from diphenolic acid and ...
    The Williamson ether synthesis has been applied to the formation of an arylene/alkylene polyether with pendant carboxyl groups from 4,4-bis(p-hydroxyphenyl) ...
  23. [23]
    Poly(aryl ethers) by nucleophilic aromatic substitution. I. Synthesis ...
    The synthesis involves the condensation of a dialkali metal salt of a dihydric phenol with an "activated" or negatively substituted aromatic dihalide in an ...
  24. [24]
    Synthesis of poly(ether sulfone)s by self-polycondensation of AB ...
    Feb 20, 2013 · In general, PESs are synthesized via a nucleophilic aromatic substitution (SNAr) reaction between dihalogenated diphenylsulfones (A2) and ...
  25. [25]
    Green polymer chemistry: Living oxidative polymerization of dithiols
    ### Summary of Oxidative Polymerization of Dithiols to Poly(disulfide)s
  26. [26]
    Poly(disulfide)s - Sulfur‐Containing Polymers - Wiley Online Library
    Mar 19, 2021 · This chapter provides a comprehensive description on various synthetic procedures for PDS including oxidative polymerization of dithiols ...<|separator|>
  27. [27]
    Pd/Cu-catalyzed sonogashira cross-coupling polycondensation
    Sep 11, 2023 · Sonogashira cross-coupling polycondensation has been one of the most useful tools for the construction of conjugated polymers with useful properties for ...Missing: polyynes | Show results with:polyynes
  28. [28]
    Reconsidering terms for mechanisms of polymer growth: the “step ...
    The terms “step-growth polymerization” and “chain-growth polymerization” are used widely in both written and oral communications to describe the two main ...<|control11|><|separator|>
  29. [29]
  30. [30]
    A parallel algorithm for step- and chain-growth polymerization in ...
    Apr 1, 2015 · The resulting concentration of monomers thus follows an exponential decay, ... growth is stopped when the polymer chain reaches n monomers.
  31. [31]
    Going Beyond the Carothers, Flory and Stockmayer Equation ... - NIH
    Jul 22, 2021 · Principle of step-growth polymerization for (a) two bifunctional monomers (A2: white and B2: black) and (b) two multifunctional monomers (A4: ...
  32. [32]
  33. [33]
    Step-Growth Polymerization - an overview | ScienceDirect Topics
    Step-growth polymerization is a key technique used to create sulfur-based polymers through the reaction of sulfur-containing monomers with functional groups.
  34. [34]
    Crystallinity in Polymers
    Intermolecular forces can be a big help for a polymer if it wants to form crystals. A good example is nylon. You can see from the picture that the polar amide ...
  35. [35]
    Making Nylon 6,6 - Polymer Science Learning Center
    To start off, nylon is made by a reaction which is a step-growth polymerization, and a condensation polymerization. Nylons are made from diacids and diamines.
  36. [36]
    Telechelic Polymer - an overview | ScienceDirect Topics
    Telechelic polymers are linear polymers with functional groups at both ends, enabling chemical reactions. They are also called macromolecules with two reactive ...
  37. [37]
    A Simple Mathematical Model for the Structural Description of ...
    Aug 17, 2025 · In general, the polymerization of glycerol and dicarboxylic acids produces branched polyesters consisting of terminal units 1T and 2T ...
  38. [38]
    Network Formation and Physical Properties of Epoxy Resins for ...
    Jun 9, 2022 · (24) Thus, epoxy monomers can generate a cross-linked network structure by either chain-growth ring-opening polymerization or step-growth ...
  39. [39]
    Synthesis and characterization of glycerol-adipic acid ...
    Sep 26, 2014 · This A2 + B3 system forms a HBPE by a step-growth polymerization reaction, which if performed in equimolar quantities of functional groups, ...<|separator|>
  40. [40]
    [PDF] 1 Introduction to Synthetic Methods in Step-Growth Polymers
    In general, step-growth polymers such as polyesters and polyamides possess more robust mechanical properties, including toughness, stiffness, and higher.
  41. [41]
    STUDIES ON POLYMERIZATION AND RING FORMATION. I. AN ...
    STUDIES ON POLYMERIZATION AND RING FORMATION. I. AN INTRODUCTION TO THE ... Wallace Carothers and Polymer Chemistry: A Partnership Ended Too Soon. 2017 ...
  42. [42]
  43. [43]
    Kinetic Studies of Polyethylene Terephthalate Synthesis with ...
    Aug 6, 2025 · The catalytic activity of these systems was investigated and their effect on the kinetics of the synthesis process was revealed.
  44. [44]
    Molecular Size Distribution in Linear Condensation Polymers 1
    Polymer molecular weight distributions: a tutorial on transformations between number density, mass density, and linear/logarithmic axes.Missing: intramolecular | Show results with:intramolecular<|control11|><|separator|>
  45. [45]
    [PDF] Step-Growth Polymerization - The Gu Lab
    What will be the expected degree of polymerization and extent of reaction if the reaction is carried out in a closed system without removal of the by ...Missing: approximation exponential
  46. [46]
    Step-Growth Polymerization - an overview | ScienceDirect Topics
    Step-growth polymerization is defined as a process involving one or more monomers with at least two reactive functional groups that react with each other, ...
  47. [47]
  48. [48]
  49. [49]
  50. [50]
    Poly-Phenylene Sulphide - an overview | ScienceDirect Topics
    Polyphenylene sulphide (PPS) is noted for its exceptional resistance to acid attack. Results of tests in 85% sulphuric acid at 120 °C for up to 5000 h,
  51. [51]
    Telechelic Dithiol Copolymers as Tunable Building Blocks for ...
    Dec 8, 2024 · For multiblock copolymers, the use of thiols in AA-step growth copolymerization requires the synthesis of telechelic prepolymers containing ...
  52. [52]
    Step-Growth “Click” Coupling of Telechelic Polymers Prepared by ...
    Procedures for the preparation of propargyl 2-bromoisobutyrate and homo- and heterotelechelic polystyrenes and their step-growth click coupling.Missing: advantages | Show results with:advantages
  53. [53]
    [PDF] Synthesis of Telechelic Polyolefins
    Aug 20, 2021 · Telechelic polyolefins are synthesized using dienes or cyclic olefins, and ring-opening metathesis polymerization (ROMP) of cyclic olefins.  ...
  54. [54]
    Click Step-Growth Polymerization and E/Z Stereochemistry Using ...
    Aug 25, 2022 · Step-growth polymerization and E/Z stereochemistry using nucleophilic thiol–yne/–ene reactions: applying old concepts for practical sustainable (bio)materials.
  55. [55]
    Click Step-Growth Polymerization and E/Z Stereochemistry ... - NIH
    Aug 25, 2022 · Click step-growth polymerization using the thiol–yne Michael addition, a reaction first reported over a century ago, has emerged as an extremely mild and atom- ...
  56. [56]
    Click chemistry strategies for the accelerated synthesis of functional ...
    Apr 29, 2021 · Click chemistry is one of the most powerful strategies for constructing polymeric soft materials with precise control over architecture and functionality.
  57. [57]
    Synthesis of biobased and versatile monomers from itaconic acid ...
    Aug 19, 2024 · New monomers are synthesized by functionalizing itaconic acid with thiolactone, and then used in step-growth and radical polymerization.
  58. [58]
    Green polymer chemistry: lipase-catalyzed synthesis of bio-based ...
    Jul 17, 2013 · This study employs itaconic anhydride (IAn) as a novel monomer for lipase-catalyzed ring-opening addition condensation polymerization (ROACP) ...
  59. [59]
    Integrating Computational and Experimental Methods for the ... - NIH
    Sep 19, 2025 · In particular, we have demonstrated that the vinyl group of itaconic acid (IA) is preserved when lipase-catalyzed polycondensation is carried ...
  60. [60]
    Enzyme-Catalyzed Synthesis of Unsaturated Aliphatic Polyesters ...
    Bio-based commercially available succinate, itaconate and 1,4-butanediol are enzymatically co-polymerized in solution via a two-stage method.<|separator|>
  61. [61]
    [PDF] investigating the effect of dynamic covalent bonds on the - IDEALS
    (a) Step growth polymerization of telechelic alkane diols and boric acid to make PE vitrimers. By controlling the length of the diols, networks with precise ...
  62. [62]
    Dynamic Covalent Bonds in Vitrimers Enable 1.0 W/(m K) Intrinsic ...
    Feb 16, 2023 · In this work, vitrimers with a fixed chemical structure and slow crystallization kinetics are investigated to reveal the relationships between crystallinity ...
  63. [63]
    Elastomeric vitrimers from designer polyhydroxyalkanoates with ...
    Nov 22, 2023 · We develop biodegradable and recyclable elastomers by chemically installing the covalent adaptable network within biologically produced mcl-PHAs.Missing: 2020s | Show results with:2020s
  64. [64]
    Shape-Memory and Self-Healing Polymers Based on Dynamic ...
    Shape-memory and self-healing polymers can complete programed shape transformation, and they can spontaneously repair damage, thereby extending the life of the ...
  65. [65]
    Synthesis and healing properties of poly(arylene ether sulfone)‐poly ...
    Apr 17, 2018 · The high content of disulfide bonds in the multiblock copolymer resulted in a lower self-healing energy. This study clearly shows that a high ...