Radical polymerization

Last updated

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

Contents

Free-radical polymerization is a key synthesis route for obtaining a wide variety of different polymers and materials composites. The relatively non-specific nature of free-radical chemical interactions makes this one of the most versatile forms of polymerization available and allows facile reactions of polymeric free-radical chain ends and other chemicals or substrates. In 2001, 40 billion of the 110 billion pounds of polymers produced in the United States were produced by free-radical polymerization. [1]

Free-radical polymerization is a type of chain-growth polymerization, along with anionic, cationic and coordination polymerization.

IUPAC definition

A chain polymerization in which the kinetic-chain carriers are radicals.

Note: Usually, the growing chain end bears an unpaired electron. [2]

Initiation

Initiation is the first step of the polymerization process. During initiation, an active center is created from which a polymer chain is generated. Not all monomers are susceptible to all types of initiators. Radical initiation works best on the carbon–carbon double bond of vinyl monomers and the carbon–oxygen double bond in aldehydes and ketones. [1] Initiation has two steps. In the first step, one or two radicals are created from the initiating molecules. In the second step, radicals are transferred from the initiator molecules to the monomer units present. Several choices are available for these initiators.

Types of initiation and the initiators

Thermal decomposition
The initiator is heated until a bond is homolytically cleaved, producing two radicals (Figure 1). This method is used most often with organic peroxides or azo compounds. [3]
Figure 1: Thermal decomposition of dicumyl peroxide Initiation - thermal decomp.png
Figure 1: Thermal decomposition of dicumyl peroxide
Photolysis
Radiation cleaves a bond homolytically, producing two radicals (Figure 2). This method is used most often with metal iodides, metal alkyls, and azo compounds. [3]
Figure 2: Photolysis of azoisobutylnitrile (AIBN) Initiation - photolysis.png
Figure 2: Photolysis of azoisobutylnitrile (AIBN)
Photoinitiation can also occur by bi-molecular H abstraction when the radical is in its lowest triplet excited state. [4] An acceptable photoinitiator system should fulfill the following requirements: [4]
  • High absorptivity in the 300–400 nm range.
  • Efficient generation of radicals capable of attacking the alkene double bond of vinyl monomers.
  • Adequate solubility in the binder system (prepolymer + monomer).
  • Should not impart yellowing or unpleasant odors to the cured material.
  • The photoinitiator and any byproducts resulting from its use should be non-toxic.
Redox reactions
Reduction of hydrogen peroxide or an alkyl hydrogen peroxide by iron (Figure 3). [3] Other reductants such as Cr2+, V2+, Ti3+, Co2+, and Cu+ can be employed in place of ferrous ion in many instances. [1]
Figure 3: Redox reaction of hydrogen peroxide and iron.
Persulfates
The dissociation of a persulfate in the aqueous phase (Figure 4). This method is useful in emulsion polymerizations, in which the radical diffuses into a hydrophobic monomer-containing droplet. [3]
Figure 4: Thermal degradation of a persulfate Initiation - persulfates.png
Figure 4: Thermal degradation of a persulfate
Ionizing radiation
α-, β-, γ-, or x-rays cause ejection of an electron from the initiating species, followed by dissociation and electron capture to produce a radical (Figure 5). [3]
Figure 5: The three steps involved in ionizing radiation: ejection, dissociation, and electron-capture Initiation - ionizing radiation.png
Figure 5: The three steps involved in ionizing radiation: ejection, dissociation, and electron-capture
Electrochemical
Electrolysis of a solution containing both monomer and electrolyte. A monomer molecule will receive an electron at the cathode to become a radical anion, and a monomer molecule will give up an electron at the anode to form a radical cation (Figure 6). The radical ions then initiate free radical (and/or ionic) polymerization. This type of initiation is especially useful for coating metal surfaces with polymer films. [5]
Figure 6: (Top) Formation of radical anion at the cathode; (bottom) formation of radical cation at the anode Initiation - electrochemical.png
Figure 6: (Top) Formation of radical anion at the cathode; (bottom) formation of radical cation at the anode
Plasma
A gaseous monomer is placed in an electric discharge at low pressures under conditions where a plasma (ionized gaseous molecules) is created. In some cases, the system is heated and/or placed in a radiofrequency field to assist in creating the plasma. [1]
Sonication
High-intensity ultrasound at frequencies beyond the range of human hearing (16 kHz) can be applied to a monomer. Initiation results from the effects of cavitation (the formation and collapse of cavities in the liquid). The collapse of the cavities generates very high local temperatures and pressures. This results in the formation of excited electronic states, which in turn lead to bond breakage and radical formation. [1]
Ternary initiators
A ternary initiator is the combination of several types of initiators into one initiating system. The types of initiators are chosen based on the properties they are known to induce in the polymers they produce. For example, poly(methyl methacrylate) has been synthesized by the ternary system benzoyl peroxide and 3,6-bis(o-carboxybenzoyl)-N-isopropylcarbazole and di-η5-indenylzirconium dichloride (Figure 7). [6] [7]
Figure 7: benzoyl peroxide + 3,6-bis(o-carboxybenzoyl)-N-isopropylcarbazole + di-e -indenylzicronium dichloride Benzoyl peroxide + 3,6-bis(o-carboxybenzoyl)-N-isopropylcarbazole + di-e5-indenylzicronium dichloride.svg
Figure 7: benzoyl peroxide + 3,6-bis(o-carboxybenzoyl)-N-isopropylcarbazole + di-η -indenylzicronium dichloride
This type of initiating system contains a metallocene, an initiator, and a heteroaromatic diketo carboxylic acid. Metallocenes in combination with initiators accelerate polymerization of poly(methyl methacrylate) and produce a polymer with a narrower molecular weight distribution. The example shown here consists of indenylzirconium (a metallocene) and benzoyl peroxide (an initiator). Also, initiating systems containing heteroaromatic diketo carboxylic acids, such as 3,6-bis(o-carboxybenzoyl)-N-isopropylcarbazole in this example, are known to catalyze the decomposition of benzoyl peroxide. Initiating systems with this particular heteroaromatic diket carboxylic acid are also known to have effects on the microstructure of the polymer. The combination of all of these components—a metallocene, an initiator, and a heteroaromatic diketo carboxylic acid—yields a ternary initiating system that was shown to accelerate the polymerization and produce polymers with enhanced heat resistance and regular microstructure. [6] [7]

Initiator efficiency

Due to side reactions, not all radicals formed by the dissociation of initiator molecules actually add monomers to form polymer chains. The efficiency factor f is defined as the fraction of the original initiator which contributes to the polymerization reaction. The maximal value of f is 1, but typical values range from 0.3 to 0.8. [8]

The following types of reactions can decrease the efficiency of the initiator.

Primary recombination
Two radicals recombine before initiating a chain (Figure 8). This occurs within the solvent cage, meaning that no solvent has yet come between the new radicals. [3]
Figure 8: Primary recombination of BPO; brackets indicate that the reaction is happening within the solvent cage Initiation - primary recombination.png
Figure 8: Primary recombination of BPO; brackets indicate that the reaction is happening within the solvent cage
Other recombination pathways
Two radical initiators recombine before initiating a chain, but not in the solvent cage (Figure 9). [3]
Figure 9: Recombination of phenyl radicals from the initiation of BPO outside the solvent cage Initiation - other recombination.png
Figure 9: Recombination of phenyl radicals from the initiation of BPO outside the solvent cage
Side reactions
One radical is produced instead of the three radicals that could be produced (Figure 10). [3]
Figure 10: Reaction of polymer chain R with other species in reaction

Propagation

During polymerization, a polymer spends most of its time in increasing its chain length, or propagating. After the radical initiator is formed, it attacks a monomer (Figure 11). [9] In an ethene monomer, one electron pair is held securely between the two carbons in a sigma bond. The other is more loosely held in a pi bond. The free radical uses one electron from the pi bond to form a more stable bond with the carbon atom. The other electron returns to the second carbon atom, turning the whole molecule into another radical. This begins the polymer chain. Figure 12 shows how the orbitals of an ethylene monomer interact with a radical initiator. [10]

Figure 11: Phenyl initiator from benzoyl peroxide (BPO) attacks a styrene molecule to start the polymer chain. Initiation - part 2.png
Figure 11: Phenyl initiator from benzoyl peroxide (BPO) attacks a styrene molecule to start the polymer chain.
Figure 12: An orbital drawing of the initiator attack on ethylene molecule, producing the start of the polyethylene chain. Initiation - orbitals.png
Figure 12: An orbital drawing of the initiator attack on ethylene molecule, producing the start of the polyethylene chain.

Once a chain has been initiated, the chain propagates (Figure 13) until there are no more monomers (living polymerization) or until termination occurs. There may be anywhere from a few to thousands of propagation steps depending on several factors such as radical and chain reactivity, the solvent, and temperature. [11] [12] The mechanism of chain propagation is as follows:

Figure 13: Propagation of polystyrene with a phenyl radical initiator. Propagation.png
Figure 13: Propagation of polystyrene with a phenyl radical initiator.

Termination

Chain termination is inevitable in radical polymerization due to the high reactivity of radicals. Termination can occur by several different mechanisms. If longer chains are desired, the initiator concentration should be kept low; otherwise, many shorter chains will result. [3]

Chain transfer

Contrary to the other modes of termination, chain transfer results in the destruction of only one radical, but also the creation of another radical. Often, however, this newly created radical is not capable of further propagation. Similar to disproportionation, all chain-transfer mechanisms also involve the abstraction of a hydrogen or other atom. There are several types of chain-transfer mechanisms. [3]

Effects of chain transfer: The most obvious effect of chain transfer is a decrease in the polymer chain length. If the rate of transfer is much larger than the rate of propagation, then very small polymers are formed with chain lengths of 2-5 repeating units (telomerization). [14] The Mayo equation estimates the influence of chain transfer on chain length (xn): . Where ktr is the rate constant for chain transfer and kp is the rate constant for propagation. The Mayo equation assumes that transfer to solvent is the major termination pathway. [3] [15]

Methods

There are four industrial methods of radical polymerization: [3]

Other methods of radical polymerization include the following:

Reversible deactivation radical polymerization

Also known as living radical polymerization, controlled radical polymerization, reversible deactivation radical polymerization (RDRP) relies on completely pure reactions, preventing termination caused by impurities. Because these polymerizations stop only when there is no more monomer, polymerization can continue upon the addition of more monomer. Block copolymers can be made this way. RDRP allows for control of molecular weight and dispersity. However, this is very difficult to achieve and instead a pseudo-living polymerization occurs with only partial control of molecular weight and dispersity. [16] ATRP and RAFT are the main types of complete radical polymerization.

Kinetics

In typical chain growth polymerizations, the reaction rates for initiation, propagation and termination can be described as follows:

where f is the efficiency of the initiator and kd, kp, and kt are the constants for initiator dissociation, chain propagation and termination, respectively. [I] [M] and [M•] are the concentrations of the initiator, monomer and the active growing chain.

Under the steady-state approximation, the concentration of the active growing chains remains constant, i.e. the rates of initiation and of termination are equal. The concentration of active chain can be derived and expressed in terms of the other known species in the system.

In this case, the rate of chain propagation can be further described using a function of the initiator and monomer concentrations [21] [22]

The kinetic chain length v is a measure of the average number of monomer units reacting with an active center during its lifetime and is related to the molecular weight through the mechanism of the termination. Without chain transfer, the kinetic chain length is only a function of propagation rate and initiation rate. [23]

Assuming no chain-transfer effect occurs in the reaction, the number average degree of polymerization Pn can be correlated with the kinetic chain length. In the case of termination by disproportionation, one polymer molecule is produced per every kinetic chain:

Termination by combination leads to one polymer molecule per two kinetic chains: [21]

Any mixture of both these mechanisms can be described by using the value δ, the contribution of disproportionation to the overall termination process:

If chain transfer is considered, the kinetic chain length is not affected by the transfer process because the growing free-radical center generated by the initiation step stays alive after any chain-transfer event, although multiple polymer chains are produced. However, the number average degree of polymerization decreases as the chain transfers, since the growing chains are terminated by the chain-transfer events. Taking into account the chain-transfer reaction towards solvent S, initiator I, polymer P, and added chain-transfer agent T. The equation of Pn will be modified as follows: [24]

It is usual to define chain-transfer constants C for the different molecules

, , , ,

Thermodynamics

In chain growth polymerization, the position of the equilibrium between polymer and monomers can be determined by the thermodynamics of the polymerization. The Gibbs free energy (ΔGp) of the polymerization is commonly used to quantify the tendency of a polymeric reaction. The polymerization will be favored if ΔGp < 0; if ΔGp > 0, the polymer will undergo depolymerization. According to the thermodynamic equation ΔG = ΔH – TΔS, a negative enthalpy and an increasing entropy will shift the equilibrium towards polymerization.

In general, the polymerization is an exothermic process, i.e. negative enthalpy change, since addition of a monomer to the growing polymer chain involves the conversion of π bonds into σ bonds, or a ring–opening reaction that releases the ring tension in a cyclic monomer. Meanwhile, during polymerization, a large amount of small molecules are associated, losing rotation and translational degrees of freedom. As a result, the entropy decreases in the system, ΔSp < 0 for nearly all polymerization processes. Since depolymerization is almost always entropically favored, the ΔHp must then be sufficiently negative to compensate for the unfavorable entropic term. Only then will polymerization be thermodynamically favored by the resulting negative ΔGp.

In practice, polymerization is favored at low temperatures: TΔSp is small. Depolymerization is favored at high temperatures: TΔSp is large. As the temperature increases, ΔGp become less negative. At a certain temperature, the polymerization reaches equilibrium (rate of polymerization = rate of depolymerization). This temperature is called the ceiling temperature (Tc). ΔGp = 0. [25]

Stereochemistry

The stereochemistry of polymerization is concerned with the difference in atom connectivity and spatial orientation in polymers that has the same chemical composition.

Hermann Staudinger studied the stereoisomerism in chain polymerization of vinyl monomers in the late 1920s, and it took another two decades for people to fully appreciate the idea that each of the propagation steps in the polymer growth could give rise to stereoisomerism. The major milestone in the stereochemistry was established by Ziegler and Natta and their coworkers in 1950s, as they developed metal based catalyst to synthesize stereoregular polymers. The reason why the stereochemistry of the polymer is of particular interest is because the physical behavior of a polymer depends not only on the general chemical composition but also on the more subtle differences in microstructure. [26] Atactic polymers consist of a random arrangement of stereochemistry and are amorphous (noncrystalline), soft materials with lower physical strength. The corresponding isotactic (like substituents all on the same side) and syndiotactic (like substituents of alternate repeating units on the same side) polymers are usually obtained as highly crystalline materials. It is easier for the stereoregular polymers to pack into a crystal lattice since they are more ordered and the resulting crystallinity leads to higher physical strength and increased solvent and chemical resistance as well as differences in other properties that depend on crystallinity. The prime example of the industrial utility of stereoregular polymers is polypropene. Isotactic polypropene is a high-melting (165 °C), strong, crystalline polymer, which is used as both a plastic and fiber. Atactic polypropene is an amorphous material with an oily to waxy soft appearance that finds use in asphalt blends and formulations for lubricants, sealants, and adhesives, but the volumes are minuscule compared to that of isotactic polypropene.

When a monomer adds to a radical chain end, there are two factors to consider regarding its stereochemistry: 1) the interaction between the terminal chain carbon and the approaching monomer molecule and 2) the configuration of the penultimate repeating unit in the polymer chain. [5] The terminal carbon atom has sp2 hybridization and is planar. Consider the polymerization of the monomer CH2=CXY. There are two ways that a monomer molecule can approach the terminal carbon: the mirror approach (with like substituents on the same side) or the non-mirror approach (like substituents on opposite sides). If free rotation does not occur before the next monomer adds, the mirror approach will always lead to an isotactic polymer and the non-mirror approach will always lead to a syndiotactic polymer (Figure 25). [5]

Figure 25: (Top) formation of isotactic polymer; (bottom) formation of syndiotactic polymer. Stereochemistry - iso and syn.png
Figure 25: (Top) formation of isotactic polymer; (bottom) formation of syndiotactic polymer.

However, if interactions between the substituents of the penultimate repeating unit and the terminal carbon atom are significant, then conformational factors could cause the monomer to add to the polymer in a way that minimizes steric or electrostatic interaction (Figure 26). [5]

Figure 26: Penultimate unit interactions cause monomer to add in a way that minimizes steric hindrance between substituent groups. (P represents polymer chain.) Stereochemistry - steric hindrance.png
Figure 26: Penultimate unit interactions cause monomer to add in a way that minimizes steric hindrance between substituent groups. (P represents polymer chain.)

Reactivity

Traditionally, the reactivity of monomers and radicals are assessed by the means of copolymerization data. The Q–e scheme, the most widely used tool for the semi-quantitative prediction of monomer reactivity ratios, was first proposed by Alfrey and Price in 1947. [27] The scheme takes into account the intrinsic thermodynamic stability and polar effects in the transition state. A given radical and a monomer are considered to have intrinsic reactivities Pi and Qj, respectively. [28] The polar effects in the transition state, the supposed permanent electric charge carried by that entity (radical or molecule), is quantified by the factor e, which is a constant for a given monomer, and has the same value for the radical derived from that specific monomer. For addition of monomer 2 to a growing polymer chain whose active end is the radical of monomer 1, the rate constant, k12, is postulated to be related to the four relevant reactivity parameters by

The monomer reactivity ratio for the addition of monomers 1 and 2 to this chain is given by [28] [29]

For the copolymerization of a given pair of monomers, the two experimental reactivity ratios r1 and r2 permit the evaluation of (Q1/Q2) and (e1 – e2). Values for each monomer can then be assigned relative to a reference monomer, usually chosen as styrene with the arbitrary values Q = 1.0 and e = –0.8. [29]

Applications

Free radical polymerization has found applications including the manufacture of polystyrene, thermoplastic block copolymer elastomers, [30] cardiovascular stents, [31] chemical surfactants [32] and lubricants. Block copolymers are used for a wide variety of applications including adhesives, footwear and toys.

Academic research

Free radical polymerization allows the functionalization of carbon nanotubes. [33] CNTs intrinsic electronic properties lead them to form large aggregates in solution, precluding useful applications. Adding small chemical groups to the walls of CNT can eliminate this propensity and tune the response to the surrounding environment. The use of polymers instead of smaller molecules can modify CNT properties (and conversely, nanotubes can modify polymer mechanical and electronic properties). [30] For example, researchers coated carbon nanotubes with polystyrene by first polymerizing polystyrene via chain radical polymerization and subsequently mixing it at 130 °C with carbon nanotubes to generate radicals and graft them onto the walls of carbon nanotubes (Figure 27). [34] Chain growth polymerization ("grafting to") synthesizes a polymer with predetermined properties. Purification of the polymer can be used to obtain a more uniform length distribution before grafting. Conversely, “grafting from”, with radical polymerization techniques such as atom transfer radical polymerization (ATRP) or nitroxide-mediated polymerization (NMP), allows rapid growth of high molecular weight polymers.

Figure 27: Grafting of a polystyrene free radical onto a single-walled carbon nanotube. Nanotube grafting 1.jpg
Figure 27: Grafting of a polystyrene free radical onto a single-walled carbon nanotube.

Radical polymerization also aids synthesis of nanocomposite hydrogels. [35] These gels are made of water-swellable nano-scale clay (especially those classed as smectites) enveloped by a network polymer. Aqueous dispersions of clay are treated with an initiator and a catalyst and the organic monomer, generally an acrylamide. Polymers grow off the initiators that are in turn bound to the clay. Due to recombination and disproportionation reactions, growing polymer chains bind to one another, forming a strong, cross-linked network polymer, with clay particles acting as branching points for multiple polymer chain segments. [36] Free radical polymerization used in this context allows the synthesis of polymers from a wide variety of substrates (the chemistries of suitable clays vary). Termination reactions unique to chain growth polymerization produce a material with flexibility, mechanical strength and biocompatibility.

Figure 28: General synthesis procedure for a nanocomposite hydrogel. NC Gel synthesis figure.png
Figure 28: General synthesis procedure for a nanocomposite hydrogel.

See also

Related Research Articles

A chain reaction is a sequence of reactions where a reactive product or by-product causes additional reactions to take place. In a chain reaction, positive feedback leads to a self-amplifying chain of events.

<span class="mw-page-title-main">Polymerization</span> Chemical reaction to form polymer chains

In polymer chemistry, polymerization, or polymerisation, is a process of reacting monomer molecules together in a chemical reaction to form polymer chains or three-dimensional networks. There are many forms of polymerization and different systems exist to categorize them.

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

In polymer chemistry, emulsion polymerization is a type of radical polymerization that usually starts with an emulsion incorporating water, monomers, and surfactants. The most common type of emulsion polymerization is an oil-in-water emulsion, in which droplets of monomer are emulsified in a continuous phase of water. Water-soluble polymers, such as certain polyvinyl alcohols or hydroxyethyl celluloses, can also be used to act as emulsifiers/stabilizers. The name "emulsion polymerization" is a misnomer that arises from a historical misconception. Rather than occurring in emulsion droplets, polymerization takes place in the latex/colloid particles that form spontaneously in the first few minutes of the process. These latex particles are typically 100 nm in size, and are made of many individual polymer chains. The particles are prevented from coagulating with each other because each particle is surrounded by the surfactant ('soap'); the charge on the surfactant repels other particles electrostatically. When water-soluble polymers are used as stabilizers instead of soap, the repulsion between particles arises because these water-soluble polymers form a 'hairy layer' around a particle that repels other particles, because pushing particles together would involve compressing these chains.

Chain-growth polymerization (AE) or chain-growth polymerisation (BE) is a polymerization technique where unsaturated monomer molecules add onto the active site on a growing polymer chain one at a time. There are a limited number of these active sites at any moment during the polymerization which gives this method its key characteristics.

<span class="mw-page-title-main">Step-growth polymerization</span> Type of polymerization reaction mechanism

In polymer chemistry, step-growth polymerization refers to a type of polymerization mechanism in which bi-functional or multifunctional monomers react to form first dimers, then trimers, longer oligomers and eventually long chain polymers. Many naturally-occurring and some synthetic polymers are produced by step-growth polymerization, e.g. polyesters, polyamides, polyurethanes, etc. Due to the nature of the polymerization mechanism, a high extent of reaction is required to achieve high molecular weight. The easiest way to visualize the mechanism of a step-growth polymerization is a group of people reaching out to hold their hands to form a human chain—each person has two hands. There also is the possibility to have more than two reactive sites on a monomer: In this case branched polymers production take place.

<span class="mw-page-title-main">End group</span> Functional group at the extremity of an oligomer or other macromolecule

End groups are an important aspect of polymer synthesis and characterization. In polymer chemistry, they are functional groups that are at the very ends of a macromolecule or oligomer (IUPAC). In polymer synthesis, like condensation polymerization and free-radical types of polymerization, end-groups are commonly used and can be analyzed by nuclear magnetic resonance (NMR) to determine the average length of the polymer. Other methods for characterization of polymers where end-groups are used are mass spectrometry and vibrational spectrometry, like infrared and raman spectroscopy. These groups are important for the analysis of polymers and for grafting to and from a polymer chain to create a new copolymer. One example of an end group is in the polymer poly(ethylene glycol) diacrylate where the end-groups are circled.

In polymer chemistry, anionic addition polymerization is a form of chain-growth polymerization or addition polymerization that involves the polymerization of monomers initiated with anions. The type of reaction has many manifestations, but traditionally vinyl monomers are used. Often anionic polymerization involves living polymerizations, which allows control of structure and composition.

Atom transfer radical polymerization (ATRP) is an example of a reversible-deactivation radical polymerization. Like its counterpart, ATRA, or atom transfer radical addition, ATRP is a means of forming a carbon-carbon bond with a transition metal catalyst. Polymerization from this method is called atom transfer radical addition polymerization (ATRAP). As the name implies, the atom transfer step is crucial in the reaction responsible for uniform polymer chain growth. ATRP was independently discovered by Mitsuo Sawamoto and by Krzysztof Matyjaszewski and Jin-Shan Wang in 1995.

<span class="mw-page-title-main">Reversible addition−fragmentation chain-transfer polymerization</span>

Reversible addition−fragmentation chain-transfer or RAFT polymerization is one of several kinds of reversible-deactivation radical polymerization. It makes use of a chain-transfer agent (CTA) in the form of a thiocarbonylthio compound to afford control over the generated molecular weight and polydispersity during a free-radical polymerization. Discovered at the Commonwealth Scientific and Industrial Research Organisation (CSIRO) of Australia in 1998, RAFT polymerization is one of several living or controlled radical polymerization techniques, others being atom transfer radical polymerization (ATRP) and nitroxide-mediated polymerization (NMP), etc. RAFT polymerization uses thiocarbonylthio compounds, such as dithioesters, thiocarbamates, and xanthates, to mediate the polymerization via a reversible chain-transfer process. As with other controlled radical polymerization techniques, RAFT polymerizations can be performed under conditions that favor low dispersity and a pre-chosen molecular weight. RAFT polymerization can be used to design polymers of complex architectures, such as linear block copolymers, comb-like, star, brush polymers, dendrimers and cross-linked networks.

In polymer chemistry, chain termination is any chemical reaction that ceases the formation of reactive intermediates in a chain propagation step in the course of a polymerization, effectively bringing it to a halt.

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

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

In polymer chemistry, the kinetic chain length of a polymer is the average number of units called monomers added to a growing chain during chain-growth polymerization. During this process, a polymer chain is formed when monomers are bonded together to form long chains known as polymers. Kinetic chain length is defined as the average number of monomers that react with an active center such as a radical from initiation to termination.

<span class="mw-page-title-main">Radical (chemistry)</span> Atom, molecule, or ion that has an unpaired valence electron; typically highly reactive

In chemistry, a radical, also known as a free radical, is an atom, molecule, or ion that has at least one unpaired valence electron. With some exceptions, these unpaired electrons make radicals highly chemically reactive. Many radicals spontaneously dimerize. Most organic radicals have short lifetimes.

Living free radical polymerization is a type of living polymerization where the active polymer chain end is a free radical. Several methods exist. IUPAC recommends to use the term "reversible-deactivation radical polymerization" instead of "living free radical polymerization", though the two terms are not synonymous.

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

Radical disproportionation encompasses a group of reactions in organic chemistry in which two radicals react to form two different non-radical products. Radicals in chemistry are defined as reactive atoms or molecules that contain an unpaired electron or electrons in an open shell. The unpaired electrons can cause radicals to be unstable and reactive. Reactions in radical chemistry can generate both radical and non-radical products. Radical disproportionation reactions can occur with many radicals in solution and in the gas phase. Due to the reactive nature of radical molecules, disproportionation proceeds rapidly and requires little to no activation energy. The most thoroughly studied radical disproportionation reactions have been conducted with alkyl radicals, but there are many organic molecules that can exhibit more complex, multi-step disproportionation reactions.

In polymer chemistry, reversible-deactivation radical polymerizations (RDRPs) are members of the class of reversible-deactivation polymerizations which exhibit much of the character of living polymerizations, but cannot be categorized as such as they are not without chain transfer or chain termination reactions. Several different names have been used in literature, which are:

In polymer chemistry, ionic polymerization is a chain-growth polymerization in which active centers are ions or ion pairs. It can be considered as an alternative to radical polymerization, and may refer to anionic polymerization or cationic polymerization.

In organosulfur chemistry, the thiol-ene reaction is an organic reaction between a thiol and an alkene to form a thioether. This reaction was first reported in 1905, but it gained prominence in the late 1990s and early 2000s for its feasibility and wide range of applications. This reaction is accepted as a click chemistry reaction given the reactions' high yield, stereoselectivity, high rate, and thermodynamic driving force.

References

  1. 1 2 3 4 5 6 7 Odian, George (2004). Principles of Polymerization (4th ed.). New York: Wiley-Interscience. ISBN   978-0-471-27400-1.
  2. Jenkins, A. D.; Kratochvíl, P.; Stepto, R. F. T.; Suter, U. W. (1996). "Glossary of basic terms in polymer science (IUPAC Recommendations 1996)". Pure and Applied Chemistry . 68 (12): 2287–2311. doi: 10.1351/pac199668122287 . S2CID   98774337.
  3. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 Cowie, J. M. G.; Arrighi, Valeria (2008). Polymers: Chemistry and Physics of Modern Materials (3rd ed.). Scotland: CRC Press. ISBN   978-0-8493-9813-1.
  4. 1 2 Hageman, H. J. (1985). "Photoinitiators for Free Radical Polymerization". Progress in Organic Coatings. 13 (2): 123–150. doi:10.1016/0033-0655(85)80021-2.
  5. 1 2 3 4 5 Stevens, Malcolm P. (1999). Polymer Chemistry: An Introduction. New York: Oxford University Press. ISBN   978-0-19-512444-6.
  6. 1 2 Islamova, R. M.; Puzin, Y. I.; Kraikin, V. A.; Fatykhov, A. A.; Dzhemilev, U. M. (2006). "Controlling the Polymerization of Methyl Methacrylate with Ternary Initiating Systems". Russian Journal of Applied Chemistry. 79 (9): 1509–1513. doi:10.1134/S1070427206090229. S2CID   94433190.
  7. 1 2 Islamova, R. M.; Puzin, Y. I.; Fatykhov, A. A.; Monakov, Y. B. (2006). "A Ternary Initiating System for Free Radical Polymerization of Methyl Methacrylate". Polymer Science, Series B. 48 (3): 130–133. doi:10.1134/S156009040605006X.
  8. Fried, Joel R. Polymer Science & Technology (2nd ed., Prentice-Hall 2003) p.36 ISBN   0-13-018168-4
  9. "Addition Polymerization". Materials World Modules. June 2009. Retrieved 1 April 2010.
  10. 1 2 "Polymer Synthesis". Case Western Reserve University. 2009. Archived from the original on 7 February 2010. Retrieved 10 March 2010.
  11. Leach, Mark R. "Radical Chemistry". Chemogenesis. Retrieved 2 April 2010.
  12. Pojman, John A.; Jason Willis; Dionne Fortenberry; Victor Ilyashenko; Akhtar M. Khan (1995). "Factors affecting propagating fronts of addition polymerization: Velocity, front curvature, temperature profile, conversion, and molecular weight distribution". Journal of Polymer Science Part A: Polymer Chemistry. 33 (4): 643–652. Bibcode:1995JPoSA..33..643P. doi:10.1002/pola.1995.080330406.
  13. Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) p.220 ISBN   0-12-601680-1
  14. Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) p.212 ISBN   0-12-601680-1
  15. The Mayo equation for chain transfer should not be confused with the Mayo–Lewis equation for copolymers.
  16. 1 2 Colombani, Daniel (1997). "Chain-Growth Control in Free Radical Polymerization". Progress in Polymer Science. 22 (8): 1649–1720. doi:10.1016/S0079-6700(97)00022-1.
  17. Kato, M; Kamigaito, M; Sawamoto, M; Higashimura, T (1995). "Polymerization of Methyl Methacrylate with the Carbon Tetrachloride / Dichlorotris-(triphenylphosphine)ruthenium(II) / Methylaluminum Bis(2,6-di-tert-butylphenoxide) Initiating System: Possibility of Living Radical Polymerization". Macromolecules . 28 (5): 1721–1723. Bibcode:1995MaMol..28.1721K. doi:10.1021/ma00109a056.
  18. Wang, J-S; Matyjaszewski, K (1995). "Controlled/"living" radical polymerization. Atom transfer radical polymerization in the presence of transition-metal complexes". J. Am. Chem. Soc. 117 (20): 5614–5615. doi:10.1021/ja00125a035.
  19. "The 2011 Wolf Prize in Chemistry". Wolf Fund. Archived from the original on 17 May 2007. Retrieved 21 February 2011.
  20. 1 2 "Stable Free Radical Polymerization". Xerox Corp. 2010. Archived from the original on 28 November 2003. Retrieved 10 March 2010.
  21. 1 2 Cowie, J. M. G. (1991). Polymers: Chemistry and Physics of Modern Materials (2nd ed.). Blackie (USA: Chapman & Hall). pp.  58–60. ISBN   978-0-216-92980-7.
  22. Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) pp.195-9 ISBN   0-12-601680-1
  23. Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) p.209 ISBN   0-12-601680-1
  24. Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) p.214 ISBN   0-12-601680-1
  25. Fried, Joel R. Polymer Science & Technology (2nd ed., Prentice-Hall 2003) p.39 ISBN   0-13-018168-4
  26. Clark, Jim (2003). "The Polymerization of Alkenes". ChemGuide. Retrieved 1 April 2010.
  27. Alfrey, Turner; Price, Charles C. (1947). "Relative reactivities in vinyl copolymerization". Journal of Polymer Science. 2 (1): 101–106. Bibcode:1947JPoSc...2..101A. doi:10.1002/pol.1947.120020112.
  28. 1 2 Allcock H.R., Lampe F.W. and Mark J.E. Contemporary Polymer Chemistry (3rd ed., Pearson Prentice-Hall 2003) p.364 ISBN   0-13-065056-0
  29. 1 2 Rudin, Alfred The Elements of Polymer Science and Engineering (Academic Press 1982) p.289 ISBN   0-12-601680-1
  30. 1 2 Braunecker, W. A.; K. Matyjaszewski (2007). "Controlled/living radical polymerization: Features, developments, and perspectives". Progress in Polymer Science. 32 (1): 93–146. doi:10.1016/j.progpolymsci.2006.11.002.
  31. Richard, R. E.; M. Schwarz; S. Ranade; A. K. Chan; K. Matyjaszewski; B. Sumerlin (2005). "Evaluation of acrylate-based block copolymers prepared by atom transfer radical polymerization as matrices for paclitaxel delivery from coronary stents". Biomacromolecules. 6 (6): 3410–3418. doi:10.1021/bm050464v. PMID   16283773.
  32. Burguiere, C.; S. Pascual; B. Coutin; A. Polton; M. Tardi; B. Charleux; K. Matyjaszewski; J. P. Vairon (2000). "Amphiphilic block copolymers prepared via controlled radical polymerization as surfactants for emulsion polymerization". Macromolecular Symposia. 150: 39–44. doi:10.1002/1521-3900(200002)150:1<39::AID-MASY39>3.0.CO;2-D.
  33. Homenick, C. M.; G. Lawson; A. Adronov (2007). "Polymer grafting of carbon nanotubes using living free-radical polymerization". Polymer Reviews. 47 (2): 265–270. doi:10.1080/15583720701271237. S2CID   96213227.
  34. Lou, X. D.; C. Detrembleur; V. Sciannamea; C. Pagnoulle; R. Jerome (2004). "Grafting of alkoxyamine end-capped (co)polymers onto multi-walled carbon nanotubes". Polymer. 45 (18): 6097–6102. doi:10.1016/j.polymer.2004.06.050. hdl: 2268/8255 .
  35. Haraguchi, K. (2008). "Nanocomposite hydrogels". Current Opinion in Solid State and Materials Science. 11 (3–4): 47–54. Bibcode:2007COSSM..11...47H. doi:10.1016/j.cossms.2008.05.001.
  36. Haraguchi, K.; Takehisa T. (2002). "Nanocomposite hydrogels: a unique organic-inorganic network structure with extraordinary mechanical, optical, and swelling/de-swelling properties". Advanced Materials. 14 (16): 1120–1123. doi: 10.1002/1521-4095(20020816)14:16<1120::AID-ADMA1120>3.0.CO;2-9 .