Emulsion polymerization

Emulsion polymerization is a type of radical polymerization that usually starts with an emulsion incorporating water, monomer, and surfactant. The most common type of emulsion polymerization is an oil-in-water emulsion, in which droplets of monomer (the oil) are emulsified (with surfactants) 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 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 stopped 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.

Emulsion polymerization is used to manufacture several commercially important polymers. Many of these polymers are used as solid materials and must be isolated from the aqueous dispersion after polymerization. In other cases the dispersion itself is the end product. A dispersion resulting from emulsion polymerization is often called a latex (especially if derived from a synthetic rubber) or an emulsion (even though "emulsion" strictly speaking refers to a dispersion of an immiscible liquid in water). These emulsions find applications in adhesives, paints, paper coating and textile coatings. They are often preferred over solvent-based products in these applications due to the absence of VOCs (Volatile Organic Compounds) in them.

IUPAC definition

Emulsion polymerization: Polymerization whereby monomer(s), initiator, dispersion
medium, and possibly colloid stabilizer constitute initially an inhomogeneous system
resulting in particles of colloidal dimensions containing the formed polymer.

Note: With the exception of mini-emulsion polymerization, the term “emulsion polymerization”
does not mean that polymerization occurs in the droplets of a monomer emulsion.[1]

Batch emulsion polymerization: Emulsion polymerization in which all the ingredients are
placed in a reactor prior to reaction.[2]

Advantages of emulsion polymerization include:[3]

Disadvantages of emulsion polymerization include:

History

The early history of emulsion polymerization is connected with the field of synthetic rubber.[4][5] The idea of using an emulsified monomer in an aqueous suspension or emulsion was first conceived at Bayer, before World War I, in an attempt to prepare synthetic rubber.[6][7] The impetus for this development was the observation that natural rubber is produced at room temperature in dispersed particles stabilized by colloidal polymers, so the industrial chemists tried to duplicate these conditions. The Bayer workers used naturally occurring polymers such as gelatin, egg albumin, and starch to stabilize their dispersion. By today's definition these were not true emulsion polymerizations, but suspension polymerizations.

The first "true" emulsion polymerizations, which used a surface-active agent and polymerization initiator, were conducted in the 1920s to polymerize isoprene.[8][9] Over the next twenty years, through the end of World War II, efficient methods for production of several forms of synthetic rubber by emulsion polymerization were developed, but relatively few publications in the scientific literature appeared: most disclosures were confined to patents or were kept secret due to wartime needs.

After World War II, emulsion polymerization was extended to production of plastics. Manufacture of dispersions to be used in latex paints and other products sold as liquid dispersions commenced. Ever more sophisticated processes were devised to prepare products that replaced solvent-based materials. Ironically, synthetic rubber manufacture turned more and more away from emulsion polymerization as new organometallic catalysts were developed that allowed much better control of polymer architecture.

Theory

The first successful theory to explain the distinct features of emulsion polymerization was developed by Smith and Ewart,[10] and Harkins[11] in the 1940s, based on their studies of polystyrene. Smith and Ewart arbitrarily divided the mechanism of emulsion polymerization into three stages or intervals. Subsequently it has been recognized that not all monomers or systems undergo these particular three intervals. Nevertheless, the Smith-Ewart description is a useful starting point to analyze emulsion polymerizations.

Schematic diagram of emulsion polymerization

The Smith-Ewart-Harkins theory for the mechanism of free-radical emulsion polymerization is summarized by the following steps:

Smith-Ewart theory does not predict the specific polymerization behavior when the monomer is somewhat water-soluble, like methyl methacrylate or vinyl acetate. In these cases homogeneous nucleation occurs: particles are formed without the presence or need for surfactant micelles.[12]

High molecular weights are developed in emulsion polymerization because the concentration of growing chains within each polymer particle is very low. In conventional radical polymerization, the concentration of growing chains is higher, which leads to termination by coupling, which ultimately results in shorter polymer chains. The original Smith-Ewart-Hawkins mechanism required each particle to contain either zero or one growing chain. Improved understanding of emulsion polymerization has relaxed that criterion to include more than one growing chain per particle, however, the number of growing chains per particle is still considered to be very low.

Because of the complex chemistry that occurs during an emulsion polymerization, including polymerization kinetics and particle formation kinetics, quantitative understanding of the mechanism of emulsion polymerization has required extensive computer simulation. Gilbert has summarized a recent theory.[13]

Process

Emulsion polymerizations have been used in batch, semi-batch, and continuous processes. The choice depends on the properties desired in the final polymer or dispersion and on the economics of the product. Modern process control schemes have enabled the development of complex reaction processes, with ingredients such as initiator, monomer, and surfactant added at the beginning, during, or at the end of the reaction.

Early styrene-butadiene rubber (SBR) recipes are examples of true batch processes: all ingredients added at the same time to the reactor. Semi-batch recipes usually include a programmed feed of monomer to the reactor. This enables a starve-fed reaction to ensure a good distribution of monomers into the polymer backbone chain. Continuous processes have been used to manufacture various grades of synthetic rubber.

Some polymerizations are stopped before all the monomer has reacted. This minimizes chain transfer to polymer. In such cases the monomer must be removed or stripped from the dispersion.

Colloidal stability is a factor in design of an emulsion polymerization process. For dry or isolated products, the polymer dispersion must be isolated, or converted into solid form. This can be accomplished by simple heating of the dispersion until all water evaporates. More commonly, the dispersion is destabilized (sometimes called "broken") by addition of a multivalent cation. Alternatively, acidification will destabilize a dispersion with a carboxylic acid surfactant. These techniques may be employed in combination with application of shear to increase the rate of destabilization. After isolation of the polymer, it is usually washed, dried, and packaged.

By contrast, products sold as a dispersion are designed with a high degree of colloidal stability. Colloidal properties such as particle size, particle size distribution, and viscosity are of critical importance to the performance of these dispersions.

Living polymerization processes that are carried out via emulsion polymerization such as iodine-transfer polymerization and RAFT have been developed.

Ingredients

Monomers

Typical monomers are those that undergo radical polymerization, are liquid or gaseous at reaction conditions, and are poorly soluble in water. Solid monomers are difficult to disperse in water. If monomer solubility is too high, particle formation may not occur and the reaction kinetics reduce to that of solution polymerization.

Ethylene and other simple olefins must be polymerized at very high pressures (up to 800 bar).

Comonomers

Copolymerization is common in emulsion polymerization. The same rules and comonomer pairs that exist in radical polymerization operate in emulsion polymerization. However, copolymerization kinetics are greatly influenced by the aqueous solubility of the monomers. Monomers with greater aqueous solubility will tend to partition in the aqueous phase and not in the polymer particle. They will not get incorporated as readily in the polymer chain as monomers with lower aqueous solubility. This can be avoided by a programmed addition of monomer using a semi-batch process.

Ethylene and other olefins are used as minor comonomers in emulsion polymerization, notably in vinyl acetate copolymers.

Small amounts of acrylic acid or other ionizable monomers are sometimes used to confer colloidal stability to a dispersion.

Initiators

Both thermal and redox generation of free radicals have been used in emulsion polymerization. Persulfate salts are commonly used in both initiation modes. The persulfate ion readily breaks up into sulfate radical ions above about 50 °C, providing a thermal source of initiation. Redox initiation takes place when an oxidant such as a persulfate salt, a reducing agent such as glucose, Rongalite, or sulfite, and a redox catalyst such as an iron compound are all included in the polymerization recipe. Redox recipes are not limited by temperature and are used for polymerizations that take place below 50 °C.

Although organic peroxides and hydroperoxides are used in emulsion polymerization, initiators are usually water soluble and partition into the water phase. This enables the particle generation behavior described in the theory section. In redox initiation, either the oxidant or the reducing agent (or both) must be water-soluble, but one component can be water-insoluble.

Surfactants

Selection of the correct surfactant is critical to the development of any emulsion polymerization process. The surfactant must enable a fast rate of polymerization, minimize coagulum or fouling in the reactor and other process equipment, prevent an unacceptably high viscosity during polymerization (which leads to poor heat transfer), and maintain or even improve properties in the final product such as tensile strength, gloss, and water absorption.

Anionic, nonionic, and cationic surfactants have been used, although anionic surfactants are by far most prevalent. Surfactants with a low critical micelle concentration (CMC) are favored; the polymerization rate shows a dramatic increase when the surfactant level is above the CMC, and minimization of the surfactant is preferred for economic reasons and the (usually) adverse effect of surfactant on the physical properties of the resulting polymer. Mixtures of surfactants are often used, including mixtures of anionic with nonionic surfactants. Mixtures of cationic and anionic surfactants form insoluble salts and are not useful.

Examples of surfactants commonly used in emulsion polymerization include fatty acids, sodium lauryl sulfate, and alpha olefin sulfonate.

Non-surfactant stabilizers

Some grades of polyvinyl alcohol and other water-soluble polymers can promote emulsion polymerization even though they do not typically form micelles and do not act as surfactants (for example, they do not lower surface tension). It is believed that growing polymer chains graft onto these water-soluble polymers, which stabilize the resulting particles.[14]

Dispersions prepared with such stabilizers typically exhibit excellent colloidal stability (for example, dry powders may be mixed into the dispersion without causing coagulation). However, they often result in products that are very water sensitive due to the presence of the water-soluble polymer.

Other ingredients

Other ingredients found in emulsion polymerization include chain transfer agents, buffering agents, and inert salts. Preservatives are added to products sold as liquid dispersions to retard bacterial growth. These are usually added after polymerization, however.

Smith-Ewart Mechanism

Stage 1

When radicals generated in the aqueous phase encounter the monomer within the micelle, it imitates the polymerization process. The conversion of monomer to polymer within the micelle lower the monomer concentration and generates monomer concentration gradient. Consequently, the monomer from monomer droplets and uninitiated micelles begin to diffuse to this growing, polymer-containing micelles. Those micelles that did not encounter radical during the earlier stage of conversion begin to disappear, losing their monomer monomer and surfactant to those that have been initiated. At the end of this stage, the number of growing polymer particles (originally micelles) stays the same.

Stage 2

This stage is also known as steady state reaction stage. In this stage, monomer droplets act as reservoirs supplying monomer to the growing polymer particle by diffusion through the water. In this stage, the ratio of growing free-radical per particles can be divided into three cases. The ratio of growing free-radicals per micelles is (first case), (second case), and (third case), where the second case happens often in practice. A monomer-swollen micelle that has been struck by a radical contains one growing chain. With this only one radical per particle, the chain cannot terminate, and it will continue to grow until a second initiator radical enters the particle. For the second case, where particles are tiny and radical are highly reactive, the rate of termination is much greater than the rate of propagation, so the chain growth is terminated immediately after the entrance of the second initiator radical. The particle then remains dormant until a third initiator radical enters, initiating the growth of a second chain. Consequently, The polymer chain in this case either have zero radical (dormant state), or 1 radical (polymer growing state) and a very short period of 2 radical (terminating state) which could be safely ignored for the free-radicals per particle calculation. At any given time (for case 2), a micelles contains either one growing chain or no growing chains (assumed to be equally probably). Thus, on average, there is around 1/2 radical per particle. The polymerization rate in this stage can be expressed by

where is the homogeneous propagation rate constant for polymerization within the particles and is the equilibrium monomer concentration within a particle. represents concentration of polymerizing radical in the reaction. For the case 2, where free radicals per micelles is . could be calculated in following expression:

where is number concentration of micelles (number of micelles per unit volume), and is Avogadro's constant (radicals/mol radicals). Consequently, the rate of polymerization is then

Stage 3

Monomer droplet begins to deplete as reaction continues. Micelles with polymer in this stage maybe sufficiently large enough that they contain more than 1 radical per micelles.

Applications

Polymers produced by emulsion polymerization can be divided into three rough categories.

See also

References

  1. "Terminology of polymers and polymerization processes in dispersed systems (IUPAC Recommendations 2011)" (PDF). Pure and Applied Chemistry. 83 (12): 2229–2259. 2011. doi:10.1351/PAC-REC-10-06-03.
  2. "Terminology of polymers and polymerization processes in dispersed systems (IUPAC Recommendations 2011)" (PDF). Pure and Applied Chemistry. 83 (12): 2229–2259. 2011. doi:10.1351/PAC-REC-10-06-03.
  3. Odian, G, Principles of Polymerization, Wiley, New York
  4. Whitby, G. S.; Katz, M. Ind. Eng. Chem., (1933), 25, 1338.
  5. Hohenstein, W. P.; Mark, H. J. Polym. Chem., (1946), 1, 127.
  6. German patent 250690 (Sept. 12, 1909)
  7. U.S. Patent 1149577, filed Jan. 6, 1913.
  8. German patent 558890 (Jan. 8, 1927)
  9. U. S. Patent 1732795, filed Sept.13, 1927.
  10. Smith, W. V.; Ewart, R. H. J. Chem. Phys., (1948), 16, 592.
  11. Harkins, W. D. J. Am. Chem. Soc., (1947), 69, 1428.
  12. Fitch, R. M. Polymer Colloids, Plenum, NY 1971.
  13. Gilbert, R. G. Emulsion Polymerization: a Mechanistic Approach Academic Press, London, 1996.
  14. Kim, N., Sudol, E., Dimonie, V. and El-Aasser, M. (2004). Grafting of PVA in Miniemulsion Copolymerizations of n -Butyl Acrylate and Methyl Methacrylate Using Water-Soluble, Partially Water-Soluble, and Oil-Soluble Initiators. Macromolecules, 37(9), pp.3180-3187. DOI: 10.1021/ma035153w
This article is issued from Wikipedia - version of the 12/1/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.