You are on page 1of 45

9

Paints, Varnishes, and Related Products


K. F. Lin

1. RELATIONSHIP OF FATS AND OILS TO THE PAINT-COATING INDUSTRY Historically, drying oils have been the major lm formers of coatings, including paints, varnishes, and inks. Although it is not certain whether linseed oil was used in paints in ancient Egypt, ax was grown and ax seeds were collected at that time. The early Renaissance was probably the real beginning of paints as we know them today in the West. The Van Eyck brothers (13881441) are said to be the rst to use linseed oil as a binder (1). Whereas in China, tung oil has been used for centuries as waterproong caulking or as coating for wood objects including boats, houses, and furniture. The rst American paint factory was opened in Boston in 1737 by Thomas Childs (2). The pigment and oil were placed on a granite trough, and a granite ball, known as the Boston Stone, was then rolled over the mixture to make the paint. The ball is now preserved and serves as the symbol for the Federation of Societies for Paint Technology. The drying oils owe their value as raw materials for decorative and protective coatings to their ability to polymerize and cross-link, or dry, after they have been applied to a surface, to form tough, adherent, impervious, and abrasion-resistant

Baileys Industrial Oil and Fat Products, Sixth Edition, Six Volume Set. Edited by Fereidoon Shahidi. Copyright # 2005 John Wiley & Sons, Inc.

307

308

PAINTS, VARNISHES, AND RELATED PRODUCTS

lms. Their lm-forming properties are closely related to their degree of unsaturation, since it is through the unsaturated centers or double bonds that polymerization and cross-linking take place. With one exception (to be noted later) the oils used in paints varnishes and similar products are relatively high in iodine value. In any given product, there is an optimum degree of reactivity in the oil; the speed with which the oil dries must be balanced against such factors as elasticity and durability in the paint lm. In general, however, unsaturation is at a premium in paint and varnish oils, and the oils in greatest demand are those in which drying takes place most readily. The form in which drying oils are used in coating applications has gone through an evolutionary change over time. The simplest and most primitive way is to use them directly as the lm former of a coating. It was discovered that drying oils may be made more useful by altering their natural state. By aging in vats, by heating, or by blowing air through, the viscosity and drying characteristics of the drying oil may be changed enough to improve its general properties for coating applications. In the case of fast-drying oils with conjugated double bonds, such as tung, oiticica, and dehydrated castor oil, heat treatment is necessary to gas-proof them, so that the oils do not dry into undesirable wrinked and/or frosted lms. The oils are not necessarily used in their original form of triglycerides for coating applications. It has become a common practice to hydrolyze them rst, and the free fatty acids are then used to synthesize coating resins with certain advantages. Therefore, the term fats and oils includes fatty acids for the purpose of the discussions in this Chapter, unless otherwise specied. Through progress, people found that sometimes mixtures of different oils could be used to greater advantage and that natural gums could be added; thus oleoresinous varnishes were born. Thereafter, human creativity started to make rapid and diversed progress in the development of new coating materials, many of which have departed completely away from the drying oil base. When the drying characteristics of oils were relied on as the sole (or major) cause for a varnish-based coating lm to dry, those oils belonging to the linolenic or conjugated acid groups, such as linseed, perilla, tung, oiticica, and highly unsaturated winterized sh oils were of the prime interest to coating formulators. Since about the 1950s, with the advent of synthetic resins, particularly, alkyd resins, it has become possible to make considerable use of oils with poorer drying characteristics. Semidrying oils such as soybean oil, safower oil, and sunower seed oil have become viable as raw materials for making drying paints. Nondrying oils, such as coconut oil, are also used in coating materials. However, their function is primarily that of a plasticizer rather than of an active component for air drying. In addition, castor oil, a nondrying oil, has been converted chemically by dehydration to give excellent drying property. There is no denying that the once prominent position of fats and oils as the most important raw material for coatings has been greatly eroded by other materials. The total U.S. consumption of drying and semidrying oils in coating and allied applications peaked around 1950 at about 1.2 billion lb. It went down steadily thereafter. According to SRI International (3), the direct use of drying oils accounted for only

A BRIEF OVERVIEW OF THE COATINGS TECHNOLOGY

309

about 4% of the total lm formers consumed in the United States in 1990, at about 98 million lb, whereas the consumption of alkyds, urethane alkyds, and epoxy esters was estimated at 645, 40, and 12 million lb, respectively (46). Very little drying oil is used in paints at present. Drying oils and oxidizing alkyds have been studied as binders for organic inorganic coatings (4). Urethane coatings are the fastest growing sector. Use in 2003 was 1:6 106 t (5). Epoxy resins are among the most widely used (6).

2. A BRIEF OVERVIEW OF THE COATINGS TECHNOLOGY Modern requirements in protective coatings are extremely diverse and exacting. They go far beyond the mere necessity of protecting the nished surface from weather or from ordinary wear or abrasion. Some coatings (e.g., those employed as electrical insulation) must possess extreme resistance to high temperatures or to penetration by moisture. Others (e.g., marine varnishes and the enamels for coating the interior of cans) must withstand prolonged contact with water or aqueous solutions. Modern assembly-line methods of manufacture produce many particular requirements and have created a special demand for quick-drying nishes. The wide distribution of illustrated journals, the proliferation of advertising matter, and the development of high speed printing processes have greatly elaborated the requirements of users of printing inks. Tung and other conjugated oils are particularly suitable for manufacture of fast-drying nishes, and for a time, the consumption of these oils increased signicantly in response to more exacting requirements for specialized nishes. New systems based on epoxy resins, urethane polymers, silicones, and other synthetic intermediates have greatly decreased dependency on tung oil, and use has shrunk signicantly. The complex and diversied requirements of modern industry have to a large extent removed the manufacture of paints and varnishes from the category of an art to that of a science. In most plants, the manufacturing processes are now carried out under careful laboratory control and are freely modied or revised, whenever revision is indicated, in accordance with known scientic principles. As a result, the industry has been able to offer a succession of constantly improved products through periods of uctuation in the availability of many important raw materials, pressures for solvent replacement to meet emerging air quality standards, and extensive pigment reformulation to replace mercury and lead to conform to new federal regulations on toxicity. A most important development in the modern paint and varnish industry has been the introduction of synthetic resins as replacements for natural resins in the manufacture of varnishes and enamels. By using synthetic resins it has been possible to produce a variety of coatings that, in many cases, have important points of superiority over any of those compounded from natural resins. The synthetic vehicles are particularly distinguished by their hardness and durability and their high degree of resistance to the action of water, alkalies, and other chemical agents.

310

PAINTS, VARNISHES, AND RELATED PRODUCTS

New methods for application of paint lms and new procedures for curing have placed challenging demands on the resourcefulness of the resin chemist. For years brushing was virtually the only method of application; later, spraying, was used. Now a host of new application and curing techniques are commonplace. These include roller coating, dipping, coil coating, powder coating, electrodeposition, hot spray, uid bed coating, electrostatic spray, two-component spray, ultraviolet cure, and electron beam cure. The two foremost reasons for the decline of the direct use of oils, including oleoresinous varnishes, for coating applications are performance and environment. Drying oils by themselves, or even in the form of oleoresinous varnishes, do not give the drying speed and, sometimes, lm properties that would satisfy the modern needs. The ease in the application and cleaning of latex paints caused oil-based paint to lose most, if not all, of the trade-sales market. The implementation of Rule 66 in California was the opening salvo for protecting the environment against solvent-based coatings of which practically all oil-based coatings belong. Thus the emphasis has switched to the development and commercialization of other coating materials or systems that are more environment friendly than oil-based materials. Indeed, it is quite remarkable that in the face of such severe odds, oil-based materials have been able to hold ground as well as they have. There is an extraordinary body of terms used to dene various features of protective coatings. Before discussing paints and varnishes and the particular function of fats and oils in coatings, it is desirable to review and dene some of the language of the industry. Protective coatings protect (or decorate) surfaces. Greases, mineral oils, plastic web coats, and mastic compositions may be used for protection, particularly of metal surfaces; but in the usual sense, protective coatings are materials that form durable lms adhered to the surface to provide protection. A varnish is a solvent-thinned combination of a drying oil and a hard resin. Also, a varnish is the clear lm obtained using a varnish as a coating vehicle. By extension, vehicles used for clear lms are called varnishes although the vehicle may be a true varnish, an alkyd resin solution, a urethane-modied oil, or even a lacquer. A paint is a pigmented system applied to hide and protect a surface. Paints contain a wide range of ingredients as follows: The vehicle is the carrier for pigment, consisting of combinations of oils, resins, polymers, and solvents; the nonvolatile portion is commonly called the binder. Prime pigments are used for their ability to hide or cover the surface. The term hiding power is used to describe the relative ability of xed amounts of different pigments to cover a surface and depends on the difference in refractive indices between the pigment and the binder; thus primary pigments usually have high values in refractive index. Extender pigments give relatively little hiding, but their cost is lower than that of prime pigments; they provide control of such properties as ow consistency, durability, and adhesion.

A BRIEF OVERVIEW OF THE COATINGS TECHNOLOGY

311

Driers are metal salts, especially of cobalt, manganese, zirconium, calcium, and iron, that accelerate the conversion of the liquid lm to a solid; lead was commonly used as a primary drier, but due to its toxicity, it is rarely used now. Solvents or thinners control paint consistency and application properties. Slow solvents evaporate slowly and leave the lm open (workable) for longer periods than fast solvents, which evaporate rapidly; in water-thinned paints, water is the thinner and there is no control over rate of evaporation. A variety of other materials may be added for special effects. Antiskin agents minimize skin formation on the top of the can during use or storage. Mildewcides protect the applied lm from fungus growth. Wetting agents or griding aids promote wetting of pigment particles by the vehicle. Antiood and antioat agents minimize ooding and oating, deciencies characterized by separation of colored pigments during drying. Antisetting agents minimize separation of pigment into a rm or hard mass in the bottom of the can. Antisag agents minimize sagging or curtaining of wet lms during application. Pufng agents, or thixotropic agents, increase the paint consistency and minimize sagging by giving a thixotropic consistency to paint (a type of behavior in which the viscosity of the system decreases when agitated, as under the shear of brushing, and increases when allowed to stand). Paints are made by grinding pigments in the vehicle. Actually the term grinding is somewhat inaccurate. The pigments are received from the manufacturer are already as ne in particle size as they will be in the nished paint. The grinding operation is designed to break up the aggregates of pigment particles and to disperse them in the vehicle so that each particle is wetted. Griding is usually carried out in roller mills, in which shear between steel rollers disperses pigments, in ball mills or pebble mills, in which steel balls or pebbles rotating and rubbing against each other in a closed cylinder to produce the shear for dispersing the pigment, or in sand mills, in which agitation of sand causes pigment separation and dispersion. For products in which a ne grind (ne pigment dispersion) is not required, as in barn paints or house paints, high speed rotors may be used to grind the paint. In typical paint manufacture, a paste prepared from all of the pigment and a portion of the vehicle is subjected to the appropriate grinding technique until the desired neness of grind is attained, and the resultant paste is let down (diluted) with the remaining portion of vehicle and solvent. Fineness of grind is commonly expressed numerically using a grind gauge, which is a shallow wedge cut from polished metal. Paint is lled into the wedge

312

PAINTS, VARNISHES, AND RELATED PRODUCTS

and a bar is drawn across the surface. With ne grinds, the paint lls the wedge even to the shallowest part. With coarse grinds, the paint is pulled away from the shallow edge of the wedge of the deeper end. The line of demarcation ranging from 0 (very coarse grind) to 8 (ultrane grind) designated the quality of the pigment dispersion. An enamel is a paint based on a vehicle that dries to a considerably harder lm than paints derived from unmodied drying oils. Paints and enamels are classied by type of nish as follows: Flat paints or enamels dry to a velvety nonglossy or matte surface. Semigloss paints or enamels dry to an intermediate gloss range between at and glossy. Gloss paints or enamels dry to a highly reecting surface. There are many variations in the nomenclature, and lms that are called gloss lms by one observer might be classied as a semigloss by the individual who demands mirrorlike surfaces. Other designations might be used also such as eggshell (between at and semigloss) or full gloss to differentiate mirror gloss from normal gloss. The degree of gloss is measured by a glossmeter, which measures light reectance at a low angle from the horizontal (20 gloss) or high angle (60 gloss). The 60 reectance is most common, and although ranges of values are not sharply divided, the general consensus is as follows:

Type of Paint Flat Eggshell Semigloss Low gloss High or full gloss

60 Reectance 4 or 5 maximum 520 2060 (up to 80) 80 to 90 90 to 98

Since the ability of a surface to reect light, which gives the gloss measurement, depends on the smoothness of the surface, one can readily visualize that a coating with a greater surplus of binder over its pigment content will have a greater ability to produce a smooth surface, thus high in gloss. Conversely, with a high pigment content, there will be more pigment particles or aggregates at or near the lm surface to cause a scattering of light, thus resulting in low gloss. The amount of pigment in the paint is measured by the pigment volume concentration (PVC), i.e., the volume percent of pigments in the dried paint lm. In solvent-based systems, products of low PVC (up to 2025%) are glossy, products in the middle range

A BRIEF OVERVIEW OF THE COATINGS TECHNOLOGY

313

(2550%) have a semigloss nish, and products in the high range (4570) have a at or matte surface. Gloss in water-thinned systems does not correspond to the above, particularly in emulsion systems, because the pigment particles are not wetted uniformly by the binder. Solvent-thinned paints containing certain atting agents such as extremely ne silica do not conform to the normal pattern. Coating systems are divided into two general classications, depending on the point of application. Trade sales nishes are purchased by the user in a paint store or hardware store (or today even in a drugstore or a supermarket) and are applied by the purchaser, usually by brush or roller. Included are barn paints, house paints, trim paints, varnishes, porch and deck enamels; wall paints, architectural enamels, and similar user-applied nishes. Industrial nishes are applied to objects by the manufacturer, usually by spraying, dipping, roller coating, air knife, or other high speed production application methods, and they are usually force dried by baking. Implement, automotive, appliance, and furniture nishes are typical industrials. Finishes are also described by the function of the paint. A primer is used to coat the original surface. Its major functional property is good adhesion. Protection against corrosion is an especially important characteristic of metal primers. Hiding is a secondary function. A sealer is a primer whose major function is covering a porous surface such as plaster, gypsum board, or paperboard with a surface coating that exhibits a minimum penetration into the surface. Good sealing prevents ghosting, a lm defect in which variation in penetration causes gloss differences and visual color differences in the nal coat. A typical ghosting effect is obtained in painting wallpaper in which the pattern can show through multiple coats because of the variability in porosity of the substrate. A sanding sealer is designed for easy sanding after short dry so that smoother top coats can be attained. Sanding sealers are most important in furniture nishes and industrial nishes such as automotive systems. An undercoat is another name for primer or sealer, especially as an enamel undercoat, which serves to supply a uniform base for an enamel so that there will not be a wide variation in gloss. Undercoats may be applied over sealers. The top coat or nish coat is the outside layer of paint applied over the primer or sealer. The specication of products for coating applications involves a large number of factors, including color and color retention of lms, rate of setup, bakability, rate of cure of lms, hardness of lms, adhesion, wetting action in grinding with pigments, exibility and retention of exibility on aging, reactivity with pigments, reactivity with driers, water resistance, alkali resistance, solvent resistance, viscosity, viscosity stability in the package as clear products or in pigmented systems, thermoplasticity, durability, compatibility with other lm-forming agents, mar resistance, abrasion resistance, stain resistance, performance in the varnish kettle, gloss, gloss retention, etc. Obviously, no single product can be optimum in all of these characteristics, and in each use a compromise must be made to provide the best performance in the intended usage.

314

PAINTS, VARNISHES, AND RELATED PRODUCTS

3. FILM DRYING PROCESS OF OIL-BASED COATING MATERIALS 3.1. Drying of a Nonconjugated System As mentioned earlier, it is through the unsaturated centers or double bonds of the fatty acids in drying oils that polymerization and cross-linking, i.e., drying, takes place. Hence, oils are conventionally classied, based on their iodine values into three groups: drying, semidrying, and nondrying. The generally accepted demarcations are, respectively, >140, 140125, and <125. These numbers are, more or less, arbitrarily assigned. When the oil contains conjugated double bonds, the iodine values determined are usually low due to incomplete halogen absorption. Such a classication can only be used for a rough guidance. Wicks and Jones (7) suggested that the methylene groups between two double bonds, i.e., the CH2 groups allylic to two C C groups, are much more reactive than those being allylic to only one C C group and are mostly responsible for the drying of the nonconjugated oil. Thus the average number of such groups fn in an oil molecule serves as a better indicator for the drying characteristics of the oil. An oil with an fn value of greater than 2.2 is a drying oil, those with fn values somewhat less than 2.2 are semidrying, and there is no sharp dividing composition between semidrying and nondrying oils, according to the authors. While such a classication system does have merits over the conventional way (based on iodine value), it does not provide a rule for classifying oils with conjugated unsaturation. The chemical mechanism of drying has been established as an oxidative radical chain reaction process, which has been summarized as follows (8): 1. A period of induction at the beginning of the reaction during which no visible change in physical or chemical properties in the oil is noticed; natural antioxidant compounds are consumed during this period. 2. The reaction becomes perceptible and oxygen uptake is considerable; discrete interaction of oxygen and olens takes place followed by the formation of hydroperoxides. 3. Conjugation of double bonds occurs accompanied by isomerization of cis to trans unsaturation. 4. The hydroperoxides start to decompose to form a high free-radical concentration; the reaction becomes autocatalytic. 5. Polymerization and scission reactions begin and yield high molecular weight cross-linked products and low molecular weight carbonyl and hydroxy compounds; carbon dioxide and water are also formed and are present in the volatile products of lm formation. It is now generally believed that the induction is slow at rst but is autocatalytic and the rate increases steadily. The rate depends on the reaction conditions such as temperature, light, and traces of heavy metals or inhibitors in the oil or coating (9).

FILM DRYING PROCESS OF OIL-BASED COATING MATERIALS

315

The active sites are the allylic carbon a to a double bond, especially those a to two double bonds with one on each side, such as carbon number 11 in a 9,12-octadecadienoic (linoleic) acid and proceeds through the following mechanism:
CH CH CH2 CH CH
H

Abstraction of a hydrogen atom

CH CH CH CH CH CH CH CH CH
O2

CH CH CH CH CH CH

Resonance hybrid free radicals

CH CH CH OO CH CH CH CH CH CH OO
H

CH CH Three possible peroxy radicals CH CH OO CH CH

Addition of hydrogen atom abstracted from another linoleate molecule CH CH CH CH OOH CH CH Three possible hydroperoxides, two of which are conjugated

CH CH CH OOH CH CH CH CH CH CH OOH

The initial step is believed to be the dehydrogenation from the a-methylene group to form a radical. Since such a hydrogen extraction would require a considerable amount of energy, a number of investigators proposed that the hydrogen is removed through reaction with a free radical. Thus a radical, A , abstracts a hydrogen from a molecule of linoleate, RH, to form the radical R , RH A ! R AH Since the radical is allylic to the double bonds on either side of it, resonance hybrid free radicals are formed resulting in shifting the double bonds to a conjugated position. This is then followed by: R O2 ! RO2

316

PAINTS, VARNISHES, AND RELATED PRODUCTS

and RO2 RH ! ROOH R The net reaction is hydroperoxide formation: RH O2 ! ROOH During the oxidation to form hydroperoxides, the natural cis,cis unsaturation of linoleate is converted to cis, trans and trans, trans isomers. Privett and co-workers (10) concluded that at least 90% of linoleate hydroperoxide preparations are conjugated. When the oxidation is conducted at 0 C the hydroperoxides are predominately cis, trans isomers, but room temperature oxidation produces a large amount of trans, trans unsaturation (11, 12). Ethyl or methyl linoleate hydroperoxides are relatively low melting and as a result purication by crystallization is difcult. Bailey and Barlow (13) prepared high melting p-phenylphenacyl linoleate, oxidized the ester in benzene solution, and isolated virtually pure hydroperoxide by crystallization. Infrared spectra of the 99% purity p-phenylphenacyl linoleate hydroperoxide correspond to a trans, trans conjugated isomer. The autoxidation of linoleate described above shows the characteristic features of a chain reaction involving free radicals. Materials that decompose to form free radicals catalyze the reaction even when present in very low concentrations to produce high yields of hydroperoxides; initiation of the reaction by light can produce quantum yields much greater than unity and easily oxidized substances that consume free radicals, but do not themselves undergo signicant autoxidation, can markedly inhibit the chain reaction. Although there is quite general agreement on the mechanism of the chain propagation reaction, there is much less unanimity of opinion on the primary reaction to produce the radicals (indicated as A above) responsible for the initiation of the chain reaction. Originally, it was proposed that hydroperoxides are the initial products of autoxidation (14, 15). Primarily because of the high energy requirement for rupture of the a-methylenic carbonhydrogen bond several authors (1619) almost simultaneously concluded that the initial point of oxidative attack was the double bond and not the a-methylene group, although some (16) proposed a limited attack at the double bond to produce radicals in sufcient amount to initiate the chain reaction through the a-methylenic carbon. Kahn (20) questioned the formation of a diradical and proposed direct addition of oxygen to a double bond to form a cyclic transition state, which breaks down to yield the hydroperoxide. The theory of oxidation has received little support, because it does not explain the inhibitory effect of free-radical acceptors in the initial stages of autoxidation. It has been contended that the direct attack of oxygen on the double bond has low thermodynamic probability (21, 22), and it has been considered that trace metal contaminants catalyze the initiation of autoxidation by producing free radicals through electron transfer. Alternative pathways are as follows, using cobalt as an

FILM DRYING PROCESS OF OIL-BASED COATING MATERIALS

317

example of a metal that can facilely shift valence states in oxidationreduction reactions: 1. Reduction activation of trace hydroperoxides in the system yields free radicals. Co2 ROOH ! Co3 OH RO 2. Direct reaction of a metal ion with oxygen: Co2 O2 ! Co3 O 2 The O radical ion reacts readily with a proton to form the HO2 radical, 2 which can initiate the chain reaction of oxidation. 3. Complex reaction of metal compounds with oxygen and subsequent formation of an HO2 radical. Co2 O ! Co3 O2 2 Co3 O2 XH ! Co3 X HO2 4. Oxidation by electron transfer of the a-methylenic group by the metal ion. Co3 RH ! Co2 H R According to Uri (21) the kinetic and thermodynamic probabilities for formation of free radicals by the metal-catalyzed initiation reaction are considerably more favorable than the Bolland and Gee (16) proposal of diradicals by direct oxidation of a double bond. Once hydroperoxides are formed, even in trace amounts, they can play a profound role in the autocatalysis. Monomolecular decomposition yields two free radicals:
ROOH RO + HO

A bimolecular reaction, perhaps proceeding through intermediate hydrogen bonding, is more probable:
ROO + HOOR H [ROO H HOOR] HOH + RO + RO 2

Either the monomolecular or the dimolecular decomposition serves to feed new radicals into the reaction to initiate the chain reaction of autoxidation. These radicals may further react through different paths. They may follow a radical chain mechanism or other well-known radical reactions, such as coupling or disproportionation.

318

PAINTS, VARNISHES, AND RELATED PRODUCTS

The reactions may lead to the formation of dimers or polymers or may achieve cross-linking, resulting in an insoluble, infusible lm (i.e., drying). Apparently, the dominant reaction path depends on the temperature. At room temperature, mostly C C bonds are produced, whereas C bonds are predominantly O C formed under baking conditions. The free radicals may also undergo chain cleavage reactions. Low molecular weight by-products, such as water, carbon dioxide, aldehydes, ketones, and alcohols may be formed, which cause the odor and taste of the oils. The strong odor of rancid soybean oil was shown to be caused by 2-pentylfuran found in oxidized oil in storage (23). Chemically, the air-drying of a nonconjugated oil such as linseed is characterized by the adsorption of 1216% by weight of oxygen. The reactivity of drying oils is based on the mesomeric stabilization of the radical intermediate: the unpaired electron is delocalized over several carbon atoms, and less energy is required to eliminate the proton as illustrated below (24).
Activation Energy (kJ/mol) 415 335 289 168 Relative Rate of Oxidation 0 1 120 330

Triglycerides Stearate Oleate Linoleate Linolenate


a

Mesomers a 2 5 11

Saturated molecules.

3.2. Drying of a Conjugated System Tung oil, whose dominant feature is the conjugated cis, trans, trans-9,11,13-octadecatrienoic acid, a-eleostearic acid, dries to a coherent lm with absorption of only 5% by weight of oxygen. Privett (25) suggested oxidation through 1,2- or 1,4-addition to the diene system to yield noncyclic peroxides. Faulkner (26) identied 1,6-peroxide in addition and suggested that the autoxidation does not proceed via hydroperoxide groups but rather via cyclic peroxides. It has also been found that the triene content decreased and the diene content increased in proportion to the absorption of oxygen (27, 28). The main reaction is believed to consist of a direct attack by oxygen on the C C double bonds to form cyclic peroxides and dienes. The peroxides then react with allylic methylene groups or thermally dissociate to give radicals, initiating a radical chain reaction mechanism, forming polymers via C or C C bonds (29). C O

4. OLEORESINOUS VARNISHES As noted, coating systems were advanced from oil-only vehicles to oleoresinous varnishes for improved performances. These are basically oils that have been

ALKYD RESINS

319

hardened or modied by treatment with one or more suitable resins, natural or synthetic. The oils and resins are combined, usually by heating together at temperatures of 250 C or above, until a homogeneous mixture is formed. In most cases, it is simply a case of dissolving the resin or resins in the oil. In some cases, chemical reaction may have taken place between the resin and the oil, such as that between the methylol groups of a heat-reactive phenolic resin and the double bonds of a drying oil in forming a chroman ring structure as shown below (30):
R OH HOH2C CH2OH + R R CH CH R R Chroman ring structure O HOH2C CH CH C H R

The major improvements obtained by incorporating resins into drying oils are faster drying, greater lm hardness, higher gloss, better water and chemical resistance, and greater durability. The degree of property change depends on the type and the amount of the resin incorporated in the oil. Varnish makers express the oil to resin ratio in terms of oil length, which is dened as the number of gallons of oil used per 100 lb of the resins in the varnish. Varnishes are categorized according to their oil length as short- (515 gal), medium- (1630 gal), and long-oil (30 gal) varnishes. These demarcations are somewhat arbitrary and not universally agreed. From oleoresinous varnishes, the coatings industry progressed into alkyd resins. While one might say that this was only an evolutionary change, it nevertheless did open a new horizon for coating technologists and has been responsible for the longevity of oil-based coating materials. It behooves us to take a more comprehensive look at the various aspects of alkyd resins.

5. ALKYD RESINS Alkyd resins have been the workhorse for the coatings industry over the last half century. The term alkyd was coined to dene the reaction product of polyhydric alcohols and polybasic acids, in other words, polyesters. However, its denition has been narrowed to include only those polyesters containing monobasic acids, usually long-chain fatty acids. Thus thermoplastic polyesters typied by polyethylene terephthalate (PET) used in synthetic bers, lms, and plastics and unsaturated polyesters typied by the condensation product of glycols and unsaturated dibasic acids (which are widely used in conjunction with vinylic monomers in making sheet molding compounds or other thermosetting molded plastics) are not considered as part of the alkyd family and are beyond the scope of the present discussion.

320

PAINTS, VARNISHES, AND RELATED PRODUCTS

The rst appearance of the term alkyd resin in the subject index of Chemical Abstracts was in 1929, under resins. It was not until 1936 that alkyd resins was listed in its alphabetical place, but still appeared as see resinous products. The proliferation of literature on alkyd resins peaked in the 1940s through the 1960s. Research activities on alkyds in the United States, as indicated by the number of publications, has apparently tapered off in the last two decades. Readers who are alkyd history buffs can nd more detailed historical reviews (3134). In spite of the challenges from many new coating resins developed over the decades, alkyd resins, as a family, have maintained a prominent position even until today. There are two major reasons for such sustained popularity. First, alkyds are extremely versatile. An alkyd technologist can choose from a large variety of reaction ingredients and at widely different ratios to tailor the structure and properties of the resin or to obtain similar resin properties from different ingredients, as their availability or cost may sometimes so dictate. For almost any given coating application, from baking enamels for appliances to at house paints to clear wood nishes, one can design an alkyd resin to meet the property requirements. The second reason is that alkyd resins can be made at relatively low cost. Most of the raw materials are fairly low cost commodity items, and major capital investment and high processing cost are not needed to produce the resins. 5.1. Basic Reactions and Resin Structure The main reactions involved in alkyd resin synthesis are polycondensation by esterication and ester interchange. If one uses the following symbols to represent the basic components of an alkyd resin: O R O , a polyol molecule or radical;
O

X X, a polybasic acid molecule or radical; and X a mono-basic acid moleA F, cule or radical, a schematic representation of the resin molecule can be given (Figure 1). As Figure 1 implies, there is usually some amount of residual acidity along with free hydroxyl groups left in the resin molecules. The structure-property relationship and the principles commonly followed to design the resin structure will be discussed below. 5.2. Classication of Alkyd Resins Alkyd resins are usually referred to by a shorthand description based on a certain way of classication or a combined classication, from which the general properties

Figure 1. Schematic representation of an alkyd resin molecule.

ALKYD RESINS

321

of the resin become immediately apparent. The commonly used bases for classication are as follows. Drying versus Nondrying, and the Specic Source of Fatty Acids. Alkyd resins can be broadly classied into the drying type and the nondrying type, depending on the ability of their lm to dry by air oxidation. This drying ability is derived from the polyunsaturated fatty acids in the resin composition. If drying oils, such as linseed oil, are the sources of the fatty acids for the alkyd, the resin would belong to the drying type and is usually used as the lm former of coatings or inks. On the other hand, if the fatty acids come from nondrying oils, such as coconut oil, the resin would be a nondrying alkyd. They are used either as plasticizers for other lm-formers, such as in nitrocellulose lacquers, or are cross-linked through their hydroxyl functional groups to become part of the lm former. More frequently, an alkyd resin is classied by the source of the fatty acids, e.g., a linseed alkyd, a tung oil-modied soy alkyd, and a coconut alkyd. Classied by Oil Length or Fatty Acid Content. Probably inherited from oleoresinous varnish practice but with a different way of expression, alkyd resins are also classied by their oil length. For an alkyd resin, the oil length is dened as the weight percent of oil or triglyceride equivalent, or alternatively, as the weight percent of fatty acids in the nished resin, for example, the resin represented in Figure 1. The structure indicates that the molar ratio of these three ingredients is 4 : 4 : 3. Assume that the polyol is glycerol, the polybasic acid is phthalic anhydride, and the fatty acids came from soybean oil with an average molecular weight of 280. The formula weight of the resin would be 1674 and the triglyceride equivalent of the fatty acids would be 878, thus the oil length would be 52.4%. Alternatively, the above resin would be described as one having 50.2% fatty acids. Since the overwhelming majority of alkyd resins are based on phthalic anhydride, it is also customary to describe an alkyd in terms of its phthalic anhydride content in percents based on the nished resin. By this approach, alkyd resins are classied into four classes:
Percent Fatty Acids >68 5368 4352 <42 Percent Phthalic Anhydride <20 2030 3035 >35

Resin Class Very long oil Long oil Medium oil Short oil

Percent Oil >70 5670 4655 <45

It should be noted that these demarcations are arbitrary and may vary from author to author. Furthermore, the boundaries are usually not clear-cut. More frequently, alkyd resins are described by a combined classication in terms of their oil length, the type of fatty acids, and any unusual ingredients. Such descriptions as an isophthalic, very long tall oil alkyd or a medium oil dehydrated castor-PE (the PE refers to pentaerythritol, not polyethylene) alkyd or a short oil lauric-benzoic alkyd would immediately project the general properties of the resin.

322

PAINTS, VARNISHES, AND RELATED PRODUCTS

5.3. Oil LengthResin Property Relationship Obviously, the oil length of an alkyd resin has profound effects on the properties of the resin. A few of these effects are discussed below. Effect on Solubility. At long oil lengths, the aliphatic hydrocarbon chains of the fatty acids constitute the major portion of the mass of the resin molecules; therefore, the resin would be soluble in nonpolar aliphatic solvents. Conversely, as the oil length decreases and the phthalic content increases, the aromaticity of the resin molecules increases, and the aromaticity and/or the polarity of the solvent will also need to be increased to dissolve the resin effectively. Effect on Drying Characteristics. Alkyd resin molecules have a comblike structure, with a thermoplastic polyester backbone and dangling fatty acid side chains. Each of these two fractions contributes to the drying, or lm-forming, characteristics of the resin. The backbone fraction dries by solvent release, similar to a lacquer material, whereas the side chain fraction dries in a manner similar to the oil from which the fatty acids came. Therefore, short oil alkyds develop a surface dryness relatively quickly due to a faster solvent release, which is often further facilitated by the fact that the solvents used have high volatility. However, their through-dry in air is usually slower, because the fatty acid side chains are fewer in numbers and more scattered in space to cross-link with each other through the action of oxygen, and the dry surface would impede the transportation of air oxygen to reach down into the lm. On the other hand, long oil alkyds are relatively slow in reaching the set-to-touch stage of surface drying, but the greater abundance of fatty acid side chains and the relative openness of the lm surface would facilitate the lm to reach through-dry. Other trends of changing properties (Table 1) would become obvious, considering how the structure of resin molecules would change with oil length. Theoretically, one could design and make alkyd resins at almost any oil length. However, for any given set of starting ingredients, as the oil length goes up, it will reach a point

TABLE 1 Trends of Property Changes with Oil Length of Alkyd Resinsa. Oil Length Requirement of aromatic, polar solvents Compatibility with other lm formers Viscosity Ease of brushing Air dry time, set-to-touch, Through-dry Film hardness Gloss Gloss retention Color retention Exterior durability
a

Long

Medium

Short

! ! ! ! ! ! ! ! !

Primarily referring to drying-type alkyds.

ALKYD RESINS

323

where the maximum extent of fatty acid modication of the polyester molecules has been achieved, and any additional amounts of fatty acids or oil remain as separate entities, blended with the polyester molecules. Refer again to the resin structure in Figure 1. If the molar ratio is 1 : 1 : 1 among glycerol, phthalic anhydride, and soy acids and the reaction was carried to completion, the resin would have an oil length of 60.5%, or 57.9% of fatty acids. There is no more room in the resin structure to accommodate any additional amount of fatty acid. Therefore, with those three ingredients, if the oil length exceeds 60.5%, the excess amount of oil would only be retained in the resin as a blend. Obviously, the very long oil types of alkyd resins would almost certainly be resinoil blends. The maximum oil length of an alkyd resin (before it becomes a resinoil blend) depends on the molecular weight of the ingredients as well as the functionality of the polyol. If the C18 soy fatty acids in the above example is replaced with C12 lauric acid, the transition would be reached at 52.6% oil. On the other hand, if a tetra-hydroxyl polyol, such as pentaerythritol, replaces glycerol, the stoichiometry would allow 2 moles of fatty acids, for every 1 mole of phthalic anhydride and pentaerythritol. Thus theoretically, the maximum amount of soy fatty acids that may be chemically combined in the resin structure would be 70.9%, equivalent to a 74.1% oil length. 5.4. Major Ingredients Each of the three principal components of alkyd resinsthe polybasic acids, the polyols, and the monobasic acidshas a large variety to be chosen from. The selection of each one of these ingredients will affect the properties of the resin. As will be shown later, the choice of ingredients may even affect the choice of manufacturing processes. To both the resin manufacturers and the users, the selection of the proper ingredients is a major decision. Polybasic Acids and Anhydrides. The major types of polybasic acids used in alkyd preparation are as follows.
Molecular Weight 148 166 98 116 146 160 174 371 192 Equivalent Weight 74 83 49 58 73 80 87 185.5 64

Type Phthalic anhydride Isophthalic acid Maleic anhydride Fumaric acid Adipic acid Azelaic acid Sebacic acid Chlorendic anhydride Trimellitic anhydride

Phthalic anhydride is by far the most important dibasic acid used in alkyd preparation, because of its low cost and the excellent overall properties it imparts to the

324

PAINTS, VARNISHES, AND RELATED PRODUCTS

resin. Its anhydride structure allows a fast esterication to form half-esters at relatively low reaction temperatures without liberating water, thereby avoiding the danger of excessive foaming in the reactor. However, since the two carboxyl groups of phthalic anhydride are in the ortho position to each other on the benzene ring, cyclic structure may and does occur in the resin molecules. Consequently, the development of chain length of the polymer would be restricted, and the average molecular weight would tend to be low. Phthalic anhydride has a tendency to sublime. (Heat of sublimation: 143 g-cal/g; heat of vaporization: 87.2 g-cal/g.) Therefore, care must be taken to prevent its loss. Isophthalic acid is the meta isomer of phthalic acid. Since the two carboxyl groups are adequately separated, the chances of forming a cyclic structure in the resin molecules are greatly diminished. Therefore, isophthalic alkyds usually attain higher molecular weight and show much higher viscosity than their phthalic counterparts at the same oil length. This is the major motivation for resin manufacturers to use isophthalic acid for the preparation of long oil alkyds. Another major advantage of isophthalic acid is that the resultant alkyd resins show much higher thermal stability than the phthalic type (35). In spite of these advantages, isophthalic acid has not gained the same popularity as phthalic anhydride in alkyds, because the resin making process is much more complicated and difcult than that with phthalic anhydride. Its melting point (350 C) is much higher than that of phthalic anhydride (131 C), and it has a low solubility in the initial alkyd reactants, which causes the reactants to stay as a two-phase solidliquid system and does not become clear until the reaction is near complete (36, 37). The diacid does not readily form half-esters at relatively low reaction temperature as would the anhydride, and twice as much water will be formed and needs to be removed from esterication. Usually, additional care and equipment are needed for the higher processing temperature required for isophthalic acid. The para isomer terephthalic acid may also be used for making alkyds. The resultant resins showed even better thermal stability than isophthalic alkyds (35). However, it has all the disadvantages of isophthalic acid and is more expensive. It is rarely used in making alkyd resins. Maleic anhydride is sometimes used for partial substitution of phthalic anhydride in making alkyds. It imparts vinylic unsaturation functionalities in the backbone chain of the resin molecules, which allows the resin to be grafted with styrene, acrylic esters, or other vinyl monomers. The presence of a small amount of maleic anhydride, up to 10% on molar base of the total dibasic acids, in the resin formulation would accelerate the viscosity increase during the resin manufacturing process. The resins usually dry more rapidly and give harder lms with improved color, adhesion, water resistance, alkali resistance, and exterior durability. However, the resin cooking process needs to be monitored and controlled with greater care, particularly when it is near the desired end point, to prevent gelation. Fumaric acid, the trans isomer of maleic acid, may be used in an equivalent manner. Maleic and fumaric acid can also be, and are often, incorporated in alkyd resins in the form of the Diels-Alder adduct of rosin. The adducts are tribasic acids. They provide one of the means to impart pendant carboxyl groups in the resin molecules,

ALKYD RESINS

325

which can then be saponied to give ionic and, in turn, water-soluble characteristics to the resin. Alkyds containing maleicrosin adducts often have poorer color retention, toughness, gloss retention, and exterior durability. Aliphatic dibasic acids, such as succinic, adipic, azelaic, and sebacic acids have also been used to make alkyd resins. Their linear and exible chain structure lends higher exibility and lower viscosity to the resin than the rigid aromatic rings of phthalic acids. Chlorendic anhydride is the Diels-Alder adduct of maleic anhydride and hexachlorocyclopentadiene. It is also known as hexachloro-endo-tetrahydrophthalic (HET) anhydride. The major interest of the alkyd industry in this material is that the resultant resins contribute to the ame retardancy of the coatings. It has been reported to give a greater reaction rate than phthalic anhydride, such that at 204 210 C (400410 F) the reaction rate approximates that of phthalic anhydride at a temperature of 238 C (460 F) (38). However, the resins are prone to develop darker color, particularly at high processing temperature. Tetrachlorophthalic anhydride, made by conventional chlorination of phthalic anhydride, would also impart ame retardancy to its alkyds. However, it is appreciably less soluble in the usual processing solvents than is phthalic anhydride and is reported to be of appreciably lower chemical reactivity (39). Trimellitic anhydride (TMA), 1,2,4-benzenetricarboxylic acid anhydride, has gained greater prominence in recent years due to the greater interest in watersoluble alkyds. A partial substitution of the phthalic anhydride with TMA gives a measured quantity of pendent carboxyl groups for water solubilization with ammonia or other suitable base. The anhydride hydrolyzes to the acid form simply by allowing it to stand in open containers. Premature cross-linking of alkyd resins formulated with a high content of TMA would occur at high acid numbers when large amounts of trimellitic acid are present (40). Polyhydric Alcohols. The major types of polyol used in alkyd synthesis are as follows.
Molecular Weight 136 92 134 120 62 104 Equivalent Weight 34 31a 44.7 40 31 52

Type Pentaerythritol Glycerol Trimethylolpropane Trimethylolethane Ethylene glycol Neopentyl glycol

a Since glycerol is usually supplied at 99% purity (1% moisture), its equivalent weight is commonly assumed to be 31 in recipe calculations.

Pentaerythritol (PE) is one of the most important polyols used in alkyd resins. Its molecular structure, four methylol groups (CH2OH) surrounding a center carbon atom, is the basis for its many interesting attributes. The four equal and highly

326

PAINTS, VARNISHES, AND RELATED PRODUCTS

reactive primary hydroxyl groups make it versatile for designing resin structures, and the neopentyl core structure lends stability against heat, light, and moisture. As a result, alkyds based on PE usually are superior to their counterparts based on glycerol in viscosity, drying properties, lm hardness, gloss retention, color and color stability, humidity resistance, thermal stability, and exterior durability. On the other hand, its high functionality demands that the resin composition be more carefully designed and the synthesizing process be more carefully monitored and controlled to reduce or eliminate the tendency of gelation. Dipentaerythritol and tripentaerythritol are linear dimer and trimer of PE. They are hexa- and octafunctional polyols, respectively. Technical grades of PE usually contain small or trace quantities of di- and tri-PE that were not completely removed in the manufacturing process. The high functionality of these materials makes them impractical to be considered as the sole or major polyol of an alkyd resin. Among the triols, glycerol is undoubtedly the most important one in alkyd technology. Natural fats and oils are triglycerides. Therefore, whenever oils are used directly as the source of fatty acids in an alkyd resin, glycerol will automatically be a part of the polyols of the resin. Besides the difference in functionality, the major difference between glycerol and PE is that one of the hydroxyl groups in glycerol is secondary, which has lower reactivity than primary hydroxyl groups. This often manifests itself as if glycerol had a de facto functionality of less than 3. Consequently, a larger excess of glycerol would be required in the resin formulas, which would result in poorer resin properties as a coating material. At high temperatures, the proton on the secondary carbon in glycerol may undergo a dehydration reaction with one of the primary hydroxyl groups on the adjacent carbon atom to give water and acrolein, whereas such reaction is not possible with PE. Glycerol alkyds are more prone to thermally decompose to give color bodies, resulting in darkening of the resin. Trimethylolethane and trimethylolpropane are synthetic triols. Like PE, they have the neopentyl structure and equivalent primary hydroxyl functional groups. Therefore, they also yield alkyds with better resistance to heat, light, moisture, and alkali than glycerol. They have one less hydroxyl group than PE, and the equivalent weights of these polyols are higher than that of PE. Trimethylolethane has been reported to give alkyds that are faster drying and higher in lm hardness than trimethylolpropane (41), whereas trimethylolpropane was claimed to give alkyds with better water and alkaline resistance, color and color retention, and impact resistance than trimethylolethane (42). Diols such as ethylene glycol, propylene glycol, and neopentyl glycol are sometimes used as part of the polyols in alkyd formulations primarily for the purpose of regulating the functionality of the reaction system. Their relatively low boiling points, (197, 188, and 207 C), respectively, for the above three glycols) require that special precautionary measures be taken during the resin manufacturing process. Analogous to the use of linear a,o-dibasic acids (such as adipic and sebacic), polyols with long, exible chains between hydroxyl groups (such as 1,4-butanediol, 1,6-hexanediol, and diethylene glycol) may also be used to impart greater exibility in the resin.

ALKYD RESINS

327

It should be pointed out that under high temperatures, such as those used for alkyd resin synthesis, and in the presence of high acidity, etherication between the hydroxyl groups of two polyol molecules may condense them into a new polyol with a functionality of n n0 2, where n and n0 are the numbers of hydroxyl groups of the two original molecules. The introduction of such high functionality polyols plus the net reduction of total available hydroxyl groups can lead to an increased danger of gelation during the poly-condensation process. Monobasic Acids. The overwhelming majority of monobasic acids used in alkyd resins are long-chain fatty acids of natural occurrence. They may be used in the form of oil or free fatty acids. Free fatty acids are usually available and classied by their origin, viz., soy fatty acids, linseed fatty acids, coconut fatty acids, etc. The fatty acid composition of various types of fats and oils that are commonly used in alkyd resins are given in Table 2. The drying property of alkyd resins reects directly that of the oil or fatty acids in the resin structure, discussed earlier. It should be pointed out that alkyds based on conjugated unsaturated fatty acids, such as those from tung and oiticica oils, dry so fast that if not properly moderated, the surface layer will dry long before the underlayer, resulting in a wrinkled surface due to the stresses created in the dried surface layer. Therefore, in alkyd resins, tung oil and oiticica oil are primarily used to furnish a minor portion of the fatty acids to improve drying properties. Even so, greater care must be exercised during the manufacturing process to avoid gelation, which is caused by the dimerization of the fatty acid chains through a Diels-Alder addition between the conjugated diene structure on one molecule and a double bond on another molecule. It should be noted that nonconjugated diene groups, such as those in linoleic and linolenic acids, may undergo isomerization to become conjugated. Furthermore, ene-reaction could also occur between two unsaturated fatty acid chains, which leads to gelation. Rheineck and co-workers (44) have found that linolenic acid is responsible for the high yellowing tendency of alkyds based on linseed oil fatty acids. Therefore, alkyds intended for making white or light color enamels should avoid high linolenic content fatty acids by choosing soy oil, safower oil, or dehydrated castor oil (DCO). Alkyds made with nondrying oils or their fatty acids have excellent color and gloss stability. They are frequently the choice for white industrial baking enamels and lacquers. Since the mid-1950s, tall oil fatty acids (TOFA) have become available in good quality and large quantities. Rened grades of TOFA have degrees of unsaturation rivaling that of soy acids. Since it is a year-round by-product from the paper industry, its supply and price are more stable than agricultural products like soy fatty acids. It is used extensively in medium- to long-oil alkyds, virtually as equivalent to soy fatty acids. Although the minor quantities of rosin acids in TOFA may impart some yellowing tendency, its lack of linolenic acid may be more than enough to give as good or even better color retention than soy fatty acids. The typical properties of rened grades of commercial TOFA are given in Table 3. A number of monobasic acids that are not derived from fats and oils have been used in alkyd resins. However, except in the rare cases of making the so-called

TABLE 2. Fatty Acid Compositions of Fats and Oils Commonly Used in Alkyd Resins. Fatty Acid Composition (Percent by Weight) Oil Castor Castor, dehydrated Coconut Iodine Value 85.8 125135 8.7 Saponication Value 195 191 257 Saturateda 2.4 2.4 76.6 Palmitoleic Oleic 7.4 8.0 5.7 Linoleic 3.1 86 2.6 Linolenic 3.0 Other Ricinoleic, 87.0; dihydroxystearic, 0.6 About 33% of the linoleic is 9,11-conjugated. Caprylic, 7.9; capric, 7.2; of the saturated, lauric, 48.0 and myristic, 17.5 Tetradecenoic, 0.1 Lignoceric, 0.2 Highly unsaturated C20H2(20 x)O2, 19; and C22H2(22 x)O2, 12.b Licanic, 82.5 Arachidic, 2.4; behenic 3.1: lignoceric, 1.1 Behenic, 0.7: lignoceric, 0.8; eicosenole, 4.8; erucle, 47.8; docosendienole, 1.5 Saturated C20-C24, 2.4 Lignoceric, 0.4 Eleostearic, 87

Cottonseed Linseed Menhaden

105.0 180 148185

196 191 191

27.2 9.3 24.0

2.0 15.0

22.9 19.0 30.0

47.8 24.1

47.4

Oiticica Peanut Rapeseed

93.3 102.3

192 190 175

11.3 13.8 6.1

1.7 1.5

6.2 54.3 12.3

26.0 15.8

8.7

Safower Soybean Sunower Tung


a b

136.2 132.6 130.8

191 193 188 192

6.0 13.4 7.1 5.0

1.0

32.8 23.5 34.0 5.0

61.1 51.2 57.5 3.0

1 8.5

Aliphatic monocarboxylic acids, C12 to C20, principally palmitic and stearic. x 4 10. Owing to incomplete halogen absorption, iodine values for conjugated acid oils by the usual methods (Wijs, Hanus, etc.) are both low and variable. The true value of fresh tung oil, as determined by special method, is 248252: that of oiticica oil is 205220 (43).

ALKYD RESINS

329

TABLE 3. Typical Properties of Rened Tall Oil Fatty Acidsa. Grade Designation Pamak 1 193 125 96.8 2.0 51.0 41.0 6.0 1.4 1.8 3 5 Pamak 2 192 128 95.9 1.8 2.3 3 5 Pamak Pamak Pamolyn Pamolyn 4A 4 200b 300a 191 130 94.1 3.5 2.4 4 5 188 131 91.5 4.0 51.0 39.0 6.0 4.0 4.5 6 6 195 162 97 <1 22.0 68.0 10.0 1.5 1.5 3 28 196 156 97 <1 21.0 39.0 40.0 1.5 1.5 3 28

Characteristic Acid number Iodine number Total fatty acids, % Saturated acids, % of free fatty acids Oleic Linoleic, nonconjugated Linoleic, conjugated Linolenic Rosin acids, % Unsaponiables, % Color, Gardner Titer,  C
a b

Data from Hercules, Inc. (Wilmington, Del.). Enriched polyunsaturated fatty acids from highly rened TOFA. c Same as Pamolyn 200, with further treatment to isomerize the nonconjugated linoleic acid.

oil-free alkyds for special purposes, they are used in conjunction with fatty acids to modify resin properties. Rosin acids, primarily in the form of abietic acid, are the most common type of such acids. They may be used in neat form or be brought in as a part of TOFA. Presumably, the fused ring structure of rosin contributes to the lm hardness, initial gloss, and water resistance of the alkyd. However, color and color retention, and exterior durability will be adversely affected if the rosin content goes much above 56%. The drying rate of alkyds usually appears to be improved with rosin modication. However, since rosin does not participate in the oxidative drying mechanism that applies to polyunsaturated fatty acids, the true drying rate of the alkyd resin would be reduced due to a reduction of the fatty acid unsaturation. Synthetic saturated carboxylic acids (such as pelargonic acid, 2-ethylhexanoic acid, and isoctanoic acid) and aromatic monobasic acids (such as benzoic acid and p-alkyl-benzoic acids) can improve color retention, gloss retention, and exterior durability even better than those based on castor or coconut fatty acids. The aromatic acids, similar to rosin, also give higher lm hardness and faster apparent drying rate. 5.5. The Concept of Functionality and Gelation The concept of functionality and its relationship to polymer formation was rst advanced by Carothers (45) in 1929. Flory (46) greatly expanded the theoretical consideration and mathematical treatment of polycondensation systems. Thus if a dibasic acid and a diol are reacted to form a polyester, assuming there is no possibility of other side reactions to complicate the issue, only linear polymer molecules

330

PAINTS, VARNISHES, AND RELATED PRODUCTS

will be formed. When the reactants are present in stoichiometric amounts, the average degree of polymerization n follows the equation: x n 1=1 p x 1

where p is the extent of reaction, in fractions. Thus when the reaction is driven to completion, theoretically, the molecular weight would approach innity and the whole mass would form one giant polymer molecule. Although the material should theoretically be still soluble and fusible, it is considered and dened as a gel, and this would be the only time that difunctional monomers could be polymerized to gelation. The functionality of the system f is the sum of all of the functional groups, i.e., equivalents, divided by the total number of moles of the reactants present in the system. Thus in the above equimolar reaction system: f 1 2 1 2=1 1 2 However, when there are reactants with three or more functionalities participating in the polymerization, branching and the formation of intermolecular linkages (i.e., cross-linking of the polymer chains) become denite possibilities. If extensive cross-linking occurs in a polymer system to form network structures, the mobility of the polymer chains is greatly restricted. Then the system would lose its uidity and transform from a moderately viscous liquid to a gelled material with innite viscosity. The experimental results of several such reaction systems reported by different investigators are collected in Table 4. The data in Table 4 show that when the reactants are present in stoichiometric proportions, gelation occurs before the completion of esterication, and the extent of reaction p reached at the gel point depends on the functionality of the system. Carothers (47) showed that at the gel point, p 2=f . Thus, to avoid premature gelation, the polymerization system should have an average functionality of no more than 2. This can be accomplished by adding low functionality reactants and/or adding an excess amount of one of the reactants, usually the one with high functionality constituents. The latter has the net effect of reducing the functionality of the reactant. For example, if a 20% excess of glycerol over the stoichiometric

TABLE 4. Gel Points of Polyesterication Reaction Systems with Stoichiometric Reactants (46). Polybasic Acid (COOH) Adipic (0.707) tricarballylic (0.293) Dibasic (1.0) Adipic (1.0)
a b

Polyol (OH) di-EG (1.0)a glycerol (1.0) PE (1.0)b

f 2.103 2.400 2.667

Percent Esterication at Gel Point 0.911 0.765 0.578

Diethylene glycol. Pentaerythritol.

ALKYD RESINS

331

amount required to esterify all of the carboxyl groups present in the formula is added, the glycerol would have an effective functionality of 3/1.2, or 2.5. Frequently, both of these measures are taken to safeguard against premature gelation. Patton (48) showed that for alkyd resins, the extent of the reaction at gel point was pc 2=f 2mo =eo where eo is the total effective equivalents of all of the reactants present at the beginning of the reaction (i.e., the excess reactants are discounted in the manner discussed above), mo is the total number of moles of all reactants at the beginning, and f is the effective average functionality of the formulation. 5.6. Microgel Formation and Molecular Weight Distribution Bobalek and co-workers (49) observed that the behavior of alkyd resin reaction often deviates from that predicted by the theory of Flory. They proposed a mechanism of microgel formation by some of the alkyd molecules at a relatively early stage of the reaction. The microgel particles would be dispersed and stabilized by smaller molecules in the remaining reaction mixture. As polyesterication proceeds, more microgel particles would be formed, until nally a point is reached at which they could no longer be kept separated. The microgel particles would then coalesce or occulate, phase inversion would occur, and the entire reaction mass would be gelled. They showed that the drying capability of an alkyd resin comes primarily from the microgel fraction and, when the highest molecular weight fraction representing about 20% of the total was removed through fractionation, the residual linoleic alkyd lost all ability to air dry to a hard lm. Solomon and co-workers (50, 51) further elaborated on the microgel theory by proposing the formation of micelles as precursors of microgels. They proposed that when some of the molecules have grown to reach certain fatty acid: polyester ratios, surface activity develops to form micelles. The polyesterication reactivity at the surface of the micelles would be preferentially greater, which would lead to the eventual formation of microgels. From electronmicroscopy evidence, they observed that the size of microgels increased with reaction time, and particle diameters as large as 2m have been reported. Functional groups such as OH groups in the microgel particles are believed to be buried in the structure and not available for reactions. Whereas in polyesterication reactions without fatty acids, at all stages of the reaction up until the physical gelation of the reaction mixture, the hydroxyl values corresponded with the calculated values. Furthermore, the oil-free systems showed no sign of microgel formation under electron microscope, and the reaction mixture would undergo a sudden change from a soluble polymer to a gelled mass. The reaction temperature for alkyd preparation in both of the above references was kept at no more than 200 C, well short of what would normally be required for the bodying of unsaturated oils in the absence of an oxidizing reagent. This indicates that polymerization between unsaturated fatty acids of the resin molecules is not necessary for the formation of microgels.

332

PAINTS, VARNISHES, AND RELATED PRODUCTS

Kumanotani and co-workers (5255) further conrmed the formation of microgels by characterizing fractions of alkyd resins from preparative GPC columns. Their results showed that the presence of microgel can be detected even in low molecular weight fractions. Colloidal gel particles up to 10 mm in diameter were observed in the high molecular weight (>105) fractions, with or without unsaturated fatty acids in the alkyd formulation. However, the unsaturated fatty acids made a signicant contribution to the formation of the colloidal particles. Higher reaction temperature led to higher molecular weight and broader molecular weight distribution. Acid value and hydroxyl value each went through a minimum in the middle fractions (molecular weight about 103104), whereas the polyester: monoacid ratio increased with the molecular weight of the fraction. Cured lms from alkyds with greater amount of colloidal fractions gave better thermomechanical properties. Finally, the high molecular weight colloidal fractions were preferentially adsorbed by pigment particles and would thus stabilize the pigment dispersion in the coating formulation. 5.7. Basic Principles for the Designing of Alkyd Resins The process of alkyd resin designing should begin with the following question: what would be the intended application(s) of the resin? The application would dictate property requirements, such as solubility, viscosity, drying characteristics, compatibility, lm hardness, lm exibility, acid value, water resistance, chemical resistance, environmental endurance, etc. With these targets in mind, a selection on oil length, and a preliminary list of alternative choices of ingredients can then be made. For commercial production, the raw material list is screened based on considerations in material cost, availability, yield, impact on processing cost, and potential hazard to health, safety, and the environment. The list may be further narrowed by limitations imposed by the production equipment or other considerations. Once the oil length and ingredients are chosen, the rst draft of a detail formulation for the resin can then be made. It would be highly desirable that one could rely on a simple equation or formula to obtain the optimum formulation of the alkyd resin with the chosen ingredients, and several approaches have been proposed for such purpose (5659). However, the complexity of the alkyd reaction system has rendered these equations to be of no more value than providing a rst approximation of a starting formula. The causes of the complexities include the formation of intramolecular cyclic structures, which would reduce the chance of gelation; the etherication of polyols, which would increase the chance of gelation by forming higher functionality materials and reducing the number of hydroxyl groups available for esterication; the cross-linking between unsaturation groups, especially the conjugated double bonds, which would increase the chance of gelation; and the phenomenon of microgel formation. Except when nondrying alkyds are used strictly as plasticizers for other thermoplastic polymers, alkyd resins do not remain as a thermoplastic material in their ultimate application. The lm integrity is largely derived after the resin molecules have been cross-linked, either through the unsaturation functionalities on their fatty

ALKYD RESINS

333

acid side chains or through the reactions of their residual hydroxyl or carboxyl functionalities with such cross-linking agents as amino resins or polyisocyanate materials. In a sense, alkyds are usually made and applied as B-stage resins. Therefore, it is not necessary to build the molecules of alkyd resins to huge molecular weights, as one would for thermoplastic polymers. In fact, too high a molecular weight would lead to poor solubility and high solution viscosity and would be undesirable for practical applications. Most of the published data show that the average molecular weight of alkyds is less than 10,000. Nevertheless, within the practical limits, it is still preferred to have a linear backbone structure and high molecular weight to give the best lm-forming and lm properties. Alkyd formulations with an equimolar ratio of dibasic acids: polyols tend to have the best chance of achieving a linear molecular structure and high molecular weight. Thus a simple molecular approach is favored by some of the alkyd chemists for deriving the starting formulation. The basic premise of this approach is that when the total number of moles of polyols is equal to or slightly larger than that of the dibasic acids and the hydroxyl groups are present in an empirically prescribed excess amount, the probability for gelation to occur would be small. Table 5 lists the empirical requirements of excess hydroxyl groups based on the oil (fatty acid) length of the alkyd. The values were developed based on experimental experience (57, 58). With the new understanding that some of the hydroxyl groups would be buried in microgel structures (4955), such requirements may be better rationalized. The procedure of this method for formulating alkyd resins will be illustrated with examples. The rst example demonstrates the formulations of a 50% soy oil alkyd for baking enamels. The preliminary selection of ingredients would be alkaline rened soy oil, phthalic anhydride (PA), and pentaerythritol. The basis for calculation is 1 mole of PA. From Table 5, the excess OH recommended at 50% oil length is 25%. Therefore, the quantity of PE required, in equivalents, would be EPE 1 2 1 0:25 2:5 eq: 0:625 moles

TABLE 5. Excess Hydroxyl Content Required in Alkyd Formulations. Oil Length, Percent Fatty Acida 62 or more 5962 5759 5357 4853 3848 2938 Percent Excess OH Based on Diacid Equivalents 0 5 10 18 25 30 32 Percent Excess OH in Finished Resin 0 05 510 1015 1520 2025 2530

a Based on C18 fatty acids with average equivalent weight of 280. If the average equivalent weight of the monobasic acids is signicantly different, adjust accordingly.

334

PAINTS, VARNISHES, AND RELATED PRODUCTS

Since the total polyol is to be equimolar to PA, the glycerol from the soy oil will, therefore, be 1 0:625 0:375 moles, which gives 3 0:375 1:125 moles of soy fatty acids. The ingredients can be listed as follows:

Ingredient PA Soy oil Tech-PE Total Water Resin

M 1.0 0.375 0.625 2.000 1.0

COOH 2.000 1.125 3.125

OH 1.125 2.500 3.625

Weight (g) 148.0 330.0 88.5 566.6 (18.0) 548.5

The above formulation does not meet the test of 50% oil length. The oil content must reduced. A reduction in oil would cause a corresponding reduction in glycerol, consequently, free glycerol is added to make up the loss. Let MPE X, MGly Y, and Moil Z. Since the total polyols is to be equimolar to dibasic acids, X Y Z 1. The 25% excess OH requirement denes 4X 3Y 2 1:25 2:5, and the 50% oil length requirement gives the following: 880Z=148 141:6X 93Y 880Z 18 0:5 where 880, 148, 141.6, 93, and 18 are the molecular weights of the oil, PA, PE, glycerol, and water, respectively. When solve the simultaneous equations to nd X 0:221, Y 0:539, and Z 0:240. Thus the nal formulation is listed as follows:

Ingredient PA Soy oil Glycerol Tech-PE Total Water Resin

M 1.0 0.240 0.539 0.221 2.000 1.0

COOH 2.000 0.720 2.720

OH 0.720 1.617 0.884 3.221

Weight (g) 148.0 211.2 50.1 31.3 440.6 (18.0) 422.6

The above formulation meets all of the requirements of the resin design, i.e., equimolar PA and polyols, 25% excess OH, and 50% oil. The next example shows the formulation of a 50% TOFA alkyd for baking enamels. Assume that PA, PE, ethylene glycol (EG), and rened TOFA with 4% rosin acids are the chosen ingredients. From the given constraints, the

ALKYD RESINS

335

following simultaneous equations can be established. Let MEG X, MPE Y, and MTOFA Z. XY 1 2X 4Y 2 1 0:25 Z 295Z 148 141:8Y 62X 295Z 18 1 Z 0:5 Solve the equations, to nd X 0:362, Y 0:638, and Z 0:776. Therefore, the nal formulation can be listed as follows:

Ingredient PA Tech-PE EG TOFA-4 Total Water Resin

M 1.000 0.638 0.362 0.776 2.776 1.776

COOH 2.000 0.776 2.776

OH 2.552 0.724 3.276

Weight (g) 148.0 90.3 22.4 228.9 489.6 (32.0) 457.6

The percent excess OH 3:276 2:776=2 25%, and the oil length 228:9=457:6 50% TOFA. The nal formulations derived in the above examples are meant to be only the starting formulations. They should be ne-tuned based on small-scale laboratory experiments before being used in plant production. Since the molecular chain length or the degree of polymerization is a function of the extent of the reaction as shown in equation 1, the alkyd reaction is usually carried to a point short of completion, i.e., to a nite acid number to guard against premature gelation. It has been shown that the esterication of phthalic anhydride was slower and showed higher temperature dependence, i.e., higher activation energy, than that of fatty acids (6062). Therefore, one may assume that the residual acidity belongs to unreacted dibasic acids, which contributes to the limiting of chain growth. In real practice, an additional safety margin against premature gelation is provided by having a slight molar excess of polyols over dibasic acids in the alkyd formulation. If the molar ratio between the polyols and the dibasic acids is r, equation (1) may be rewritten as: n 1 1=r=2=r1 p 1 1=r x 2

which indicates that a fractional increment in r and/or a fractional reduction in p would give a substantial reduction in n . Generally, the value of r is chosen between x 1 and 1.05.

336

PAINTS, VARNISHES, AND RELATED PRODUCTS

5.8. Chemical Procedures for Alkyd Resin Synthesis Different chemical procedures may be used for the synthesis of alkyd resins. The choice is usually dictated by the choice of the starting ingredients. Alcoholysis Process. Cost and availability often dictate that oil, rather than free fatty acids, be used as raw material for alkyd synthesis. Since oil, in the form of triglycerides, is essentially inert and would not participate in the polyesterication reaction, heating the oil with the polyol and the dibasic acid would result in the formation of seedy polycondensates between the polyol and the polybasic acids leaving the oil unreacted. The two phases thus would be incompatible with each other. Therefore, the triglycerides must rst react with additional polyol to redistribute the fatty acids among all of the polyols, thereby liberating free hydroxyl groups from the oil for further reaction with the dibasic acids. The reaction is alcoholysis. It is usually catalyzed by basic compounds such as metal oxides, hydroxides, salts, or soaps such as naphthanates. In the past, litharge was the most popular choice as the catalyst. It was found that on a molar basis lead compounds were the most efcient among the 36 that were included in the study (58). In recent years, due to the concern of lead poisoning from the resultant coatings, lithium hydroxide, sodium hydroxide, or calcium oxide have been commonly used. The dosage of these catalysts usually ranges from 0.01 to 0.06% metal based on the weight of the oil for lead and 0.008 to 0.02% for lithium or calcium. The amount of catalyst added should be kept at the minimum required for completing the reaction in an acceptable batch time. They may cause poor color, poor water resistance, or haziness in the nal resin. Ideally, 2 moles of polyol would react with 1 mole of triglyceride to form 3 moles of monoester. In reality, the reaction would reach an equilibrium, whereby some amount of diesters and triesters and neat polyol, including glycerol and the added polyol, would coexist in the reaction mixture. The composition of the alcoholysis product at equilibrium from soy oil and glycerol (1 : 2 mole ratio), and soy oil and monopentaerythritol have been reported as follows (63):

Component Glycerol Monoester Diester Triester Free glycerol PE Monoester Diester Triester Tetraester Free PE

Mole Percent Oil-Glycerol Oil-Mono-PE 42 21 3 33 20.6 6.2 2.0 14.0 29.0 16.8 5.1 Negligible 6.3

ALKYD RESINS

337

Diols, such as ethylene glycol, are usually not added during the alcoholysis step. This is because their monoesters have only one remaining hydroxyl group and would function as chain stoppers, thus severely limiting their utility in the structure design of the resin molecules. In general practice, the oil is rst heated to 230250 C under an inert gas blanket and agitation. The catalyst, usually predispersed in a small quantity of the oil, and the polyol, usually at 2 times the molar quantity of the oil present in the reactor, are added. The batch is reheated to and maintained at the desired temperature, usually in the 230250 C range. The progress of the reaction is monitored by periodical sampling from the reactor and checking miscibility with anhydrous methanol. This is because triglycerides are not soluble in methanol, whereas monoglyceride is. When a volume of the alcoholysate can tolerate three or more volumes of methanol without becoming turbid, the alcoholysis process is considered complete. Acidic contaminants are poisonous to the alcoholysis catalysts and must be avoided. If the oil has a high acid number, or there are high acidity residues left in the reactor from the previous batch, such as sublimed phthalic anhydride condensed under the dome of the reactor, the reaction can be severely retarded. A longer batch time or additional amount of the catalyst would then be required. Both are undesirable. When the alcoholysis step is complete, the polybasic acid(s) and the balance of polyol, if any, are added. The batch is reheated to and maintained at about 250 C to carry out the polycondensation step to the desired endpoint, usually a combination of the acid value and viscosity of the resin. Fatty Acid Process. When free fatty acids are used instead of oil as the starting component, the alcoholysis step is avoided. All of the ingredients can, therefore, be charged into the reactor to start a batch. The reactants are heated together, under agitation and inert blanket, until the desired end point is reached. Chen and Kumanotani (64) reported that alkyds prepared by the fatty acid process have narrower molecular weight distribution and give lms with better dynamicmechanical properties. A modied form of the fatty acid process, dubbed high polymer alkyd technique was reported (65). A portion of the fatty acids is withheld in the rst stage of the process to allow the polycondensation between the dibasic acid and the polyol to have a better chance of extending the polyester chain without being terminated by the monoacids. After the acid value of the reactant has reached a desired low level, indicating the completion of the poly-condensation, the remaining portion of fatty acids is then added to complete the process. The resins prepared by this technique have more linear backbone chain structure, higher molecular weight, and higher viscosity than the corresponding ones with identical formulation but prepared by the conventional process. Fatty AcidOil Process. When oil represents only a minor portion (33% or less) of the total furnish of fatty acids in an alkyd formulation, the alcoholysis step may be avoided. All of the ingredients, dibasic acid, polyol, oil, and free fatty acids may be charged together into the reactor and proceed as in the fatty acid process. Apparently, the oil is incorporated into the resin by ester interchange at the reaction

338

PAINTS, VARNISHES, AND RELATED PRODUCTS

temperature. The resultant resins give higher viscosity (4), faster surface drying, and slower through-dry (3). If the oil content is too high, not enough of it may be incorporated in time, then it, would result in a partial gelation to form seeds. Acidolysis Process. As mentioned previously, isophthalic and terephthalic acids are difcult to process in ordinary alkyd preparation methods, due to their high melting point and low solubility in the reaction mixture. An acidolysis process was developed for this purpose (6). The dibasic acid is heated together with the oil in the resin formulation under agitation and inert gas blanket to about 280 C, holding for about 40 min. In this reaction, which is self-catalyzed by the acidity of the reaction mixture, an ester interchange occurs. A carboxyl group of the dibasic acid displaces that of a fatty acid on the oil molecule and splits off the fatty acid. The completeness of the acidolysis reaction is determined by a tedious extraction of the oil phase and analysis of its free fatty acid content by titration. The analysis takes several hours to complete. Rapid test methods, comparable to the methanol miscibility test for alcoholysis, that could be used for process control of the acidolysis reaction have yet to be developed. Therefore, the process is normally controlled by reaction time and temperature, based on experience. After acidolysis, the reactant temperature is dropped to about 230 , the polyol is charged and heated back up to the desired temperature to bring the esterication step to the desired end point. The acidolysis process is not suitable for phthalic anhydride or other dibasic acids with a high tendency to sublime. Alkyd Resin Production Processes. Parallel to the above chemical procedures, the processing method may also be varied with different mechanical arrangements to remove by-product water, to drive the esterication reaction toward completion. Fusion Process. In the fusion process, also frequently referred to as fusion cook, inert gas is continuously sparged from the bottom of the reactor to carry away water vapor from the reaction mixture. The exhaust is then either vented away or sent to a fume scrubber, which is frequently a small vessel with water atomizing nozzles. After the reaction is completed, the nished resin may be discharged, ltered, and packaged without solvent. More frequently, it is cooled to a safe temperature, then dissolved with the desired type and amount of solvent in a thinning tank, ltered, and packaged, or pumped to a storage tank. The reactor usually needs to be cleaned by charging a small volume of solvent into the vessel and heated to reux for an appropriate time period. If deemed necessary, the vessel is further cleaned by digesting with caustic soda solution. The fusion process has the advantage of simplicity in mechanical arrangement. However, it has several signicant disadvantages. Low boiling and/or subliming ingredients, such as glycols and phthalic anhydride, would be lost during the reaction causing the product composition and its properties to deviate from the design. The material loss causes an increase in the cost of the resin. Reactants as well as the product may adhere to the reactor walls above the surface level of the charge, which will contaminate or even become catalyst poisons to the subsequent batches. And the resin produced from a fusion cook is more prone to develop dark colors. For these reasons, most of the manufacturers have discontinued the practice of fusion cook, unless it is dictated by the existing equipment.

ALKYD RESINS

339

Figure 2. Equipment for solvent processing of alkyd resins. Courtesy of Hercules Inc. (Wilminton, Del.).

Solvent Process. In the solvent process, or solvent cook, the water formed from the reaction is removed from the reactor as an azeotropic mixture with an added solvent, typically xylene. Usually between 3 to 10 weight percent, based on the total charge of the solvent, are added at the beginning of the esterication step. The mixed vapor passes through a condenser. The condensed water and solvent have low solubility in each other, and phase separation is allowed to occur in an automatic decanter. The water is removed, usually to a measuring vessel. The amount of water collected can be monitored as one of the indicators of the extent of the reaction. The solvent is continuously returned to the reactor to be recycled. A typical equipment for this process is shown in Figure 2. The reactor temperature is

340

PAINTS, VARNISHES, AND RELATED PRODUCTS

modulated by the amount and type of reuxing solvent. Typical conditions are as follows.

Solvent Xylene Xylene Xylene High ash naphtha

Weight Percent 3 4 7 10

Temperature ( C) 251260 246251 204210 204210

The solvent vapor also serves as a blanket in the reactor. The processing solvent is usually left in the product as part of the dilution solvent. The reuxing solvent provides a constant wash to the reactor, and brings back the reactants that had escaped out of the reaction mixture. The reaction temperature is better controlled by the constant reuxing, and the viscosity of the reaction mixture is lower, which improves the effectiveness of the agitation. The product usually has better color and is more uniform than those made by the fusion process. Ordinarily, the reactor requires no more than a solvent wash to be clean enough for the next batch. These advantages far outweigh the higher cost of the production facility. Therefore, few would consider building a new alkyd plant without solvent process capability. When low boiling ingredients such as ethylene glycol are used, a special provision in the form of a partial condenser will be needed to return them back into the reactor. Otherwise, not only would the balance of the reactants be upset and the raw material cost of the resin be increased, they would also become part of the pollutant in the waste water and incur additional water treatment costs. Usually, a vertical reux condenser or a packed column is used as the partial condenser, which is installed between the reactor and the overhead total condenser (Figure 3). The temperature in the partial condenser is monitored and maintained at a level to effect a fractionation between water, which is to pass through the reactor, and the glycol or other materials, which is to be condensed and returned to the reactor. If the fractionation is poor and water vapor is also condensed and returned, the reaction will be retarded and result in a loss of productivity. As the reaction proceeds toward completion, water evolution slows down, and most of the glycol will have been combined into the resin structure. The temperature in the partial condenser may then be raised to facilitate the removal of water vapor. 5.9. Process Control The progress of the alkyd reaction is usually monitored by periodical determinations of the acid number and the viscosity (solution in a suitable solvent and at an appropriate concentration) of samples taken from the reactor. The frequency of sampling is commonly every half hour. The general practice is to plot the determined values separately against time on semilogarithmic coordinants (Figure 4).

ALKYD RESINS

341

Figure 3. Solvent-processing equipment using a partial condenser. Courtesy of Hercules Inc. (Wilmington, Del.).

Toward the end of the reaction, the resin viscosity tends to increase exponentially. Gelation in the reactor is always a threat, due either to what the formulation would theoretically allow by the completion of the polyesterication or to the occurrence of some of the side reactions. After the onset of gelation, it would progress extremely rapidly and would be almost impossible to arrest. Therefore, it is routine to

342

PAINTS, VARNISHES, AND RELATED PRODUCTS

Figure 4. Alkyd reaction control plots.

extrapolate the plots in Figure 4 when predicting the point at which to terminate the reaction in time to prevent gelation. If gelation should occur in the reactor, it would cause not only the loss of product but also signicant down time for cleaning the reactor. Some alkyd practitioners have found that a rapid addition of a large quantity of raw oil quenches the runaway gelation, disperses the gel, and signicantly eases the cleanup operation. A technique of injecting water or steam during the condensation process to reverse or retard gelation has been reported in the patent literature (66).

MODIFICATION OF ALKYD RESINS BY BLENDING WITH OTHER POLYMERS

343

6. SAFETY AND ENVIRONMENTAL PRECAUTIONS The manufacturing of alkyd resins involves a wide variety of organic ingredients. While most of them are relatively mild with low toxicity, some of them such as phthalic anhydride, maleic anhydride, solvents, and many of the vinyl (especially acrylic) monomers are known irritants, or skin sensitizers, and are poisonous to humans. Persons involved should be thoroughly familiar with the hazard potential of each and every one of the chemicals by consulting the material safety data sheets provided by the suppliers and practicing the recommended safety precautions in handling the materials. The use of personal safety equipment such as protective goggles, gloves, clothing, and respiratory devices should be diligently observed. Since large quantities of highly ammable solvents are routinely handled in an alkyd plant, re safety should be the utmost in everyones mind. Electrical equipment and power circuitry in the plant should conform to all applicable codes, and all equipment should be properly grounded. The areas for the reactors and storage tanks should be separated by re walls and must be adequately ventilated. The storage tanks should be blanketed by inert gas. A slight positive pressure of inert gas should be maintained in the reactor or storage tanks during discharge of the resin or resin solution to prevent air from being sucked into the vessel to form an explosive mixture with the solvent vapor. With the ever increasing awareness of the need for environmental protection, the emission of solvent vapors and organic fumes into the atmosphere should be prevented by passing the exhaust through a proper scrubber. The solvent used for cleaning the reactor is usually consumed as part of the thinning solvent. Aqueous efuent should be properly treated before discharge.

7. MODIFICATION OF ALKYD RESINS BY BLENDING WITH OTHER POLYMERS As mentioned earlier, one of the important attributes of alkyds is their good compatibility with a wide variety of other coating polymers. This good compatibility comes from the relatively low molecular weight of the alkyds and the fact that the resin structure contains, on the one hand, a relatively polar and aromatic backbone and, on the other hand, many aliphatic side chains with low polarity. The alkyd resin involved in a blend with another coating polymer may serve as a modier for the other lm former, or it may be the major lm former and the other polymer may serve as the modier for the alkyd to enhance certain properties. The following describes some of these compatible blends. Nitrocellulose-based lacquers often contain a fair amount of short- or mediumoil alkyds to improve exibility and adhesion. The most commonly used are short-oil nondrying alkyds. Amino resins or urethane resins with residual isocyanate functional groups may be added to cross-link the coating lm for improved

344

PAINTS, VARNISHES, AND RELATED PRODUCTS

solvent and chemical resistance. The major applications are furniture coatings, top lacquer for printed paper, and automotive renishing primers. Amino resins are probably the most important modiers for alkyd resins. Butylated urea- or melamineformaldehyde resins are compatible with alkyds. They react with the free hydroxyl groups of the alkyd to effect cross-linking and to impart hardness, mar resistance, chemical resistance, and durability to the coating. Shortor medium-oil alkyds of both drying and nondrying types are frequently used. Color and color retention requirements often dominate the choice of the alkyd. Many industrial baking enamels, such as those for appliances, coil coatings, and automotive nishes (especially renishing enamels), are based on alkyd-amino resin blends. Some of the so-called catalyzed lacquers for nishing wood substrate require low bake or no bake at all. Chlorinated rubber is often used in combination with medium-oil drying type alkyds. The alkyd gives better toughness, exibility, adhesion, and durability, and the chlorinated rubber contributes to faster dry and better resistance to water and chemicals. The major applications are highway trafc paint, concrete oor, and swimming pool paints. Vinyl resins of the type that are the copolymers of vinyl chloride and vinyl acetate and that contain a fair amount of hydroxyl groups (from the partial hydrolysis of vinyl acetate) and/or carboxyl groups (e.g., from copolymerized maleic anhydride) may be formulated with alkyd resins to improve application properties and adhesion. The blends are primarily used in making marine topcoat paints. Synthetic latex house paints sometimes contain emulsied long-oil or very long oil drying alkyds to improve adhesion to chalky painted surfaces. Silicone resins with high phenyl contents may be used with medium- or short-oil alkyds as blends in air-dried or baked coatings to improve heat and weather resistance, whereas the alkyd component contributes to adhesion and exibility. Major applications include insulation varnishes, heat-resistant paints, and marine coatings. 7.1. Chemically Modied Alkyd Resins While blending with other coating resins provides a variety of ways to improve the performance of alkyds, or vice versa, chemically combining the desired modier into the alkyd structure would eliminate the compatibility problem and give a more uniform product. Several such chemical modications of the alkyd resins have gained commercial importance, and are described below. Vinylated alkyds are alkyd resins that have been incorporated with a signicant amount (2060% by weight) of vinyl monomers (such as styrene, vinyl toluene, and methyl methacrylate) by grafting the monomers through a free-radical mechanism onto unsaturated reaction sites in the resin molecules. The modied resin embodies the good attributes of ease of application, good wetting, and adhesion from the alkyd as well as fast solvent release, hardness, and weather resistance from the vinyl modication. The common objective of such a modication is for achieving a drying rate comparable to that of lacquer materials. The reaction sites on alkyd resin molecules are primarily the allylic carbons on unsaturated fatty acid chains

MODIFICATION OF ALKYD RESINS BY BLENDING WITH OTHER POLYMERS

345

and the double bond of a,b-unsaturated dibasic acids. Free radicals, generated from the thermolysis of such free-radical initiators as benzoyl peroxide, dicumyl peroxide, and di-t-butyl peroxide are usually required to kick off the reaction. Ideally, the initiating species would attack the active sites on the resin molecules and all of the added monomers would be evenly distributed in grafted side chains. In reality, it is inevitable that part of the monomers would engage in homopolymerization, and some of the resin molecules would remain unmodied. The presence of a large amount and high molecular weight homopolymer of the vinyl monomer would lead to incompatibility and result in a hazy product. Methods that have been used to minimize homopolymerization include using fatty acids having conjugated diene structure, using maleic anhydride as part of the dibasic acids for the alkyd, choosing initiators such as peroxides or hydroperoxides that tend more to extract allylic protons than to add to double bonds, avoiding initiators that would decompose at very low temperatures, adding the monomer gradually along with an appropriate amount of the initiator, choosing monomers that would have a more favorable tendency to copolymerize with the active species on the alkyd resin, and properly maintaining the reaction temperature. Chain transferring is the preferred mechanism for terminating chain growth from the addition polymerization of the monomer. Usually, the solvent, the fatty acid chain, and the monomer are effective chain-transfer agents. If an additional transferring agent is used, care must be exercised, or too much of it could cause the formation of a large amount of very low molecular weight homopolymers, and would result in poor lm properties. Occasionally, vinylation is rst performed on the fatty acids or the oil before the alkyd reactions. It should be emphasized that the presence of a large amount of either conjugated fatty acids or maleic anhydride in the alkyd formulation gives rise to a high degree of probability of premature gelation during the alkyd reaction. An allowance must be made in the alkyd formulation, and the polyesterication is frequently terminated at a relatively high acid number (about 15), to avoid gelation. It has been reported that the optimum amount of maleic anhydride in the alkyd is an amount having a maleic group in one-third of the resin molecules (17). A common procedure for the preparation of vinylated alkyds is as follows: rst, a base alkyd resin is brought to the desired end point. The resin is then cooled to about 160 C and often diluted with aromatic thinner. Next, the desired monomer is added, usually at about 2060% based on the nal product, followed by an appropriate amount of a free-radical initiator. Alternatively, a premix of the monomer and the initiator is added at a controlled rate over most of the reaction. Then the reaction is brought to monomer reux, until the residual monomer content has dropped below a specied level. The residual monomer, if any, is stripped away before the product is diluted in a solvent, ltered, and packaged. Silicone alkyds are etherication products of alkoxy-polysiloxane oligomers and the free hydroxyl groups of alkyd resins (67, 68). The property improvements and applications are similar to those of the alkyd-silicone blends, with the added advantage of incorporating the stable O structure into the alkyd moleSi C cules. The preferred silicone oligomers are those with high phenyl contents, and

346

PAINTS, VARNISHES, AND RELATED PRODUCTS

the alkyds at long- or medium-oil length based on polyols with primary hydroxyls. To improve the thermal stability effect imparted by the silicone modication further, one can use isophthalic alkyds rather than the phthalic type. The etherication reaction may be carried out on an alkyd resin designed for the purpose or on the polyol before it is used in alkyd preparation. The silicone content of the modied alkyd lies usually between 20 and 60% of the total product. Urethane alkyds, or uralkyds, are alkyds with a part or even all of the dibasic acids replaced by diisocyanates. The isocyanate group, C O, reacts with N the hydroxyl group of a polyol, at low temperature, to form a urethane linkage, NHC(O)O without spitting out water as a by-product. Toluene diisocyanate , (TDI) is commonly used for such modication. It is commercially supplied as an 80 : 20 mixture of 2,4- and 2,6-isomers. The NCO group para to the methyl group has about 8 times greater reactivity than the one on the ortho position, which aids greatly the control of the reaction. Since the NCO group is reactive with labile protons, water must be excluded from the reaction system. The esterication reaction of the base alkyd must be brought to the desired end point with the by-product water removed, and the temperature lowered to about 100 C to prevent any continuation of the esterication before the introduction of the NCO reactant. TDI is highly toxic and is usually handled in a closed system under a dry inert gas blanket. Metallic soaps, such as dibutyltin dilaurate, stannous octoate, and calcium naphthanate, are used as reaction catalysts. The reaction is vigorous and exothermic. Therefore, the reaction temperature is maintained under 135 C, and great care must be exercised to bring the reaction under proper control. Uralkyds have superior adhesion, hardness, abrasion resistance, durability, and chemical resistance to the unmodied alkyds. They nd major applications in wood oor nishes, marine coatings, metal primers, and maintenance paints. Phenolic resins are well known for their contribution in improving hardness, gloss, and water and chemical resistance in oleoresinous varnishes. Those based on p-alkyl-substituted phenols and with heat-reactive methylol groups have also been incorporated into alkyd resins. The reaction has not been well studied. Presumably, the methylol group would react with the unsaturation functionality on the fatty acid chain to form the chroman structure, similar to what is believed to have occurred in the varnish. Etherication between the methylol group and free hydroxyl of the alkyd resin, catalyzed by the residual acidity in the resin, would be another possible reaction. Polyamide modied alkyds show a special rheological behaviorthey are thixotropic (69, 70). Typically, the polyamide resin would be of the type based on dimer acids, i.e., dimerized unsaturated fatty acids, and aliphatic diamines, such as ethylene diamine. These would react to form polyamide resins with low acid and amine values. The alkyd resin would be a medium- or long-oil drying alkyd. The reaction products from the polyamide and the alkyd are gel-like materials that undergo a time-dependent shear thinning and recover to the gel-like state after the shearing action is stopped. This allows the preparation of no-drip paints, which are easy to brush and can be applied at high lm thickness from a single coat with little or no danger of sagging. Pigment settling during storage of the paint is also

MODIFICATION OF ALKYD RESINS BY BLENDING WITH OTHER POLYMERS

347

minimized. The major applications are at oil-based architectural paints and maintenance paints. Generally, up to 10% of the polyamide based on the weight of the alkyd is added to the alkyd and heated at normal alkyd reaction temperature under agitation. Ester interchange reaction takes place, and fragments of the polyamide resin become chemically bonded to the alkyd. The modied alkyd serves as a compatibilizer for the mutually insoluble unreacted polyamide and alkyd to form a gellike structure. The stiffness of the structure decreases with the increasing amount of the compatibilizer. If the reaction is allowed to continue, and there is no unreacted polyamide left in the system, little or no thixotropy will be exhibited. Therefore, the reaction must be precisely controlled to give the desired degree of thixotropy. Aromatic solvents such as xylene tend to destroy the thixotropic structure. Therefore, they must be reduced to less than 0.5% in the product. Polyamide modied alkyd resins are available commercially to be used as additives for making thixotropic alkyd paints.

7.2. High Solids Alkyds There has been a strong trend in recent years to increase the solids level of all coating materials, including alkyds, to reduce solvent vapor emission. To raise the solids level and still maintain a manageable viscosity, the molecular weight of the resin must be reduced. Consequently, lm integrity must be developed through further chain extension and/or cross-linking of the resin molecules during the drying step. A high cross-linking density necessitated by the lower molecular weight of the resin would build a high level of stress in the lm, and cause it to be prone to cracking. Therefore, adequate exibility should be designed into the resin structure. This means that the distance between the hydroxyl groups of the polyol and the carboxyl groups of the dibasic acid would need to be lengthened by linear linkages. Thus long-chain diols, polyether polyols, and linear a,o-dibasic acids would not only build in more exibility but also reduce the viscosity for high solids alkyds, due to the greater spacing of polar ester groups and the reduction of aromaticity in the resin structure. In addition to the manipulation of resin molecular structure for increasing coating solids, the use of more active, though more expensive, oxygenated solvents also serves to reduce the viscosity of resin solutions. Chain extension and cross-linking of high solids alkyd resins are typically achieved by the use of polyisocyanato oligomers or amino resins. An adequate amount of excess hydroxyl groups must be designed into the alkyd structure to provide reaction sites for these modiers. To limit the molecular weight of the alkyd resin, the molar ratio between polyols and dibasic acids should be greater than 1. The hydroxyl functionality of the formulation should be controlled by a careful selection of polyols to avoid an overpresence of free hydroxyl groups in the product, which would adversely affect water resistance and other properties of the coating lm. Most of the high solids alkyd systems are used in industrial baking nishes. For air drying applications, higher doses of driers are usually needed to achieve acceptable drying rate (71).

348

PAINTS, VARNISHES, AND RELATED PRODUCTS

7.3. Water-Reducible Alkyds Replacing solvent-borne coatings with water-borne coatings would not only reduce solvent vapor emission but also improve safety against the re and health hazards of organic solvents. Alkyd resins may be rendered water-reducible by either converting the resin into an emulsion form or by incorporating water-soluble groups in the molecules. The latter will be the subject for further discussion. The most common approach for imparting water solubility in an alkyd resin is to leave enough pendent carboxyl groups in the resin and to neutralize them with a fugitive base, such as ammonia or low molecular weight amines, to build ionic characteristics into the resin. Nonfugitive base materials, such as caustic soda, would leave the salt in the coating lm and damage its water- and corrosion-resistance. Trimellitic anhydride (TMA) is the most frequent choice of ingredient to provide the pendent carboxyl groups. It was reported that glycerol gives resins with poor hydrolytic stability (72). Therefore, polyols with primary hydroxyl groups are preferred for the preparation of water-soluble alkyds. The recommended procedure (37) for the preparation of water-soluble alkyds is to hold off the TMA in the initial stage of the alkyd reaction so that the high functionality of TMA would not be a cause of gelation. When the reaction has progressed to a desired low acid number, i.e., the building of the polymer chain is completed, the temperature is lowered to 180 C; the TMA is then added and maintained at that temperature until a desired acid number, usually about 5060, is reached. At such a temperature, only the anhydride group of the TMA would react to form half esters, and the remaining two carboxyl groups would essentially remain unreacted. If one desires to have the TMA participating in the backbone structure of the resin, a part of the TMA is charged in the beginning of the alkyd reaction, often with the presence of an appropriate amount of a monohydric alcohol, such as benzyl alcohol, to balance the functionality of the system. The remaining TMA is then added in the same manner as described to end cap the resin and to provide pendent carboxyl groups for water solubilization. The nished resins are usually dissolved in oxygenated coupling solvents, such as glycol ethers, to improve the solubilization of the resin in aqueous media and the handling of the resin. Water and the base are premixed and added to the resin solution when needed. The coupling solvents usually have higher boiling points than water. During the drying process, the solvent would be enriched in the coating lm as water evaporates preferentially. The resin molecules would become better solubilized, i.e., molecular chains would be extended, and result in better formation and integrity of the lm.

8. ECONOMIC ASPECTS Alkyd resins as a family have remained the workhorse of the coatings industry for decades. In the United States, the total consumption of alkyds increased from about 200,000 t in the mid-1950s to more than 300,000 t in the mid-1960s. It peaked in

REFERENCES

349

1973 at about 345,000 t, constituting about 33% of all synthetic coating resins. In 1980, alkyds still accounted for 30% of the 1,090,000 t of all resins consumed for coatings. From 1987 to 1989, although the consumption maintained at about 300,000 t/year, its market share among all coating resins was reduced to 26% in 1987 and 25% in 1989. At present, 5560% of the alkyd resins consumed in the United States are used for architectural coatings. A decline in consumption is expected because of regulations involving VOCs (volatile organic emissions). California and the Northeastern United States are expected to adopt regulations that will severely restrict the use of solvent-borne coatings (73). The overall demand in Europe is expected to decrease at the rate of 3%/yr over the next ve years. Environmental regulations are expected by 2007. Japans consumption has declined, but has stabilized because high performance alternatives have already replaced the coatings in question. The industry was hard hit in 20032004 with higher prices for raw materials such as linseed and soybean oil. Alkyd producers are already developing new water-borne products (73). Other uses for alkyds are in general industrial coatings such as machinery and metal furniture. Alkyd resin-chlorinated rubber based coatings are used in trafc paints, but use is declining because of VOC concerns. Some alkyds are still used in renish paints for automobiles. Uralkyds are used as a vehicle for urethane varnishes for the do-it-yourself market. 9. FUTURE PROSPECTS Stemming from the drive by the coatings industry to reduce solvent emission, there has been a clear trend of gradual decline in the market share of alkyd resins. However, their versatility and low cost will undoubtedly continue to keep alkyds as major players in the coatings arena. Alkyds are much more amenable to move toward higher solids than most other coating resins. Great strides in the development of water-borne types have also been made in recent years. There is one more good reason to remain optimistic about alkyds for the futurea signicant portion of their raw material, fatty acids, is renewable. REFERENCES
1. 2. 3. 4. 5. 6. 7. A. E. Rheineck, J. Paint Technol. 44(566), 3554 Apr. 1972. A History of Paint and Color, Pittsburgh Plate Glass Co., 1951. An Overview of the Paint and Coatings Industry, SRI International, Aug. 1992. D. Deffar and M. D. Soucek, J. Coat. Technol, 73 (919), 95 (2001). Urethane Surface Coatings, SRI Consulting, Oct. 2004. Epoxy Surface Coatings, SRI Consulting, Oct. 2004. Z. W. Wicks, Jr. and F. N. Jones in Z. W. Wicks, F. N. Jones, and S. P. Pappas, eds., Organic Coatings Science & Technology, Vol. 1, John Wiley & Sons, Inc., New York, 1992, Chapt. 9.

350

PAINTS, VARNISHES, AND RELATED PRODUCTS

8. A. E. Rheineck and R. O. Austin in R. R. Meyers and J. S. Long, eds., Treatise on Coatings, Vol. 1, Part 2, Marcel Dekker, Inc., New York, 1968, Chap. 4. 9. D. N. Rampley and J. A. Hasnip, J. Oil Colour Chem. Assoc. 59, 356362 (1976). 10. O. S. Privett, W. O. Lundberg, N. A. Khan, W. E. Tolberg, and D. H. Wheeler, J. Am. Oil Chem. Soc. 30, 61, (1953). 11. S. A. Harrison and D. H. Wheeler, J. Am. Chem. Soc. 76, 2379 (1954). 12. O. S. Privett and C. J. Nickell, J. Am. Oil Chem. Soc. 33, 156 (1956). 13. W. J. Bailey and G. L. Barlow, J. Paint Technol. 42(544), 287 (1970). 14. E. H. Farmer and A. Sundralingam, J. Chem. Soc. 1942, 121. 15. E. H. Farmer and D. A. Sutton, J. Chem. Soc. 1942, 139. 16. J. L. Bolland and G. Gee, Trans. Faraday Soc. 42, 236, 244 (1946). 17. E. H. Farmer, Trans. Faraday Soc. 42, 228 (1946). 18. E. H. Farmer, Trans. Inst. Rubber Ind. 21, 122 (1945). 19. F. D. Gunstone and T. P. Hilditch, J. Chem. Soc. 1946, 1022. 20. N. A. Khan, Can. J. Chem. 32, 1149 (1954). 21. N. Uri in W. O. Lunberg, ed., Autoxidation and Antioxidants, Vol. 1, Wiley-Interscience, New York, 1961, Chapt. 2. 22. F. W. Heaton and N. Uri, J. Lipid Res. 2, 152 (1961). 23. T. H. Smouse and S. S. Chang, J. Am. Oil Chem. Soc. 44, 509 (1967). 24. H. P. Kaufmann, Fette Seifen Anstrichmit. 59, 153162 (1957). 25. O. S. Privett, J. Am. Oil Chem. Soc. 36, 507 (1959). 26. R. N. Faulkner, J. Appl. Chem. 8, 448458 (1958). 27. R. R. Allen and F. A. Kummerow, J. Am. Oil Chem. Soc. 28, 101 (1951). 28. A. E. Rheineck and S. C. G. Peng, Ofcial Digest 36, 878 (1964). 29. L. A. ONeill, Chem. Ind. 1954, 384387. 30. K. Hultzsch, J. Prakt. Chem. 158, 275 (1941). 31. R. H. Kienle and C. S. Ferguson, Ind. Eng. Chem. 21, 349 (1929). 32. R. H. Kienle, Ind. Eng. Chem. 41, 726 (1949). 33. R. G. Mraz and R. P. Silver in Encyclopedia of Chemical Technology, 2nd ed., Vol. 1, John Wiley & Sons, Inc., New York, 1963, pp. 851880. 34. Z. W. Wicks, Jr., in Encyclopedia of Chemical Technology, 5th ed., Vol. 2, John Wiley & Sons, Inc., Hoboken, New Jersey, 2004, pp. 147169. 35. C-D-I-C Society for Paint Technology, Ofcial Digest 430, 1477 (Nov. 1960). 36. Bulletin IP4, Amoco Chemicals Corp., Chicago, Oct. 1960. 37. Bulletin IP-65b, Amoco Chemicals Corp., Chicago, 1990. 38. Technical Bulletin No. 524-5, Velsicol Chemical Corp., Chicago. 39. R. I. Stirton in Encyclopedia of Chemical Technology, 1st ed., p. 595, Interscience Publishers, New York, 1953. 40. Bulletin TMA 27, Amoco Chemicals Corp., Chicago. 41. IMC Polyols: A Complete Guide, International Minerals & Chemicals Corp., Des Plaines, Ill., 1980, p. 9. 42. Trimethylolpropane in Alkyd Coating Resins, Celanese Chemical Co., New York, 1961, p. 4.

REFERENCES

351

43. L. Klee and G. H. Benham, J. Am. Oil Chem. Soc. 27, 130133 (1950). 44. A. E. Rheineck, F. D. Williamson, and D. DeClerk, Am. Chem. Soc. Div. Org. Coat. Plast. Chem. 22(1), 2434 (1962). 45. W. H. Carothers, J. Am. Chem. Soc. 51, 1548 (1929). 46. P. J. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca, N.Y., 1953, Chapt. 9. 47. W. H. Carothers, Trans Faraday Soc. 32, 43 (1936). 48. T. C. Patton, Alkyd Resin Technology, Formulating Techniques and Allied Calculations, Wiley Interscience, New York, 1962. 49. E. G. Bobalek, E. R. Moore, S. S. Levy, and C. C. Lee, J. Appl. Polym. Sci. 8, 625657 (1964). 50. D. H. Solomon, B. C. Loft, and J. D. Swift, J. Appl. Polym. Sci. 11, 15931602 (1967). 51. D. H. Solomon and J. J. Hopwood, J. Appl. Polym. Sci. 10, 18931914 (1966). 52. J. Kumanotani, H. Hata, and H. Masuda, Org. Coat. 6, 3554 (1984). 53. H. Kiryu, K. Horiuchi, K. Sato, and J. Kumanotani, Shikizai Kyokaishi 57(11), 597601 (1984). 54. H. Hata, H. Tomita, Y. Nishizawa, and J. Kumanotani, Fatipec Cong. 15(3), 111-485/509 (1980). 55. H. Hata, J. Kumanotani, Y. Nishizawa, and H. Tomita, Fatipec Cong. 14, 359364 (1978). 56. N. M. Wiederhorn, Am. Paint J. 41(2), 106 (1956). 57. R. P. Silver, Alkyd Report No. 8, Hercules Inc., Wilmington, Del. 58. R. P. Silver, Alkyd Report No. 1.2, Hercules Inc., Wilmington, Del. 59. T. C. Patton, Off. Digest Fed. Paint Varnish Prod. Clubs 430, 1544 (1960). 60. Y. Tanizaki and T. Yoshida, Shikizai Kyokaishi 45(2), 8387 (1972). 61. T. Nagata, J. Appl. Polym. Sci. 13, 26012619 (1969). 62. T. Yoshida, Kobunshi Kagagu 22, 769 (1965). 63. Alkyd Report No. 1.5, Hercules Inc., Wilmington, Del. 64. L. W. Chen and J. Kumanotani, J. Appl. Polym. Sci. 9, 36493660 (1965). 65. W. M. Kraft, G. T. Roberts, E. G. Janusz, and J. Weisfeld, Am. Paint J. 41(28), 96 (1957). 66. W. B. Callahan and B. F. Coe (to E. I. du Pont de Nemours & Co.), U.S. 4,045,932 (Aug. 30, 1977). 67. A. F. Karim, B. Golding, and R. A. Morgan, J. Chem. Eng. Data 5, 117 (1960). 68. Silicone Notes 7-701, 7-702a, 7-703, Technical Bulletin, Dow Corning Corp., Midland, Mich. 69. A. G. North, J. Oil Color Chem. Assoc. 39(9), 696 (1956). 70. P. Oldring and G. Hayward, eds., Resins for Surface Coatings, SITA Technology, London, 1987. 71. J. Larson, Am. Paint Coat. J., 5661 (Apr. 4, 1984). 72. M. Lerman, J. Coat. Technol. 48(12), 3742 (1976). 73. Alkyd/Polyester Surface Coatings, SRI Consulting, Jan. 2005.

You might also like