Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Progress in Organic Coatings, 20 (1992) 139-167 139

Developments in the field of rosin chemistry and its


implications in coatings

Gan&Fung Chen
Research and Development, P.O. Box 2219, Ashland Chemical, Inc., Columbus,
OH 43216 (USA]

Contents

1 Introduction .............................................. 139


2 Wood and rosin ............................................ 140
3 Chemical composition of rosin .................................. 141
4 Isomerization of resin acids .................................... 143
4.1 Isomerization of abietic-type resin acids ......................... 143
4.1.1 Acid-catalyzed isomerization of resin acids ................... 143
4.1.2 Thermal isomerization of resin acids ....................... 144
4.2 Isomerization of pimaric-type resin acids. ........................ 149
5 Other reactions of resin acids. .................................. 149
5.1 Oxidation and disproportionation. ............................. 150
5.2 Hydrogenation and dehydrogenation ............................ 150
5.3 Dimerization and polymerization .............................. 151
6 Modifications of rosin for coating applications ........................ 154
6.1 Carboxy functionality and esterification .......................... 154
6.2 Condensation with dienophiles ............................... 155
6.2.1 With maleic anhydride ................................ 155
6.2.2 With other dienophiles ................................ 157
6.2.3 Coatings applications ................................. 158
6.3 Condensation with formaldehyde .............................. 159
6.4 Condensation with phenohc resin. ............................. 162
7 Conclusion ............................................... 164
Acknowledgements ............................................ 164
References ................................................. 164

1 Introduction

The composition of the old Italian varnish used during the 200 year
period from AD 1550 to AD 1750, and by Stradivarius has been speculated
upon for many years. It is generally thought [ 1,2] that the major components
of the varnish probably contained linseed oil hardened by natural resins such
as rosin derivatives. Throughout history, the early Egyptians and Persians
were known to employ gum arabic as a binder for preparing their varnishes,
and the ancient Chinese and Japanese were also known to use sap tapped
from trees as vehicles in the preparation of protective lacquers [3 1. Thus

0033-0655/92/$5.00 0 1992 - EIsevier Sequoia. AU rights reserved


140

the coatings industry has historically been a large user of renewable resources
such as seed oils and wood derivatives primarily because of availability.
The Arab-Israeli war and the subsequent oil embargo of 1973 resulted
in dramatic OPEC oil price increases. The recent war in the Gulf was also
in part due to the supply of World oil. Both events reflect the fact that we
are becoming increasingly dependent on the use of fossil fuels despite the
continued decrease [ 41 in global oil reserves. On the other hand, the extensive
use of fossil fuels is a major contributor to global warmings [5]. It has been
estimated [5] that if present trends in greenhouse-gas emissions continue,
the earth could be committed to a warming of 1.5-4.5 “C by the middle of
the next century. The key to preventing global warming is to reduce World
consumption of fossil fuels by 50% or more over the next several decades
[5]. With the exhaustion of readily accessible petroleum and natural gas
looming in the future, and the increasing environmental concern, a shift back
to an emphasis on renewable resources for raw materials is becoming imminent.
It will be too late if and when the coatings industry has to be regulated by
another ‘Rule’ or ‘Act’ to restrict the use of raw materials when coal and
petroleum are near exhaustion. Thus, the need to seriously consider renewable
resources in coating technologies is clear [ 6, 71.
Rosin, being one of the important renewable resources, has in the past
found its use in many coating applications such as adhesives, varnishes and
printing inks [S]. This is largely due to factors such as fast drying properties,
excellent solubility and compatibility with other resins and oils, economy
and ready availability. However, the use of this renewable raw material has
declined slowly since the 1950s (9, lo]. This decline can be attributed to
the increased supplies of petroleum-based resins.
Contemporary polymer synthesis places a strong demand on good control
of molecular architecture. This requires a good understanding of rosin
chemistry if it is to be considered as a polymer intermediate for protective
coating applications. The large amount of work done in the early to mid
parts of this century on the chemistry of rosin, such as isomerization upon
heating, can be complicated and easily forgotten. The purpose of this
manuscript is to re-explore knowledge of rosin chemistry and to discuss
various aspects of rosin modification and its implications for protective
coatings.

2 Wood and rosin

The chemical composition of various wood species generally consists


of 35-45% cellulose, 25-35% lignin, 20-30% hemicellulose and 2-5% ex-
tractives where rosin is found. The content of extractives and their composition
vary greatly among different wood species and also within different parts of
the same tree. They consist mainly of rosin, fatty acids, phenolic compounds
and a large variety of terpene derivatives [ 111. Furthermore, the extractives
content increases after the death of a living tree due to chemical changes.
141

Up to 12-l 4% extractives were found in some dead pine species [ 111 and
up to 27% of extractives were reported for an aged virgin pine stump [ 121.
There are three important methods for obtaining rosin commercially:
(1) distillation of volatile turpentine from oleoresin exuded from the wound
of living pine trees to obtain gum rosin; (2) solvent extraction of pine stump
wood along with the removal of the turpentine by steam distillation to obtain
wood rosin; and (3) separation of tall oil to obtain tall oil rosin. However,
since 1949 new facilities designed to recover and fractionate practically all
of the crude tall oil in the pulp industry has made tall oil rosin the major
source of rosin [ 12, 131. The name ‘tall oil’ originates from the Swedish
word ‘tall’ which means pine [ 111.
Traditionally, pine wood is the major source of rosin. The pulp and
paper industry is the largest consumer of pine. Their principal method of
separating extractives from the cellulose fibers of wood is by the sulfate or
Kraft process. Generally this process involves the digestion of wood chips
with sodium hydroxide and sodium sulfide under heat and pressure to form
a highly alkaline solution of sodium salts. After a few washings, the fatty
acid and rosin soaps are separated by skimming. Acidification with sulfuric
acid yields a crude mixture of rosin and fatty acid which then undergoes
acid refining or fractionation [ll-141. The acid relining process removes
odor bodies, such as mercaptans and disulfides, and a large portion of colored
materials. It also dimerizes a portion of the rosin and fatty acids, which
eliminates the crystallizing tendencies of the rosin. In the fractionation process,
distillation temperatures of up to 280 “C [ 151 are used. Two major fractions
are obtained: a fatty acid fraction and a rosin fraction. The reported contents
of rosin fractions in softwood, such as slash pine, are 47% [ 161. Very little
rosin is found in hardwood such as birch [ 111. The rosin obtained from the
above process is called tall oil rosin.
At present, the US gum rosin industry is almost extinct since it is a
labor-intensive industry. As a result, since the 196Os, the US imports a great
deal of its gum rosin primarily from Portugal, Brazil and the People’s Republic
of China. Similarly, only one company in the US, Hercules Inc., produced
wood rosin by 1989 [lo]. It should be noted that all rosin sold in the US
must be described by reference to the US Standards, and subject to grading
(ASTM D-509-70) based on its color or appearance (e.g., grade X, WW, WG,
N and others) prior to such sale.

3 Chemical composition of rosin

The principal components of rosin are resin acids, which are mono-
carboxylic acids of allqd hydro-phenanthrene. Resin acids occur in pine in
a number of isomeric forms having the molecular formula C20H3002. Sixteen
resin acids have been identified by gas chromatography-mass spectrometry
(GC-MS) in Finnish tall oil rosin [ 151. The composition of rosin also depends
on where the pine trees are grown and the separation process used [ 171.
142

Generally, the resin acids can be divided into two subgroups: the pimaric-
type acid, characterized by both methyl and vinyl substituents at the C-7
position; and the abietic-type acid, bearing only a single isopropyl group at
this position as shown in 1. The numbering system used in this review
adheres to the rules for naming resin acids as derivatives of abietane and
pimarane from Klyne [ 181.

15
(1)

The structures of some of the important resin acid isomers including


abietic (2) levopimaric (3), palustric (4), neoabietic (5), dehydroabietic (6),
dihydroabietic (7) and tetrahydroabietic (8) acids are shown in Table 1.
The distribution of these isomers found in pine trees varies depending
on their geographic location and, perhaps equally important, on the thermal
history of the rosin [ 15, 19, 201. WG grade gum rosin has total abietic-type
acids of 60-65% [ 211, but Finnish tall oil rosin has only 39-47% [ 151. It
has been shown [ 151 that resin acid composition between crude tall oils
and tall oil rosin changes substantially during the high-temperature frac-
tionation process. One of the most obvious changes is the large increase in
the more stable dehydroabietic acid from 17.S18.2% to 30.2-37.6% and
a corresponding decrease in abietic acid from 52.9-54.9% to 32.5-39.8%
[ 11, 201. Furthermore, some of the resin acids originally not present in
crude tall oil are found after distillation [ 111. This is mainly due to the
decomposition, dehydrogenation, isomerization and disproportionation re-

TABLE 1
Structures of major resin acid isomers

Isomers Double-bond positions

abietic acid (2) C-7, C-8 and C-9, C-14


levopimaric acid (3) C-6, C-7 and C-8, C-14
paIustric acid (4) C-8, C-9 and C-13, C-14
neoabietic acid (6) C-S, C-14 and C-7, C-18
dehydroabietic acid (6) aromatic in the ring bearing Isopropyl group
dihydroabietic acid (7) one double bond among C-9, C-14, C-13
and C-S (three probable)
tetrahydroabietic acid (8) none
143

TABLE 2
Composition of tall oil [22], gum [ 191 and wood [23] rosin

Resin acids Melting Tall oil GUtl Wood Ref.


point rosin rosin rosin
(“Cl (wt.%> (wt.%) (wt.%>

abietic 172-175 27-37 18-33 39 24


neoabietic 167-169 4-5 14-16 10 24
palustric 162-167 10-14 6-35 12 25
pimaric 217-219 l-2 5-6 7 26, 27
isopimaric 160 8-15 18 20 28
dehydroabietic 171-172 29 6-9 8 29
levopimaric 150-152 - 30

actions taking place during fractionation. The melting points of resin acids
and the compositional data for different rosins are given in Table 2.

4 Isomerization of resin acids

The pimaric-type acids, including pimaric and isopimaric acids, contain


non-conjugated double bonds. They are more stable to heat than the abietic-
type resin acids including levopimaric, palustric, abietic and neoabietic acids
which contain two conjugated double bonds. Migration of the whole diene
system to an adjacent position should be expected when these resin acids
are heated to above their melting point.

4.1 IscmLerizatim of abietic-type resin acids


The establishment of the isomerization among the various resin acids
was a result of much work during the 1940s and 1950s with the use of
improved separation techniques, such as chromatographic and ultraviolet
absorption analysis [ 311. This led to the identification of neoabietic acid
[24], isopimaric acid [27] and palustric acid [25].

4.1.1 Acid-catalyzed isomerkatim of resin acids


In the presence of strong acid catalysts, such as hydrochloric acid and
p-toluenesulfonic acid, levopimaric acid [32] and neoabietic acid [ 331 were
found to isomerize rapidly to reaction mixtures which contain over 90% of
abietic acid and a small amount of starting material In both cases it was
demonstrated [ 32, 331 that the isomerizations were catalyzed by the solvated
protons in anhydrous ethanol with the same activation energy (21.4 + 0.4
kcal mol- ‘) involved, being llrst order with respect to the catalyst and the
resin acids. It was rationalized [33] that the rate-controlling step in the
isomerization was the addition of solvated proton to either levopimaric or
neoabietic acid to form a resonance hybrid of tertiary carbonium ions 9 and
10.
144

COOH
(9) (10)
It is to be noted that 9 can lose its charge by expulsion of a proton
either at C-6 or C-18 to regenerate levopimaric acid and neoabietic acid
[25]. On the other hand, 10 can lose its charge by expelling a proton from
C-9 to form the abietic acid. However, if 10 loses its charge by expelling
a proton from C-13, the result would be a new unknown resin acid. Later,
with the use of partition chromatography it was discovered [25] that this
unknown resin acid was palustric acid 4.
The formation of resonance hybrid 9 and 10 proposed by Ritchie and
McBurney [32, 331 implies that all four conjugated resin acids are inter-
exchangeable upon acid-catalyzed isomerization. This was demonstrated by
Baldwin et al. 1341, Takeda et al. [35] and Enoki [36], i.e. in the presence
of acid catalyst, all four resin acids eventually reached the same equilibrium
distribution of resin acids (92-95% abietic, 3-5% palustric and 2-3% neo-
abietic acids) in dilute (0.7-2.03/o) ethanolic solutions at room temperature.
In all the isomerization studies, no detectable amount of levopimaric acid
was found in the final equilibrium mixture.
Based upon the above information, a generalized reaction path may be
drawn [33] and shown in Scheme 1.
It can be seen that levopimaric, palustric, neoabietic and abietic acids
can all yield the same resonance carbonium ions upon attack of a proton
at appropriate locations. The subsequent or concurrent loss of a proton from
appropriate positions (Y to the electron-deficient carbon atoms can lead to
the formation of all four acids in question. In the case of palustric acid,
protonation must occur from the (Y side exclusively at C-13 or else abietic
and neoabietic acid would be formed [35].
The relative instability of a homoannular diene system, coupled with the
lack of full substitution at C-6, would explain the considerable reactivity of
levopimaric acid. The heteroannular distribution of the diene system in abietic
acid may well be responsible for its stability [37]. It was determined [ 361
from NMR spectra that the rate of isomerization of levopimaric acid to
palustric acid was faster than that of neoabietic acid to levopimaric acid,
and that both levopimaric and neoabietic acids isometied to abietic acid
through palustric acid [ 361.

4.1.2 Therm& isornerization of resin acicik


It has been shown [38-401 that levopimaric acid, originally present in
commercial pine oleoresins, is completely isomerized to other resin acids
during gum distillation and that the major end-product is abietic acid. In
145

Neoabietlc acid (5) Levopimaric acid (3) Palustrlc acid (4)

Abietic acid (2)

I'kOOH

Scheme 1.

the mid-1950s to early 1960s there was a substantial increase in tall oil
and gum rosin production utilizing high-temperature distillation and frac-
tionation [20]. Thermal isomerization of rosin represents a practical interest
because it defines the product distribution.
When levopimaric acid was heated at 155 “C, just above its melting
point, for 4 h, it was completely isomerized [30]. After 221 h isomerization,
the product contained 7% pahrstric, 86% abietic and 8% neoabietic acids.
When pahrstric acid was heated at 170 “C for 24 h, the product contained
13% pahrstric, 80% abietic and 7% neoabietic acids [41]. Similarly, when
neoabietic acid was heated at 200 “C for 72 h, the product approached an
equilibrium mixture of 13% pahrstric acid, 82% abietic acid and 5% neoabietic
acid [42]. While abietic acid was found to be the most dominant product
of the thermal isomerization of levopimaric, neoabietic and pahrstric acids,
Lawrence et al. [35] discovered that abietic acid also isomerized thermally.
At 200 “C abietic acid undergoes a rapid isomerization after 1 h to give an
equilibrium mixture of 81% abietic acid, 14% pahrstric acid and 5% neoabietic
acid [ 351. No levopimaric acid was found in any of the final product equilibrium
146

TABLE 3
Product distribution between acid-catalyzed and thermal isomerization of resin acids

Temperature Abietic acid Palustric acid Neoabietic acid Ref.


(“C) (wt.%) (wt.%) (wt.%)

Acid-catalyzed isowwrizatim
room temp. 92-95 3-5 2-3 34-36

Thermal isomerizatim
155 86 7 8 30
170 80 13 7 41
200 80 11 7 41
200 82 13 5 42
200 81 14 5 35

mixtures. These results of the thermal as well as the acid-catalyzed iso-


merizations of resin acids are summarized in Table 3.
Data shown in Table 3 indicate that the abietic acid content in the acid-
catalyzed isomerization is higher than in the thermal isomerization at different
temperatures. The difference has been attributed [ 351 to the different rates
of reaching dynamic equilibrium during each isomerization. Thus, it is not
surprising to see that in the absence of a catalyst, all four resin acids
isomerized in dilute (0.67%) 99:l chlorofo&ethanol solutions at 25 “C in
the dark under a nitrogen atmosphere to give a final equilibrium mixture of
92% abietic acid, 5% palustric acid and 3% neoabietic acid determined by
specific rotation, NMR and GC-mass spectral techniques [36].
It should be noted that levopimaric acid was shown [30] to exhibit no
observable isomerization at temperatures below its melting point (150-152
“C). However, it completely isomerized at 155 “C to abietic, palustric and
neoabietic acids which possess a higher melting point. It can be reasonably
assumed that at 155 “C, levopimaric acid probably isomerized directly to
other resin acids irreversibly. This may be supported by the fact that, at
155 “C, the abietic acid content in the final product distribution was higher
(86%) than at 170-200 “C (80-829/o) as shown in Table 3.

Carboxylic acid-catalyzed thermal Qornerizatim - In attempts to


determine whether or not the isomerization of resin acids in the absence of
acid catalysts was due to hydrogen ions from the carboxy group, the use
of methyl esters of resin acids provide a convenient way for such investigations.
When methyl levopimarate was isomerized at 155 “C for 18.5 h, it was
found [30] that about 50% methyl levopimarate was still present. At the
same temperature, it requires only 4 h for levopimaric acid to completely
isomerize to other acids. When methyl neoabietate was isomerized at 200
“C, it was found [42] that, after 168 h, 88% of the ester remained unchanged.
This compares to 8 h at the same temperature for neoabietic acid to completely
isomerize to the final equilibrium mixture containing 13% neoabietic acid.
147

Similarly, it was estimated [41] that the isomerization rate of palustric acid
at 200 “C was about 2000-times as fast as that of methyl ester. Consequently,
the extent of isomerization of resin acids was controlled by adding an alkali
such as sodium hydroxide, sodium carbonate or potassium hydroxide to the
crude oleoresin before it was processed into rosin [43].
In dilute chloroform solutions without the presence of a catalyst (in the
dark under nitrogen), levopimaric, palustric and neoabietic acids were found
[36] to begin isomerizing after 6, 7 and 80 h, respectively, and rapidly
approached an equilibrium state. However, in the presence of hydrochloric
acid (7 ppm), the above resin acids were found to isomerize immediately
under similar conditions [ 361.
Clearly, the so-called thermal isomerization of resin acids is actually
ionic in character, being catalyzed by the proton of the carboxy group [351.

Isomerixatim temperature and product distr&utim - The results


of the thermal isomerization of resin acids indicate that the relative stability
of resin acids upon heating is in the order abietic acid> palustric
acid > neoabietic acid > levopimaric acid, and that the iinal product distribution
is dependent on the isomerization temperature. Data shown in Table 4
demonstrate that as the isomerization temperature increases, the abietic acid
content decreases and the palustric acid content increases in the final product
distribution. While this may be due to the sensitivity of the resin acids towards
acid catalysis [30, 34, 371 at higher isomerization temperatures, reactions
other than isomerization may be involved. It was suggested by Radbil and
Kushnir [44] that at higher isomerization temperatures (> 210 “C), other
processes, such as decarboxylation and disproportionation, were involved.
These authors reported that when abietic acid was heated at 260 “C for 3
h, the product consisted of 64.1% abietic, 19.9% palustric, 8.3% neoabietic,
1.1% dihydroabietic and 6.6% dehydroabietic acid.
It has been long established that the isomerizations of abietic [35 1,
palustric [ 411 and levopimaric [30, 411 acids at temperatures between 155
“C and 200 “C are &&-order reactions. However, neoabietic acid did not
give a reaction curve characteristic of a first- or second-order reaction [41,
421. This was attibuted [41] to the back-reaction of neoabietic acid (5)

TABLE 4
Isomerization temperatures and final product distribution [44]

Isomerization Abietic Palustric Neoabietic


temp. (“C) acid acid acid
W.%) ht.%) ht.%)

190 83 11 5 0.20 0.13


200 81 13 5 0.23 0.16
210 78 15 5.5 0.28 0.19
220 76 16 6 0.32 0.21
148

forming the carbonium species (see Scheme l), which tends to progressively
retard the reaction rate, and therefore the isomerization fails to behave in
a first-order kinetic manner. This explanation is also partially reflected in
the data shown in Table 4, which indicate that the concentration of neoabietic
acid remains small and constant (c. 5%) at 190-220 “C.
As can be seen from the data listed in Table 4, the calculated equilibrium
rate constants for the isomerization of abietic acid (I&,) and the formation
of palustric acid (I&J increase as the isomerization temperature increases.
This suggests that at higher temperatures the reverse reaction is important,
as was also suggested from thermodynamic equilibrium considerations [45].
When the data shown in Table 3 were plotted as In K vs. UT according to
the van’t Hoff equation, it was found [44] that both the plots corresponding
to abietic acid isomerization and palustric acid formation have the same
slope. Since the formation of neoabietic acid remains constant, a similar
slope indicates that the thermal isomerization of abietic acid is reversible.

Formation of levopinuwic acid during thermal is- kation -


Fomin et al. [46] have studied the thermal isomerization of abietic acid at
180 “C and 200 “C and found that levopimaric acid is present in the product
mixture throughout the isomerization. The 200 “C isomerization data are
shown in Table 5
Data shown in Table 5 contradict the results of all previous investigations
which indicate that levopimaric acid is not present in the equilibrium mixtures.
Levopimaric acid, for example, was found [30] to completely isomerize to
other acids at 155 “C after 4 h, and at 200 “C after 0.5 h.
The thermal isomerization data shown in Table 3 were generally acquired
by optical rotation measurements, UV absorption and chromatographic anal-
yses. Data shown in Table 5 were largely dependent on the use of improved
GLC methods. Fomin et al. [46] showed that methyl abietate (obtained

TABLE 5

Isomerization of abietic acid at 200 “C (CO,) [46]

Time (min) Abietate Palustrate Levopimarate Neoabietate


ht.%) (wt.%) (wt.%) (wt.%)

0 100.0
5 92.3 1.0 4.6 2.1
10 82.0 2.4 12.6 2.9
15 78.9 3.8 11.6 4.6
20 80.1 3.1 11.6 4.2
25 76.1 4.3 14.3 4.8
40 74.5 5.3 14.8 4.7
60 74.9 4.1 13.0 4.2
90 78.5 4.7 11.8 4.3
540 77.0 5.4 13.2 4.4
600 78.7 2.8 14.0 4.5
149

through room-temperature reaction with diazomethane [47]) gave a single


peak in the chromatogram before isomerization. When abietic acid was
isomerized followed by methylation, it gave four resolved peaks in the
chromatogram. The peak of levopimarate was identified by UV and IR
spectroscopy. To further confirm the levopimarate peak, these authors added
maleic anhydride to the methyl esters of isomerized samples at room tem-
perature for 24 h and found that the peak for levopimarate decreased with
the appearance of an additional new peak. This new peak was attributed to
the methyl maleopimarate. Since only levopimaric acid reacts with maleic
anhydride at room temperature [48, 491, this confirms that the original peak
in the GLC chromatogram was methyl levopimarate.
It is difficult to judge the results shown in Table 5. Brooks et al. [50]
showed that methyl levopimarate selectively isomerized to palustrate as a
result of acid catalysis in a GLC column (Versamide 900). This caused an
unresolved levopimarate and palustrate peak in the chromatogram. It is
possible that either the palustrate or neoabietate of the isomerized products
shown in the work of Fomin et al. [ 461 may have selectively isomerized to
levopimarate in the GLC column. It is also possible that the palustric and
neoabietic acids of the isomerization products may have isomerized to
levopimarate selectively during the methylation process using diazomethane.
More work is needed in this area.

4.2 Isomerization of pimaric-type resin acids


An important explanation of the relative stability of pimaric-type acids
toward thermal isomerization is that the 18,19-double-bond migration is
prevented by tetrasubstitution at C-7 [51]. In earlier studies, it was reported
[20] that pimaric and isopimaric acids remain unchanged by the processing
of gum and tall oil rosins at temperatures normally encountered in tall oil
distillation. Although it has been shown [ 151, during the distillation of crude
Finnish tall oil rosin at 280 “C, that pimaric-type acids also undergo iso-
merization, the extent was small (l-2%).

5 Other reactions of resin acids

Since gum and tall oil rosins come from living trees, the different methods
involved in obtaining these products account for most of the differences in
their composition. In the case of gum rosin, these changes involve the
isomerization of levopimaric acid to palustric, neoabietic and abietic acids.
The resin acids from which tall oil rosin is obtained are exposed to oxidation
and high temperatures. The former is responsible for the increased con-
centrations of the dehydroabietic acid (6) and the latter is responsible for
the increase of the abietic acid content in rosin. Since at high reaction
temperatures, other reactions such as decarboxylation, disproportionation
and polymerization can also occur, it will be helpful to discuss these reactions.
150

5.1 Oxidation and disproportionation


It is known that when rosin is exposed to air over a prolonged period
of time, the exposed surface darkens. When pure abietic acid is exposed to
air, ultraviolet absorption characteristics indicate the disappearance of the
double-bond chromophore and the saturation of the conjugated double-bond
system by oxygen [ 121. This is because abietic-type acids with conjugated
double-bond systems readily take up oxygen. In studies of the photosensitized
oxidation of pine gum rosin, Schuller et al. [52, 531 showed that levopimaric
acid, palustric acid and neoabietic acid all gave transannular peroxide with
an oxygen content in the neighborhood of 3%. Undoubtedly, the oxidation
of rosin dienes takes place readily at high temperatures.
As was described earlier, the fractionation of tall oil rosin at high
temperatures results in the formation of large amounts of dehydroabietic
acid and a significant decrease in the abietic acid content. Because of greater
stability due to the formation of a rigid aromatic ring system in dehydroabietic
acid, the oxidation products of rosin have been the subject of much inves-
tigation. In the thermal dimerization of tall oil, gum and wood rosin, Par-kin
et al. [54] found that the dehydroabietic acid content increased to 28.3%
(from 6-9%, see Table 2) when gum rosin was heated at 230 “C for 16 h
in air. When the reaction was conducted at 275 “C for 29.4 h, the dehydroabietic
acid content further increased to 49.2%. In the presence of an oxidizing
agent such as iodine, abietic acid was also shown to oxidize and give a
product mixture consisting of dehydroabietic acid (40%) and dihydroabietic
acid (7) (20%) (at temperatures of 160-170 “C [55]).
Hence, an important consequence of the oxidation of resin acid is the
formation of dehydroabietic acid. It has been shown [ 36 1that dilute chloroform
solutions of homoannular diene, such as levopimaric and palustric acids,
isomerize in the dark under aeration to dehydroabietic acid in about 50%
yield. The formation of dehydroabietic acid presumably occurs by hydrogen
abstraction from palustric or levopimaric acids [36, 561. Although hetero-
annular dienes such as abietic or neoabietic acids do not form dehydroabietic
acid under similar conditions, they give unidentified oxygen adducts [36].
Because of the transfer of part of the hydrogen from other resin acids,
the reaction leading to the formation of dehydroabietic acid frequently has
been described as a disproportionation. When the abstracted hydrogen is
consumed by the pimaric-type acids present in the rosin, it leads to the
formation of stable dihydropimaric acid [ 121. Similarly, in the study of the
conditions affecting the dimerization of pure resin acids and rosin, part of
the by-product was found [57] to be dehydro- and dihydro-abietic acids.

5.2 Hydrogenation and dehydrogenation


Hydrogenation and dehydrogenation are important methods for modifying
the resin acids in rosin to reduce its air-oxidation tendencies. The former
involves the addition of hydrogen to double bonds in the resin acid, typically
catalyzed by nickel compounds [58, 591 or noble metals [60], to form
saturated ring structures. The latter consists in the removal of two atoms
151

of hydrogen from the abietic-type acids and the rearrangement of the double-
bond system, catalyzed by iodine [55] to form an aromatic nucleus such as
dehydroabietic acid. Both methods are typically carried out above 250 “C.
In ordinary hydrogenation reactions of rosin, it is generally diflicult to
completely hydrogenate the rosin without causing decarboxylation and the
formation of dehydroabietic acid [58, 611. This is because endocyclic double
bonds are generally more difficult to hydrogenate than exocyclic double
bonds [581. However, since the residual double bond is highly resistant to
air oxidation, much of the desired stability is obtained even by hydrogenation
to the dihydro stage. Analysis of the various types of resin acids in highly
hydrogenated rosin, such as Foral@ AX (Hercules Inc.), showed the complete
absence of abietic-type acids with the presence of c. l-14% dihydroabietic
acids and 66-80% tetrahydroabietic acids [12]. Because the hydrogenated
product is so resistant to oxidation, it is used as a tackifier and plasticizer
in pressure-sensitive adhesives [ 621.
On the other hand, dehydrogenation of abietic acid was studied at 250
“C in the presence of platinum catalysts and found to lead to disproportionation
to dehydroabietic and dihydroabietic acids plus a small amount of tetra-
hydroabietic acid [ 611. At an even higher temperature (275 “C), the main
product of the reaction was dehydroabietic acid [61]. A similar conclusion
was reached when palladium was used as a catalyst at 230 “C [63]. Since
the formation of dehydroabietic acid requires the loss of hydrogens from
abietic acid, it should come as no surprise that hydrogenation and dehy-
drogenation in abietic acid occur through disproportionation [37].

5.3 Di merization and polymm-izatiim


The dimerization and polymerization of rosin has been subjected to
investigations from the 1940s to the 1960s. Much work [64-661 has been
centered around the reaction conditions necessary for ‘dimerizing’ or ‘po-
lymerizing’ the resin acids. This was due to the increasing commercial use
of dimerized rosin in adhesives, lacquers, varnishes and printing inks. These
applications generally require rosin molecules to have good solubility and
compatibility with high Tg values. A variety of reactions were investigated
for optimizing the yield of dimerized rosin. Generally, mineral acids are
required to dimerize the tall oil, wood and gum rosin in 50-60% yield [64,
671. However, weaker acids such as a sulfonic acid were also found by
Olechowski and Lawson [68] to yield 32% dimer.
In the presence of acid catalysts such as sulfuric acid [ 651, boron
trifluoride, phosphoric acid or aluminum chloride, the dimerization products
of rosin can be complicated [ 121. One should not forget that isomerization
reactions can also occur during the diierization process. Su and Han 1561
have studied the polymerization process of resin acids in the presence of
sulfuric acid in chloroform solution. These authors found that the rate of
isomerization of abietic-type acids is faster than the rate of polymerization,
and that isomerization equilibrium among abietic, palustric and neoabietic
acids remains the same throughout the reaction. The reaction consisted of
152

a fast isomerization step and a rate-determining dimerization step in agreement


with earlier results obtained by Sinclair et al. 1671.
Sinclair et al. [57] dimerized abietic acid, levopimaric acid and methyl
abietate in the presence of a mineral acid catalyst, such as sulfuric acid, at
44 “C using chloroform as the solvent. These authors found [57, 671 that
the rate of dimer formation was dependent on the concentration of the
monomer and the acid catalyst:

fast
AA+ (H+) ‘fast (AA-H+)
slow
y (AA-AA)
fast
--H; (dim--)

Rate=k[AA](AA-Hf]

where AA denotes abietic acid.


The product mixture consisted of 80% dimer with a loss of 20% of the
original acid functionality. When gum rosin was used, about 60% dimer was
obtained with a loss of about 25% adid functionality. Furthermore, dehy-
droabietic (6) and dihydroabietic (7) acids have been identified among the
by-products. The loss of acid functionality suggests that either ester formation
or decarboxylation may have occurred. The presence of dehydroabietic and
dihydroabietic acids is probably the result of disproportionation [57] and,
more importantly, oxidation [ 561 of the resin acids. A general chemical
kinetic scheme consistent with the kinetic data has been proposed 1671 and
is shown in Scheme 2.
Thus the structures of dimerized resin acids can be mixtures of various
polymeric forms depending on the reaction conditions. Parkin et al. [54]
studied the thermal dimerization of gum rosin under various conditions.
These authors isolated c. l&30% of a polymeric material. For example,
when gum rosin was heated at 300 “C for 10 min, it yielded 16% isolated
polymeric residue and dehydroabietic acid. The polymeric residue had an
acid number equivalent weight of 645, a saponification equivalent weight of
408 and a molecular weight (determined by vapor pressure osmometry) of
654. The above information suggests that the polymerized materials were
essentially a mixture of monobasic dimer acids. This was further supported
by the strong peaks at 1700 cm-’ and 1730 cm-’ in the IR spectrum, which
indicated the presence of carboxylic acid and ester groups. Presumably the
reaction was the result of the addition of the rosin carboxylic acid across
one of the double bonds of another resin acid, such as abietic acid, to form
ester dimers. In the dimerization of abietic acid or levopimaric acid, Sinclair
et al. 1671 determined by MS and GC methods that there were at least 12
structures. The most important dimerization conditions were found to be the
solvent (CHCla) employed, the acid strength of the catalyst and the
CHCla:H2S0,&4 ratio. Under the most favorable dimerization conditions, the
153

COOH

Scheme 2.

resin-acid dimer was limited to BO-90% yield by side-reactions. The remaining


tmdimerized materials were inert to further dimerization because they lacked
conjugated diene structures.
It is clear from the above information that the important changes in
composition during the polymerization of rosin are the decrease in carboxy
functionality, the disappearance of a large portion of the abietic-type acid
and the increase in dimer and dehydroabietic acid content. These changes
make the dimerized rosin mixture less susceptible to attack by atmospheric
oxygen so that the desired stabilization of the abietic-type acids is produced.
154

A polymerized rosin has the further advantage of requiring less maleic


anhydride [ 12 1, or phenol-formaldehyde condensate, to yield maleic-modified
or phenolic-modified rosin esters to achieve a given hardness and viscosity.
Consequently, it also requires less polyol for esterification than the unmodified
rosin to achieve low acid number and high softening point.

6 Modifications of rosin for coating applications

Rosin has been recognized to have excellent solubility and compatibility


with a variety of other synthetic resins. This is because resin acids have a
hydrophobic skeleton in combination with hydrophilic carboxy groups. It is
advantageous to modify these resin acids for many coatings applications.
For example, when rosin or its adduct were incorporated into alkyds and
varnishes for coatings applications, faster drying and better chemical resistance
have been reported [60, 69-711. The faster drying property is primarily due
to the higher Tg value of rosin and its adducts in coatings that impart a
lacquer-dry characteristic. The versatility of rosin applications has been
summarized in the report by Hedrick et al. [ 721.

6.1 Carboxy functionality and esterjficatim


Esterification is an important part of rosin modification. The carboxy
group of resin acids is bonded to a tertiary carbon and is therefore sterically
hindered. Furthermore, three rings in the resin acids are fused together to
form relatively rigid structures which make rosin possess a relatively high
melting point (see Table 1). Consequently, high temperatures are normally
required for esterification to occur. Ester gum finds its use in adhesive
applications and is obtained by modifying rosin with glycerol or pentaerythritol
at temperatures in the range 200-240 “C. Even higher reaction temperatures
(260-275 “C) and longer reaction times are required for some rosin ester
preparations. Much information may be found in the work by Ellis [ 73 1.
As would be expected, the high-temperature esterification process results
in a decrease in abietic acid and an increase in dehydroabietic acid due to
isomerization and oxidation. However, this is not the only consequence. In
an attempt to prepare a Diels-Alder adduct of abietic acid and fumaric acid
(200 “C for 2.5 h), Halbrook and Lawrence [74] noted a 2% weight loss
based on abietic acid. These authors found that this was mostly due to
decarboxylation as evidenced by the carbon dioxide trapped as barium
carbonate from the exhaust gas. This leads to the formation of abietenes
or rosin oil [ 751, which generally results in products with a lower softening
point. The decarboxylation products at least explain, in part, the yellow,
gummy, neutral materials which often accumulate in the condenser during
a typical rosin ester preparation. In comparison to Su and Han’s finding
[ 561 that the isomerization equilibrium among abietic, palustric and neoabietic
acids was not disrupted throughout the acid-catalyzed polymerization of resin
acids, the presence of esters during esterification was also found [76] to
155

have no effect on the isomer redistribution of resin acids during the high-
temperature (290 “C) esteriIlcation of rosin.
In order to facilitate the esterification, Arimoto and Zinkel [77, 781
showed that a rosin ester with an acid number of 5.1 can be prepared at
40 “C by reacting a rosin acid quaternary ammonium salt, such as tetra-
butylammonium dehydroabietate, with all@ or alkenyl halides. More com-
monly, catalysts such as zinc dust or boric acid are used [79]. On the other
hand, the carboxy group in a resin acid can be converted to a variety of
other functional groups which are less sterically hindered. For example, the
carboxy group can be converted into a hydroxy group by hydrogenolysis
[ 75, 801 or by reaction with ethylene oxide [ 8 11. Blazhko [82 ] showed that
a flexible transparent film was obtained from products of reactions of
dehydroabietic acid or disproportionated resin acids with glycidyl ether.

6.2 Condensatti with dienophiles


One of the important synthetic reactions of rosin is the formation of
an adduct with dienophiles such as maleic anhydride or fumaric acid, via
the Diels-Alder reaction at temperatures of 180-210 “C. Portions of the
rosin are converted to tribasic acids capable of being reacting with polyols
to form high molecular weight rosin esters. Generally, the optimum Diels-Alder
reaction conditions for rosin are temperatures of c. 180-190 “C for 2 h
[21]. Above 200 “C, evaporation of maleic anhydride, decarboxylation and
darkening of the product were reported [21].

6.2.1 With makeic anhydride


In a Diels-Alder reaction between resin acids and maleic anhydride,
abietic-type acids are first isomerized [21, 831 to levopimaric acid before
forming the maleopimaric acid (Scheme 3).
It was reported by Bacon and Ruzicka [48], and by Wienhaus and
Sandermann [49], that with abietic acid maleic anhydride addition products
formed only at temperatures above 100 “C, and were identical to those
obtained in quantitative yield from levopimaric acid in benzene at room
temperature. Similarly, palustric acid and neoabietic acid were shown [37]
to be inert toward maleic anhydride at room temperature, but reacted at
150-200 “C to yield maleopimaric acid. It has been assumed by Fieser and
Campbell [29] that as the trace quantities of levopimaric acid present in
tall oil rosin react with maleic anhydride, the equilibrium is displaced until
eventually all of the abietic-type acids capable of isomerizing to levopimaric
acid are converted to the Diels-Alder adduct. Thus, when commercial rosin
is heated together with maleic anhydride above 180 “C, it has been commonly
accepted [21] that the reaction is driven to completion due to the continuous
supply of levopimaric acid (originally present in trace quantities) by maintaining
an active isomerization of the other abietic-type acids. Of course the assumption
made by Fieser and Campbell [29] may not be entirely correct if the data
shown in Table 5 are true. It is also possible that maleic anhydride may
induce resin acid isomerization to form dipolar or zwitterionic intermediates
156

Abletic acid
Neoabletic acld

4
COOH
Scheme 3.

based on the known [84] mechanism of the Diels-Alder reaction. More work
is needed in this area.
It should be noted that pure maleopimaric acid is converted to the
monoether ester in alcoholic KOH solution and may be titrated as a dicarboxylic
acid. This shows an acid number of 280, but this is not due to the third
carboxylic group, which is sterically protected, as reported by Shukla et al.
]85].
Generally in Diels-Alder reactions, the original geometry in the dienophile
is retained. When a cyclic diene and a dienophile are allowed to react via
the Die&Alder reaction, the endo isomer is generally favored [84 1. Thus,
the adduct from anthracene and maleic anhydride is the cis-anhydride, and
fumaric acid is a trans-dicarboxylic acid [ 861. Although it has been established
in the reaction between maleic anhydride and furan that the isolated product
was exo, the reaction was kinetically controlled. The initial rate of endo-
adduct formation was 500-times faster than that of the exo adduct but the
former was less stable by 1.9 kcal mol- ’ 1841.
In the Die&Alder reaction between abietic-type acids and maleic an-
hydride, it should be noted that only one isomer (m-p., 225-227 “C; [(Y],
-25O) of the product, maleopimaric acid, has been reported [49, 87, 881.
However, in the studies of the reaction between maleopimaric acid and 3-
phenyl-l-propanol at 240 “C, Molhoek [89] has observed that the anhydride
functionality in maleopimaric acid exhibits a large difference in reactivity
toward- hydroxy groups. The difference in reactivity was attributed [89] to
the suspected presence of isomers of maleopimaric acid, but the author
failed to isolate them. Molhoek’s observations of different anhydride reactivities
were based on measurements of the changes in the peak height at 1782
cm-’ (C=O stretch-vibration) in the IR spectrum, which has long been
157

established [90, 911. However, data shown in this author’s work are largely
incomplete, and it is probably best not to speculate.
Burgstahler and coworkers [92] have shown that the isopropyl-bearing
ring in levopimaric acid. is skewed in such a manner that the /3 face is
shielded by the angular methyl group at C-12, while the CIface is free for
attack by the dienophile. This would indicate a back-side approach in the
Diels-Alder reaction between levopimaric acid and maleic anhydride. This
was also reflected in the low Arrhenius kinetic parameters A (105.‘) and E
(3.6 kcal mol-‘) for the reaction between levopimaric acid and maleic
anhydride, which suggest the absence of a strain [87]. Thus, the anhydride
moiety of maleopimaric acid was shown [93] to be cis to the double bond.
Similarly, with the intention of confirming the a configuration at C-13 in
levopimaric acid, Ayer and McDonald [94] examined the NMR spectra of a
series of 22 compounds derived from maleopimaric acid. From a comparison
of the proton shifts and the observed long-range shielding effect of the olelinic
bond, these authors concluded that the anhydride group is in the endo
position shown in 11. One should not that in 11, the carbonyls in the
anhydride ring shown are oriented differently; one is in the endo position,
and the other is in the exo position.

COOR
is
(11)

6.2.2 With other dimophiles


When pure abietic acid, or a mixture of abietic, neoabietic, palustric
and levopimaric acids in a ratio commonly found for rosin, was reacted with
excess fumaric acid at 200 “C for 2-3 h, fumaropimaric acid in two isomeric
forms was produced [ 74 1. They were crystalline fumaropimaric acid (major,
50-72% yield) and non-crystalline tricarboxylic acid. The crystalline fumaro-
pimaric acid was shown 1931 to have structure 12.

COOH

COOH COOH

(12)
158

The non-crystalline portion of the product was separated and found to


contain 9% maleopimaric acid and 16% non-crystalline fumaropimaric acid
based on the weight of the abietic acid used. The presence of maleopimaric
acid was due [74] to the isomerization of fumaropimaric acid. It was
demonstrated [74, 93, 951 that pure fumaropimaric acid will isomerize to
maleopimaric acid (m.p., 228-229 “C [19]) in 62% yield by heating at its
melting point (280-282 “C) under nitrogen. The non-crystalline form of
fumaropimaric acid was presumably [93 ] the mixture of the additional isomers.
Similarly, reaction of levopimaric acid or rosin with acrylic acid [93],
P-propiolactone [93], methyl acrylate [93] and acrylonitrile [ 961 through the
Diels-Alder reaction all resulted in isomeric products. In the reaction between
levopimaric acid and acrylonitrile, the resulting isomers were isolated and
found [96 ] to possess a large difference in melting point. Adduct 1, crystallized
from alcohol-water, had a melting point of 191.5-l 92 “C. Adduct 2, crystallized
from benzene, had a melting point of 318.9 “C. The two adducts were shown
[93] to be structural isomers differing in the position of the nitrile group.

62.3 Coatings applications


Depending on the sources and application requirements, various amounts
of maleic anhydride or fumaric acid are used to react with rosin. A typical
commercial rosin adduct for coating applications is prepared by incorporating
rosin with c. 4-6 wt.% maleic anhydride [ 19, 2 11. High concentrations of
adducts in the rosin generally lead to high melting points, viscosity and
gelation upon esterification with polyols. Generally, fumaric acid-rosin adducts
have about a 20 “C higher softening point than maleic anhydride-rosin
adducts. For example, Halbrook and Lawrence [95] studied the thermal
reactions of fumaric and maleic anhydride with gum, wood and tall oil rosin
under various conditions. When 100 parts of gum rosin were reacted with
20 parts of fumaric acid for 3 h, the softening point was 138 “C, while the
maleic anhydride-rosin adduct had a softening point of 117 “C. It should
be realized that the maximum molar yield of such adducts produced in typical
rosin is approximately equivalent to the total mole fraction of the abietic-
type acids present.
In recent years, trifunctional anhydrides such as trimellitic anhydride or
oligomers terminated by anhydride groups have been used [97,98] to crosslink
epoxy resins in powder coating compositions. The rosin-maleic adducts or
rosin-fumaric adducts present an interesting possiblity in this area. Penczek
et al. [99] reported that acid anhydrides obtained from rosin were used as
epoxy crosslinkers for curing at 210 “C, with phenol as an accelerator, to
give hlms with good mechanical strength for stove varnishes. Conceivably,
these adducts may be used in many alkyd formulations [70, 731 to replace
phthalic acid for fast drying. Shukla and Vasishtha reported [ 711 that fast-
drying coatings and industrial finishes were obtained with styrenated alkyds
based on maleopimaric acid. The maleopimaric acid was also reported [ 100,
1011 to react with hexamethylenediamine to form a polyamideimide with
excellent thermal stability. Similarly, coatings based on unsaturated polyesters
159

derived from rosin modification with P-propiolactone were shown [ 1021 to


have excellent resistance to water, base and acid. It should be noted that,
due to the shortage of phthalic anhydride during World War II, the maleinization
and styrenation of rosin was developed and subsequently produced in tank
car volumes [103].

6.3 Condensation with formaldehyde


While the Die&Alder reaction of abietic-type acids with maleic anhydride
produces tribasic acid, a considerable effort was also made in the early 1960s
to make hydroxy-functional resin acids. This can be accomplished by con-
densing resin acids with formaldehyde.
The reaction of formaldehyde with unsaturated hydrocarbons to produce
unsaturated alcohols is well known [104]. However, the condensation of
formaldehyde and resin acids usually gives a complex mixture of hydroxy-
methyl-substituted resin acids. Parkin et al. [ 105, 1061 reacted levopimaric
acid with paraformaldehyde under reflux conditions (100 “C) to yield the
cyclic ether of the levopimaric acid-formaldehyde adduct 13 shown in
Scheme 4.
Treatment of adduct 13 with dilute mineral acid gave 6-hydroxymethy-
labietic acid (14), quantitatively. To increase the thermal stability, the 6-
hydroxymethylabietic acid can be hydrogenated by palladium on carbon to
give 6-hydroxymethyldihydroabietic acid (15), or by copper-chromite re-
duction to give 6-hydroxymethyltetrahydroabietinol (16) [ 1051.
The reaction between levopimaric acid and formaldehyde is
time-temperature-sensitive as a result of competition with the isomerization
of levopimaric acid, [30], condensation with formaldehyde and further con-
version of the adduct to 6-hydroxymethylabietic acid [ 1061. However, in the
reaction between abietic-type resin acids and formaldehyde in the presence
of acetic acid, the products were shown [107] to be mixtures of various
methylol, dimethylol and dioxan derivatives. The heterogeneous nature of
the various condensation products suggest that reactions other than
Die&Alder are involved in the acid-catalyzed condensation reactions.
The paraformaldehydes commonly used in the trade have a degree of
polymerization ranging from 10 to 100 and contain up to 7% water. At
temperatures above 100 “C, paraformaldehyde depolymerizes to release
methylene glycol rather than formaldehyde. This is because in aqueous media,
the dissociation constant of methylene glycol to formaldehyde is 1.4 X lo-‘*
[ 108 1. In acidic media, such as resin acids, the formation of the highly
reactive hydroxyalkylating [ +CHa-OH] species occurs:

HO-(CHa-O),-H+H20 L HO-(CH2-0),-1-H+HO-CH2-OH
HO-CHa-OH = [ +CH,OH] +H,O

One can expect that this highly reactive intermediate is capable of


alkylating resin acids to form a mixture of hydroxymethyl- and dihydroxy-
methyl-substituted resin acids 14, 17 and 18 with small amounts (5%) of
160

I3O.C
20-30
MIN
CH,O ,,& 1 Hs/Pd

COOli
(13)
4khJ
COOii (15)

Ii,/ch-00,
CH,OH

&d/
CHIOH

(16)
c h,Oli

Scheme 4.

unreacted resin acids [ 60,109]. The reactions involved, with either a resonance-
stabilized carbonium ion attack or a formation of a six-member cyclic
intermediate via an ‘ene’-type reaction mechanism, are shown in Scheme 5
[ 1091.
Of course the condensation of resin acids with formaldehyde is not the
only method to produce hydroxymethyl derivatives. A so-called ‘0x0’ process
can also be used to increase the functionality of resin acids. This process
is generally used in the production of alcohols and aldehydes by passing
olefin hydrocarbon vapors over cobalt catalysts in the presence of CO and
H2 under high pressure. Levering and Glasebrook [ 110 ] used the ‘0x0’ process
to prepare hydroxymethylated rosin with 89.5% yield. The actual product
composition of oxonated rosin was studied by Rohde and Hedrick [ 1111.
These authors used gum rosin with CoC03 as a catalyst to produce a mixture
of products consisting of the saturated monohydroxy-/dihydroxy-addition
products 19,20 and 21 in a molar ratio of 44:21:35, together with unoxonated
material.
161

COOH

COOH
4 COOH H
(18)
Scheme 5.

;yl’l
( .....
COOH
R*
R'

where (19) R’ = H, R2 = CHzOH


(20) R’=CH,OH, R2=H
(21) R’, R2 =CH20H

The unoxonated material was presumably due to the unreactive resin


acids, e.g., dehydro- and tetrahydro-abietic acids [ 1121. It should be realized
that, under oxonation, the pimaric-type acids are ako capable of converting
to hydroxymethylated resin acids [ 110, 1111.
The resulting polyhydroxy-functional rosin can be reacted with a variety
of fatty acids or drying oils in an wd formulation to obtain improved air-
drying properties for varnishes [ 113 1. These hydroxy-functional rosins were
also shown [81] to control the crosslink density in film-forming reactions
with isocyanates to form urethane foams [ 811.
6.4 Condematicm with pherwlic resin
One important area of rosin modification is phenolic-modified rosin
esters, which are frequently used in printing inks and varnishes. These
materials offer improved drying, gloss and wetlability. Generally, two processes
are used to prepare phenolic-modified rosin esters.
In the direct condensation process, the reaction is brought about by
reacting rosin ester with alkaline-condensed reactive phenolic resin at tem-
peratures of 250-270 “C. In this type of reaction, it was at one time suggested
by Krumbhaar [ 1141 that there was no chemical linkage between rosin and
phenolic resin and that self-condensation of the phenolic portion was re-
sponsible for the increased softening point. Krumbhaar’s comments reflected
the uncertainty during the late 1940s on the reaction between phenolic
derivatives and drying oils. However, the reaction of o-hydroxymethyl phenols
with unsaturated compounds such as oleic acid were found [ 1151 to occur
at 180 “C and resulted in the formation of the chroman ring 23, as shown
in Scheme 6.
The quinone methide 22, formed from either a hydroxy compound or
hydroxymethylphenols, is highly reactive towards electrophiles as well as
nucleophiles. The formation of the chroman ring has in the past been considered
important in the modification of resole with drying oils for protective coatings.
The rigidity of the molecular structure was assumed to be responsible for
the increase in melting point when rosin is combined with a phenolic resin
[ 1161. However, one should note that in order for rosin to form a chroman
ring with reactive phenols, two ‘active methylene hydrogens’ from resin acids
are needed. This appears to be possible only with pimaric-type acids. Since

(23)

Scheme 6.
163

the pimaric-type acid concentration in rosin is usually low, and the con-
centration of phenol derivatives employed in the mo~cation of rosin esters
is normally less than 15%, it appears unlikely that the formation of the
chroman ring would play a major role in the reactions of phenolic-modified
rosin esters.
The other process for the preparation of phenolic-modified rosin ester
involves condensing rosin (in which a portion is a maleic anhydride adduct)
with p~afo~~dehyde and phenol at loo-130 “C. Following the addition
of polyols, the reaction temperature is increased to 260-270 “C for ester-
iflcation to occur and is maintained until the desired properties are obtained.
One can see that, in this process, the reactions can be even more complex
than in the one previously described. The highly reactive hydroxyalkylating
agent not only reacts with resin acids to form the hy~o~e~yl-substituted
resin acids 17 and 18, but also reacts with phenol derivatives to form the
reactive benzylic carbonium ion 24 as shown in Scheme 7 [ 1081.
In addition to the phenolic condensation, the benzylic carbonium ions
are very likely to alkylate the abietic-type acids which are normahy present
in high concentration. As a result, the structure for phenol&modified rosin
esters, having a methylene linkage such as 25, is generally accepted (1171.
One should realize that phenolic condensation by itself is a complex phe-

+ H+
it

(20

AUIETIC
ACID

Scheme 7.
164

nomenon since it is pH-, catalyst- and time-dependent. Although the general


kinetics are relatively well-researched, the actual constitution of the hy-
droxyalkylating agent in the alkali-catalyzed reaction is not yet fully understood
[ 1081. Therefore, depending on the process and reaction conditions used,
the properties of phenolic-modihed rosin esters can be significantly different.

7 Conclusion

For centuries, rosin has been proven to be a highly versatile raw material
for use in coatings. The unique properties of rosin arise from its hydrophobic
skeleton and its hydrophilic carboxy groups which contribute to its excellent
solubility and compatibility with a variety of other synthetic resins. Modi-
fications of rosin such as isometiation, disproportionation and hydrogenation
to produce new materials can be advantageously used in meeting the demanding
applications of today’s coatings.
This review provides fundamental information for a basic understanding
of rosin chemistry. With appropriate modification of rosin, many different
multifunctional rosin derivatives can be obtained. A renewable resource is
therefore available for creative individuals to develop new resins for coatings
applications.

Acknowledgements

The author wishes to express his gratitude to Profs. Frank N. Jones.and


Xiao-Yin Hong, and Dr Robert Solimeno for their comments and criticism
of this manuscript. Permission for publication of this manuscript by Ashland
Chemical, Inc. is appreciated.

References

1 J. Michelman, Violin Varnish, Micheiman, Cincinnati, OH, 1946.


2 J. Nagyvary, C & EN, (May 23, 1988) 24.
3 J. J. MattieIlo, Protective and Decorative Coatings, Vol. 1, Wiley, New York, 1941,
p. 3.
4 World Oil, February 1990.
5 M. Brower, Cool Energy: TheRenewable Solutionto GlobalWarming, Union of Concerned
Scientists, Cambridge, MA, 1990.
6 Chicago Society for Coatings Technology, Renewable Resources Subcommittee, J. Coat.
Technol., 53 (1981) 45.
7 S. Borman, C & EN, (Sept. 10, 1990) 19.
8 H. I. Enos, Cum-erUand Potential Use of Rosin, presented at Paper and Textile Chemistry
Division Meeting, Am. Chem. Sot., March 23, 1977.
9 Naval Stores Annual Summary, Crop Reporting Board, StatisticaI Reporting Services, US
Department of Agriculture, May, 1979.
10 Chemical Economics Handbook - SRI International, October 1989.
165

11 E. Sjostrom, Wood Chemistry: Fundamentals and Applications, Academic Press, New


York, 1981.
12 H. I. Enos, G. C. Harris and G. W. Hedrick, in H. F. Mark, J. J. McKetta Jr. and D. F.
Othmer (eds.), Encyclopedia of Chemical Technology, 2nd edn., Wiley, New York, 1968,
Vol. 17, p. 475.
13 J. C. Weaver, in H. F. Mark, D. F. Othmer, C. G. Overberger and G. T. Seaborg (eds.),
Encyclopedia of Chemical Technalom, 3rd edn., Wiley, New York, 1982, Vol. 20, p. 197.
14 L. G. Zachary, H. W. Bajak and F. J. Eveline (eds.), Tall Oil and Its Uses, McGraw-Hill,
New York, 1965.
15 B. HoImbom, E. Avela and S. PekkaIa, J. Am. Oil Chem. Sot., 51 (19 74) 397.
16 L. E. Wise and S. T. Moore, J. Org. Chem., 10 (1945) 516.
17 M. Joye Jr. and R. V. Lawrence, J. Chem. Eng. Data, 12 (1967) 279.
18 W. J. Klyne, J. Chem. Sot., (1953) 3072.
19 S. C. Saksena, H. Panda, Ahisanuddin and H. Rakhshinda, J. Oil CoZour Chem. Assoc.,
64 (1981) 299.
20 R. V. Lawrence, TMPI, 45 (1962) 654.
21 S. C. Saksena, H. Panda and H. Rakhshinda, J. Oil Colour Chem. Assoc., 65 (1982)
317.
22 J. Drew and M. Propst (eds.), Tall Oil, Pulp Chemicals Associations, New York, 1981.
23 J. Brady, Am. Inkmaker, (Sept. 1976) 34.
24 G. C. Harris and T. F. Sanderson, J. Am. Chem. Sot., 70 (1948) 334, 339.
25 V. M. LoebIich, D. E. Baldwin and R. V. Lawrence, J. Am Chem. Sot., 77 (1955) 2823.
26 E. E. Fleck and S. Pa&n, J. Am. Chem. Sot., 62 (1940) 2044.
27 G. C. Harris and T. F. Sanderson, J. Am. Chem. Sot., 70 (1948) 2079, 2081.
28 W. Antkowlak, 0. E. Edwards, R. Howe and J. W. ApSimon, Can, J. Chem., 43 (1965)
1257.
29 L. F. Fieser and W. P. Campbell, J. Am. Chem. Sot., 60 (1938) 159.
30 V. M. LoebIich, D. E. Baldwin, R. T. O’Connor and R. V. Lawrence, J. Am. Chem. Sot.,
77 (1955) 6311.
31 J. A. Hudy, Anal. Chem., 31 (1959) 1754.
32 P. F. Ritchie and L. F. McBurney, J. Am. Chem. Sot., 71 (1949) 3736.
33 P. F. Ritchie and L. F. McBumey, J. Am. Chem. Sot., 72 (1950) 1197 .
34 D. E. Baldwin, V. M. LoebIich and R. V. Lawrence, J. Am. Chem. Sot., 78 (1956) 2015.
35 H. Takeda, W. H. SchuIIer and R. V. Lawrence, J. Org. Chem., 33 (1968) 1683.
36 A. Enoki, Wood Res., 59-60 (1976) 49.
37 W. H. SchuIIer, R. N. Moore and R. V. Lawrence, J. Am. Chem. Sot., 82 (1960) 1734.
38 B. L. Davis and E. E. Fleck, Ind. Eng. Chem., 35 (1943) 171.
39 J. KohIer, J. Pmkt. Chem., 85 (1912) 534.
40 L. Ruzicka, F. Balas and F. ViIim, Helv. Chim. Acta, 7 (1924) 458.
41 N. M. Joye and R. V. Lawrence, J. Org. Chem., 26 (1961) 1024.
42 V. M. LoebIich and R. V. Lawrence, J. Am. Chem. Sot., 79 (1957) 1497.
43 V. M. LoebIich and R. V. Lawrence, Ind. Eng. Chem., 50 (1958) 619.
44 B. A. RadbiI and S. R. Kushnir, Khim. Drev., 1 (1976) 105.
45 S. W. Benson, Thermochemical Kinetics, Wiley, New York, 1976, p. 1.
46 A. S. Fomin, M. I. Arkhipov, M. Poldsaar and G. N. Shutova, Izv. Vgssh. U&&n. Zaved.,
Khim. Khim. Tekhnol., 19 (1976) 1355.
47 J. Churacek, M. Drahkoupilova, P. Matousek and K. Komarek, Chromatographia, II (1969)
493 [Chem. Abstr., 72 (1970) 50 733~3.
48 R. G. R. Bacon and L. Ruzicka, Cha. Ind. @ondon), (1936) 546.
49 H. Wienhaus and W. Sandermann, Ber., 69 (1936) 2202.
50 T. W. Brooks, G. S. Fisher and N. M. Joye Jr., Anal. Chem., 37 (1965) 1063.
51 W. H. SchuIIer and R. V. Lawrence, J. Am. Chem. Sot., 83 (1961) 2563.
52 W. H. SchuIIer, J. C. Minor and R. V. Lawrence, In&. Eng. Chern., Prod. Res. Dev., 3
(1964) 97.
53 W. H. SchuIIer and R. V. Lawrence, Ind. Eng. Chem., Prod. Res. Dev., 6 (1967) 266.
54 B. A. Parkin, W. H. SchuIler and R. V. Lawrence, Ind. Eng. Chem., Prod. Res. Dev., 8
(1969) 304.
55 T. Hasselstrom, E. A. Brennan and S. Hopkins, J. Am. Chem. SOC., 63 (1941) 1759.
56 2. Su and K. Han, Linchan Huoxue Yu Gongye, 5 (1985) 1.
57 R. G. Sinclair, D. A. Berry, W. H. Schuller and R. V. Lawrence, Am. Chem. Sot., Div.
Org. Coat. Plast. Chem., Pap, 29 (1969) 455.
58 J. B. Montgomery, A. N. Hoffmann, A. L. Glasebrook and J. I. Thigpen, Ind. Eng. Chem.,
50 (1958) 313.
59 R. J. By&it, US Pat. 2094 117 (1937).
60 A. L. Glasebrook, A. N. Hoffmann and J. B. Montgomery, US Put. 2 776 276 (1957).
61 E. E. Fleck and S. Palkin, J. Am. Chem. Sot., 60 (1938) 921.
62 P. A. MancinelIi, Adhesive Age, (Sept. 1989) 18.
63 E. R. Littmann, J. Am. Chem. Sot., 60 (1938) 1419.
64 B. A. Parkin and W. H. SchuIler, Ind. Eng. Chem., Prod. Res. Dev., 11 (1972) 156.
65 A. L. Rummelsburg, US Pat. 2 136 525 (1938).
66 T. L. Chang, Anal. Chem., 40 (1969) 989.
67 R. G. Sinclair, D. A. Berry, W. H. Schuller and R. V. Lawrence, Ind. Eng. Chem., Prod.
Res. Dew, 9 (1970) 60.
68 J. R. Olechowski and N. E. Lawson, US Pat. 4414146A (1983).
69 B. H. Hingorani, A. Uddin, H. Panda and S. C. Saksena, J. Oil CoZour Chem. Assoc., 63
(1980) 407.
70 H. Panda and R. Panda, J. Oil Colour Chem. Assoc., 67 (1984) 8.
71 M. C. Shukla and A. K. Vasishtha, J. Oil Colour Chem. Assoc., 68 (1985) 71.
72 G. W. Hedrick, R. V. Lawrence, H. B. Sum&ers Jr. and L. W. Mazzeno Jr., Levopimuric
Acid, USDA Rep. ARS 72-35, Jan. 1965.
73 C. EIlis, The Chemistry of Synthetic Resins, Reinhold, New York, 1935, Vol. 1 and 2, p.
747.
74 N. J. Halbrook and R. V. Lawrence, J. Am. Chem. Sot., 80 (1958) 368.
75 G. C. Harris, in R. E. Kirk and D. F. Othmer (eds.), Encyclopedia of Chemical Technology,
Wiley, New York, 1953, Vol. 11, p. 779.
76 I. I. Bardyshev, R. G. Shlyashinskii and A. N. Bulgakov, Vesti Akud. Navuk B SSR, Ser.
Khim. Nuvuk, 1 (1976) 84 [Chem, Abstr., 84 (1976) 181 798u].
77 K. Arimoto and D. F. Zinkel, J. Am. Oil Chem. Sot., 59 (1982) 166.
78 D. F. Zinkel and K. Arimoti, US Put. 4405514A (1983).
79 L. N. Bent and A. C. Johnson, US Pat. 1820265 (1931).
80 A. N. Hoffmann, W. Manor and J. B. Montgomery, US Put. 2 727 885 (1955).
81 G. W. Hedrick, Ind. Eng. Chem., Prod. Res. Dev., 12 (1973) 246.
82 E. V. Blazhko, Siti. Fiz.-Khim. Polim., 14 (1974) 164.
83 V. M. LoebIich, D. E. Baldwin and R. V. Lawrence, J. Am. C&m. Sot., 77 (1955)
2823.
84 B. C. Trivedi and B. M. Culbertson, Muleic Anhydride, Plenum, New York, 1982, p. 103.
85 M. C. ShukIa, V. S. Gandotra, K. L. Bedi and A. K. Vasishtha, J. Oil Colour Chem. Assoc.,
68 (1985) 147.
86 M. C. Kloetzel, Org. React., 4 (1948) 13.
87 W. D. Lloyd and G. W. Hedrick, J. Org. Chem., 26 (1961) 2029.
88 W. Sandermann, Be-r., 71B (1938) 2005.
89 Ir. L. J. Molhoek, XIXth FATIPEC Congress, Congress book, Vol. 1, (1988) p. 391.
90 S. Seltzer, J. Am. Chem. Sot., 83 (1961) 1861.
91 F. H. Marquardt, J. Chem. Sot. B, (1966) 1242.
92 A. W. BurgstahIer, H. Ziffer and U. Weiss, J. Am. Chem. Sot., 83 (1961) 4660.
93 N. J. Halbrook, R. V. Lawrence, R. L. Dressler, R. C. Blackstone and W. Hen, J. Ch+g.
Chem., 29 (1964) 1017.
94 W. A. Ayer and C. E. McDonald, Can. J. Cha., 41 (1963) 1113.
95 N. J. Halbrook and R. V. Lawrence, Ind. Eng. Chem., 50 (1958) 321.
96 N. J. Halbrook, J. A. Wells and R. V. Lawrence, J. Org. Chem., 26 (1961) 2641.
167

97 Sumitomo Durez Co., Ltd., Jpn. Kokai Tokkyo Koho 8586175 (1985) [ChemAbstr., 103
(1985) 89 155~1.
98 Y. Mukoyama and H. Nishizawa, Kokasei Jushi, 6 (1985) 1 [Chem. Abstr., 103 (1985)
142 821n].
99 P. Penczek, B. Kielska, H. Staniak and C. Czamecki, Polimery, 15 (1970) 598 [Chem.
Abstr., 75 (1971) 37 969x].
100 W. H. Schuller, R. V. Lawrence and B. M. Culbertson, J. Polym. Sci., Part A-l, 5 (1967)
2204.
101 S. S. Ray, A. K. Kundu, M. Maiti, M. Ghosh and S. Maiti, Angew. Makromol. Chem.,
122 (1984) 153.
102 N. J. Halbrook, R. V. Lawrence, M. D. DaIluge and G. A. Stein, Ind. Eng. Chem., Prod.
Res. Dev., 2 (1963) 182.
103 J. C. Weaver, personal communication.
104 P. R. Stapp, Ind. Eng. Chem., Prod. Res. Dev., 15 (1976) 189.
105 B. A. Parkin, H. B. Summers, R. L. Settine and G. W. Hedrick, Ind Eng. Chem., Prod.
Res. Dew, 5 (1966) 257.
106 B. A. Parkin and G. W. Hedrick, J. Org. Chem., 30 (1965) 2356.
107 E. E. Royals and I. T. Greene Jr., J. Org. Chem., 23 (1958) 1437.
108 A. Knop and W. Scheib, Chemistry and Application of Phenolic Resins, Springer-Verlag,
New York, 1979, p. 34.
109 D. K. Black and G. W. Hedrick, J. Org. Chem., 32 (1967) 3758, 3763.
110 D. R. Levering and A. L. Glasebrook, Ind. Eng. Chem., 50 (1958) 317.
111 W. A. Rohde and G. W. Hedrick, Ind. Eng. Chem., Prod. Res. Dev., 10 (1971) 447.
112 N. M. Joye Jr. and R. V. Lawrence, J. Chem. Eng. Data, 12 (1967) 279.
113 J. C. Minor and R. V. Lawrence, Ind. Eng. Chem., 50 (1958) 1127.
114 W. Krumbhaar, Coating and Ink Resins, Reinhold, New York, 1947, Chap. 11.
115 G. P. Springhng, J. Am. Chem. Sot., 74 (1952) 2937.
116 R. R. Myers and J. S. Long, Treatise on Coatings: Film-Forming Compositions, Marcel
Dekker, New York, 1972, Vol. 1, Part 3, p. 165.
117 S. Paul, Su@ace Coatings, Wiley, New York, 1985, p. 166.

You might also like