You are on page 1of 17

Progress in Polymer Science 33 (2008) 1199–1215

Contents lists available at ScienceDirect

Progress in Polymer Science


journal homepage: www.elsevier.com/locate/ppolysci

Condensation polymers from natural oils


Vinay Sharma, P.P. Kundu ∗
Department of Chemical Technology, Sant Longowal Institute of Engineering and Technology, Sangrur, Punjab 148106, India

a r t i c l e i n f o a b s t r a c t

Article history: Innovative technologies and competitive industrial products are reducing the dependence
Received 23 August 2007 on petrochemicals for the production of polymers. Increasing concerns about the deteri-
Received in revised form 1 June 2008
orating environment caused by conventional polymers have directed worldwide research
Accepted 1 July 2008
toward renewable resources. Vegetable oils are one of the most readily available alterna-
Available online 23 August 2008
tive renewable resources. The functional groups present in natural oils can be activated
for condensation polymerization. Accordingly, various types of useful condensation poly-
Keywords:
Soybean oil mers, such as polyurethanes, polyesters and polyethers, are being produced by this route.
Castor oil The incorporation of natural oils into the polymer chain allows tailoring the properties of
Nahar seed oil polyurethane products, for their widespread applications.
Polyols © 2008 Published by Elsevier Ltd.
Condensation polymers
Polyurethanes
Polyesters

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1199
1.1. Polyols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1200
1.2. Polyurethanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1200
2. Polymers based on soybean oil polyols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1200
3. Polymers and IPNs based on castor oil polyols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1205
3.1. Polymers based on castor oil polyols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1205
3.2. Interpenetrating and semi-interpenetrating networks based on castor oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1209
4. Polymers based on nahar seed oil polyol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1212
5. Other polymers based on oil polyols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1213
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1214
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1214

1. Introduction (i.e., polyester or polyether polyol) and a chain extender


(usually a low molecular weight diol or diamine). Cur-
Since the synthesis of polyurethanes by Bayer in rently, the majority of polyols (polyether and polyester
1937 [1], their utilization has become ubiquitous. These polyols) is derived from petrochemicals, a resource sub-
polymers are synthesized by reacting three basic compo- ject to depletion. Hence, bio-based materials are receiving
nents: polyisocyanate, polyhydroxyl containing polymer wide attention as the oil crisis and threat of global warming
deepen. The synthesis of bio-based materials from natural
oils affords an alternative route [2–4] to natural oil-based
∗ Corresponding author. addition polymers, which we discussed in an earlier review
E-mail address: ppk23@yahoo.com (P.P. Kundu). [5].

0079-6700/$ – see front matter © 2008 Published by Elsevier Ltd.


doi:10.1016/j.progpolymsci.2008.07.004
1200 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Vegetable oils are becoming extremely important as polyester polyol (such as adipic polyester resin) and rigid
renewable resources for the preparation of polyols required foam is prepared from aromatic polyester polyol, e.g.
for the polyurethane industry. Polyols from natural oils, depolymerized PET.
such as soybean, castor, and palm oils are increasingly being The biggest challenge for the production of PU using
viewed by industry as a viable alternative to hydrocarbon- polyols from natural oils is the variation of the unsatura-
based feedstocks. These oils are annually renewable, and tion content among natural oils. If the double bond content
are cost-competitive as well as environment friendly. in natural oils is not properly controlled, the nature of the
According to a market summary published by the United polyol will be changed, affecting the performance of the
Soybean Board in February 2000, vegetable oil-based final polyurethane product. Polyurethane foam finds appli-
polyurethanes are best suited to three markets namely: cation as one of the most effective insulating materials.
polyurethane foams, polyurethane binders and agricultural Blowing agents are generally used to create a fine cellu-
films (the last may not be polyurethanes). Currently, the lar foam structure. Upon dissociation the blowing agent
total U.S. annual market size is approximately 3000 million generates gases, which are trapped within the small cells
pounds for polyurethane foams and 400 million pounds (or closed cell) of the foam. This enhances the insulating
for polyurethane binders and fillers. The introduction of properties of the PU foam.
natural oils as polyols into the polyurethane supply chain
can provide an opportunity for polyurethane suppliers and 2. Polymers based on soybean oil polyols
customers to reduce their dependence on natural gas and
crude oil, whose highly volatile and increasing costs con- Petrovic and co-workers [7] studied polyurethane
tinue to make it difficult for them to compete. foams-based on soybean oil. They prepared both
Vegetable oils are excellent renewable source of raw hydrochlorofluorocarbon (HCFC) and pentane blown
materials for the manufacture of polyurethane components rigid polyurethane (PU) foams from polyols derived from
such as polyols. The transformations of the double bonds soybean oil. The effect of process variables on foam prop-
of triglycerides of oils to hydroxyls and their application in erties was studied by varying the amounts and types of
polyurethanes have been the subject of many studies [6]. catalyst, crosslinker, blowing agent, surfactant and water.
The main technological advantages of these polyurethanes The foams prepared from these polyols were found to have
from vegetable oils are high strength as well as stiffness, mechanical and insulating properties such as compressive
environmental resistance and long life. strength and thermal strength comparable to those of
commercially available polypropylene oxide (PPO)-based
1.1. Polyols foams. The commercial PU foams derived from PPO-based
triols have the disadvantages of thermal degradation
Oil-based polyols are often oligomers with a wide dis- and thermal oxidation. It was observed that the soybean
tribution of molecular weights and a considerable degree polyol derived PU foams were more stable. Comparative
of branching, which affect the viscosity and processing thermo-gravimetric analysis (TGA) of PU foams-based on
properties of polyurethane foams produced from them. PPO triols and soybean polyols in air and in nitrogen N2 is
Precise characterization of the polyol composition and its shown in Fig. 1. In air (Fig. 1b), both the PU samples from
properties are very important for understanding synthetic PPO and soy polyol show higher weight loss, compared
processes as well as for quality control. Polyols are a compo- with the thermo-gravimetric loss in nitrogen (Fig. 1a).
nent in the production of polyurethanes used in appliances, The weight loss of up to 5% is important as PUs are never
automotive parts, adhesives, building insulation, furniture, used beyond this weight loss level because of substantial
bedding, footwear and packaging. Although, polyols are deterioration in mechanical and other properties. The PUs
currently produced from petroleum, vegetable oils are also from soy polyol were observed to be more stable than
used extensively for their production. The vegetable oil those from PPO-based polyol at around 200 ◦ C, where the
molecules must be chemically transformed in order to weight loss was less than 5%.
obtain hydroxyl moieties. For example, soybean oil does Petrovic and co-workers [8] studied the thermal sta-
not contain hydroxyl groups, but it has an average of 4.6 bilities of PU-based on various vegetable oils. They used
double bonds per molecule. The unsaturated portion of the TGA and Fourier-transform infra-red (FTIR) spectroscopy
oil can be converted to hydroxyl groups. for thermal analysis of the polymers. In TGA, PPO-based
PU showed single-stage degradation, while the vegetable
1.2. Polyurethanes oil-based PUs degraded in two stages. The isothermal (at
250 ◦ C) thermo-gravimetric loss of PPO-PU, SOY-PU and
Polyurethanes (PU) are polymers containing urethane castor-PU versus time is shown in Fig. 2. The PU from PPO
linkages (–NHCOO–) in the main polymer chain. They can (arcol, as shown in Fig. 2) was more stable at 250 ◦ C than
be classified in the following major groups: (a) flexible the PUs from soy and castor polyol. The degradation pat-
foams, (b) rigid foams, (c) elastomers, (d) fibers, (e) molding tern with temperature of these three PUs derived from PPO,
compositions, (f) surface coatings and (g) adhesives. castor and soybean oils are shown in Fig. 3. This figure
Depending upon the degree of cross-linking, shows four derivative peaks for oil-based and two peaks
polyurethane foams may be flexible or rigid. Polyester for PPO-based PUs. The temperature for initial degradation
polyols used in the foams are based on industrial waste up to 10% by weight was almost the same for all PUs. The
streams of polyester and depolymerized PET from scrap residue at 500 ◦ C appeared to correlate with the amount of
bottles. Flexible foam is based on a flexible aliphatic isocyanate in the polymer, except for PPO-based PU. From
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1201

Fig. 1. Thermal behavior of soybean oil polyol based foam versus PPO-based foam: (a) TGA in N2 . (b) TGA in air. Reproduced from Guo et al. [7] with
permission of John Wiley and Sons, Inc.

Fig. 3, the PPO-based PU showed the fastest rate of weight


loss and the smallest residue at 300 ◦ C. On the other hand,
at the same temperature the oil-based PUs were thermally
more stable than PPO-based PU. From these studies, Petro-
vic et al. proposed three mechanisms of decomposition of
urethane bonds, as shown in Scheme 1. All three reactions
may proceed simultaneously.
The Petrovic group [9] also studied the structure and
properties of PUs prepared from halogenated as well as
non-halogenated soybean polyols. They determined the
structure and properties of these polymers by spectro-
scopic, chemical and physical methods. In this study, they
modified epoxidized soybean oil (ESBO) with hydrochloric
acid, hydrobromic acid, methanol and hydrogen. The effects
of these modifications on PUs were studied by infra-red
(IR) spectroscopy, nuclear magnetic resonance (NMR), gel
permeation chromatography (GPC) and rheological meth-
Fig. 2. Isothermal TGA curve at 250 ◦ C of three polyurethanes in nitrogen. ods. The properties of polyols obtained from ESBO are
Reproduced from Javni et al. [8] with permission of John Wiley and Sons,
given in Table 1. From these polyols, four types of polyol
Inc.
polyurethanes were prepared [10]. The properties of these
polyurethanes followed the same pattern as those of the
polyols. For example, the brominated PU had higher den-
sity, followed by chlorinated, then by methoxylated and,
finally, hydrogenated PUs. For the preparation of PUs, two
types of isocyanates PAPI-2143L (a crude MDI) and Isonate
2143L (a liquid MDI prepolymer containing carbodiimide

Fig. 3. Derivative TGA curves of PPO-based PU, castor-based PU and


soybean-based PU in air. Reproduced from Javni et al. [8] with permission
of John Wiley and Sons, Inc. Scheme 1. Mechanisms of decomposition of urethane bonds. Reproduced
from Javni et al. [8] with permission of John Wiley and Sons, Inc.
1202 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Table 1
Properties of polyols derived from epoxidized soybean oila

Polyol Reagent Yield (%) Hydroxyl no. Equivalent weight Functionality Polyol, Mn Physical state at
(mg KOH/g) (g/equivalent) room temperature

Soy-H2 H2 89 212 265 3.5 938 Grease


Soy-Met CH3 OH 93 199 282 3.7 1053 Liquid
Soy-HClb HCl 94 197 285 3.8 1071 Grease
Soy-HBrb HBr 100 182 308 4.1 1274 Grease
a
Reproduced from Guo and Petrovic [9] with permission of John Wiley and Sons, Inc.
b
All values were calculated on the basis of the analyzed Cl and Br contents and under the assumption that each halogen is accompanied by a hydroxyl
group.

bonds) were used. The PUs from both of these isocyanates while polyols from a cobalt-catalyzed reaction (67%
showed comparable thermal stability, mechanical strength conversion) resulted in hard PU rubbers with lower
and dielectric properties. mechanical strength. The soybean-based polyurethanes
The PU networks from soybean oil had numerous valu- were prepared from aliphatic, cycloaliphatic and aromatic
able properties. These properties solely depended upon the isocyanates [13]. The isocyanates used in this study were
chemical composition and cross-link density. The NCO/OH 4,4 -diphenylmethane diisocyanate (MDI), 2,4:2,6-toluene
mole ratios (isocyanate index) in the polyol directly affected diisocyanate (TDI), hydrogenated MDI (RMDI), isophorone
the physical and mechanical properties of the resulting diisocyanate (IPDI), hexamethylene diisocyanate (HDI),
PU networks [11]. NCO/OH molar ratios were varied from Desmodur N-100 and Desmodur N-3300 (triisocyanates
1.05 to 0.4 with the resulting properties shown in Table 2. derived from 1,6-hexamethylene diisocyanate), Desmodur
With a decrease in the NCO/OH mole ratio, the glass tran- RF-E (a tris (p-isocyanato-phenyl)-thiophosphate) and
sition temperature Tg decreased linearly from 64 to −3 ◦ C. Desmodur CB 75N a (trimethylol propane TDI-based pre-
The tensile strength of the networks decreased from 47.3 polymer). A higher cross-link density of the PU networks
to 0.3 MPa with a decreasing NCO/OH mole ratio. The resulted for triisocyanates than for diisocyanates.
polyurethane networks, prepared with lowering of the The density, glass transition temperature [measured by
molar ratio of NCO and OH groups, had an increasing differential scanning calorimetry (DSC), thermal mechan-
amount of imperfections in the form of dangling chain ical analysis (TMA) and dynamic mechanical analysis
ends. Therefore, the properties of the networks deterio- (DMA)] and the degree of swelling in toluene are
rated with a decrease in the NCO/OH molar ratio. Petrovic reported in Table 3. In comparison with diisocyanates,
et al. [12] also studied the properties of vegetable oil-based the triisocyanates resulted in better overall properties
PUs through hydroformylation. Hydroformylation is an such as density, Tg , swelling and mechanical properties.
aldehyde synthesis process that falls under the general clas- The triisocyanates RF-E [a tris (p-isocyanato-phenyl)-
sification of a Fischer–Tropsch reaction. Hydroformylation, thiophosphate-based prepolymer] and CB 75N (a trimethy-
also known as the oxo synthesis, is an important industrial lol propane TDI-based prepolymer) showed flexural moduli
process for the production of aldehydes from alkenes. This of the order of 2480 and 2040 MPa and tensile strengths of
chemical reaction entails the addition of a formyl (CHO) 48 and 65 MPa, respectively.
group and a hydrogen atom to a carbon–carbon double Petrovic et al. [14,15] also studied the properties of
bond. polyisocyanurate cast resins and foams from soybean
The double bonds of the soybean oil were first converted oil-based polyols. Polyisocyanurate cast resins are heat
to aldehydes using either rhodium or cobalt as catalyst. resistant materials obtained by polycyclotrimerization
The aldehydes were hydrogenated to alcohols forming a of diisocyanates or isocyanate-terminated prepolymers
triglyceride polyol. Polyols from a rhodium-catalyzed reac- (Scheme 2). The formation of polyisocyanurate from
tion (95% conversion) resulted in a rigid polyurethane, the prepolymer took place at higher temperatures com-

Table 2
Molecular weights Mc between crosslinks and concentration e of ENAC for polyurethanes (based on soybean polyol and MDI) withNCO/OH mole ratios
(between 1.05 and 0.4)a

NCO/OH mole ratio Density (g/cm3 ) Degree of swelling (W1 /W0 ) Sol fraction (%) Affine model Phantom model

e (mol/cm3 ) Mc (g/mol) e (mol/cm3 ) Mc (g/mol)

1.05 1.104 1.5199 0.82 2.73 × 10−3 404 4.30 × 10−3 257
1 1.104 1.5528 1.02 2.45 × 10−3 451 3.91 × 10−3 282
0.95 1.104 1.5911 1.29 2.17 × 10−3 509 3.52 × 10−3 313
0.9 1.101 1.6334 1.55 1.92 × 10−3 574 3.17 × 10−3 348
0.8 1.095 1.7299 2.78 1.48 × 10−3 741 2.52 × 10−3 435
0.7 1.088 1.8480 5.20 1.12 × 10−3 981 1.96 × 10−3 556
0.6 1.083 2.0165 8.68 7.65 × 10−4 1415 1.41 × 10−3 769
0.5 1.074 2.350 11.67 4.13 × 10−4 2601 8.09 × 10−4 1372
0.4 1.064 3.060 24.78 1.49 × 10−4 7118 3.19 × 10−4 3333
a
Reproduced from Petrovic et al. [11] with permission of Plenum Publishing Corp.
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1203

Table 3
Density, Tg (measured by DSC, TMA, and DMA), and degree of swelling in toluene of soybean oil-based polyurethanesa

Isocyanate Density (kg/m3 ) Tg by DSC (◦ C) Tg by TMA (◦ C) Tg by DMA (◦ C) Swelling degree (%)

MDI 1104 59 55 74 55
TDI 1104 47 53 62 64
RMDI 1062 50 47 69 75
IPDI 1061 48 48 68 79
HDI 1066 10 15 22 79
N100 1082 26 25 41 52
N3300 1096 25 25 37 59
RFE 1272 NA 64 84 22
CB 75N 1186 93 85 92 12
a
Reproduced from Javni and Petrovic [13] with permission of John Wiley and Sons, Inc.

pared to the reaction temperature for the formation of were semi-rigid at an isocyanate index of 110, but the soy
polyurethane [14]. As a result, the isocyanate terminated polyol foams had a higher density at an index of 110 and
prepolymers (PU) also reacted with themselves, producing showed higher compressive stress.
strong isocyanurate chains (Scheme 2). These cross-links The thermal and mechanical properties of glass rein-
were stronger than normal bonds in polyurethane. For forced soybean oil-based polyurethane composites were
example, polyisocyanurate foam from soy polyol (59 MPa) also studied by Petrovic [16]. PUs in this study were
showed nearly double the stress at break than PPO-based derived from a soybean oil-based polyol (Soypolyol 204)
polyisocyanurate foam (31 MPa). Thus, the resulting poly- and petrochemical polyol (Jeffol G30-650). Although the Tg
isocyanurate foam was chemically and thermally more of jeffol composites was slightly higher than that of soy
stable (disintegration started above 400 ◦ C). Polyisocyanu- polyol composites, the oxidative, thermal and hydrolytic
rates are used in applications where the material is exposed stability of the latter was superior to propylene glycol-
for an extended period of time at an elevated temperature. based polyurethanes [17–19], suggesting that it could
Two polyols (PPO 168 and Soypolyol 173) having the same find increasing applications in the composites area. The
hydroxyl number, functionality and molecular weight, but structure–property relationships were further studied by
with significant structural differences, were used. PPO 168 Petrovic et al. [20]. They prepared two types of soybean
had terminal OH groups, which on reaction with isocyanate polyols, one from epoxidation of soybean oil, followed
of polyisocyanurate formed polyurethane. Soypolyol 173 by methanolysis and the other from hydroformylation of
had internal hydroxyl groups positioned in the middle of soybean oil, followed by hydrogenation [18,19]. The mech-
the fatty chain. Thus, Soypolyol 173 exhibited higher cross- anistic paths of these processes are shown in Scheme 3.
link density of the polyisocyanurate network than PPO The polyol II in Scheme 3 was more reactive than polyol
168. The soybean polyol-based polyisocyanurate showed I, because polyol II contains a primary hydroxyl group
a higher Tg (about 30 ◦ C higher) than PPO 168-based poly- as well as one additional carbon atom in the crosslink-
isocyanurates. ing chain after hydroformylation, while polyol I contains
All the PU foam samples showed a ␤-transition at about a secondary hydroxyl group and methoxy groups as side
−30 to −20 ◦ C [15]. The PPO 168-based PUs decomposed chains.
at a higher rate at 235 ◦ C, while the maximum rate of
decomposition for soy polyol-based PUs was 370 ◦ C. The
compression strength (measured at a stress of 10, 20 and
30% strains with applied force parallel to the foam rise of
soy polyol-based PU foams was much higher than for PPO-
based PU foams, as shown in Fig. 4. The PPO-based foams

Fig. 4. Compression stress of soybean-based foams and PPO-based foams.


Scheme 2. Schematic representation of polyisocyanurate structure. Reproduced from Javni et al. [15] with permission of Plenum Publishing
Reproduced from [14] with permission of John Wiley and Sons, Inc. Corporation.
1204 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Scheme 3. Schematic representation of methanolysis and hydroformylation of soybean oil. Reproduced from Guo et al. [20] with permission of Springer
Science and Business Media, LLC.

Dunjic and co-workers [21] prepared a series of ure- acid and its macrostructure, such as the contents of
thane acrylates derived from soybean fatty acid modified monomer and trimer acids, directly affected the resulting
hyperbranched aliphatic polyesters through a two step polymer properties.
process. The material obtained was used as oligomer in Deng et al. [23] reported the synthesis of soy-
radiation curable compositions. The UV cured acrylates based copolyamides with different ␣-amino acids. The
exhibited good mechanical and thermal properties. The contents of amino acids were varied, keeping the con-
tensile strengths of these samples ranged from 15.73 to tents of dimer acid and phenylenediamine constant. The
72.53 MPa and they showed 5% weight loss at around 300 ◦ C copolyamides thus prepared showed a drastic decrease
during thermal degradation. Fan et al. [22] reported soy- in mechanical strength. For instance, Young’s modu-
based polyamide resins via condensation polymerization lus for homopolymer was 2116.0 MPa; and that for
from different dimer acids and diamines. The polyamide copolymers containing tyrosine, glutamic acid and pheny-
prepared from 1,4-phenylenediamine showed a rapid lalanine were 330.8, 242.9 and 183.2 MPa, respectively.
increase in molecular weight above 260 ◦ C. Polyamide from John et al. [24] studied PU foams derived from soybean
1,4-phenylenediamine showed higher Tg , melting temper- oil. The reaction was monitored by FTIR spectroscopy.
ature (Tm ), decomposition temperature and mechanical The introduction of soybean oil caused an increase in
strength than aliphatic polyamides. The flexibility of dimer hydroxyl values and acid values as well as molecular
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1205

polyester with a Tg of −5 ◦ C, which indicated the rubbery


nature of the polyester. With maleic and phthalic anhy-
drides, stiffer polyesters with flexural moduli in the range of
500–1000 MPa and Tg between 43 and 75 ◦ C were obtained.
Eren et al. [27] polymerized maleinized soybean oil with
different alcohols such as low molecular weight polyols
and long chain diols [glycerol, pentaerythritol, polyethy-
lene glycol (PEG) 160, PEG 400, PEG 600, etc.]. The material
obtained was highly crosslinked and become jellylike in
common solvents like acetone and chloroform.
Mohanty and co-workers [28] evaluated the thermo-
mechanical properties of PU-based on soybean phosphate
ester polyols and polymeric dipenylmethane diisocyanate
(pMDI). The glass transition temperature of the PUs var-
ied from 69 to 82 ◦ C and the values of storage moduli were
between 4 × 108 and 1.3 × 109 Pa. The PU samples were
postcured at 100 and 150 ◦ C. The crosslink density e of
the samples postcured at 100 ◦ C was lower than for those
postcured at 150 ◦ C, despite a longer postcure time. Mon-
teavaro et al. [29] investigated the thermal properties of
soybean oil-based PUs by thermogravimetric analysis. The
TGA was carried out under nitrogen at a heating rate of
10 ◦ C min−1 for dynamic TG, and isothermal measurements
were carried out at 230, 240 and 250 ◦ C. Fig. 6a presents
the rate of weight loss in isothermal measurements at 230,
240 and 250 ◦ C for the PU samples. The curves show an
increase in the rate of weight loss with an increase in tem-
perature. The rate of weight loss was highest in the initial
period (up to approximately 100 min). From Fig. 6a, one can
measure the time for weight loss of 4, 5 and 7% at 230, 240
and 250 ◦ C. Fig. 6b shows the log–log plot of time versus
Fig. 5. SEM micrographs of (a) soybean polyol foam at 3-php water and
(b) voranol polyol foam at 3-php water. Reproduced from Singh and Bhat- temperature for 4, 5 and 7% weight loss of PU. The verti-
tacharya [25] with permission of John Wiley and Sons, Inc. cal lines in Fig. 6b correspond to iso-conversional (for same
weight loss) plots and the plots along the temperature axis
weight, whereas iodine values and fatty acid content correspond to isothermal lines. If the PU had followed the
decreased. same decomposition pattern at different temperatures, the
Singh and Bhattacharya [25] also studied PU foams isothermal lines would be expected to be parallel. Instead,
from soybean oil. They studied the viscoelastic changes they are not parallel, indicating different decomposition
and cell opening by using vane geometry in a strain- processes at different temperatures.
controlled rheometer. They identified four stages of Pechar et al. [30] synthesized PU networks from soybean
modulus development during the foaming reaction namely, oil polyol, petroleum-based polyols and their blends. The
(i) bubble nucleation and growth, (ii) packing of bubble net- soybean polyols were prepared by air oxidation of raw soy-
work/liquid foam, (iii) urea microphase separation and cell bean oil and hydroxylation of epoxidized soybean oil. The
opening, and (iv) final curing. In their study, they further Tg of the PU samples ranged from −21 to +83 ◦ C and was
investigated the effects of temperature, voranol (a synthetic linearly related with the hydroxyl numbers (55–237 mg
polyol), water and straining frequency on the properties of KOH/g) of the soybean oil polyols. Quintero et al. [31]
PU foam. A SEM micrograph for soybean oil polyol foam developed the vegetable oil macro-monomer (VOMM)
and voranol polyol foam at 3-php (parts per hundred parts) technology for application to aqueous coatings. This would
water is shown in Fig. 5. The size and shape of the cells was help to reduce the volatile organic content in waterborne
less uniform for soybean polyol foams compared to syn- coatings. They used soybean acrylated macro-monomer
thetic voranol foams (Fig. 5). Soybean polyol-based foams (SAM) as a copolymerizable hydrophobe for miniemulsion
had more pinhole openings and more partially open cell polymerization. This system allowed an improvement in
structures rather than completely open cells. flow and smoothing of the coated film before crosslinking
Rosch and Mulhaupt [26] synthesized polyester resins and cessation of flow.
and blends from anhydride-cured epoxidized soybean oil.
The mechanical and thermal properties of the casting resins 3. Polymers and IPNs based on castor oil polyols
were dependent upon the type of anhydride used. At low
epoxy conversions, i.e., using low accelerator content or 3.1. Polymers based on castor oil polyols
less reactive anhydrides such as cycloaliphatic anhydrides,
the crosslinked polyesters were highly flexible. For exam- The use of castor oil in the polyurethane industry adds
ple, norbornene dicarboxylic acid anhydride resulted in value to the crop. Bao et al. [32] studied the effect of the
1206 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Scheme 4. Preparation of modified castor oil polyurethane dispersions. Reproduced from Bao et al. [32] with permission of Iran Polymer and Petrochemical
Institute.
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1207

Fig. 7. Avrami plots for castor oil, sodium benzoate and related PET
compositions during isothermal crystallization from the melt at 220 ◦ C.
Reproduced from Barrett et al. [33] with the permission of John Wiley and
Sons, Inc.

tiveness of castor oil. Avrami plots for castor oil, sodium


benzoate and related PET compositions during isothermal
crystallization from the melt at 220 ◦ C are shown in Fig. 7.
In the presence of castor oil, a high curvature appeared in
the Avrami plot for the crystallization of PET. This showed
a smooth transition from the primary to secondary crys-
tallization. The sodium benzoate blends displayed a rapid
transition from primary to secondary crystallization. The
mixture of castor oil and sodium benzoate showed a grad-
ual transition.
Kansara et al. [34] statistically studied the reaction
kinetics of castor oil (CO)-based polyol and TDI. They used
Fig. 6. (a) Isothermal TG curves for PU04 (from toluenediisocyanate and
soy polyol with NCO/OH ratio 0.80) at 230, 240 and 250 ◦ C in nitro-
a second-order rate expression for the equimolar reaction
gen (b) Variation of degradation time (isothermal) with temperature between polyol and TDI. For a specific case, a moles of both
(dynamic TGA) in nitrogen to achieve the same conversion level at three of the reactants are present at the start of the reaction and
temperatures (isothermal and isoconversion lines) for PU04. Reproduced at a specific instant t, x moles of the reactants are reacted.
from Monteavaro et al. [29] with permission of Associacao Brasileira de
Then, one can find from the bimolecular rate expression
Polimeros.
that the quantity x/a(a − x) is directly proportional to the
time t. Thus, plots of x/a(a − x) versus time were linear.
NCO/OH mole ratio on the structure and properties of aque- They studied the reactivity of ortho and para—NCO groups
ous PU from modified castor oil (MCO). The synthesis of the in TDI with varying temperature, catalyst ratio and polyol
PU from modified castor oil (MCPU) is shown in Scheme 4 chain length of. The prepolymers R60 and R92 were polyols-
where acid-ended modified castor oil used for the produc- based on castor oil. The hydroxyl value, equivalent weight
tion of PU-polyamide forms salts with a triamine, leading and Brookfield viscosity (cP) of R60 and R92 were reported
to the formation of a water-soluble polymer. A three step to be 5, 220, 620 and 6, 190, 700, respectively. R60 and R92
degradation was observed for all the PU-polyamide sam- showed higher conversion compared with the parent castor
ples. In the first stage, a weight loss of less than 5% occurred oil at all temperatures and catalyst concentrations.
at about 210–240 ◦ C. In the second stage, a rapid weight loss Kansara and co-workers [35] also studied castor oil-
started at 250 ◦ C and continued up to 360 ◦ C. based polyurethane adhesives. They used two teak wood
Barrett et al. [33] studied the crystallization kinet- pieces and applied the adhesive solution to a thickness of
ics of poly ethylene terephthalate (PET) from a mixture 0.1 mm and joined the pieces together under a load of 2.5 kg
containing naturally functionalized triglyceride oil. The for 12 h. These samples were found to have 10 times more
crystallization of PET was dependent on several factors such lap shear strength as compared with commercially avail-
as plasticization due to the presence of the triglyceride oil, able wood adhesives. The adhesive films were also tested
nucleation from added agents, bond interchange reactions for thermal stability and chemical resistance.
and the formation of a crosslinked triglyceride oil network. Yeganeh and Mehdizadeh [36] investigated millable
The identification of such factors allowed some control PU elastomers from castor oil-based polyols. Millable PU
over microstructure of these products and similar blends elastomers (MPE) are a special type of synthetic rub-
and semi-IPNs. The Avrami analysis supported the effec- ber. MPEs can serve the rubber industry as they can be
1208 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Each sample has NCO/OH mole ratio 1.0. Quantities in parentheses are in g. PHC-I, PHC-II and PHC-III contain castor oil, polyethylene glycol, polyethylene terephthalate and zinc acetate. The curing agent is
mixed, extruded, calendared, and compression or injec-

Dielectric strength
tion molded in conventional rubber processing equipments
[37]. The MPEs were prepared by the reaction of difunc-
tional castor oil and TDI dimer in an inert atmosphere. The

(V/mil)a
synthesis of difunctional castor as reported by Yeganeh and

2159
2117
1926
Mehdizadeh [36] is as follows. In a three-necked, round-
bottomed flask equipped with a stirrer, dropping funnel

Q factor
and thermometer, dried castor oil (100 g) was added. The

121.95
44.24
35.33
temperature was brought to 100 ◦ C, followed by dropwise
addition of phenyl isocyanate (8.33 g) under stirring. The
reaction was continued for 2 h, before cooling.

Dielectric constant
These MPEs showed two-step degradations, one at
270 ◦ C and the second one at 380 ◦ C. The mechanical prop-
erties of these MPE were dependant on the contents of
castor oil and the chain extenders. On comparison with

toluene diisocyanateand trimethylol propane dissolved in a minimum amount of dimethyl formamide and N-methyl aniline. Sample thickness: 17 mil (0.017 in.).
(60 Hz)
a commercially available elastomer (Urepan 600), these

5.00
5.45
5.10
MPEs exhibited comparable tensile strength, compression
set, resilience and slightly inferior abrasion resistance and
elongation at break.

Dissipation factor
Novel PU insulating coatings from glycolyzed PET
(GPET) and castor oil were also investigated by Yeganeh
and Shamekhi [38]. First, they synthesized polyhydroxy

(60 Hz)

0.0082
0.0226
0.0283
compounds from different compositions of CO and GPET.
Several types of reactions can occur during the prepara-
tion of polyhydroxy compounds (PHC). So, proper analysis
such as measurement of hydroxyl values and acid val-

Char (%)@600
ues was carried out to ascertain the transesterification
as the major reaction. The thermal and electrical prop-
erties of the PUs obtained from the PHCs are shown in

(o C)

8.0
3.8
8.7
Table 4. The Tg of the samples ranged from 47 to 61 ◦ C. The
tensile strength was between 19 and 47 MPa and the elon- T 50% (◦ C)
gation was 8–24%. It was observed that with a decrease

Reproduced from Yeganeh and Shamekhi [38] with permission of John Wiley and Sons, Inc. [38].
in the hydroxyl value of PHC, the crosslink density also 345
326
320
decreased. Yeganeh and Hojati-Talemi [39] also studied
biodegradable PU networks obtained from the PHCs of
T 10% (◦ C)

castor oil and PEG. The PU networks obtained were char-


acterized by their physical, mechanical and viscoelastic
262
267
272

properties. The stress–strain curves for the castor oil PUs


(CPU) are shown in Fig. 8. Fig. 8 shows a smooth transition in
Tg (◦ C)

stress–strain behavior for CPU1 and CPU2 samples similar


to that in lightly crosslinked amorphous rubber. However,
61
57
47

the stress–strain behavior of CPU3 and CPU4 was different:


these showed a yield point due to presence of a crystalline
Pu4[PHC-II(1.01) + curing agent (1.3) + DMF(0.86) + xylene(2.59)]
Pu1[PHC-I(0.99) + curing agent (1.3) + DMF(0.86) + xylene(2.57)]

Pu7[PHC-III(1.1) + curing agent (1.3) + DMF(0.9) + xylene(2.7)]

phase.
Because of the higher crystallinity of CPU4, it showed
higher initial modulus and tensile strength and lower
elongation at break. The tensile strength of these
CPUs ranged from 0.66 to 2.51 MPa. The Tg of the
samples was below room temperature (between −2.0
and 5.1 ◦ C).
Preparation of PU nanoparticles by miniemulsion
Thermal and electrical propertiesa

polymerization from castor oil polyol was reported by


Zanetti-Ramos et al. [40]. The PU particle size measured
by dynamic light scattering ranged from 200 to 300 nm.
Ogunniyi et al. [41] reported the preparation and proper-
ties of PU foams from the reaction of TDI with a mixture
of castor oil polyol and polyether polyol. The introduc-
tion of castor oil polyol in the formulation of PU foams
increased the tensile strength from 3.33 to 138.89 kN/m2
Sampleb
Table 4

and compression set from 5.2 to 49.25% with increasing oil


a

content.
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1209

formed were tougher than the corresponding homopoly-


mers.
Patel and Suthar [51] synthesized liquid pre-
polyurethanes from castor oil and TDI under different
conditions and with varying NCO/OH ratios. Then, the
pre-polyurethane was reacted with poly (ethyl acrylate) to
obtain IPNs. These IPNs were characterized by chemical,
mechanical and thermal methods. The glass transition
temperature of these IPNs ranged from 38 to 41 ◦ C. The
chemical resistance of the IPNs was studied in different
reagents: 25% H2 SO4 , 15% HCl, 5% HNO3 , 5% NaOH, 25%
acetic acid, 5% H2 O2 , 40% NaCl, methylethyl ketone, dis-
tilled water, carbon tetrachloride and toluene for seven
7 days. The IPNs were stable in all the media except
methylethyl ketone, carbon tetrachloride and toluene.
The IPNs were stable up to 300 ◦ C with 7–9% weight loss;
54–57% weight loss was observed around 400 ◦ C and
decomposition was complete beyond 550 ◦ C. Mechanical
properties of these IPNs are given in Table 5.
Fig. 8. Stress–strain curves for CPU 1–4. CPU1 containing100% epoxy mod-
Suthar and Patel [52] studied IPNs from castor oil-
ified PU prepolymer-based on castor oil (EPU1) and 0% epoxy modified
PU-based on polyethylene glycol (EPU2), CPU2 containing 90% EPU1 and
based PU and poly (methyl acrylate). The castor oil was
10% EPU2, CPU3 containing 70% EPU1 and 30% EPU2 and CPU4 contain- reacted with 4,4 -diphenylmethanediisocyante to obtain
ing 50% EPU1 and 50% EPU2. Reproduced from Yeganeh and Hojati-Talemi pre-polyurethane, which was further reacted with methyl
[39] with permission of Elsevier Ltd. acrylate monomer and ethylene glycol dimethacrylate as
a crosslinker. These IPNs showed low chemical resis-
3.2. Interpenetrating and semi-interpenetrating tance towards CCl4 , methylethyl ketone and toluene. They
networks based on castor oil showed high light transmittance (49–81%). These poly-
mers were stable up to 300 ◦ C, lost 45% weight rapidly at
Polymer systems comprising of two or more networks, around 450 ◦ C and decomposed completely beyond 600 ◦ C.
which are at least partially interlaced on a molecu- The tensile strength of these polymers was between 0.43
lar scale, but not covalently bonded to each other, are and 1.87 mN/m2 and they had elongations in the range of
called interpenetrating polymer networks (IPN). These 84–102%.
polymer networks cannot be separated unless chemical Suthar et al. [53] also studied the IPNs from castor
bonds are broken. Polymer systems comprising one poly- oil-based polyesters and poly (methyl methacrylate). The
mer network and another of linear or branched polymer, polyesters were synthesized from castor oil and dibasic
characterized by the penetration on a molecular scale acids such as malonic, succinic, glutaric, adipic, suberic and
of the linear or branched macromolecules in the poly- sebacic acids. These IPNs showed two glass transitions cor-
mer networks, are called semi-interpenetrating networks. responding to their individual component networks. The
Semi-interpenetrating polymer networks are distinguished first Tg was in the range of −76 to −70 ◦ C and the second was
from interpenetrating polymer networks because the con- between 25 and 34 ◦ C. The tensile strength was between
stituent linear or branched polymers could, in principle, be 140 and 280 kPa and Young’s modulus was 730–1200 kPa.
separated from the constituent polymer network(s) with- These polymers were stable upto 300 ◦ C showing only
out breaking chemical bonds. When they can be separated, 2–4% weight loss, whereas they showed 50% weight loss
they are called polymer blends. at around 400 ◦ C. The samples decomposed completely
Attempts have been made to synthesize interpenetrat- beyond 500 ◦ C.
ing polymer networks (IPN) and a series of cross-linked Suthar et al. [54] synthesized three series of IPNs-
copolymers and IPNs from epoxidized vegetable oils and based on PU from castor oil and TDI with polystyrene,
maleinized tung oil [42]. Cross-linked copolymers were poly (methyl methacrylate) and poly (n-butyl methacry-
synthesized by mixing epoxidized vegetable oils and late). They studied the effects of structural variables such
adduct of tung oil with maleic anhydride in different as composition, type of vinyl monomer and the effect of
proportions. The IPNs were tested by dynamic mechan- interaction of phases on the dielectric properties. These
ical spectroscopy (DMS) to investigate compatibility and IPNs were characterized in terms of the variation of dielec-
damping properties. The IPNs formed by triglycerides were tric permittivity E , dissipation factor E and tan ı versus
found to increase the toughness and fracture resistance temperature. The dielectric relaxation studies showed that
in conventional thermoset polymers. There are reports of these IPNs behaved like homogeneous materials. Sperling
development of IPNs consisting of cross-linked polystyrene and co-workers [55,56] reported the preparation of inter-
and epoxidized linseed oil elastomers [43]. IPNs from cross- penetrating networks from castor oil-based polyurethanes
linked polystyrene and castor oil elastomers have also and styrene monomer. The resulting IPNs were char-
been reported [44–46]. Suthar and co-workers [47–50] acterized by electron microscopy, modulus-temperature
prepared IPNs from polyurethanes-based on castor oil measurements and stress–strain analysis. The interpen-
with other vinyl monomers. They found that the IPNs etrating networks (IPN) showed stress–strain behavior
1210 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

Table 5
Mechanical properties of IPNsa

Sampleb , c Tensile strength (mN/m2 ) Young’s modulus (mN/m2 ) Elongation @ break (%) Hardness shore A

IPN-1 (PPU-25 + EA-75) 3.41 1.64 194 79


IPN-2 (PPU-35 + EA-65) 3.66 2.17 182 83
IPN-3 (PPU-45 + EA-55) 2.85 1.90 188 95
IPN-4 (PPU-25 + EA-75) 3.37 3.67 148 81
IPN-5 (PPU-35 + EA-65) 7.10 3.01 225 73
IPN-6 (PPU-45 + EA-55) 6.38 2.99 200 80
IPN-7 (PPU-25 + EA-75) 9.38 4.67 183 86
IPN-8 (PPU-35 + EA-65) 10.60 4.13 186 87
IPN-9 (PPU-45 + EA-55) 8.67 3.84 178 79
PEAd 62.30 2500 16 92
a
Reproduced from Patel and Suthar [51] with permission of John Wiley and Sons, Inc.
b
Contents of pre-polyurethane (PPU) and ethyl acrylate (EA) are given in parenthesis.
c
NCO/OH molar ration for IPNs 1–3 is 1.6, for IPNs 4–6 is 1.8 and for IPNs 7–9 is 2.0.
d
PEA is the homopolymer of ethyl acrylate.

similar to reinforced elastomers. The properties of the IPN Kansara et al. [59,60] also studied the sorption and
showed a dependence on the NCO/OH ratio in the castor oil- diffusion behavior of IPNs-based on PU and unsaturated
based PU and the content of polystyrene. At a fixed NCO/OH polyester (UPE). The PU was prepared from castor oil polyol.
ratio of 0.75, with a decrease in the polystyrene content The sorption behavior of these IPNs was studied with
in the IPN from 53 to 50%, the IPNs exhibited an increase changes in crosslink density, NCO/OH mole ratio, composi-
in elongation from 130 to 140% and in tensile stress from tion of the constituents and hydroxyl value. Chlorobenzene
1000 to 2500 psi. For a fixed 50% content of polystyrene, was used to study the swelling behavior of the IPNs. Sorp-
a decrease in elongation (from 140 to 130%) and in tensile tion was inversely related to the NCO/OH mole ratio. The
stress (from 2500 to 1200 psi) was observed for an increase crosslink density ranged from 5 to 27 × 104 mol/g for dif-
in the NCO/OH ratio from 0.75 to 0.85. The decrease in ferent IPNs, as calculated by the Flory–Rehner equation
polystyrene content and NCO/OH ratio caused a decrease [61]. An increase in diffusion coefficient was observed with
in the amount of stiff material in the IPN, leading to an an increase in the NCO/OH molar ratio and the content of
increase in elongation. The decrease in stiff material was unsaturated polyester.
expected to cause a decrease in tensile stress; but con- Xie and co-workers [62] studied the damping behav-
trary to expectation, the tensile stress increased with a ior of grafted interpenetrating networks. They prepared
decrease in stiffer material. This phenomenon has not been IPNs from castor oil, toluene diisocyanate, monohydroxy
explained [56], but may possibly be due to synchronization terminated acrylic prepolymer and acrylic monomer in the
of the phases. presence of dibutyltin dilaurate and redox initiators. The
Prashantha et al. [57] studied IPNs-based on dynamic mechanical properties of the IPNs were observed
polyol modified castor oil polyurethane and poly (2- to exhibit high damping properties over a wide range of
hydroxyethylmethacrylate). They investigated mechanical temperature.
and thermal properties of the IPNs such as tensile strength, Xie and Guo [63] studied adhesives made from IPNs for
elongation, hardness, Tg and degradation. The tensile bonding with rusted iron without pretreatment. They pre-
strength ranged from 27 to 38 MPa and elongation at break pared two types of room-temperature curable IPNs: one
was between 58 and 120%. All the IPNs showed about 2–4% from castor oil-based PU and vinyl or acrylic polymer and
decomposition at 200 ◦ C, about 10% at 300 ◦ C and about the other from castor oil PU, unsaturated polyester and vinyl
40% at 400 ◦ C. There was a rapid weight loss from 40 to or acrylic polymer. The lap shear strength of joints between
90% in temperature ranges of 400–500 ◦ C. rusted iron plates and IPNs (adhesives) ranged from 1.51 to
Sanmathi et al. [58] synthesized and characterized IPNs 8 MPa.
from modified castor oil-based polyurethane and poly (2- Nayak and co-workers [64–68] also synthesized and
ethoxyethyl methacrylate). The preparation of these IPNs characterized castor oil-based interpenetrating networks.
is shown in Scheme 5. The triglyceride of castor oil was First, they prepared polyurethanes from castor oil and hex-
exchanged with glycerol, followed by reaction with a diiso- amethylene diisocyanate by varying the NCO/OH ratio and
cyanate to form castor-based PU (GC-PU). 2-Ethoxyethyl then prepared IPNs by reacting polyurethanes with various
methacrylate was subsequently polymerized in the pres- acrylates such as hydroxyethyl methacrylate and cardanyl
ence of GC-PU to prepare the IPNs. The presence of methacrylate by using benzoylperoxide as initiator and
H-bonding between poly (2-ethoxyethyl methacrylate) and ethylene glycol dimethacrylate as crosslinker. Infra-red IR
castor PU (GC-PU) is also shown in Scheme 5. The incor- spectroscopy, NMR spectroscopy, TGA, etc. were employed
poration of poly (2-ethoxyethyl methacrylate) in GC-PU to study the properties of resulting IPNs.
during the formation of IPN caused an increase in tensile Cunha et al. [69] employed a statistical method to
strength and decrease in elongation, when compared with accurately evaluate the properties of castor oil-based semi-
the individual components of GC-PU and (2-ethoxyethyl interpenetrating polymer networks (sIPN). They used a
methacrylate). These IPNs exhibited Young’s modulus of 23 factorial experimental design for optimization of the
the order of 2.9–84 mN/m2 . properties. The semi-interpenetrating polymer networks
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1211

Scheme 5. Preparation of GC-PU/poly (2-EOEMA) IPNs. Reproduced from Sanmathi et al. [58] with permission of John Wiley and Sons, Inc.

(sIPNs) were prepared from castor oil polyol, toluene diiso- to study the properties of the sIPNs. They showed differ-
cyanate and methyl methacrylate. The properties of these ent swelling properties for different samples. The sample
sIPNs depended upon the ratio of the contents and the containing maximum methyl methacrylate exhibited low
mole ratio NCO/OH. Various techniques were employed swelling. Mechanical properties were also largely influ-
1212 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

enced by the methyl methacrylate content. For instance, into methyl ester, and then reacted with diethanol amide
the sIPN containing 20% MMA showed 22.6% swelling and to form diethanol amide of fatty acids. The fatty acids amide
1.48 MPa elastic modulus, whereas the sIPN containing 60% was then reacted with dibasic acid or their derivatives. The
MMA showed 17.4% swelling and 6.49 MPa elastic modulus. hardness of the resins was tested as pencil hardness.
The pencil hardness test is used to test coatings for their
4. Polymers based on nahar seed oil polyol hardness and resistance to scratches and wear. The princi-
ple of operation uses a pressure applied to allow the pencil
Nahar (Mesua ferrea L.) seed oil contains mainly triglyc- lead to just crush and therefore, repeatable results for the
erides of linoleic, oleic, palmitic and stearic acids [70]. Karak hardness of the coating can be obtained. These pencils,
and co-workers [71–74] studied polyester, polyesteramide, when pressed for a specified number of times on the coat-
polyurethane and polyurethane amide resins from nahar ing, will also allow a wear factor to be determined. This
seed oil [71–74]. They prepared polyester resins from wear factor is related to the hardness of the pencil used.
nahar seed oil monoglycerides and phthalic and or maleic Pencils of different grades are used in the test, ranging from
anhydrides [71]. Nahar seed oil was first converted into 9H to 9B. [H signifies hardness and as one goes to a lower
monoglycerides by alcoholysis and then the resins were H number, the lead becomes softer, and grades HB to 9B
prepared by reacting monoglycerides with phthalic and designate increasing softness (B = blackness).]
or maleic anhydrides. The alcoholysis and polycondensa- The pencil hardness of the cured resins as reported by
tion reaction of the oil with acid anhydride is shown in Karak and Mahapatra [72] ranged from HB to 2H. The cured
Scheme 6. An increase in the maleic anhydride content in resin had a gloss between 81 and 85 at 60◦ and had 100%
the resin decreased the curing time. Resin with 50% maleic adhesion. Karak and Dutta [73] studied the effect of the
anhydride took 7 h for complete curing, whereas resin with NCO/OH ratio on the properties of nahar seed oil modified
75% maleic anhydride took 6 h. The polyester films were PU resins. The resins were prepared by varying NCO/OH
highly resistant to dilute HCl (10%), aqueous NaCl salt solu- molar ratio from 0.8 to 2.0. The coating properties such
tion (10%) and distilled water. as pencil hardness, gloss, adhesion was studied. The hard-
Polyesteramide resins from nahar oil were also studied ness of the resins ranged from HB to 3H, the gloss at a 60◦
by Karak and Mahapatra [72]. The oil was first converted angle was from 109 to 117% and the adhesive strength was

Scheme 6. Alcoholysis and poly-condensation reaction of oil with acid anhydride. Reproduced from Dutta et al. [71] with permission of Elsevier B.V.
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1213

125–336 kN/m. These properties were found to increase


with increasing NCO/OH ratio from 0.8 to 2.0. The PU resin
with equal proportions of NCO and OH showed three degra-
dation stages at temperatures of 295, 375 and 490 ◦ C. The
degradation temperatures were observed to decrease with
increasing NCO/OH ratio. Resin with a NCO/OH ratio of 1.5
showed two-stage degradation at 290 and 390 ◦ C and the
resin with a NCO/OH ratio of 2 showed a single-stage degra-
dation at 320 ◦ C.
Karak and Dutta [74] synthesized poly (urethane amide)
resins from nahar seed oil and TDI in the presence of dibutyl
tin dilaurate (DBTDL) as catalyst. The coating performance
of these resins was tested by measuring gloss, pencil hard-
ness, adhesion and chemical resistance. The gloss at a 45◦
angle ranged from 66 to 70 and the pencil hardness ranged
from HB to 2H. Chemical resistance was studied for 10 days
insolutions of NaOH, HCl, NaCl, ethyl alcohol and in dis-
tilled water. The cured PU amide resin showed average to
excellent chemical resistance.

5. Other polymers based on oil polyols

There are many vegetable oils other than soybean, cas-


tor and nahar that can be used in the synthesis of polyols
for PUs, such as rapeseed oil, palm oil, canola oil, linseed
oil, olive oil, sunflower oil, safflower oil, etc. Rapeseed
oil methyl esters have been synthesized by the reaction
of rapeseed oil and methanol at a temperature range of
60–70 ◦ C with sodium hydroxide as a catalyst. They have
been produced on an industrial scale for use as biodiesel
fuel [75].
Hu et al. [76] prepared rigid polyurethane foam from
rapeseed oil polyol. They first converted the oil into polyol
by hydroxylation, followed by alcoholysis. The alcoholysis Fig. 9. Glass transition behavior for several vernonia-sebacic polyester
of the hydroxylated rapeseed oil was required because of network/PET semi-IPN compositions. First heat data, the materials as syn-
its low hydroxyl value (ca. 100 mg KOH/g), which is not thesized (A) and second heat data, recorded after cooling rapidly from
suitable for the preparation of rigid polyurethanes. The 300 ◦ C. Reproduced from Barrett et al. [81] with permission of John Wiley
and Sons, Inc.
polyol thus prepared was compared with a commercially
available polyester polyol, Daltolac P744. The experimental
data indicate that the compression strength of the PU foam phthalic acid. The PGPEA coatings showed 10% weight loss
made from rapeseed oil polyol is lower than that of Dalto- up to 275 ◦ C, 50% weight loss up to 370 ◦ C and 78% weight
lac P744-based foam. The other properties of the rapeseed loss up to 400 ◦ C. Almost all of the physico-mechanical
oil polyol foam are similar to those of commercial foam, properties such as scratch hardness (3.5 kg), gloss (up to
Daltolac P744. 90% at a 60◦ angle) and adhesion strength (100%) were
Chian and Gan [77] reported the development of superior to those of polyesteramides obtained from other
rigid PU foams from palm oil. Their PU foams exhib- seed oils. Ahmad et al. [80] also studied ambient cured
ited high compression strength (1–35 MPa) and densities PU modified epoxy coatings from linseed oil. The linseed
(200–300 kg/m3 ). Desai and co-workers [78] reported PU oil was first hydroxylated and then linseed oil PU was
adhesive from argemone oil and castor oil for wood bond- prepared. The linseed oil PU obtained by 10% loading of
ing applications. The best performing adhesive (argemine TDI showed the best physico-mechanical and anticorrosive
oil-based) was compared with various commercially avail- properties.
able adhesives such as Dunlop adhesive, Fevicol and Barrett et al. [81] synthesized semi-IPNs from vernonia
Araldite. The average lap shear strength of PU adhe- oil–sebacic acid polyester network (VOSA) and PET. The
sive was 63.1 × 105 N/m2 , while the highest strength for semi-IPN containing 50% PET and 50% VOSA was observed
the commercially available adhesives was observed to be to be over 15 times tougher than PET and over 50 times
28.3 × 105 N/m2 for araldite. tougher than the neat vernonia oil elastomer. The semi-
Ahmad et al. [79] synthesized polyesteramide from IPN with 50% PET showed better properties than the other
Pongamia glabra oil for biologically safe anticorrosive coat- semi-IPNs, e.g. the tensile strength for the semi-IPN with
ings. The oil was first converted to N,N-bis(2-hydroxyethyl) 50% PET was 6030 MPa and the modulus was 25.8 MPa. The
P. glabra fatty amine (HEPGA) and then HEPGA was con- derivative plots for the first and the second heating scans
verted into polyesteramide (PGPEA) by reaction with in DSC heat flow around the Tg region are shown in Fig. 9.
1214 V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215

All the semi-IPNs showed two glass transitions in the first [16] Husic S, Javni I, Petrovic ZS. Thermal and mechanical properties of
scan, which transformed into a single transition in the sec- glass reinforced soy-based polyurethane composites. Compos Sci
Technol 2005;65:19–25.
ond scan after rapid cooling from 300 ◦ C. Thus the materials [17] Petrovic Z, Guo A, Javni I. Process for the preparation of vegetable
were multiphase as prepared, but after heating to above the oil-based polyols and electroinsulating casting compounds created
melting point of the PET and cooling, they became nearly from vegetable oil-based polyols, 2000; US Patent 6,107,433.
[18] Petrovic Z, Javni I, Guo A. Method of making natural oil-based polyols
single phase or microheterogeneous.
and polyurethanes there from, 2002; US Patent 6,433,121.
[19] Petrovic ZS, Javni I. Process for the synthesis of epoxidized natural
oil-based isocyanate prepolymers for application in polyurethanes,
6. Conclusions 2002; US Patent 6,399,698.
[20] Guo A, Zhang W, Petrovic ZS. Structure-property relation-
ships in polyurethanes derived from soybean oil. J Mater Sci
Waste plastic materials are the major concerns of envi- 2006;41:4914–20.
ronmentalists. Thus, renewable resources are now greatly [21] Dzunuzovic E, Tasic S, Bozic B, Babic D, Dunjic B. UV-curable hyper-
favored for the production of polymers. The wide range branched urethane acrylate oligomers containing soybean fatty
acids. Progr Org Coat 2005;52:136–43.
of mechanical properties and the ability for easy machin-
[22] Fan XD, Deng Y, Waterhouse J, Pfromm P. Synthesis and characteri-
ing and forming cause the plastic foams and elastomers to zation of polyamide resins from soy-based dimer acids and different
find wide industrial and consumer applications. In partic- amides. J Appl Polym Sci 1998;68:305–14.
[23] Deng Y, Fan XD, Waterhouse J. Synthesis and characterization of soy-
ular, urethane foams and elastomers have been found to
based copolyamides with different ␣-amino acids. J Appl Polym Sci
be well suited for many applications. The vegetable oils 1999;73:1081–8.
provide a large variety of options for the preparation of [24] John J, Bhattacharya M, Turner RB. Characterization of polyurethane
polyurethanes. All the vegetable oils are triglycerides of foams from soybean oil. J Appl Polym Sci 2002;86:3097–107.
[25] Singh AP, Bhattacharya M. Viscoelastic changes and cell open-
fatty acids and most contain unsaturated groups. Only a few ing of reacting polyurethane foams from soy oil. Polym Eng Sci
oils contain other groups such as hydroxyl in castor oil, saf- 2004;44:1977–86.
flower oil and lesquerella oil, and oxirane group in vernonia [26] Rosch J, Mulhaupt R. Polymers from renewable resources: polyester
resins and blends based upon anhydride-cured epoxidized soybean
oil. Soybean and linseed oils are also used in the prepa- oil. Polym Bull 1993;31:679–85.
ration of polyurethanes by converting them into epoxies [27] Eren T, Kusefoglu SH, Wool R. Polymerization of maleic anhydride-
or by hydroformylation. The incorporation of oils in the modified plant oils with polyols. J Appl Polym Sci 2003;90:
197–202.
polyurethanes provides a great opportunity to tailor the [28] Dwan’Isa JP, Mohanty AK, Misra M, Drzal LT, Kazemizadeh M. Novel
properties of polyurethane products. biobased polyurethanes synthesized from soybean phosphate ester
polyols: thermomechanical properties evaluations. J Polym Environ
2003;11:161–8.
References [29] Monteavaro LL, Riegel IC, Petzhold CL, Samios D. Thermal stabil-
ity of soy-based polyurethanes. Polimeros: Ciencia e Technologia
2005;15:151–5.
[1] Bayer O, Siefken W, Rinke H, Orthner L, Schild H. A process for the [30] Pechar TW, Sohn S, Wilkes GL, Ghosh S, Frazier CE, Fornof A, et
production of polyurethanes and polyureas, 1937; German Patent al. Characterization and comparison of polyurethane networks pre-
DRP 728981. pared using soybean-based polyols with varying hydroxyl content
[2] Perepelkin K. Polymeric materials of the future based on renew- and their blends with petroleum-based polyols. J Appl Polym Sci
able plant resources and biotechnologies: fibres, films, plastics. Fibre 2006;101:1432–43.
Chem 2005;37:417–30. [31] Quintero C, Mendon SK, Smith OW, Thames SF. Miniemulsion
[3] Tian D, Ross JS. Eco-innovations in floor-covering materials: bio- polymerization of vegetable oil macromonomers. Progr Org Coat
based elastomeric materials based on castor oil. Proc 11th Ann Green 2006;57:195–201.
Chem Eng Conf 2007:161. [32] Bao LH, Lan YJ, Zhang SF. Effect of NCO/OH molar ratio on the structure
[4] Ardakani AA, Gelorme JD, Kosbar LL. Cross-linked biobased materials and properties of aqueous polyurethane from modified castor oil. Iran
and uses thereof, 1998; US Patent 5,833,883. Polym J 2006;15:737–46.
[5] Sharma V, Kundu PP. Addition polymers from natural oils—a review. [33] Barrett LW, Sperling LH, Gilmer J, Mylonakis SG. Crystallization
Progr Polym Sci 2006;31:983–1008. kinetics of poly (ethylene terephthalate) in compositions con-
[6] Petrovic Z, Javni I, Guo A, Zhang W. Method of making natural taining naturally functionalized triglyceride oil. J Appl Polym Sci
oil-based polyols and polyurethanes therefrom, 2004; US Patent 1993;48:1035–50.
6,686,435. [34] Ajithkumar S, Kansara SS, Patel NK. Kinetics of castor oil based polyol-
[7] Guo A, Javni I, Petrovic Z. Rigid polyurethane foams based on soybean toluene diisocyanate reactions. Euro Polym J 1998;34:1273–6.
oil. J Appl Polym Sci 2000;77:467–73. [35] Somani KP, Kansara SS, Patel NK, Rakshit AK. Castor oil based
[8] Javni I, Petrovic ZS, Guo A, Fuller R. Thermal stability of polyurethanes polyurethane adhesives for wood-to-wood bonding. Int J Adhesion
based on vegetable oils. J Appl Polym Sci 2000;77:1723–34. Adhesives 2003;23:269–75.
[9] Guo A, Cho Y, Petrovic ZS. Structure and properties of halogenated [36] Yeganeh H, Mehdizadeh MR. Synthesis and properties of isocyanate
and nonhalogenated soy-based polyols. J Polym Sci Part A: Polym curable millable polyurethane elastomers based on castor oil as a
Chem 2000;38:3900–10. renewable resource polyol. Euro Polym J 2004;40:1233–8.
[10] Petrovic ZS, Guo A, Zhang W. Structure and properties of [37] Hepburn H. Polyurethane elastomers. Essex: Elsevier; 1992.
polyurethanes based on halogenated and nonhalogenated soy- [38] Yeganeh H, Shamekhi MA. Novel polyurethane insulating coatings
polyols. J Polym Sci Part A: Polym Chem 2000;38:4062–9. based on polyhydroxy compounds, derived from glycolyzed PET and
[11] Petrovic ZS, Zhang W, Zlatanic A, Lava CC, Ilavsky M. Effect of OH/NCO castor oil. J Appl Polym Sci 2006;99:1222–33.
molar ratio on properties of soy-based polyurethane networks. J [39] Yeganeh H, Hojati-Talemi P. Preparation and properties of novel
Polym Environ 2002;10:5–12. biodegradable polyurethane networks based on castor oil and poly
[12] Guo A, Demydov D, Zhang W, Petrovic ZS. Polyols and (ethylene glycol). Polym Degrad Stab 2007;92:480–9.
polyurethanes from hydroformylation of soybean oil. J Polym [40] Zanetti-Ramos BG, Lemos-Senna E, Soldi V, Borsali R, Cloutet E,
Environ 2002;10:49–52. Cramail H. Polyurethane nanoparticles from a natural polyol via
[13] Javni I, Zhang W, Petrovic ZS. Effect of different isocyanates miniemulsion technique. Polymer 2006;47:8080–7.
on the properties of soy-based polyurethanes. J Appl Polym Sci [41] Ogunniyi DS, Fakayejo WRO, Ola A. Preparation and properties of
2003;88:2912–6. polyurethane foams from toluene diisocyanate and mixtures of cas-
[14] Javni I, Zhang W, Petrovic ZS. Soybean oil based polyisocyanurate cast tor oil and polyol. Iran Polym J 1996;5:56–9.
resins. J Appl Polym Sci 2003;90:3333–7. [42] Yin Y, Yao S, Zhou X. Synthesis and dynamic mechanical behavior of
[15] Javni I, Zhang W, Petrovic ZS. Soybean oil based polyisocyanurate crosslinked copolymers and IPNs from vegetable oils. J Appl Polym
rigid foams. J Polym Environ 2004;12:123–9. Sci 2003;88:1840–2.
V. Sharma, P.P. Kundu / Progress in Polymer Science 33 (2008) 1199–1215 1215

[43] Carraher CE, Sperling LH. Polymer applications of renewable- [63] Xie HQ, Guo JS. Adhesives made from interpenetrating polymer net-
resource materials. New York: Plenum Press; 1983. works for bonding rusted iron without pretreatment. Int J Adhesion
[44] Devia N, Manson JA, Sperling LH, Conde A. Simultaneous interpen- Adhesives 1997;17:223–7.
etrating networks based on castor oil elastomers and polystyrene. [64] Nayak PL, Lenka S, Panda SK, Pattnaik T. Polymers from renew-
IV. Stress–strain and impact loading behavior. Polym Eng Sci able resources I Castor oil-based interpenetrating polymer net-
1979;19:878–82. works: thermal and mechanical properties. J Appl Polym Sci
[45] Devia N, Manson JA, Sperling LH, Conde A. Simultaneous interpen- 1993;47:1089–96.
etrating networks based on castor oil elastomers and polystyrene. [65] Nayak P, Mishra DK, Parida D, Sahoo KC, Nanda M, Lenka S,
III. morphology and glass transition behavior. Polym Eng Sci et al. Polymers from renewable resources IX Interpenetrating
1979;19:869–77. polymer networks based on castor oil polyurethane poly (hydrox-
[46] Devia N, Manson JA, Sperling LH, Conde A. Simultaneous yethylmethacrylate): synthesis, chemical, thermal and mechanical
interpenetrating networks based on castor oil elastomers and properties. J Appl Polym Sci 1997;63:671–9.
polystyrene. 2. Synthesis and systems characteristics. Macro- [66] Mohapatra DK, Das D, Nayak PL, Lenka S. Polymers from renew-
molecules 1979;12:360–9. able resources XX Synthesis, structure and thermal properties
[47] Patel M, Patel P, Suthar BP. Synthesis and characterization of some of semi-interpenetrating polymer networks based on cardanol-
polyurethanes. Angew Makromol Chem 1988;156:29–35. formaldehyde-substituted aromatic compounds copolymerized
[48] Patel M, Suthar B. Interpenetrating polymer networks from castor resins and castor oil polyurethanes. J Appl Polym Sci 1998;70:837–42.
oil-based polyurethanes and poly (methyl methacrylate) V. J Polym [67] Mohapatra DK, Nayak PL, Lenka S. Polymers from renewable
Sci Part A: Polym Chem 1987;25:2251–60. resources. XXI. Semi-interpenetrating polymer networks based on
[49] Patel M, Suthar B. Interpenetrating polymer networks from castor cardanol-formaldehyde-substituted aromatic compounds copoly-
oil-based polyurethanes and poly (methyl methacrylate) III. Angew merized resins and castor oil polyurethanes: Synthesis, structure,
Makromol Chem 1987;149:111–7. scanning electron microscopy and XRD. J Polym Sci Part A: Polym
[50] Patel P, Suthar B. Interpenetrating polymer networks from castor oil- Chem 1997;35:3117–24.
based polyurethanes and poly (methyl methacrylate) XV. J Polym Sci [68] Nayak P, Mishra DK, Sahoo KC, Pati NC, Jena PK, Lenka S, et al.
Part: A Polym Chem 1989;27:3053–62. Polymers from renewable resources XIII interpenetrating polymer
[51] Patel M, Suthar B. Interpenetrating polymer networks from castor networks derived from castor oil-hexamethylene diisocyanate and
oil-based polyurethanes and poly (ethyl acrylate) VII. J Appl Polym polymethacrylamide. J Appl Polym Sci 2001;80:1349–53.
Sci 1987;34:2037–45. [69] Cunha FOV, Melo DHR, Veronese VB, Forte MMC. Study of castor
[52] Patel P, Suthar B. Interpenetrating polymer networks from castor oil- oil polyurethane-poly (methyl methacrylate) semi-interpenetrating
based polyurethanes and poly (methyl acrylate). XIV. J Appl Polym polymer network (SIPN) reaction parameters using a 23 factorial
Sci 1989;37:841–50. experimental design. Mater Res 2004;7:539–43.
[53] Suthar B, Dave M, Jadav K. Sequential-interpenetrating polymer [70] Konwer D, Taylor SE, Gordon BE, Otvos JW, Calvin M. Liquid fuels from
networks from castor oil-based polyesters. XXVI. J Appl Polym Sci Mesua ferrea L. seed oil. J Am Oil Chem Soc 1989;66:223–6.
1993;50:2143–7. [71] Dutta N, Karak N, Dolui SK. Synthesis and characterization
[54] Suthar B, Klempner D, Frisch KC, Petrovic Z, Jelcic Z. Novel dielectrics of polyester resins based on nahar seed oil. Progr Org Coat
from IPNs derived from castor oil based polyurethanes. J Appl Polym 2004;49:146–52.
Sci 1994;53:1083–90. [72] Mahapatra SS, Karak N. Synthesis and characterization of
[55] Sperling LH, Yenwo GM, Manson JA, Pulido J, Conde A. Castor oil based polyesteramide resins from nahar seed oil for surface coating
interpenetrating polymer networks. IV. Mechanical behavior. Polym applications. Progr Org Coat 2004;51:103–8.
Eng Sci 1977;17:251–6. [73] Dutta S, Karak N. Effect of the NCO/OH ratio on the properties of
[56] Yenwo GM, Manson JA, Pulido J, Sperling LH, Conde A, Devia N. Castor Mesua ferrea L. seed oil-modified polyurethane resins. Polym Int
oil based interpenetrating polymer networks: synthesis and charac- 2006;55:49–56.
terization. J Appl Polym Sci 1977;21:1531–41. [74] Dutta S, Karak N. Synthesis, characterization of poly (urethane amide)
[57] Prashantha K, Pai KVK, Sherigara BS, Prasannakumar S. Inter- resins from nahar seed oil for surface coating applications. Progr Org
penetrating polymer networks based on polyol modified castor Coat 2005;53:147–52.
oil polyurethane and poly (2-hydroxyethylmethacrylate): Synthe- [75] Louwrier A. Review: industrial products—the return to carbohydrate-
sis, chemical, mechanical and thermal properties. Bull Mater Sci based industries. Biotech Appl BioChem 1998;27:1–8.
2001;24:535–8. [76] Hu YH, Gao Y, Wang DN, Hu CP, Zu S, Vanoverloop L. Rigid
[58] Sanmathi CS, Prasannakumar S, Sherigara BS. Interpenetrating poly- polyurethane foam prepared from a rape seed oil based polyol. J Appl
mer networks based on polyol modified castor oil polyurethane and Polym Sci 2002;84:591–7.
poly (2-ethoxyethyl methacrylate). Synthesis, chemical, mechan- [77] Chian KS, Gan LH. Development of a rigid polyurethane foam from
ical, thermal properties and morphology. J Appl Polym Sci palm oil. J Appl Polym Sci 1998;68:509–15.
2004;94:1029–34. [78] Desai SD, Patel JV, Sinha VK. Polyurethane adhesive system from
[59] Ajithkumar S, Patel NK, Kansara SS. Sorption behaviour of interpen- biomaterial-based polyol for bonding wood. Int J Adhesion Adhesives
etrating polymer networks based on polyurethane and unsaturated 2003;23:393–9.
polyester. Polym Gels Networks 1998;6:137–47. [79] Ahmad S, Ashraf SM, Naqvi F, Yadav S, Hasnat A. A polyesteramide
[60] Ajithkumar S, Patel NK, Kansara SS. Sorption and diffusion of from Pongamia glabra oil for biologically safe anticorrosive coating.
organic solvents through interpenetrating polymer networks (IPNs) Progr Org Coat 2003;47:95–102.
based on polyurethane and unsaturated polyester. Eur Polym J [80] Ahmad S, Ashraf SM, Sharmin E, Zafar F, Hasnat A. Studies on ambient
2000;36:2387–93. cured polyurethane modified epoxy coatings synthesized from a sus-
[61] Flory PJ, Rehner J. Statistical mechanics of cross-linked polymer net- tainable resource. Progr Crystal Growth Charac Mater 2002;45:83–8.
works. II. Swelling. J Chem Phys 1943;11:521–6. [81] Barrett LW, Shaffer OL, Sperling LH. Semi-interpenetrating polymer
[62] Xie HQ, Wang GG, Guo JS. Damping materials based on simulta- networks composed of poly (ethylene terephthalate) and vernonia
neously grafted interpenetrating polymer networks from castor oil. oil. J Appl Polym Sci 1993;48:953–68.
Angew Makromol Chem 1994;221:91–102.

You might also like