Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Effect of Structure on Properties of Polyols and

Polyurethanes Based on Different Vegetable Oils

ALISA ZLATANIĆ, CHARLENE LAVA, WEI ZHANG, ZORAN S. PETROVIĆ

Kansas Polymer Research Center, Pittsburg State University, 1501 S. Joplin Street, Pittsburg, Kansas 66762

Received 23 April 2002; accepted 15 September 2003

ABSTRACT: We synthesized six polyurethane networks from 4,4⬘-diphenylmethane di-


isocyanate and polyols based on midoleic sunflower, canola, soybean, sunflower, corn,
and linseed oils. The differences in network structures reflected differences in the
composition of fatty acids and number of functional groups in vegetable oils and
resulting polyols. The number average molecular weights of polyols were between 1120
and 1300 and the functionality varied from 3.0 for the midoleic sunflower polyol to 5.2
for the linseed polyol. The functionality of the other four polyols was around 3.5.
Canola, corn, soybean, and sunflower oils gave polyurethane resins of similar crosslink-
ing density and similar glass transitions and mechanical properties despite somewhat
different distribution of fatty acids. Linseed oil– based polyurethane had higher
crosslinking density and higher mechanical properties, whereas midoleic sunflower oil
gave softer polyurethanes characterized by lower Tg and lower strength but higher
elongation at break. It appears that the differences in properties of polyurethane
networks resulted primarily from different crosslinking densities and less from the
position of reactive sites in the fatty acids. © 2004 Wiley Periodicals, Inc. J Polym Sci Part
B: Polym Phys 42: 809 – 819, 2004
Keywords: vegetable oil; polyol; polyurethane; structure

INTRODUCTION the number of hydroxyl groups in the polyols,


their distribution in the oil (zero, one, two, three,
Oils are an excellent renewable source for chem- or more per fatty acid), and the position in the
ical industry and particularly for polymer appli- fatty acid chain (in the middle or closer to the end
cations. Because they are relatively unreactive, of the chain). Thus, the properties are expected to
they must be functionalized to serve as building depend on the type of vegetable oil. In this work
blocks for polymers. Introduction of hydroxyl we studied the effect of structure of six vegetable
groups at the position of double bonds opens the oils, canola, midoleic sunflower, soybean, linseed,
whole area of applications in polyurethanes. If the sunflower, and corn oil, on properties of polyols
triglyceride structure were preserved, polyols and resulting polyurethane networks. Four oils
usually would have functionality greater than (canola, sunflower, soybean, and corn) have sim-
two, giving with diisocyanates thermosetting ilar content but different distribution of double
polyurethane networks. Network properties bonds. Linseed oil has the highest and midoleic
would depend on the crosslinking density, that is, sunflower the lowest double bond content of all
oils used in this work.
The transformation of the double bonds of trig-
Correspondence to: Z. Petrovic (E-mail: zpetrovi@ lycerides to hydroxyls and their application in
pittstate.edu)
Journal of Polymer Science: Part B: Polymer Physics, Vol. 42, 809 – 819 (2004)
urethane chemistry have been the subjects of
© 2004 Wiley Periodicals, Inc. many studies.1–9 Our group paid special attention
809
810 ZLATANIĆ ET AL.

suring the hydroxyl content, molecular weight


and distribution, density, viscosity, and melting
points. Polyurethanes were prepared by reacting
the polyols with 4,4⬘-diphenylmethane diisocya-
nate (MDI). The reactive sites in polyols having
different fatty acids are located on the 9th or 10th
carbon in oleic, linoleic, and linolenic, but the last
two have additional OH groups on the 12th or
13th and linolenic may have an additional OH
group on the 15th or 16th carbon as depicted in
Figure 1.
Crosslinking density of polymer networks was
assessed from swelling in toluene. We used differ-
Figure 1. Reactive sites on different fatty acid ential scanning calorimetry (DSC), thermome-
chains. chanical analysis (TMA), dielectric analysis
(DEA), dynamic mechanical analysis (DMA), and
mechanical methods to study the physical, ther-
to the functionalization of soybean oil and the mal, dielectric, and mechanical properties of poly-
structure–properties relationship in polyols and urethanes.
the resulting polyurethanes.10 –16 The structure of
natural oils regarding the type, composition, and
distribution of fatty acid residues in the triglyc- EXPERIMENTAL
eride molecules is complex.17 Every oil or fat has
a characteristic fatty acid profile which is unique Materials
to the type of oil. Variability in composition is
expected even in the same type of oil, depending Six vegetable oils of refined, bleached, and de-
on the local weather conditions, soil, and plant odorized grade from canola, soybean, midoleic
characteristics. The principal variation in fatty sunflower, sunflower, corn, and linseed, together
acids composition of the oils results from varia- with fatty acid profiles (Table 1) were supplied by
tions in chain length and the degree of unsatura- Archer Daniels Midland. Hydrogen peroxide was
tion. As a consequence, there is variation in the purchased from Sigma as 30% w/w solution. Ion-
length of elastically active network chains exchange resin Amberlite IR-120 in acid form was
(EANC) and dangling chains (DC) in polyure- supplied from Supelco. Tetrafluoroboric acid (48%
thane networks, obtained from oil-based polyols. solution in water) was purchased from Aldrich.
Dangling chains are the elastically inactive part Toluene [high-performance liquid chromatogra-
of the polymer structure, which do not support phy (HPLC) grade], methanol (99%), glacial acetic
stress when the sample is under load, but rather acid (certified A.C.S. Plus), ethyl acetate (HPLC
act as a plasticizer. They result from saturated grade), methylene chloride (HPLC-GC/MS grade),
fatty acids and the portion of fatty acids between anhydrous sodium sulfite, and anhydrous sodium
the functional group (usually at the 9th or 10th C sulfate were purchased from Fisher Scientific. Ion
atom) and the chain end. The reactive sites in exchange resin Lewatite MP-64 was purchased
polyols with different fatty acids are located on from Bayer. Mondur M obtained from Bayer is
the 9th or 10th carbon in oleic, linoleic, and lin- 4,4⬘-diphenylmethane diisocyanate. It was puri-
olenic, but the last two have additional OH fied by distillation under reduced pressure.
groups on 12th or 13th, and linolenic may have an
additional OH group on the 15th or 16th carbon
Methods
as depicted in Figure 1.
Thus, to draw conclusions about the effect of We determined iodine value, as the measure of
structure on the properties, it is important to concentration of double bonds, according to the
know the exact composition of the oil. Table 1 lists Hanus method.18 Epoxy oxygen content determi-
the fatty acid profiles of the six vegetable oils used nation was carried out according to the standard
in this study. The polyols were synthesized by a procedure for oils and fats.19 The hydroxyl values
two-step process: epoxidation followed by ring of the polyols were determined according to a
opening. The polyols were characterized by mea- modified ASTM titration method (D 1957-86)
POLYOLS, POLYURETHANES FROM VEGETABLE OIL 811

Table 1. Fatty Acid Profilea in Vegetable Oils

Midoleic Soybean Linseed Sunflower


Sample ID Canola Oil Sunflower Oil Oil Oil Oil Corn Oil

Caproic C6:0 0.02


Lauric C12:0 0.02
Myristic C14:0 0.04 0.09 0.05 0.03 0.06
Palmitic C16:0 4.09 4.11 10.88 5.48 5.59 10.81
Palmitoleic C16:1 0.21 0.05 0.07 0.06 0.06 0.11
Heptadecanoic C17:0 0.05 0.04 0.09 0.05 0.05 0.09
Heptadecenoic C17:1 0.09 0.04 0.04 0.04 0.01 0.03
Stearic C18:0 1.86 3.79 4.18 3.52 4.46 1.85
Elaidic C18:1n9t 0.08 0.09 0.05 0.04
Oleic C18:1n9c 56.11 57.15 22.69 18.88 22.18 27.25
C18:1 cis 4.67 1.41 2.21 1.25 0.92 0.95
C18:2 trans 0.02 0.02
C18:2 9c12t 0.16 0.19 0.15 0.06 0.31 0.29
C18:2 12c9t 0.14 0.18 0.13 0.27 0.27
Linoleic C18:2cc 21.01 31.40 52.41 16.10 64.22 56.96
C18:2 cis 0.01 0.05
Arachidic C20:0 0.60 0.25 0.26 0.12 0.24 0.30
C18:3 trans 0.04
C18:3n6t 0.53 0.02 0.17 0.19 0.02 0.04
Eicosenoic C20:1c 1.70 0.21 0.30 0.14 0.19 0.20
Linolenic C18:3n3c 7.89 0.31 6.06 53.73 0.42 0.65
Heneicosanoic C21:0 0.01 0.01 0.02 0.02
Eicosadienoic C20:2cc 0.06 0.02
Behenic C22:0 0.17 0.46 0.14 0.04 0.41 0.03
Eicosatrienoic C20:3n6c 0.05
Erucic C22:1 0.12
Lignoceric C24:0 0.05 0.03 0.06
Nervonic C24:1 0.05
Total trans FAs 0.95 0.50 0.52 0.25 0.65 0.60
Total C18:1 FAs 60.85 58.64 24.95 20.14 23.15 28.19
Total C16:1, C20:1,
C22:1, C24:1 FAs 2.08 0.26 0.37 0.20 0.25 0.31
Total C18:2 FAs 21.33 31.79 52.72 16.21 64.80 57.52
Total C20:2, C22:2 0.06 0.00 0.02 0.00 0.00 0.00
Total C18:3 FAs 8.46 0.33 6.23 53.92 0.44 0.69
Total C20:3 FAs 0.05 0.00 0.00 0.00 0.00 0.00
Total saturated FAs 6.78 8.69 15.60 9.27 10.85 13.10
Totals 99.61 99.71 99.89 99.74 99.49 99.81

*Wt % of fatty acids.

with isopropanol instead of ethanol. The number- rate of the tetrahydrofuran eluent was 1 mL/min
average molecular weights of polyols were deter- at room temperature. Three Styragel HR columns
mined with a vapor pressure osmometer (Osmo- from Waters, covering an MW range from 102 to
mat 070; UIC, Inc.). Measurements were per- 106, were used. We performed calibration using
formed at 60 °C in toluene. Benzil was used as the nine polystyrene standards with narrow molecu-
calibration standard. lar weight distribution.
We used a Waters gel permeation chromato- The polyol viscosities were measured at 27 °C
graph, consisting of a 510 pump, 410 differential with the Rheometric SR-500 Dynamic Stress Rhe-
refractometer, and the data collection system, to ometer in the stress-controlled mode, between
determine the molecular weight distribution of parallel plates 25 mm in diameter and with a gap
epoxidized vegetable oils and polyols. The flow of 0.20 mm. We measured the densities of polyols
812 ZLATANIĆ ET AL.

excess. The polyol and MDI were preheated sep-


arately at 60 °C, then mixed and poured into the
molds preheated at 100 °C. Curing was carried
out at 100 °C overnight.

Figure 2. Conversion of epoxidized oil to a polyol.

using a Paar DMA5000 density meter at 23 °C. RESULTS AND DISCUSSION


The densities of polyurethane networks were
measured by immersion in water according to Composition of Vegetable Oils
ASTM 792. We used the Thermal Analyst system
Table 1 displays the composition of selected veg-
from TA Instruments, consisting of the Controller
etable oils, obtained from GC-MS analysis. The
3100 with a DSC 2910 module, TMA 2940 mod-
dominant constituents (over 80%) of all the oils
ule, TGA 2050 module and dielectric analyzer
are C18 fatty acids. The differences between the
DEA 2070, to study the thermal transitions and
oils arise mainly from the distribution of the C18
stability and dielectric properties of the vegetable
acids containing different numbers of double
oil– based polyols and polyurethanes. Dynamic
bonds: stearic, C18:0, oleic, C18:1, linoleic, C18:2
mechanical analysis of polyurethanes was carried
and linolenic, C18:3. Therefore, six oils could be
out on a DMA 2980 Dynamic Mechanical Ana-
divided into three groups. Canola and midoleic
lyzer from TA Instruments. The storage (E⬘) and
sunflower oil contain about 60% of oleic acid and
loss (E⬙) moduli and tan ␦ were recorded from ⫺80
lower iodine numbers (Table 2). The dominant
°C to 100 °C, at the heating rate of 5 °C/min and
fatty acid in soybean, sunflower, and corn oils is
frequency of 10 Hz.
linoleic, whereas highly unsaturated linseed oil
We employed a Q-Test 2 tensile machine from
contains more than 50% linolenic.
MTS to determine tensile strength and elongation
Using the composition of oils, we calculated
at break. The sample length was 50 mm and the
some of the parameters of oils and derivatives and
extension rate 50 mm/min. Flexural moduli were
compared them with the experimental results
obtained on the same tensile machine using 70-
(Table 2). Molecular weights of the oils differ very
mm-long, 25-mm-wide, and 2-mm-thick speci-
little because the C18 fraction is similar in all the
mens. Notched Izod impact strength was deter-
oils. Table 2 also gives molecular weights and
mined on a Resil Impactor from CEAST, Model
hydroxyl numbers of the polyols derived from
P/N 6957.000.
these oils. Theoretical MW and OH numbers are
based on assumption that no dimers and trimers
Preparation of Polyols are formed and that one hydroxyl group is formed
The polyols were synthesized by epoxidation of per double bond. Iodine values as a measure of
oils and subsequent ring opening of the epoxy the double bond content of oils, calculated from
groups to hydroxyls. The epoxidation of vegetable the oils’ composition, are in good agreement with
oils was carried out in toluene, with peracetic acid the experimentally determined ones.
formed in situ as described elswhere.16,20 Epoxy
yield was fairly uniform for all the oils (91–94%)
as determined by measuring epoxy oxygen con- Characterization of Polyols
tent. The polyols were synthesized from epoxi-
The conversion of the double bonds to epoxy
dized oils by oxirane ring opening with boiling
groups during the epoxidation reaction was rela-
methanol in the presence of a tetrafluoroboric
tively high for all the oils, varying from 91% to
acid catalyst.12,13 Total conversion of double
94%. Although canola oil is closer in unsaturation
bonds to hydroxyls varied from 75.5% to 83.7%, as
to midoleic sunflower oil, owing to more efficient
displayed later in Table 3.
conversion of double bonds to hydroxyls, canola
Figure 2 illustrates the epoxy ring opening in
polyol became closer in hydroxyl numbers to soy-
the presence of an acid catalyst and methanol.
bean, sunflower, and corn polyols. From iodine
values before and after epoxidation we found that
Synthesis of Polyurethanes double bonds reacted quantitatively. This also in-
Polyurethanes were obtained by reacting the dicated that 5– 8% of epoxy groups were lost in
polyols with MDI. MDI was used in 2% molar side reactions. Gel permeation chromatography
POLYOLS, POLYURETHANES FROM VEGETABLE OIL 813

Table 2. Calculated and Experimentally Determined Properties of Vegetable Oils and Oil-Based Polyols

Midoleic
Property Canola Sunflower Soybean Linseed Sunflower Corn

Oil MW 879 879 872 871 873 871


Functionality 3.94 3.71 4.49 6.44 4.63 4.37
(Iodine value, mg I2/100 g) calc 113.8 107.0 130.5 187.6 134.6 127.3
(Iodine value, mg I2/100 g) exp 114.1 107.6 126.6 182 133.3 125.2
Epox. oil MW 942 938 944 974 947 941
(Epoxy oxygen content, %) calc 6.69 6.32 7.60 10.58 7.82 7.43
(Epoxy oxygen content, %) exp 6.29 5.87 6.95 9.5 7.07 6.72
Polyol MW 1068 1057 1088 1180 1096 1081
(Polyol MW) exp 1123 1191 1249 1274 1302 1180
(OH#, mg KOH/g) calc 207.0 196.7 231.4 306.2 237.1 226.8
(OH#, mg KOH/g) exp 173.6 163.5 179.3 247.8 177.8 179.0

(GPC) analysis of epoxidized products did not were ascribed to different crystalline structures
show the presence of oligomers. present in the polyols. DSC thermograms of the
The number-average molecular weight of poly- polyols are displayed in Figure 4.
ols, determined by vapor pressure osmometry,
was higher than that calculated from the compo- Functionality of Polyols from Vegetable Oils
sition of oils, presumably owing to the presence of
oligomers in polyols. This was especially notice- Number-average functionalities of polyols ob-
able in the case of the polyol from sunflower oil, tained from vegetable oils were calculated from
which had the highest content of oligomers (25% the hydroxyl number and theoretical molecular
compared with 15–18% for other samples) and the weight:
lowest double bond to hydroxyl conversion (Table
3). The extent of side reactions during formation Mn䡠OH#
of polyols depends on several factors such as the fn⫽
56110
catalyst concentration, temperature, time, and
methanol– epoxy molar ratio. The side reaction in The functionality of the polyols is lower than
the hydroxylation stage is mainly polymerization, the number of double bonds per molecule of the
that is, formation of dimers and trimers through starting oils because a part of the functional
the epoxy ring opening by newly formed OH groups was lost in the side reactions during the
groups. The degree of intermolecular reactions preparation of the polyols. The true functionality
can be estimated from the GPC chromatograms of of the polyols is increased when oligomers are
polyols (Fig. 3). The percentage of high-molecular- present. Functionality of dimers and trimers is
weight species significantly affects the viscosity of almost twice and three times that of monomeric
polyols. In the case of the linseed oil-based polyol, species.
however, a high OH content was primarily re-
sponsible for the high viscosity value.
We used the group contribution method,21 Characterization of Polyurethanes
based on the additivity of molar volumes of chem-
Swelling of Networks and Sol Fraction
ical groups present in the compound, to calculate
polyol densities. We used the values of the group Swelling degree of the vegetable oil– based poly-
contribution to molar volume reported by Fedors. urethanes in toluene at room temperature varied
Theoretical polyol functionality was included in from 40% for the highly crosslinked linseed oil–
the calculation. Experimental densities of polyols based polyurethane network to 96% for the mido-
were 6.4 – 8.2% lower than calculated, as seen in leic sunflower oil– based one (Table 4).
Table 3. Crosslinking density can be extracted from
All polyols from the vegetable oils exhibited swelling data. The Flory-Rehner theory allows
similar thermal behavior. Two to three weak calculation of the concentration of EANCs, ␯ e, or
melting peaks between ⫺40 and 0 °C (Table 3) molecular weight of network chains, Mc, but the
814
ZLATANIĆ ET AL.

Table 3. Properties of Polyols Based on Vegetable Oils

OH#, Wt % of Viscosity Calc. Exp. Tm


mg Conversion Mn Mn Oligomers at 27 °C, Density Density,b (DSC),
Starting Oil KOH/g % (VPO) Calc.a Functionality from GPC Pa 䡠 s g/cm3 g/cm3 °C

Canola 173.6 83.7 1123 1068 3.30 17.04 2.94 1.0608 0.9968 ⫺36
⫺16
Midoleic
sunflower 163.5 82.7 1191 1057 3.08 16.69 1.85 1.0532 0.9885 ⫺38
⫺18
0
Soybean 179.3 79.4 1249 1088 3.47 18.56 3.35 1.0779 1.0066 ⫺32
⫺15
3.6
Linseed 247.8 82.7 1274 1180 5.21 14.85 18.2 1.1375 1.0513 ⫺24
⫺8
Sunflower 177.8 75.5 1302 1096 3.47 25.08 4.08 1.0823 1.0090 ⫺34
⫺15
2.5
Corn 179.0 80 1180 1081 3.44 15.7 2.71 1.0741 1.0032 ⫺35
⫺16
⫺0.3
a
Assuming complete double bonds to hydroxyls conversion and no formation of oligomers.
b
At 23 °C.
POLYOLS, POLYURETHANES FROM VEGETABLE OIL 815

its structure, giving meaningless values for Mc


and ␯e. The absolute values for Mc are underesti-
mated compared with ideal values being about
930 for networks from trifunctional and about 780
for four functional polyols and MDI, but they give
the sense of the order of magnitude.
The sol fraction was determined after multiple
extractions with toluene. We calculated the con-
tent of the soluble part (Ws) from the difference in
the weight of the dry sample before (w1) and after
(w2) multiple extractions:
Figure 3. GPC chromatograms of polyols from vege-
table oils.
W s ⫽ (w1 ⫺ w2 )/w1

result would depend on the adopted model of rub- The presence of the sol indicates imperfect
ber elasticity: crosslinking in polyurethanes. The sol content in
polyurethanes from canola and midoleic sun-
[⫺V1 共A␾21/3 ⫺ 2B␾2/f兲] flower oil was approximately two times lower
1/ ␯ e ⫽ Mc/␳2 ⫽ than that in the networks from soybean, sun-
关ln(1 ⫺ ␾2) ⫹ ␾2 ⫹ ␹12⌽22]
flower, and corn oil. The concentration of EANCs
Here, ␳2 is the density of the dry polymer, V1 is is sensitive to network imperfections.22 In the
the molar volume of the solvent, ⌽2 is the volume model network based on triolein, for example, the
fraction of the polymer in the swollen sample, f is presence of only 0.5% of sol translates to the de-
the functionality of the network branch points, crease of the crosslinking density of 40 –50% of its
and ␹12 is the polymer–solvent interaction param- maximum value.23 Lower crosslinking density
eter. Factors A and B within the Junction Fluc- may be caused by incomplete chemical reaction or
tuation theory of Flory (JFF theory) have the cyclization reactions during crosslinking. It has
following limits: A ⫽ (f ⫺ 2)/f, B ⫽ 0 (phantom been shown24 that the probability of the forma-
network); A⫽1, B⫽1 (affine network). tion of elastically inactive network structures
Application of this equation for determining ␯e during formation of the polyurethanes increases
and Mc requires knowledge of the interaction pa- with increasing functionality of the polyols. Table
rameter from independent measurements. The 4 presents the swelling characteristics of polymer
polymer–solvent interaction parameter ␹12 was networks from the vegetable oils, as well as cor-
estimated from the solubility parameters for the responding molecular weights of the network
solvent, ␦1, and the polymer network, ␦2: chains.

共 ␦ 1 ⫺ ␦ 2兲 2V1
␹ 12 ⫽
RT

The solubility parameter for toluene was taken to


be 18.3 J ⁄2/cm ⁄2. The solubility parameters of
1 3

the polyurethane networks were calculated using


the group contribution method21 to the molar at-
traction constant, F (Hoy), and the solvent molar
volume, V1. Table 4 presents the ␦2 and ␹12 values
for polymer networks, as well as corresponding
molecular weights of the network chains. These
values are approximate at best, but the calculated
crosslinking densities give a basis for comparison
of different samples. However, the calculated ␹12
for the linseed-based polyurethane network was Figure 4. DSC thermograms of polyols from different
very high owing to the higher content of MDI in vegetable oils.
816 ZLATANIĆ ET AL.

Table 4. Swelling Properties of Polyurethanes from Vegetable Oils in Toluene

Solubility Affine Model


Swelling Sol Parameter Polymer–Solvent
Density, Degree,a Fraction, ␦2, J1/21 Interaction ␯e, 103, Mc, g
g cm⫺3 % % cm⫺3/2 Parameter ␹12 mol cm⫺3 mol⫺1

Canola 1.087 80.98 1.11 21.42 0.422 1.764 616


Midoleic sunflower 1.078 96.49 1.47 21.19 0.361 1.573 685
Soybean 1.094 80.45 2.66 21.58 0.466 1.542 710
Linseed 1.144 40.40 0.68 23.08 0.987
Sunflower 1.101 80.16 2.52 21.58 0.466 1.534 718
Corn 1.092 85.56 2.84 21.56 0.461 1.408 776
a
Based on the weight of the swollen and original (dry) sample.

Glass Transition Temperature of Polyurethanes these polyurethanes. For all of the polymers, Tg
from Vegetable Oils values taken from the change of slope on thermo-
mechanical curves were slightly lower than those
We used four different methods to determine Tg to
determined by DSC (Table 5). The linear coeffi-
cross-check each other, because both DSC and
cients of thermal expansion (LCTE) in the rub-
TMA do not show always a clear transition. Tg
bery state (measured at 50 °C above Tg) were
values from DMA and DEA are frequency depen-
around 400 ⫻ 10⫺6 K⫺1 for all samples except for
dent and temperature control is not as good as in
DSC, but reading the position of maxima is un- the linseed oil–polyurethane (320 ⫻ 10⫺6 K⫺1),
mistakable. TMA gives values close to those of pointing to a higher crosslinking density of the
DSC. Figure 5 shows the DSC curves of the poly- latter. High LCTE values indicating high free
urethanes. The Tg, taken as a temperature of volume in these samples may be related to the
inflection on DSC curves, was around 30 °C for plasticizing effect of dangling chains. LCTE below
networks from soybean, corn, sunflower, and Tg was 85 ⫻ 10⫺6 K⫺1 for the linseed oil– based
canola oil, whereas midoleic sunflower oil– based PU and 100 –110 ⫻ 10⫺6 K⫺1 for other polyure-
polymer displayed a Tg at 24 °C. Highly thanes, which exceeds the LCTE of most linear
crosslinked polyurethane from linseed oil had a glassy thermoplastic polymers (50 –90 ⫻ 107mi-
⫺1
nus;6 K ), indicating higher free volume in the
Tg at 77 °C.
Thermomechanical analysis of polyurethanes glassy state as well, which may relax with time
showed unusually high thermal expansion of and cause physical aging and increasing density.
Glass transition temperatures, determined by
DMA from the maxima on G⬙ vs T are listed in
Table 5. DMAs generally give higher values of Tg
compared with the DSC and TMA method, but
the trend was the same: the lowest Tg was re-
corded in the polyurethane from midoleic sun-
flower oil and the highest was seen in the linseed
oil– based polymer.
Figure 6 displays tan ␦ peaks for different poly-
urethanes. The width at the half height of the
dumping factor peak, as a measure of crosslinking
density, is relatively similar for all the samples
except that from linseed oil. The shape and posi-
tion of tan ␦ peak are known to be affected by
many factors such as crosslinking density and
distribution of crosslinking densities in inhomo-
geneously crosslinked materials, content of plas-
Figure 5. DSC curves of polyurethanes from vegeta- ticizer or filler, and orientation of molecular struc-
ble oils. ture. Polymer networks based on triglycerides
POLYOLS, POLYURETHANES FROM VEGETABLE OIL 817

Table 5. Thermal, Dielectric, and Mechanical Properties of Polyurethanes from Vegetable Oils

Tg, °C
Tensile Elongation Flexural Impact
DMA DEA at E aa Strength, at Break, Modulus, Resistance,
DSC TMA G⬙ max 10 Hz kJ/mol MPa % MPa J/m

Canola 32 30 40 64 187 22.9 131 353 180


Midoleic sunflower 24 21 33 47 172 14.8 168 10 408
Soybean 31 30 35 67 181 20.2 108 312 178
Linseed 77 74 77 209 56.3 8 2015 21
Sunflower 32 29 38 69 193 21.7 107 443 198
Corn 30 29 38 64 192 17.7 122 309 221
a
Activation energy of the glass transition.

contain dangling chains arising from the polyol. higher for samples with lower Ea. As would be
Although dangling chains are uniform in length expected, the highest mobility was found in the
in the triolein-based networks,25 there is a distri- midoleic oil– based polyurethanes and the lowest
bution of lengths of dangling chains in networks in the linseed oil– based ones. A linear relation-
from natural oils as a consequence of the hetero- ship between activation energy and crosslinking
geneity of their composition density was found in similar systems.26
Figure 7 shows the change of permittivity and
loss factor for soybean-based polyurethane. The Mechanical Properties of Polyurethane Networks
activation energy of the glass transition, Ea, was
Tensile strengths of all the polyurethane samples
calculated from the loss factor–temperature
fell in the range of 15–23 MPa except for the
curves and the shift of temperature of the max-
polyurethane based on linseed oil, which dis-
ima, T, with frequency, f:
played about three times higher strength. Two
factors affected tensile strength: the crosslinking
lnf ⫽ lnf0 ⫺ Ea/RT density of samples and the physical state of the
material at the test temperature. Because the Tg
R is the gas constant. The activation energy in- of all samples except the linseed-based one was in
creased linearly with crosslinking density, from the vicinity of the testing temperature, they were
172 kJ/mol for the polyurethane from mid– oleic tested in the leathery state (transient from glassy
sunflower oil to 210 kJ/mol for the sample from to rubbery), which resulted in somewhat lower
linseed oil. Activation energy of the glass transi- strengths and moduli and high elongations at
tion reflects the mobility of the chains, which is

Figure 6. Tan ␦ vs temperature of polyurethanes Figure 7. Permittivity and loss factor vs temperature
from different vegetable oils. for polyurethane from soybean polyol and MDI.
818 ZLATANIĆ ET AL.

Figure 8. Effect of the polyol OH number on theoret-


ical ␯e and Mc of polyurethane networks, calculated
Figure 10. Effect of OH number of polyols on tensile
under the assumption of complete conversion.
strength and elongation at break of polyurethanes.

break. The linseed oil– based sample exhibited


brittle fracture and very low elongation at break relationship was observed in tensile strength and
compared with other polyurethanes, but tensile flexural modulus vs the OH number, but elonga-
strength and modulus were comparable with tion at break and impact strength displayed ex-
those of strong plastic materials. ponential decay (Figs. 10 and 11). In our case, one
sample (from midoleic sunflower oil) was in the
Dependence of Properties on Crosslinking Density rubbery state and all others in either the leathery
or glassy state at the testing temperature. Elon-
The hydroxyl number of the polyols is a linear gation at break and impact resistance are usually
function of the crosslinking density of the poly- much higher in the rubbery state.
urethanes, expressed as the number of elastically
active network chains per unit volume, and in-
versely proportional to the average molecular CONCLUSIONS
weight between crosslinks, Mc, at complete con-
version (Fig. 8) Thus, it is useful to examine the We converted six vegetable oils to polyols and
dependence of properties of the polyurethanes on used them to prepare polyurethanes. The oils
the OH number of corresponding polyols. Figure 9
shows linear relationships between Tg, deter-
mined by DSC, and OH numbers. A nearly linear

Figure 11. Effect of OH number of polyols on flexural


Figure 9. Dependence of DSC Tg of polyurethanes on modulus and Notched Izod impact strength of polyure-
the OH number of polyols. thanes.
POLYOLS, POLYURETHANES FROM VEGETABLE OIL 819

were divided into three groups: low unsaturated 6. Bilyk, A.; Monroe, H. A., Jr.; Saggese, E. J.; Zubil-
midoleic sunflower and canola, highly unsatur- laga, M. P.; Wrigley, A. N. J Am Oil Chem Soc 1975,
ated linseed, and three oils with intermediate 52, 289 –292.
unsaturation. The polyol functionality and 7. Binlayo, D. L. NSTA Tech J 1985, 13–16.
8. Madrigal, R. V.; Bagby, M. O.; Pryde, E. H. J Am
crosslinking density of polyurethanes were re-
Oil Chem Soc 1988, 65, 1508 –1510.
lated to the double bond content in oils, except 9. Dahlke, B.; Hellbardt, S.; Paetow, M.; Zech, W. H.
that canola polyol migrated to the group of inter- J Am Oil Chem Soc 1995, 72, 349 –353.
mediate functionality. Functionality of the poly- 10. Petrović, Z. S.; Javni, I.; Guo, A. Proc. Polyure-
ols resulting from oils with intermediate unsat- thanes EXPO ’98, Dallas, TX, Sept 1998, pp 559 –
uration was close to 3.5, that from linseed oil was 562.
5.2, and midoleic sunflower had functionality 11. Petrović, Z. S.; Guo, A.; Zhang, W. J Polym Sci Part
around 3.0. Canola, corn, soybean, and sunflower A: Polym Chem 2000, 38, 4062– 4069.
oils gave polyurethane resins of similar crosslink- 12. Petrović, Z. S.; Guo, A.; Javni, I. U.S. Patent 6,
ing density, and thus similar Tg values and me- 107,433, Aug 22, 2000.
chanical properties despite somewhat different 13. Guo, A.; Cho, Y.; Petrović, Z. S. J Polym Sci Part A:
Polym Chem 2000, 38, 3900 –3910.
distributions of fatty acids. Linseed oil– based
14. Petrović, Z.; Guo, A.; Javni, I.; Zhang, W. Proc
polyurethanes had higher crosslinking density Polyurethanes EXPO 2000, Boston, MA, Oct. 8 –11,
and thus higher mechanical properties, whereas 2000.
midoleic oil gave softer polyurethanes character- 15. Guo, A.; Javni, I.; Petrović, Z. J Appl Polym Sci
ized by lower Tg and lower level of strength but 2000, 77, 467– 473.
higher elongation at break. The differences in 16. Petrović, Z. S.; Zlatanić, A.; Lava, C. C.; Sinadi-
properties arose primarily from different nović-Fišer, S. Eur J Lipid Sci Tech 2002, 104,
crosslinking densities of polyurethane networks 293–299.
and less from the position of functional groups in 17. Tan, C. P.; Che Man, Y. B. J Am Oil Chem Soc
the fatty acids in polyols. Mobility of chains in the 2000, 77, 143–155.
networks was inversely proportional to the 18. In Standard Methods for the Analysis of Oils, Fats
and Derivatives; Paquot, C.; Hautfenne, A., Eds.
crosslinking density.
Blackwell Scientific Publications: London, 1987, pp
88 –93.
This material is based on work supported by the Coop- 19. In Standard Methods for the Analysis of Oils, Fats
erative State Research, Education and Extension Ser- and Derivatives; Paquot, C.; Hautfenne, A., Eds.
vice, U.S. Department of Agriculture, under Agreement Blackwell Scientific Publications: London, 1987, pp
99-35504-7873. 118 –119.
20. Crivello, J. V.; Narayan, R. Chem Mat 1992, 4,
692– 699.
REFERENCES AND NOTES 21. van Krevelen, D. W. Properties of Polymers;
Elsevier: New York, 1990.
22. Dušek, K. In Networks from Telechelic Polymers.
1. Maerker, G.; Haeberer, E. T.; Ault, W. C. J Am Oil Synthesis and Application; Goethals, E. J., Ed.;
Chem Soc 1964, 41, 585–588. CRC Press: Boca Raton, 1989, pp 289 –360.
2. Scholnick, F.; Saggese, E. J.; Wrigley, A. N.; Ault, 23. Dušek, K.; Dušková-Smrčková, M.; Zlatanić, A.;
W. C.; Monroe, H. H. J.; Zubillaga, M. J. Am Oil Petrović, Z. S. 223rd ACS National Meeting, Or-
Chem Soc 1968, 45, 76 –77. lando, FL, April 7–11, 2002, pp 381–382.
3. Scholnick, F.; Saggese, E. J.; Wrigley, A. N.; Riser, 24. Zlatanić, A.; Petrović, Z.; Dušek, K. Submitted.
G. R. J Am Oil Chem Soc 1970, 47, 180 –182. 25. Zlatanić, A.; Petrović, Z. S.; Dušek, K. Biomacro-
4. Khoe, T.; Otey, F. H.; Frankel, E. N. J Am Oil Chem molecules 2002, 3, 1048 –1056.
Soc 1972, 49, 615– 618. 26. Petrović, Z. S.; Zhang, W.; Zlatanić, A.; Lava, C. C.
5. Lyon, C. K.; Garrett, V. H.; Frankel, E. N. J Am Oil 223rd ACS National Meeting, Orlando, FL, April
Chem Soc 1974, 51, 331–334. 7–11, 2002, pp 377–378.

You might also like