You are on page 1of 6

Influence of Plasticizers on Thermal and Mechanical Properties and Morphology

of Soy-Based Bioplastics
Praveen Tummala,

Wanjun Liu,

Lawrence T. Drzal,

Amar K. Mohanty,

and Manjusri Misra*


,
Composite Materials & Structures Center and School of Packaging, Michigan State UniVersity, East Lansing,
Michigan 48824
Bioplastics from soy protein and biodegradable polyester amide with plasticizers including glycerol, D-sorbitol,
and blends of these two have been made using extrusion and injection molding. The thermal properties,
mechanical properties, and morphologies of the soy-based bioplastics were evaluated with a dynamic mechanical
analyzer (DMA), a united testing system (UTS), and an environmental scanning electron microscope (ESEM).
The influence of plasticizers on the physical properties of the soy-based bioplastics was determined. It was
found that sorbitol-plasticized soy-based bioplastic (SSBP) had a higher tensile modulus and tensile strength
than glycerol-plasticized soy-based bioplastic (GSBP). GSBP had the highest impact strength, whereas SSBP
had the highest thermal stability. The mixed plasticized soy-based bioplastic (MSBP) that was obtained using
a mixture of glycerol and D-sorbitol showed an intermediate range of tensile modulus and tensile strength
values in comparison to those of SSBP and GSBP. The ESEM results suggested that SSBP had brittle fracture
features, whereas GSBP had local ductile fracture features, which is consistent with the results obtained for
the mechanical properties. The glass transition temperatures from DMA measurements of the soy protein and
polyester amide in the soy-based bioplastics differed from those of the neat components. Using Fox equation
calculations, this was interpreted as compatibility between the soy protein and polyester amide, which causes
thermal and mechanical changes in different soy-based bioplastics.
Introduction
Soy proteins are generally regarded as one of the most
important groups of natural biopolymers used to produce
biodegradable materials, and their utilization will reduce the
dependence and consumption of nonrenewable resources. Soy
proteins are complex macromolecules containing 20 amino
acids
1
that supply available sites to interact with a plasticizer.
Soy proteins can be converted to soy protein plastics through
extrusion with a plasticizer or cross-linking agent.
2-4
Common
plasticizers used in the manufacture of soy protein plastics
include glycerol, ethylene glycerol, propylene glycerol, 1,2-
butanediol, 1,3-butanediol, poly(ethylene glycol), sorghum wax,
and sorbitol.
3-6
Recently, Wang and Zhang
7
used anionic
waterborne polyurethane as a new plasticizer to prepare soy
protein plastic that exhibited good mechanical strength, water
resistance, and thermal stability. In addition, Zhong and Sun
8,9
found that sodium dodecyl sulfate (SDS) and guanidine
hydrochloride (GuHCL) acted as plasticizers for the soy protein
11S, hence leading to an improvement of its tensile strength,
elongation, and water resistance.
Although the mechanical properties of soy protein plastic can
be controlled and optimized by adjusting processing parameters
such as the molding temperature and pressure and the initial
moisture content,
10-12
the application of soy protein plastic is
limited because of its low strength. Therefore, blending soy
protein with biodegradable polyester is a way to form more
effective soy-based bioplastics. Currently, the biodegradable
polyesters being used in blends with soy plastics include
polyester amide, polycaprolactone, and poly(tetramethylene
adipate-co-terephthalate),
13-15
whose processing windows match
that of soy protein plastic.
Because soy protein and biodegradable polyester have dif-
ferent polarities, a compatibilizer is needed to reduce the domain
size of the soy protein and increase the interfacial interactions
between the soy protein and the biodegradable polyester. Zhong
and Sun
16
used methylene diphenyl diisocyanate (MDI) as a
compatibilizer to produce blends of soy protein with poly-
(caprolactone) (PCL) because it could react with both soy protein
and PCL. As a result, the mechanical properties and the water
resistance of soy protein/PCL blends were enhanced. Mungara
et al.
17
used a poly(vinyllactam) (PVL) as a compatibilizer to
blend soy protein and polyesters and observed that soy-protein-
based blends showed high tensile strength and modulus as well
as low water absorption. John and Bhattacharya
18
used maleic
anhydride grafted polyesters as compatibilizers to enhance the
compatibility of soy protein and biodegradable polyesters. As
a result, the blended products showed enhanced mechanical
properties as well as water and oil resistance.
Natural polymer such as soy protein cannot form a plastic
material without a plasticizer, but any plasticizer is expected to
reduce the efficiency of the compatibilizer in the natural polymer
and biodegradable polymer blends.
19
Therefore, only plasticizers
and no compatibilizers were used to prepare soy-based bioplas-
tics in the present article.
Soy protein is available in three different forms: soy flour
(52% protein), soy concentrate (65% protein), and soy isolate
(90% protein). The lower the protein content, the lower the cost.
In the present article, soy flour was selected for study because
it is very cost-effective and has properties comparable to those
of the other forms of soy protein. The commercial biodegradable
polyester used to blend with the soy flour was a polyester amide,
a random copolymer of aliphatic polyester and Nylon-6 as
shown in Chart 1, because of the potential compatibility between
its amide groups and the soy protein plastic. To investigate the
influence of different plasticizers on the thermal and mechanical
* To whom correspondence should be addressed. Address: Compos-
ite Material & Structure Center, 2100 Engineering Building, Michigan
State University, East Lansing, MI 48824. E-mail: misraman@
egr.msu.edu. Tel.: +1-517-353-5466. Fax: +1-517-432-1634.

Composite Materials & Structures Center.

School of Packaging.
7491 Ind. Eng. Chem. Res. 2006, 45, 7491-7496
10.1021/ie060439l CCC: $33.50 2006 American Chemical Society
Published on Web 10/04/2006
properties and morphology of the soy-based bioplastics, glycerol,
sorbitol and a mixture of the two were used as plasticizers (as
shown Chart 1).
Experimental Section
Materials. Soy flour (defatted soy flour no. 063-130) with
52% protein was obtained from Archer Daniels Midland
Company (Decatur, IL). The other components of the flour
included 30% carbohydrate (18% dietary fiber and 12% soluble
carbohydrate), 9% moisture, 3% fat, and 6% ash. Glycerol and
D-sorbitol were obtained from J.T. Baker (Phillipsburg, NJ) and
Sigma-Aldrich (St. Louis, MO). The polyester amide (BAK
1095) was obtained from Bayer Corp. (Pittsburgh, PA).
Extrusion of Soy-Based Bioplastic. All experimental materi-
als except the plasticizer were predried in a vacuum oven at
about 80 C for at least 3 h to remove moisture. Soy flour was
premixed with the plasticizer with a kitchen mixer in a 70:30
ratio by weight percentage. This mixture was fed to a ZSK 30
Werner & Pfleiderer co-rotating twin-screw extruder. The
extruder was divided into six temperature zones, and the zone
temperatures were maintained at 95, 105, 115, 125, 130, and
130 C, respectively. The screw rotation speed on the extruder
was maintained at 100 rpm, and the torque values ranged from
50% to 60% of full scale. The die used at the extruder was a
two-strand die, and the diameter of the strands coming from
the die was 2 mm. The strands of the extrudate were collected
and chopped to form pellets using a pelletizer. This extruded
mixture was blended in the extruder with the polyester amide
in a weight ratio of 2:1. All zone temperatures were maintained
at 130 C during this blending because both soy protein and
polyester amide are thermally stable at this temperature.
Injection Molding of Soy-Based Bioplastics. The soy-flour-
based biodegradable plastics (soy-based bioplastics) were injec-
tion molded into tensile coupons using an 85-ton Cincinnati-
Millacron injection molder with four temperature zones, the last
one being the nozzle. The temperatures on all zones were
maintained at 130 C, and the screw rotation speed was constant
at 50 rpm. The mold temperature was maintained at around 15
C. The materials were injection molded into standard tensile
coupons and subsequently used for the evaluation of mechanical
and thermal properties.
Mechanical Property Measurements. The tensile properties
and thermal properties of injection-molded soy-based bioplastics
were measured with a United Testing System SFM-20 instru-
ment according to standard method ASTM D638 and ASTM
D790, respectively. System control and data analysis were
performed using Datum software. The notched impact properties
were measured with a Testing Machines Inc. 43-02-01 monitor/
impact machine according to standard method ASTM D256.
Dynamic Mechanical Properties. A dynamic mechanical
analyzer (2980 DMA, TA Instruments, New Castle, DE) was
used to measure dynamic mechanical properties of the soy-based
bioplastics. Tests were performed from -120 to 120 C at a
heating rate of 4 C/min and a frequency of 1 Hz.
Heat Deflection Temperature (HDT). The heat deflection
temperatures (HDTs) of the soy-based bioplastics were measured
with a dynamic mechanical analyzer according to standard
method ASTM D648 with a load of 66 psi under a DMA-
controlled force and three-point bending modes. The heating
rate was 2 C/min.
Thermogravimetric Analysis (TGA). A thermogravimetric
analyzer (2950 TGA, TA Instruments, New Castle, DE) was
used to measure the decomposition behavior of the soy-based
bioplastics under nitrogen gas atmosphere. The weight of
samples was 20 mg, and the tests were performed at a heating
rate of 10 C/min. The gas flow rates were 60 and 40 mL/min
for sample purge and balance purge, respectively.
Environmental Scanning Electron Microscopy (ESEM).
The tensile and impact fracture surfaces of the soy-based
bioplastics were observed by environmental scanning electron
microscopy (ESEM) with a Phillips Electroscan 2020 instrument
at an accelerating voltage of 20 kV.
Results and Discussions
Mechanical Properties. Figure 1 shows a comparison of the
tensile properties of the three different soy-based bioplastics
examined in this work. The compositions of the different soy-
based plastics are reported in Table 1. It was observed that
sorbitol-plasticized soy-based bioplastic (SSBP) showed a higher
Chart 1. Schematics of Glycerol, Sorbitol, Soy Protein, and
Polyester Amide
Figure 1. Comparison of the tensile properties of (A) GSBP, (B) SSBP,
and (C) MSBP.
Table 1. Compositions and Heat Deflection Temperatures (HDTs)
of Three Different Soy-Based Bioplastics
sample
soy flour
(wt %)
glycerol
(wt %)
sorbitol
(wt %)
polyester
amide (wt %)
HDT
(C)
SSBP 46.67 0 20 33.33 45
GSBP 46.67 20 0 33.33 35
MSBP 46.67 10 10 33.33 39
7492 Ind. Eng. Chem. Res., Vol. 45, No. 22, 2006
tensile modulus and tensile strength than glycerol-plasticized
soy-based bioplastic (GSBP), whereas MSBP exhibited a range
of tensile modulus and tensile strength values intermediate to
those of SSBP and GSBP. The tensile strengths of MSBP and
SSBP were about 45% and 50%, respectively, higher than that
of GSBP, and the tensile moduli of MSBP and SSBP increased
by about 135% and 255%, respectively, compared to that of
GSBP. Figure 2 shows the stress-strain plots of the soy-based
bioplastics. GSBP exhibited higher elongation and lower stress,
but SSBP showed higher stress and lower elongation. MSBP
showed mild elongation and a tensile stress similar to that of
SSBP. The flexural properties as shown in Figure 3 also follow
the same trend as the tensile properties. The flexural strengths
of MSBP and SSBP are about 70% and 160%, respectively,
higher than that of GSBP, and the flexural moduli of MSBP
and SSBP increased by about 100% and 235%, respectively,
compared to that of GSBP. GSBP had the highest value of
impact strength, with SSBP and MSBP exhibiting lower values
(Figure 4). In summary, GSBP had higher elongation, higher
impact strength, lower modulus, and lower strength, but SSBP
had lower elongation, lower impact strength, higher modulus,
and higher strength. MSBP had values intermediate between
these two. The differences among these soy-based bioplastics
might be caused by the interactions between the plasticized soy
protein and the polyester amide, as discussed in the next section.
Thermal Properties. Heat deflection temperature refers to
the maximum temperature at which a polymer can be used as
a rigid material. Specifically, according to standard method
ASTM D648, HDT is defined as the temperature at which the
deflection of the sample reaches 250 m under an applied load
of 66 psi. The HDT values of the soy-based bioplastics examined
in this work are listed in Table 1. It was found that SSBP plastic
has the highest value of HDT (45 C) among the three plastics.
The HDT of a plastic reflects the transitions and relaxations of
molecular segments of the plastic. As discussed below, the
polyester amide phase in SSBP had a higher glass transition
temperature because the polyester amide and soy protein
penetrated each other, which led to the highest HDT.
The storage modulus and loss factor of soy protein and
polyester amide are shown as functions of temperature in Figure
5. It was found that soy protein had two main relaxation peaks
in the loss factor plot. The peak at around 69 C was the glass
transition relaxation of the soy protein, and the other peak at
around -50 C was the relaxation of the soy protein.
20
For
the polyester amide, there was sharp decrease in modulus in
Figure 2. Comparison of the stress-strain curves of (A) GSBP, (B) SSBP,
and (C) MSBP.
Figure 3. Comparison of the flexural properties of (A) GSBP, (B) SSBP,
and (C) MSBP.
Figure 4. Comparison of the impact strengths of (A) GSBP, (B) SSBP,
and (C) MSBP.
Figure 5. Dynamic mechanical properties of the soy protein and polyester
amide.
Figure 6. Dynamic mechanical properties of (A) GSBP, (B) SSBP, and
(C) MSBP.
Ind. Eng. Chem. Res., Vol. 45, No. 22, 2006 7493
the main transition region at around -10 C that was attributed
to the glass transition of the amorphous phase of the polyester
amide.
The results for the temperature dependence of the loss factor
of the soy-based bioplastics are presented in Figure 6. A
summary of the experimental investigations on the main
relaxations is included in Table 2. It was clearly found that two
relaxation peaks appeared for all of the soy-based bioplastics.
The lower-temperature peak was attributed to the glass transition
(T
g
) of the polyester amide. The higher-temperature peak
corresponded to the glass transition (T
g
) of the soy protein.
However, compared to T
g
of the neat polyester amide, it was
found that T
g
of the polyester amide in the soy-based bioplastics
shifted to higher temperature. This indicates that the soy protein
and polyester amide are partially compatible, which might be
due to the intramolecular interactions of the amino and carboxyl
groups in the soy protein and the amide, carboxyl, and hydroxyl
groups in the polyester amide. These interactions include
hydrogen bonding and polar interactions. For GSBP, T
g
of the
polyester amide shifted to higher temperature, but the main
transition of the soy protein did not change. This means that
only the molecular segments of the soy protein can migrate into
the domain of the polyester amide, but the molecular segments
of the polyester amide cannot migrate into the domain of the
soy protein. Therefore, the soy protein and polyester amide in
GSBP had limited improvement in compatibility. However, for
SSBP, T
g
of the polyester amide shifted to higher temperature,
and that of the soy protein shifted to lower temperature. This
result reveals that the molecular segments of soy protein and
polyester amide can migrate toward each other, indicating that
the compatibility between soy protein and polyester amide in
SSBP was further improved. The changes in T
g
of MSBP
followed the same trend as that of GSBP, but the degree of
change in T
g
of the polyester amide was greater for the former.
The compatibility of the blends can be predicted from the
glass transition temperature (T
g
) of the blends, and the phase
composition can be calculated according to the Fox equation
21
where 1 and 2 represent the components and W is the weight
fraction of the corresponding component.
The calculation results of phase information related to the
interactions between the soy protein and polyester amide are
reported in Table 3. It was found that there was about 56% soy
protein in the polyester-amide-rich domain and about 7%
polyester amide in the soy-protein-rich domain in SSBP. This
result confirms the above statement about the compatibility and
migration between soy protein and polyester amide. Also, this
further points out the stronger interaction between the soy and
polyester amide in SSBP.
The enhancement in compatibility can be simply measured
by the degree of inward shifting of the glass transition
temperatures of the components. The improvement in compat-
ibility in the soy-based bioplastics is reported in Table 2. It can
be easily seen that the compatibility between the soy protein
and polyester amide in SSBP was significantly improved and
showed a broadened glass transition region, but that in GSBP
was marginally improved, and that in MSBP was in the middle.
This is consistent with the results in Table 3 from calculations
using the Fox equation. The most likely reason for the significant
improvement in compatibility is that sorbitol has more hydroxyl
groups in its molecular structure. This results in a greater
contribution to the interactions between sorbitol and soy protein
and, hence, stronger hydrogen bonding between the sorbitol-
plasticized soy protein and the polyester amide.
The compatibility between the soy protein and polyester
amide determines the mechanical properties of the soy-based
bioplastics. It is not surprising that SSBP had the best tensile
and flexural properties among all of the bioplastics examined
in this work.
The loss factor provides damping and molecular mobility
information, which is related to the impact strength of the plastic.
Because of the stronger interaction between the soy protein and
polyester amide in SSBP and the higher T
g
, the molecular
motion in this bioplastic is restricted. Therefore, this plastic
cannot absorb more energy under an impact load, resulting in
a lower impact strength. In contrast, GSBP exhibits higher
energy absorption under an impact load because of the weak
interactions between the soy protein and polyester amide and
the higher molecular mobility.
Figure 6 also shows the temperature dependence of the
storage modulus of the soy-based bioplastics. SSBP had a higher
storage modulus than GSBP and MSBP. The storage modulus
at 25 C (Table 1) indicated that SSBP had the highest modulus
and GSBP had a lower modulus. This is consistent with the
results for the tensile and flexural moduli.
TGA measures the weight loss of a substance as a function
of temperature and provides information on the decomposition
behavior of a substance. The TGA results in this study (Figure
7) showed that SSBP had a higher initial decomposition
temperature than the other two plastics, indicative of the highest
thermal stability of the three plastics. This also shows that there
Table 2. Glass Transition Temperatures (Tg), Degrees of Inward
Shifting of Tg, and Storage Moduli at 25 C of Three Different
Soy-Based Bioplastics
Tg (C)
sample
polyester
amide
soy
protein
degree of
inward shifting
of Tg (C)
modulus
at 25C
(GPa)
SSBP 29.0 61.6 47.1 1.3
GSBP 0 69.7 10.0 0.3
MSBP 10.0 68.1 21.6 0.8
Table 3. Interactions between Soy Protein (SP) and Polyester Amid
(PEA) in the Phase Structure of Soy-Based Bioplastics from the Fox
Equation
weight fraction
in the PEA-
rich phase (%)
weight fraction
in the SP-rich
phase (%) weight fraction (%)
sample PEA SP PEA SP
PEA-rich
phase
SP-rich
phase
SSBP 44 56 7 93 70 30
GSBP 84 16 - 100 40 60
MSBP 69 31 1 99 47 53
1
T
g
)
W
1
T
g1
+
W
2
T
g2
(1)
Figure 7. Thermogravimetric analyses of (A) GSBP, (B) SSBP, and (C)
MSBP.
7494 Ind. Eng. Chem. Res., Vol. 45, No. 22, 2006
is a stronger interaction between the sorbitol-plasticized soy
protein and the polyester amide, which is also consistent with
the DMA results.
Morphology. The tensile fracture surfaces as observed by
ESEM are shown in Figure 8. It was found that GSBP had
ductile fracture features with a coarse surface. Some of the
deformed materials on the fracture surface showed a plastic
deformation band that did not recover after deformation. This
indicates that this sample had higher elongation during stretching
and absorbed more energy, which resulted in a higher impact
strength. This is consistent with the results from mechanical
testing. Also, some fiber was found on the fracture surface
because soy protein contains 1-3% crude fiber. However, SSBP
showed brittle fracture features with relatively smooth surfaces.
This suggests that SSBP had higher strength, higher stiffness,
and lower elongation and absorbed less energy during stretching;
hence, it had a lower impact strength. MSBP showed a local
ductile fracture feature. Therefore, MSBP had moderate tensile
strength and stiffness, as well as elongation and impact strength.
The morphology of the tensile fracture surfaces confirms the
mechanical natures of the different plasticized soy-based bio-
plastics.
These results demonstrate that sorbitol and glycerol are good
plasticizers for soy protein. Through blending with plasticizers,
one can adjust the balance between the strength, modulus, and
toughness of soy-based bioplastics.
Conclusion
A series of bioplastics from soy protein and polyester amide
with plasticizers including glycerol, D-sorbitol, and glycerol/D-
sorbitol blends have been prepared by extrusion and injection
molding. It was found that SSBP was more rigid, with a higher
tensile modulus and tensile strength than GSBP. MSBP resulted
in an intermediate range of tensile modulus and strength values.
GSBP had the highest impact strength, whereas SSBP had the
highest thermal stability. These thermal and mechanical proper-
ties are dependent on the compatibility between the soy protein
and the polyester amide. Sorbitol-plasticized soy protein ex-
hibited a significant improvement in compatibility with polyester
amide because of the possible hydrogen bonding between them.
ESEM morphology observations showed that SSBP had brittle
fracture features, whereas GSBP had local ductile fracture
features. These results demonstrate that use of a plasticizer
allows one to control the balance between the strength, modulus,
and toughness of soy-protein-based bioplastics. These soy-based
bioplastics can be used as packaging materials to substitute
petroleum-based plastic materials.
Figure 8. Tensile fracture surfaces of (A) GSBP, (B) SSBP, and (C) MSBP.
Ind. Eng. Chem. Res., Vol. 45, No. 22, 2006 7495
Acknowledgment
Financial support for this research from USDA-NRI (Grant
No. 2001-35504-10734) and GREEEN (Generating Research
and Extension to Meet Economic and Environmental Needs;
No. GR02-066) is gratefully acknowledged. The authors also
thank ADM (Decatur, IL) and Bayer Corp. (Pittsburgh, PA)
for their supplying the soy flour and polyester amide samples,
respectively.
Literature Cited
(1) Catsimpoolas, N.; Kenney, J. A.; Meyer, E. W.; Szuhaj, B. F.
Molecular weight and amino acid composition of glycinin subunits. J. Sci.
Food Agric. 1971, 22, 448.
(2) Paetau, I.; Chen, C. Z.; Jane, J. Biodegradable plastic made from
soy products. II. Effects of cross-linking and cellulose incorporation on
mechanical properties and water absorption. J. EnViron. Polym. Degrad.
1994, 2, 211.
(3) Mo, X.; Sun, X. Plasticization of soy protein polymer by polyol-
based plasticizers. J. Am. Oil Chem. Soc. 2002, 79, 197-202.
(4) Wang, S.; Sue, H. J.; Jane, J. Effects of polyhydric alcohols on the
mechanical properties of soy protein plastics. J. Macromol. Sci. Pure Appl.
Chem. 1996, A33, 557-569.
(5) Kim, K. M.; Marx, D. B.; Weller, C. L.; Hanna, M. A. Influence of
sorghum wax, glycerin, and sorbitol on physical properties of soy protein
isolate films. J. Am. Oil Chem. Soc. 2003, 80 (1), 71-76.
(6) Wu, Q.; Zhang, L. Properties and Structure of Soy Protein Isolate-
Ethylene Glycol Sheets Obtained by Compression Molding, Ind. Eng. Chem.
Res. 2001, 40, 1879.
(7) Wang, Niangui; Zhang, Lina. Preparation and characterization of
soy protein plastics plasticized with waterborne polyurethane. Polym. Int.
2005, 54 (1), 233-239.
(8) Zhong, Z. K.; Sun, X. S. Thermal and mechanical properties and
water absorption of guanidine hydrochloride-modified soy protein (11S).
J. Appl. Polym. Sci. 2000, 78, 1063.
(9) Zhong, Z. K.; Sun, X. S. Thermal and mechanical properties and
water absorption of sodium dodecyl sulfate-modified soy protein (11S). J.
Appl. Polym. Sci. 2001, 81, 166.
(10) Mo, X.; Sun, X. S.; Wang, Y. Effects of molding temperature and
pressure on properties of soy protein polymers. J. Appl. Polym. Sci. 1999,
73, 2595.
(11) Jane, J.-L.; Wang, S. Soy protein-based thermoplastic composition
for preparing molded articles. U.S. Patent, 5,523,293, 1996.
(12) Paetau, I.; Chen, C.-Z.; Jane, J. Biodegradable Plastic Made from
Soybean Products. 1. Effect of Preparation and Processing on Mechanical
Properties and Water Absorption. Ind. Eng. Chem. Res. 1994, 33, 1821.
(13) Graiver, D.; Waikul, L. H.; Berger, C.; Narayan, R. Biodegradable
soy protein-polyester blends by reactive extrusion process. J. Appl. Polym.
Sci. 2004, 92 (5), 3231-3239.
(14) Liu, W.; Mohanty, A. K.; Askeland, P.; Drzal, L. T.; Misra, M.
Influence of fiber surface treatment on properties of Indian grass fiber
reinforced soy protein based biocomposites. Polymer 2004, 45, 2247
(15) Liu, W.; Mohanty, A. K.; Drzal, L. T.; Misra, M. Novel biocom-
posites from native grass and soy based bioplastic: Processing and properties
evaluation. Ind. Eng. Chem. Res. 2005, 44, 7105.
(16) Zhong, Z.; Sun, X. S. Properties of soy protein isolate/polycapro-
lactone blends compatibilized by methylene diphenyl diisocyanate. Polymer
2001, 42, 6961.
(17) Mungara, P. et al., Processing and physical properties of plastics
made from soy protein polyester blends. J. Polym. EnViron. 2002, 10, 31.
(18) John, J.; Bhattacharya, M. Properties of reactively blended soy
protein and modified polyesters. Polym. Int. 1999, 48, 1165.
(19) Wang, H.; Sun X.; Seib, P. Effect of moisture on properties of
starch/PLA blend containing MDI. J. Polym. EnViron. 2002, 10 (4), 133-
138
(20) Zhang, J.; Mungara, P.; Jane, J. Mechanical and thermal properties
of extruded soy protein sheets. Polymer 2001,42, 2569-2578
(21) Fox, T. G. Influence of diluent and of copolymer composition on
the glass temperature of a polymer system. Bull. Am. Phys. Soc. 1956, 1,
123.
ReceiVed for reView April 7, 2006
ReVised manuscript receiVed August 9, 2006
Accepted August 29, 2006
IE060439L
7496 Ind. Eng. Chem. Res., Vol. 45, No. 22, 2006

You might also like