Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

J Therm Anal Calorim (2017) 130:919–933

DOI 10.1007/s10973-017-6459-1

Understanding intrinsic plasticizer in vegetable oil-based


polyurethane elastomer as enhanced biomaterial
Ahmad Syafiq Ahmad Hazmi1,2 • Nik Nurfatmah Pz Nik Pauzi1 • Zulina Abd. Maurad1 •
Luqman Chuah Abdullah3,2 • Min Min Aung4 • Azizan Ahmad5 • Mek Zah Salleh6 •
Rida Tajau6 • Mohd Hilmi Mahmood6 • Syahrina Elliyana Saniman7

Received: 31 May 2016 / Accepted: 15 May 2017 / Published online: 6 June 2017
 Akadémiai Kiadó, Budapest, Hungary 2017

Abstract Renewable polyol is of increasing interest as a conversion at gel point had enhanced physical properties.
building block in biomedical elastomer for bearing Highly elastic mechanical behavior was afforded from
biodegradable ester group and immaculate functionality. combination of side chains and high molecular weight
Derived from non-edible vegetable oil, a new class of polyol. At r = 1/1.2, MDI-based elastomer showed two-
elastomer was successfully functionalized with MDI and fold improvement in Young modulus at slight expense of
TDI. Crosslink densities were varied by regulating ratio of elongation. TDI-based elastomer accomplished elongation
hydroxyl to diisocyanate (r) at 1/1.0, 1/1.1, and 1/1.2. beyond 162%. Branching allophanate and biuret resisted
Produced elastomers were examined by crosslink density, early thermal breakdown by elevating activation energy.
attenuated total reflectance Fourier transform infrared Frequency response and kinetic of thermal degradation
spectroscopy, differential scanning calorimetry, thermo- provided beneficial perspective for elastomer characteri-
gravimetric analysis, dynamic mechanical analysis, tensile zation. The vegetable oil-based polyurethane was found
testing, and scanning electron microscopy. The obtained able to resemble most of the physical properties of poly-
elastomers had subambient glass transition temperature caprolactone (PCL)-derived polyurethane.
(Tg) suggested majority soft segment that acted as a con-
tinuous phase with intermediate phase separation. Medium

& Ahmad Syafiq Ahmad Hazmi 4


Department of Chemistry, Faculty of Science, Universiti
ahmadsyafiq87@gmail.com Putra Malaysia, 43400 Serdang, Selangor, Malaysia
5
1 School of Chemical Science and Food Technology, Faculty
Process Engineering and Design Unit, Advanced
of Science and Technology, Universiti Kebangsaan Malaysia,
Oleochemical Technology Division, Malaysian Palm Oil
43600 Bangi, Selangor, Malaysia
Board, No. 6, Persiaran Institusi, Bandar Baru Bangi,
6
43000 Kajang, Selangor, Malaysia Radiation Processing Technology Division, Malaysian
2 Nuclear Agency, 43000 Bangi, Selangor, Malaysia
Department of Chemical and Environmental Engineering,
7
Faculty of Engineering, Universiti Putra Malaysia, Technology Medical Associate, 40400 Shah Alam, Selangor,
43400 Serdang, Selangor, Malaysia Malaysia
3
Institute of Tropical Forestry and Forest Products, Universiti
Putra Malaysia, 43400 Serdang, Selangor, Malaysia

123
920 A. S. Ahmad Hazmi et al.

Graphical Abstract

Medium functionality polyol


HO OH

HO
OH
HO

Diisocyanate
Dibutyltin dilaurate

Crosslinking
RU RU RA
RA
RU
RA
RU
RA
RU RA

Polyurethane Crosslinked network with allophanate group

Keywords Polyurethane elastomer  Biomaterial 


Thermal properties  Degradation kinetic  Mechanical (a) R1-NCO + OH-R2 → R1-NHCOOR2 + Heat

properties  Frequency response


(b) R-NCO + H 2 O → [R-NHCOOH] → RNH 2 + CO 2 + Heat

(c) R1-NCO + R2-NH 2 → R1-NHCONH-R2


Introduction

(d) R1-NCO + R2COOH → [R1NHCOOCOR2] → R1NHCOR2 + CO 2


Polyurethane represents a group of versatile polymer and
conceived wide range of technical applications. Poly-
urethane ranks fifth in high volume of plastic production R1 COOR2
after polyethylene, polyvinyl chloride, polypropylene, and 1
R NHCOOR 2
+ 1
R -NCO N
polystyrene [1]. Recent impetuses for steady growth of (e)
polyurethane market are emerging applications in CONHR1

biomedical field, exceptional performance in automotive


industry, and environmental sustainability. Early genera-
tion of polyol is strongly connected to the renewable R1 CONHR1
resources [2, 3]. Current generation of polyurethane is
1
R NHCONHR 1
+ 2
R -NCO N
(f)
prepared from petroleum-based polyol which associated
with environmental and sustainability concerns. Raw CONHR2

materials derived from renewable resources are regarded as Fig. 1 Reactions between isocyanates and different compounds [2].
environmental friendly and sustainable. Vegetable oil, in The individual schemes are labeled: a exothermic reaction with
particular, is a precious starting materials toward valuable alcohol which contain hydroxyl group to form urethane, b exothermic
products such as high-performance polyurethane besides reaction between isocyanate and water leads to unstable carbamic
acid that produce amine and gaseous carbon dioxide which is
being relatively inexpensive [4, 5]. convenient for cellular formation, c amine group reacts rapidly with
Versatility of polyurethane chemistry (Fig. 1) allow isocyanate to form disubstituted urea, d isocyanate react with
balance of hydrophilicity and hydrophobicity on the sur- carboxylic acid to yield unstable anhydride which stabilize as amide
face. Polyurethanes elastomers are commonly found inside and release carbon dioxide, e above 80 C and under vacuum, a
reversible reaction with urethane group leads to allophanate forma-
medical microdevices, scaffolding, and biomedical micro- tion, and f above 80 C and under vacuum, a reversible reaction with
electromechanical system (bioMEMS) as being disubstituted urea to form biuret group

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 921

Fig. 2 Chemical structures of O


vegetable oil-based elastomer
and PCL. Upon polymerization, CH3
vegetable oil-based elastomer O
will have side chains that act as OOCNH R
intrinsic plasticizer O
CH3
O

OOCNH R Side
chain

R HNCOO O CH3
O

O CH3

R HNCOO OOCNH R

Vegetable oil-based elastomer

O
n
Polycaprolactone (PCL)

biocompatible and inherit relatively lower cost than inor- uniform contain of unsaturated fatty acids, specifically as
ganic materials [6]. Advantages of polyurethane elastomer oleic acid and linoleic acid that constituted more than 70%
include higher machinability than glass and having more in crude oil [20], is a characteristic of jatropha oil and vital
adhesive property which makes it compatible to be used toward achieving desired functionality and molecular
with many materials. However, it is an inconvenient fact weight between crosslink (Mc). Jatropha oil is known for its
that the petroleum-based elastomer; for instance, poly- inherited toxic ingredients and naturally inedible and hence
caprolactone (PCL) typically requires a substantial amount its price relatively insulated from supply-and-demand
of external plasticizer to acquire good mechanical flexi- surge in food industry [21]. As vegetable oil-based polyol
bility, augment crystallinity, and provide some elastomeric and PCL share similarity in chemical structure that form
mechanical properties [7–11]. Alternatively long chains of the repeating unit in polymeric network (Fig. 1), it is
fatty acid found in vegetable oil could be manipulated to paramount to understand the physical properties of such
produce side chains which then act as intrinsic plasticizer vegetable oil-based polyurethane. In this paper, veg-
(Fig. 2). Furthermore, it was found that hydrophilicity, an etable oil-based polyurethanes were prepared by reacting
important key toward biocompatibility and minimizing multifunctional jatropha oil-based polyol with two types of
biofouling, of vegetable oil-based polyurethane could be aromatic diisocyanates namely dipheylmethane-4,4’-diiso-
improved by dictating the hydroxyl number [12]. Hence cyanate (MDI) and toluene-2,4-diisocyanate (TDI). MDI is
there are growing interests on understanding the structural highly sought in biomedical applications due it
properties relationship [13–19] in renewable polyurethane. biodegradable characteristic. On the other hand, TDI usu-
Often decent properties for polyurethane are gained when ally render superior flexibility toward mechanical load. The
employing medium functionality, between 4.5 to 5.5 OH physical properties of these elastomers were examined on
groups/mol, as it capable to give medium conversion at gel differential scanning calorimeter (DSC), thermogravimet-
point and pristine crosslink densities [2]. Nevertheless, ric analysis (TG), and derivative thermogravimetric
fabricators pay little attention to the effect of crosslink (DTG), dynamic mechanical analysis (DMA), Instron
densities which eventually govern the physical properties universal tensile machine, and scanning electron micro-
of elastomer. scopy (SEM). The crosslink densities were varied by reg-
One of promising yet underutilized among vegetable oil ulating ratio of hydroxyl to isocyanates as reflected in the
for commercialization is jatropha oil. Very high and swelling measurement in toluene and extend of reaction

123
922 A. S. Ahmad Hazmi et al.

was monitored by Fourier transform infrared (FTIR) between 0.2 and 0.6 g was immersed in 50 mL of toluene
method. for 10 days at 25 C. The samples were dried prior to mass
measurement. Mc was calculated from Flory–Rehner rela-
tionship [23]:
 
Experimental 1=3
q2 V1 /2  2/f 2
Mc ¼ ð2Þ
Materials lnð1  /2 Þ þ /2 þ v12 /22
where q2 is the density of polymer and f is the functionality
The multifunctional jatropha oil-based polyol used was of network. V1 is the molar volume of solvent and in case
having hydroxyl number 175 KOH g-1 and functionality 5.1, of toluene equal to 106.28 cm3 mol-1. /2 is the volume
prepared from crude jatropha curcas oil having iodine value fraction of polymer and calculated as follows:
104.4 g I2 (100 g-1) and functionalized via epoxidation and
ring opening [22]. Technical-grade toluene-2,4-diisocyanate m1 d1
/2 ¼ ð3Þ
(TDI) and dibutyltin dilaurate were supplied by Fluka and m 1 ð d1  d2 Þ þ m 2 d2
used as received. Synthesis grade dipheylmethane-4,40 -di- m1 is the polymer mass before immersion, m2 is the
isocyanate (MDI) was obtained from Merck and used as polymer mass after swollen in solvent and reach equilib-
received. Silicone-free defoamer Afcona 2020 was acquired rium, d1 is the density of solvent, and d2 is the density of
from AFCONA Additives and used as received. polymer. Meanwhile, polymer–solvent interaction param-
eter v12 [24] is determined from following equation:
Synthesis of jatropha oil-based polyurethane
V1 ðd1  d2 Þ2
elastomer v12 ffi b þ ð4Þ
RT
Jatropha oil-based polyurethane elastomer was prepared by where b is lattice constant and approximated as 0.34. d1 is
reacting jatropha oil-based polyol with MDI and TDI. The the solubility parameter of solvent while d2 is solubility
molar ratios of hydroxyl group (OH) to the isocyanate parameter of polymer. R and T are universal gas constant
group (NCO) were 1.0/1.0, 1.0/1.1, and 1.0/1.2 as defined and temperature, respectively. As the jatropha oil-based
in following equation: polyol used was ring-opened with majority methanol, v
OH Wpolyol =EWpolyol was deliberated as 0.5293 throughout the experiment [25].
r¼ ¼  ð1Þ
NCO WPU  Wpolyol =EWisocyanate
Fourier transform infrared spectroscopy (FTIR)
where Wpolyol is the mass of polyol, EWpolyol is the
measurement
equivalent mass of polyol, WPU is the mass of produced
polyurethane, and EWisocyanate is the equivalent mass of
The FTIR spectra were recorded on Perkin-Elmer Spec-
isocyanate. The equivalent mass as supplied for TDI and
trum 2000 spectrometer which collects mid-infrared scat-
MDI are 87 g mol-1 and 125 g mol-1, respectively. A
tered radiations using a single-beam improved Michelson
calculated amount of diisocyanate and catalyst dibutyltin
interferometer with attenuated total reflectance (ATR)
dilaurate at 0.5 mass% were added into jatropha oil-based
accessory which equipped with diamond crystal tip. The
polyol in a Teflon container 10 cm 9 10 cm 9 1.5 cm. As
spectra were recorded in transmission mode in the range of
this paper is intended to produce polyurethane elastomer,
4000–500 cm-1 with a nominal resolution of 4 cm-1.
residual water and formic acid might be a concern upon
contact with isocyanate, and hence residual gas evacuation
from casting was facilitated by the additive defoamer at Differential scanning calorimetry (DSC)
1 mass%. The mixture was stirred for 1 min and degassed measurement
at 85 C for 30 min. The produced film was immediately
post-cured at room temperature for 24 h. Six different Differential scanning calorimetry (DSC) was carried out on
formulations were used when producing jatropha oil-based DSC 823e (Mettler-Toledo), equipped with STARe soft-
polyurethane to observe the effect of crosslink densities. ware 9.10, according to ASTM E3418-03 standard practice
[26]. The DSC equipped with a refrigerated cooling system
Swelling in solvent and calibrated with indium. Five milligrams of sample was
sealed in aluminum 40 lL hermetic pan and heated from 25
Swelling studies on jatropha oil-based polyurethane was to 180 C at heating rate of 10 C min-1 to eliminate
determined by using toluene as solvent. Samples mass thermal history, cooled to -60 C at cooling rate of

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 923

5 C min-1, and heated again to 180 C min-1 at heating Results and discussion
rate of 10 C min-1 under nitrogen gas flow rate of
20 mL min-1. Physical properties of polyurethane elastomer

ATR-FTIR of polyurethane
Thermogravimetric analysis (TG/DTG)
measurement
In polymerization process, the isocyanate, which is a strong
infrared absorbent, was monitored for degree of polymer-
Thermogravimetric analysis (TG) and derivative thermo-
ization and quality of produced polyurethane [30]. The
gravimetric (DTG) were carried out on TGA/SDTA 851E
absorption intensity of the isocyanate group at 2273 cm-1
(Mettler-Toledo) according to ASTM E1641-99 standard
and hydroxyl group at 3475 cm-1 gradually decreased
practice [27]. Five milligrams of sample was put on alu-
throughout reaction time and disappeared when reaction
mina sample holder 70 lL and heated from 30 to 800 C at
time reached 5 h as shown in Fig. 3a. At the end of reac-
heating rate of 2.5, 5, 10, 15, and 20 C min-1 under
tion, four characteristic absorptions of polyurethane were
nitrogen gas flow (20 mL min-1). Small sample size
observed as NH group stretching at 3340 cm-1, urethane
equilibrated faster with furnace temperature and hence
carbonyl group C=O at 1740 cm-1, C–O stretch at
mitigated temperature deviation and noise in the mea-
1225 cm-1 and NH group bending at 1538 cm-1. It is well
surement. The TG/DTG curves were analyzed on STARe
documented that excess isocyanate and application of
software version 9.10 Mettler.
catalyst dibutyltin dilaurate would lead to formation of
allophanate and biuret [31–33]. Introduction of excess
Dynamic mechanical analysis (DMA) measurement isocyanate would enhance their formations over the ure-
thane bonds and thermodynamically favorable [34, 35].
Dynamic mechanical analysis (DMA) was carried out on This was evident from differences in wet experimental
DMA Q800 V20.24 (TA Instruments) according to ASTM
D5062-01 standard practice [28]. Instrument calibrations
were verified and a series of clamp calibrations were car- (a)
5 hour NH C = ON-H COO,
ried out on each day of measurement. A rectangular test
Transmittance/Arbitrary unit

C-O
specimen (76 by 13 by 1 mm) was clamped in tension 3 hour
mode configuration. In temperature sweep of controlled
strain program, specimens were set to -100 C using ‘‘Go 2 hour
930 cm–1
1454 cm–1
To Temperature’’ command, re-torqued, and ramped to NCO
0 hour OH
140 C at 5 C min-1 at 1 Hz according to standard
practice ASTM D5026-01. Small size specimens used in
tension mode minimize temperature gradient and allowed 4000 3500 3000 2500 2000 1500 1000 500
higher temperature scanning rate. Wavenumber/cm–1
(b)
TDI
r = 1/1.2
Tensile strength measurement
TDI
r = 1/1.1
Trasmittance/Arbitrary unit

Tensile strength of sample was analyzed using Instron 5567 TDI


universal testing machine, equipped with Bluehill version r = 1/1.0
2.14 software. For each sample, five identical dumbbell- MDI
r = 1/1.2
shaped samples were prepared by using ASTM D638-03
Type V cutter [29]. The crosshead speed was 1 mm min-1 MDI
r = 1/1.1
with 30 kN load cell. The measurement was carried out at
MDI
room temperature and 50% relative humidity. r = 1/1.0
AF2 AB2 AF1 AB1
1800 1750 1700 1650 1600 1550
Scanning electron microscope (SEM) Wavenumber/cm–1

Scanning electron microscopy were carried out using Fig. 3 ATR-FTIR spectra of jatropha oil-based polyurethane. The
individual figures are labeled: a FTIR spectra of jatropha oil-based
Hitachi S-3400 N at 5 kV with 3009 magnification with
polyurethane (TDI-based, r = 1/1.0) formation as a function of
all specimens coated with gold to eliminate charging reaction time, b FTIR from 1800 to 1550 cm-1 as a function of
effect. diisocyanate and r

123
924 A. S. Ahmad Hazmi et al.

Table 1 Effect of isocyanate ratio in jatropha oil-based polyurethane elastomers toward crosslink densities, glass transition temperature, and
mechanical properties
Polyurethane Molecular Conversion at gel Glass transition temperature/ C Young modulus/ Tensile stress at Elongation at
weight between point based on MPa break/MPa break/%
crosslink/ Macosko–Miller DSC DMA
g mol-1 onset E00 peak

MDI 842 0.42 -46.9 -50.0 18.9 ± 0.7 4.1 ± 0.2 130.9 ± 5.9
r = 1/1.0
MDI 729 0.44 -45.8 -48.0 25 ± 1 4.9 ± 0.2 120.9 ± 3.8
r = 1/1.1
MDI 712 0.46 -45.7 -45.3 37.8 ± 1 5.7 ± 0.2 109.9 ± 2.2
r = 1/1.2
TDI 1017 0.50 -49.0 -55.4 11.5 ± 0.4 2.6 ± 0.1 193.8 ± 8.5
r = 1/1.0
TDI 859 0.52 -48.3 -53.1 13.9 ± 0.6 3 ± 0.2 172.8 ± 6.4
r = 1/1.1
TDI 834 0.54 -48.2 -51.2 17.8 ± 0.7 3.4 ± 0.2 162.3 ± 7.2
r = 1/1.2

analysis of crosslink densities represented by Mc values on AB1 þ AB2


RC¼O ¼ ð5Þ
Table 1 and discussed detailed in the following sec- AF1 þ AF2
tion. Typically the formation of these groups is verified
where AB1 is the surface area of bonded hydrogen bond of
using spectroscopy methods. One way is to deconvulate
carbonyl group of urea (1680–1640 cm-1), AB2 is the
FTIR data into individual contribution. The carbonyl bands
surface area of bonded hydrogen bond of carbonyl group of
of allophanate and biuret groups is known to be between
urethane (1727–1705 cm-1), AF1 is the surface area of
1460 and 1450 cm-1, and overlapped with CH2 bending of
unbonded hydrogen bond of carbonyl group of urea (1701–
polyol. To differentiate the bands, we measured the
1690 cm-1), and AF2 is the surface area of unbonded
absorption ratio between 1454 and 930 cm-1 where
hydrogen bond of carbonyl group of urethane (1745–
930 cm-1 was attributed to ether group [36]. Formation of
1736 cm-1). Then the values of DPS can be calculated as:
branching allophanate and biuret groups in the produced
polyurethane elastomers were verified via excess FTIR RC¼O
DPS ¼ ð6Þ
absorption ratio between 1454 and 930 cm-1, i.e., above 1 þ RC¼O
1.4. Ester and ether formations were observed as strong
An appreciable amount of DPS was found in all pre-
absorption peaks at 1100–1200 cm-1. Ester group was
pared elastomers (Table 2). Phase separation phenomena
located along the fatty acid chains, urethane, allophanate,
and interpenetration in multi-component polymer are
and biuret groups. On the other hand ether group was
attributed to incompatibility between the soft and hard
produced in vicinity of side chain. This ether group
segments. DPS tended to be disproportionate with
specifically was also known as methoxy group and origi-
increasing crosslink densities. The concentration of
nated from methanol during ring-opening step in producing
hydrogen bonds in elastomers with higher isocyanate index
polyol. Besides, the casting mixtures were solvent free and
(i.e., MDI (r = 1/1.2) and TDI (r = 1/1.2)) was found to
produced transparent polyurethane sheets that reflected the
be as much as 30% higher in comparison to elastomers
color of the polyol.
with lower isocyanate index (i.e., MDI (r = 1/1.0) and TDI
To further investigate phase existed in the elastomers,
(r = 1/1.0)). TDI-based elastomer shown little increment
one need to determine the RC=O index and subsequent
due having fewer free hydrogen in single aromatic struc-
degree of phase separation (DPS). Phase separation
ture whereas MDI had more free hydrogen in its double
describes the sharing of hard segment with bounded
aromatic structures.
hydrogen bonds. By using FTIR data from each sample, the
carbonyl stretching zone between 1770 and 1640 cm-1
Crosslink densities
was corrected by subtracting the baseline. Then the mul-
tiplet bands were deconvulated (Fig. 3b) using Guassian
A crosslinked polymer is able to increase its volume by
curve-fitting into individual band by using Origin 8.5.0.
absorbing large amount of solvent without being dissolved
This allowed one to calculate RC=O index as [36, 37]:

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 925

Table 2 Deconvolution results of FTIR of jatropha oil-based polyurethane elastomers between 1745 to 1650 cm-1, and subsequent degree of
phase separation (DPS)
Polyurethane AF2 AB2 AF1 AB1 Carbonyl hydrogen Degree of phase
1745–1736 cm-1 1727–1705 cm-1 1701–1690 cm-1 1680–1650 cm-1 bonding index, RC=O separation, DPS/%

MDI 0.06294 0.02641 0.02458 0.12137 1.6885 62.8


r = 1/1.0
MDI 0.10824 0.03863 0.05817 0.15703 1.1758 54.0
r = 1/1.1
MDI 0.07487 0.02139 0.05223 0.09751 0.9355 48.3
r = 1/1.2
TDI 0.07187 0.01716 0.04304 0.13791 1.3494 57.4
r = 1/1.0
TDI 0.09836 0.00186 0.07049 0.21273 1.2709 56.0
r = 1/1.1
TDI 0.08213 0.07575 0.06309 0.10832 1.2675 55.9
r = 1/1.2

as the crosslink sites comprehend the increased inter- represented Tg and increased with increasing crosslink
molecular strands. Crosslink density is defined as 1/(2 Mc) densities. The successive small exothermic peak appeared
where Mc is the number average molecular weight between between -16 and 2 C reflects small amount of rear-
crosslink. A crosslinked network has nearly infinite rangement within network structure and probably origi-
molecular weight, and hence Mc value is more specific in nated from the rotational motion of fatty acid [41].
denoting network properties. A minimal value of Mc Meanwhile, in the temperature sweep of DMA, the storage
indicates tighter packing density with fewer polymers per modulus (E0 ) decreased with decreasing crosslink density
crosslink. As shown in Table 1, jatropha oil-based poly- (Fig. 5a). The loss modulus (E00 ) shows peak in the range
urethane elastomer at higher isocyanate ratio showed lower from -55 to -45 C, and the peak height decrease with
Mc and hence higher crosslink densities. This is also decreasing crosslink density (Fig. 5b). Broader area under
reported by Dey et al. [38] where lower Mc was achieved the E00 peak suggests a softer material and wider molecular
when more isocyanate incorporated into polymeric net- weight distribution. The TDI-based exhibited longer phase
work. MDI-based polyurethane had higher crosslink den- lag than MDI-based attributed by less-ordered packing
sities than TDI-based polyurethane due to the symmetrical structure. The difference between Tg values obtained from
structure and allowed tighter packing organization. Due to DSC and DMA as shown in Table 1 originated from the
the nature of high reactivity of aromatic diisocyanate [31], choice of frequency employed in DMA measurement. Note
excess isocyanates may react with urethane and urea that in the measurement range there was no apparent
groups to afford branching allophanate and biuret groups, melting point observed in the DSC curves which suggested
which increased overall crosslink densities [39]. Based on the samples were thermoset. The majority amorphous
Macosko–Miller polyfunctional equation [40], conversions domain probably explained why the samples were
in the range from 0.42 to 0.54 as obtained in these exper- transparent.
iments (Table 1) were close to midpoint and typically yield There are two postulates to the subambient Tg obtained.
polymer with enhanced physical properties [2]. First is the application of high molecular weight of multi-
functional polyol. Introduction of long-chain polyol
Glass transition temperature (Tg) by DSC and DMA reduced the concentration of hard segments like urethane
and urea bonds, thus lowering the cohesive interaction
The glass transition temperature (Tg) analysis was done between these hard segments and subsequently lowered the
using DSC (Fig. 4) and DMA (Fig. 5). Values of Tg Tg of network formed. Second as a consequence for sig-
observed in both DSC and DMA were comparable and in nificant amount of side chains (Fig. 2) that reduced inter-
the range from -55 to -45 C (Table 1). Tg obtained in chain interaction, minimized crystallization of soft seg-
DSC and DMA were the a-relaxation where phase transi- ments, and consequently lowered the tensile strength and
tion occurred in bulk. It is anticipated that the b-relaxation modulus. Furthermore MDI-based polyurethane showed
occurred further down from the DMA measurement range, higher Tg than TDI-based. The latter exhibited a larger
which was below -100 C. The onset of DSC curves range of Tg and suggests a wider distribution of crosslink

123
926 A. S. Ahmad Hazmi et al.

(a) (a)
0.5 8000

7000 MDI (r = 1/1.0)

Storage modulus/MPa
Heat flow/mW Endo↔Exo

0 MDI (r = 1/1.0) MDI (r = 1/1.1)


6000
MDI (r = 1/1.1) MDI (r = 1/1.2)
–0.5 5000
TDI (r = 1/1.0)
MDI (r = 1/1.2) 4000 TDI (r = 1/1.1)
–1 TDI (r = 1/1.2)
3000

–1.5 2000

1000
–2 0
–100 –50 0 50 100 150
–2.5 Temperature/°C
–60 –10 40 90 140 190
(b)
Temperature/°C 400
(b) 350
2 MDI (r = 1/1.0)

Loss modulus/MPa
300
MDI (r = 1/1.1)
Heat flow/mW Endo↔Exo

1
250 MDI (r = 1/1.2)
0 TDI (r = 1/1.0)
TDI (r = 1/1.0)
TDI (r = 1/1.1) 200
TDI (r = 1/1.1)
–1 TDI (r = 1/1.2) 150 TDI (r = 1/1.2)
–2 100
–3 50

–4 0
–100 –50 0 50 100 150
–5 Temperature/°C

–6
–60 –10 40 90 140 190
Fig. 5 DMA response of jatropha oil-based polyurethane from -100
Temperature/°C to 140 C at 5 C min-1 and 1 Hz. The individual curves are labeled
a the storage modulus (E0 ) of jatropha oil-based polyurethane as a
Fig. 4 DSC curve of jatropha oil-based polyurethane elastomers at function of temperature at 1 Hz, b the loss modulus (E00 ) of jatropha
10 C min-1 in nitrogen atmosphere. The individual curves are oil-based polyurethane as a function of temperature at 1 Hz
labeled a MDI-based polyurethane, b TDI-based polyurethane
alignment and higher degree of crystallization as value of
densities. The symmetrical structure MDI increased the Mc is known sensitive to network imperfection. Further
tendency of polyurethane to crystallize and had better higher isocyanate ratio led to higher concentration of
packing of rigid segment, hence the enhanced tensile branching allophanate and biuret. This in turn produced
strength and modulus. Fox–Loshaek equation anticipates elastomer with higher mechanical modulus and improved
linear relationship between glass transition of crosslinked thermal stability. Due to the nature of high reactivity of
polymer (Tg), uncrosslinked polymer (Tgo), a constant (k), aromatic diisocyanate [31] and high temperature reaction,
and crosslink density as follow [13, 42]: excess isocyanates may reacted with urethane and urea
k groups to afford branching allophanate and biuret groups
Tg ¼ Tgo þ ð7Þ (Fig. 1e, f), which increased overall crosslink densities
Mc
(Table 1). According to Eq. (7), sufficiently very large
An increase in the molar ratio of isocyanate to hydroxyl value of Mc would lead decrease value of Tg to approach
had induced the increment of Tg values as isocyanate Tgo. From Fig. 6, both MDI and TDI-based polyurethanes
groups were the main constituents of rigid segments. For a have Tgo about -72 C. One way to approach this lower
lightly crosslinked polymer such as the elastomers pro- value of Tg is by extending the side chain of polyol as
duced in this study, it was observed that there was a linear demonstrated by Clark AJ and Hoong SS [43]. Note that
increase of Tg with increasing crosslink density (Fig. 6). they employed a high mass percent of tetrahydrofuran
The slopes for MDI- and TDI-based polyurethane elas- (THF) incorporation in the side chain and had obtained Tg
tomers were 18.2 and 17.1 C mol kg-1, respectively. In as low as -76 C.
other word, the Tg of MDI-based polyurethane elastomers The effect of these supplementary branching groups
showed more dependent on crosslink density than TDI- probably can be explained from difference found in qual-
based polyurethane elastomers. This higher sensitivity itative approach of frequency response in DMA where E0
inherited by MDI-based polyurethane is probably owed to and complex viscosity (g*) change as a function of fre-
symmetrical structure of MDI that allow better molecular quency (Fig. 7) based on Doi-Edwards theory [44]. It is

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 927

–44 –50 1000 100


MDI-based TDI-based MDI -based (r = 1/1.2)

10
–46

Log η*/MPa S–1


–52 100

Log E' /MPa


Tg /°C

Tg /°C
1

–48
10
–54
0.1
E' of MDI (r = 1/1.2) TDI -based (r = 1/1.2)
E' of TDI (r = 1/1.2)
–50
n* of MDI (r = 1/1.2)
n* of TDI (r = 1/1.2)
1 0.01
–56 0.1 1 10 100
1.2 1.4 1.0 1.2
Log frequency, Hz
1/MC /mol kg–1 1/MC /mol kg–1
Fig. 7 Crossovers between storage modulus (E0 ) and complex
Fig. 6 A linear dependence between glass transitions temperature viscosity (g*) for MDI-based and TDI-based elastomers at r =
(Tg) and crosslink density 1/1.2 which correspond to relative molecular weight and molecular
weight distribution

known that the difference found in frequency response


7
arise from the network structure. At very low frequency, MDI
TDI
r = 1/1.2
MDI-based samples had higher g* than TDI-based which 6 r = 1/1.2
TDI
suggests more branching across the polymeric chain. The MDI
r = 1/1.1
5 r = 1/1.1
crossover between the E0 - g* for MDI-based samples was TDI
Tensile stress/MPa

MDI r = 1/1.0
higher than TDI-based, indicating qualitatively higher 4 r = 1/1.0
molecular weight. Meanwhile, the crossover between the
3
E0 - g* of TDI-based occur at much higher frequencies
which suggest broader molecular weight distribution. 2
Long-branched polymeric chains such as MDI-based
1
(r = 1/1.2) and TDI-based (r = 1/1.2) showed more fre-
quency dependent and reflected in the E0 curve as reported 0
by Kevin [45]. In retroperspective, higher E0 in MDI-based 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Tensile strain/mm mm–1
also reflect higher ability to store energy.
Fig. 8 Effect of crosslink densities on mechanical tensile stress–
Tensile stress–strain behavior of polyurethane elastomer strain curves for jatropha oil-based polyurethane elastomers at 25 C

As the Tg for all jatropha oil-based polyurethane well energy storage, contributed by the crosslinked polymeric
below the room temperature, they exhibited highly elastic network and entrapped in the chain entanglement, was able
mechanical properties (Fig. 8). Polyurethane elastomers to compensate physical yield in the load direction without
derived from MDI showed higher tensile strength than TDI appreciable deformation in transverse direction. Other type
(Table 1). The ultimate elongation before break for speci- of vegetable oil such as palm oil had also demonstrated
mens in Table 1 was up to 131 and 194% for MDI-based similar mechanical capability as elastomer [43]. The stiff
and TDI-based, respectively. Very low Tg and decent initial response at the beginning of loading is a character-
mechanical modulus are sought in producing scaffolding istic of jatropha oil-based polyurethane, corresponding to
intended for biomedical engineering [46] or in engineering the cleavage of the hydrogen bonds and other secondary
sectors to avoid unanticipated failure. Jatropha oil-based forces in the soft segments [33]. Further the application of
polyurethane showed nonlinear brittle-like tensile stress– aromatic diisocyanate had enhanced the tensile strength.
strain curve with greater flexibility. It is worthy to note that On the other hand, one can increase the elongation at break
the produced elastomers had exceeded 100% tensile elon- by employing aliphatic diisocyanate [47]. This, however,
gation despite behaving as brittle-like in the tensile stress– may depreciate mechanical modulus of the elastomers.
strain. It was observed that, under tensile strain, there was In an unstrained state, the polymeric chains are highly
little visible necking which showed the complexity of coiled with random dispersion. As the polyurethane
mechanical integrity. It was thought that the available underwent strain, the soft segment disentangled and

123
928 A. S. Ahmad Hazmi et al.

straightened. Additional elongation diminishes the intrinsic


cohesive energy and structures of polyol and rigid seg-
ments reoriented normal to the applied stress. The ori-
entability depends on the isocyanate ratio as well as the
symmetry of isocyanate and confers capability to oppose
further elongation. Application of long chain in the soft
segment had afforded extensive plastic deformation. One
remarkable feature of the produced polyurethane in this
paper, which basis of chemical structure akin to PCL
(Fig. 2), was able to show highly elastic mechanical
deformation. Typically PCL content in polyurethane
deterred this type of mechanical response. Polyurethane
elastomers in this study had somewhat opposite mechanical
behavior with PCL-derived elastomers [7, 48, 49], albeit
having comparable Tg and tensile strength. Fig. 9 Scanning electron fractrograph showing brittle fracture result-
ing from uniaxial tensile load of jatropha oil-based polyurethane
Polyurethanes network with higher crosslink densities elastomer (magnification 3009, 5 kV). The individual fractographs
had increased mechanical Young modulus and strength at are labeled a MDI-based with r = 1/1.0, b MDI-based with r =
the expense of elongation. MDI-based elastomers, which 1/1.1, c MDI-based with r = 1/1.2, d TDI-based with r = 1/1.0,
had higher Tg and lower values of DPS, exhibited e TDI-based with r = 1/1.1, f TDI-based with r = 1/1.2. The arrows
indicate the ridges which propagated from the origins of the crack
approximately twofold increment in tensile modulus and with radial fan-shaped patterns corresponded to change in orientation
strength and shorter elongation at break. On the other hand, of the cleavage of atomic bonds at specific planes
TDI-based elastomers had higher values of DPS, suggest-
ing less hard domain crystallinity and attained outstanding
elongation at break. The crosslink sites are analogous to Table 3 Onset degradation temperature and DTG temperature at
maximum decomposition rate for jatropha oil-based polyurethane
vulcanization and act as strong anchor points to restrain elastomers
neighboring polymeric chains from slippage during defor-
Polyurethane MDI MDI MDI TDI TDI TDI
mation and depend primarily on the polyol characteristic
r = 1/ r = 1/ r = 1/ r = 1/ r = 1/ r = 1/
[2]. The majority soft segment comprised ether group 1.0 1.1 1.2 1.0 1.1 1.2
where C–O bond is easier to rotate and confer the molec-
ular flexibility. On the other hand, aromatic rings in the Onset of thermal degradation/C
hard segment limit molecular flexibility due to crystal- T2% 200 222 249 190 211 222
lization and steric hindrance. The existence of side chains T5% 264 286 291 248 254 275
which typically associated with impaired inter-molecular T10% 291 312 323 275 280 296
crystallization was found not to be detrimental to the ten- DTG temperature at maximum decomposition rate/C
sile strength. The fractographic examination (Fig. 9) using Step 1 293 301 303 281 287 286
scanning electron microscopy (SEM) revealed that the Step 2 374 383 399 392 388 391
fracture surfaces were relatively rough and irregular which Step 3 465 471 476 – – –
typically associated with brittle fracture. Appreciable gross
deformation such as twisting and tearing became apparent
for higher crosslink densities elastomers (Fig. 9c, f). MDI- based polyurethanes showed higher onset degradation and
based elastomers showed lower degree of tearing and rid- more receptive with isocyanate ratio than TDI-based
ges due to higher rigidity of molecular arrangement, while (Fig. 10). All samples showed increasing onset degradation
TDI-based elastomers showed more tearing due to greater temperature with increasing crosslink densities (Table 3).
plastic deformation and appeared clearly with increasing MDI-based elastomer had onset degradation temperature at
isocyanate ratio. 2% mass loss (T2%) remarkably increased by 49 C for
higher crosslink densities MDI-based (r = 1/1.2) against
Thermal properties of polyurethane elastomer lower crosslink densities MDI-based (r = 1/1.0). This was
due to less urethane concentration in the higher crosslink
TG and DTG analysis densities samples. MDI-based samples exhibited higher
onset temperature than TDI-based due to better contribution
The produced elastomer demonstrated relatively high ther- of crosslink densities. On the other hand, TDI-based samples
mal stability with onset degradation temperature at 5% mass exhibited less dependent on isocyanate ratio as illustrated by
loss (T5%) ranging from 248 to 291 C (Table 3). MDI- a small shift in maximum decomposition temperatures.

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 929

Fig. 10 TG curves of jatropha 200 0.0000


oil-based polyurethane at constant 0.0000

heating rate of 10 C/min under


180
nitrogen atmosphere. DTG is the –0.0005 –0.0005
derivative of TG curve, indicating
thermal decomposition rate. Inset 160

DTG/s–1
–0.0010 3
is the deconvulation DTG curve of –0.0010
TDI-based polyurethane 140 –0.0015 2
elastomer. The individual numbers 1
of inset are labeled: 1 multiplet –0.0015

DTG/s–1
120 –0.0020

band, 2 cumulative fit peak of 200 300 400 500

Mass/%
Temperature/°C
Gaussian, and 3 individual
Gaussian peaks 100 –0.0020

80
–0.0025

60 MDI (r = 1/1.0)
MDI (r = 1/1.1) –0.0030
40 MDI (r = 1/1.2)
TDI (r = 1/1.0)
TDI (r = 1/1.1) –0.0035
20 TDI (r = 1/1.2)

0 –0.0040
0 200 400 600
Temperature/°C

Application of long-chain polyol derived from vegetable oil D


R0 NHCOOR ! R0 NHR00 þ CO2 ð10Þ
relinquished inferior thermal stability typically associated
with soft segment. Thermal dissociation of allophanate where R0 is aromatic substituent and R00 and R¢¢¢ are ali-
reproduced urethane and isocyanate while dissociation of phatic substituent.
biuret regenerated disubstituted urea and isocyanate. It was The DTG curves were a combination of such individual
suggested that the branching allophanate and biuret formed profile of each component and could be deconvoluted
in this study was capable to resist initial thermal breakdown. (inset of Fig. 10) to separate peak maxima by using Origin
In DTG curves (Fig. 10), MDI-based samples had three 8.5.0. The cumulative fit peak agreed well with DTG curve
degradation steps while TDI-based had two which agreed with coefficient of determination (COD) above 0.99. There
to other findings [36, 50, 51]. Table 3 shows the temper- are three important findings as illustrated in Table 4. First
ature at maximum decomposition rate as obtained in DTG MDI-based elastomer showed substantial improvement at
curves. In the first step, low rate of the mass loss was higher crosslink densities (r = 1/1.2). The peak area shif-
associated with depolymerization including dissociation of ted to higher temperature and MDI-based (r = 1/1.2) had
branching allophanate and biuret groups. The second and more peak area at the third step as indicated by having
third step involved higher rate of the mass loss with evo- highest normalized peak area. In other word MDI-based
lution of volatile compounds such as carbon dioxide, car- (r = 1/1.2) had enhanced concentration of rigid segment
bon monoxide, and nitriles [52]. Initial mass loss is usually bounded with hydrogen bond and better molecular
began with the decomposition of urethane bond into carbon arrangement as suggested by DPS (Table 2). Hence MDI-
dioxide and isocyanate components (Eqs. 9–11) [51]. It is based exhibited exceptional improvement in physical
known that the main thermal breakdown mechanism of properties. Secondly, in contrast to MDI-based, there was
urethane is by scission of polyol-isocyanate bond. At no appreciable shift of peak area for TDI-based elastomer
higher temperature, formations of primary and secondary which suggested that thermal stability of TDI-based was
amine occur and associated with isocyanurate rings and not a strong function of crosslink densities. Thirdly, it is
carbodiimide linkage [33, 53]. Typically the decomposition worthy to note that deconvolution of TDI-based elastomers
of urethane is described as following reactions [33]: prepared in this paper required three peaks to achieve
COD [ 0.99. This was not clearly presented in the original
D
R0 NHCOOR00 ! R0 NCO þ R00 OH ð8Þ DTG curves. The existence of the third peak for TDI-based
D was probably owed to functionality and heterogeneity in
R0 NHCOOCR002 CH2 R000 ! R000 CH functional group positions.
¼ CR002 þ R0 NH2 þ CO2 ð9Þ

123
930 A. S. Ahmad Hazmi et al.

Table 4 Deconvolution results of DTG curves of jatropha oil-based polyurethane elastomers


MDI r = 1/1.0 MDI r = 1/1.1 MDI r = 1/1.2 TDI r = 1/1.0 TDI r = 1/1.1 TDI r = 1/1.2

Peak temperature/C
Step 1 290 299 314 288 293 296
Step 2 373 390 398 391 392 391
Step 3 459 469 473 453 453 452
Peak areaa/%/Normalized peak areab
Step 1 15/1 8/0.5 14/1.5 19/1.5 19/1.4 21/1.6
Step 2 40/2.8 47/3.2 29/3.1 60/4.6 63/4.6 59/4.5
Step 3 45/3.1 45/3.0 57/6.1 21/1.6 18/1.3 20/1.5
Peak width/C
Step 1 54 35 54 62 64 69
Step 2 62 74 58 54 51 49
Step 3 46 34 45 34 36 35
CODc 0.9962 0.9971 0.9976 0.9960 0.9948 0.9957
a
Peak area = Peak area*100/Total peak area
b
Normalized Area = Peak area/first peak area of MDI r = 1/1.0
c
Coefficient of determination (COD) is the measurement of goodness of fit

400 400 400


MDI MDI MDI
350 r = 1/1.0 350 r = 1/1.1 350 r = 1/1.2
Activation energy/kJ mol–1

Activation energy/kJ mol–1

Activation energy/kJ mol–1


300 300 300

250 250 250

200 200 200

150 150 150

100 100 100


0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Conversion Conversion Conversion
300 300 300
TDI TDI TDI
r = 1/1.0 r = 1/1.1 r = 1/1.2
Activation energy/kJ mol–1
Activation energy/kJ mol–1

Activation energy/kJ mol–1

250 250 250

200 200 200

150 150 150

100 100 100


0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Conversion Conversion Conversion

Fig. 11 Dependence of activation energy (Ea) on conversion (a) as calculated based on Ozawa method

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 931

Kinetic of thermal degradation The decomposition of elastomers in this paper under the
given experimental condition can be best described by a
Via isoconversional method, the temperatures correspond model assuming a restricted nucleation growth mechanism
to constant conversion are defined from the resultant mass such as n-dimensional Avrami-Erofeev solid-state model
loss curves. The extent of conversion (a) at a given tem- [54] which integrated ingestion of phantom nuclei and
perature is defined as: coalescence as summarized in Table 5. The shape of Ea-a
mi  mT curve (Fig. 11) typically reflects the associated kinetic
a¼ ð11Þ
mi  mf mechanism. At the beginning was diffusion-controlled
where mi is the initial mass of step changes, mT is the mass regime with concave-upward curve and had n less than 1 for
of isoconversion at temperature T, and mf is the final mass a \ 0.1. Diffusion is a rate-limited step governed by trans-
of step changes. The rate equation is given by: port phenomena which included formation of product at the
solid–gas interface. It was perceived at the same a, elastomer
da (r = 1/1.2) had higher reaction order indicating higher order
¼ kðT Þf ðaÞ ð12Þ
dt of chemical cleavage of branching groups along the poly-
where k(T) is rate constant and f(a) is the reaction model. meric network. Diffusion type of thermal degradation, onset
Temperature dependence of Arrhenius expression gives degradation temperature, and activation energy for the first
k(T) = A exp(-Ea/RT) where A is the pre-exponential step (a = 0.1) were characteristic of thermally unsta-
factor, Ea is the activation energy, R is the universal gas ble urethane, allophanate and biuret groups. A wide region
constant, and T is temperature at conversion. Ea, A, and between 0.1 \ a \ 0.6 indicated a gradual transition from
reaction order (n) are usually derived in kinetic analysis. diffusion-type to pseudo-type. This transition was evidenced
The experimental procedures were carried out based on on increasing values of Ea and n. Acceleration on Ea-a curve
isoconversional at five different heating rates. Then the typically associated with thermal breakdown along weaker
kinetic parameters were determined by applying Ozawa bond and formation of free radical. Higher values of Ea
[54], Flynn–Wall, and Kissinger [55] methods. Different beyond the middle step (a [ 0.6) with concave-upward
values of Ea at different a verified that jatropha oil-based curve were possibly contributed from relatively medium
polyurethane decomposed in multiple steps in Ozawa plot network functionality. Nevertheless, high molecular weight
(Fig. 11). of soft segment had contributed to enhanced Ea.

Table 5 Kinetic of thermal degradation analysis based on isoconversional method for jatropha oil-based polyurethane elastomers
Polyurethane Activation energy (Ea)/kJ mol-1/pre-exponential factor (ln A)/min-1/reaction order (n)a
Flynn–Wallb Ozawab Kissingerc
a = 0.05 Step 1, a = 0.1 Step 2, a = 0.6 Step 3, a = 0.9 Step 1, a = 0.1 Step 2, a = 0.6 Step 3, a = 0.9

MDI 136/14.5/0.49 126/30.0/0.56 199/37.3/1.13 247/43.4/2.36 156/32.7/1 194/35.5/1 252/40.4/1


r = 1/1.0
MDI 170/17.6/0.44 165/38.5/0.59 302/54.6/1.27 325/55.6/2.65 164/34.6/1 220/39.8/1 292/46.7/1
r = 1/1.1
MDI 208/21.2/0.66 184/43.2/0.54 343/62.8/0.97 348/59.7/2.84 176/36.5/1 258/47.4/1 327/52.9/1
r = 1/1.2
TDI 144/15.4/0.41 145/34.4/0.69 181/35.3/0.88 261/46.4/2.42 141/29.9/1 185/32.8/1 –
r = 1/1.0
TDI 148/15.9/0.45 164/38.5/0.64 186/38.3/1.05 277/49.5/2.46 139/29.4/1 175/31.2/1 –
r = 1/1.1
TDI 154/16.3/0.75 174/40.6/0.67 201/39.4/0.89 273/49.5/2.26 124/26.4/1 152/26.7/1 –
r = 1/1.2
 
a 1 Ea
 1 da

Determined from n-dimensional Avrami-Erofeev model and approximated as ln n1 ¼ R T þ ln A  ln dt
½nð1aÞ½ lnð1aÞ n
b
Isoconversional method of Ozawa and Flynn–Wall apply DoylE0 s approximation as lnðbi Þ ¼ const  RTEaa;i . The Ea is obtained as the slope in
the plot of ln qi versus 1/Ta,i as described in ASTM E1641-99
   
c Ea b n1 Ea
To determine A and n, the rate equation based on Kissinger was solved as RT 2 ¼ Anð1  aÞm exp RT m
and Ea is represented by the slope of
m

the plot of ln (b/T2m) versus 1/Tm

123
932 A. S. Ahmad Hazmi et al.

Furthermore, the side chains and heterogeneity in polyol 5. Malaysian Palm Oil Board. Malaysian Oil Palm Statistics 2012.
structure may introduced steric hindrance and made 32 ed. Bangi; 2013.
6. Saliterman S. Fundamentals of BioMEMS and medical
decomposition processes more difficult. The last degrada- microdevices. Washington: SPIE Press; 2006.
tion step (0.7 \ a \ 0.8) was dominated by second-order 7. Loh XJ, Colin Sng KB, Li J. Synthesis and water-swelling of thermo-
reaction and closely related to C–C bond dissociation. De- responsive poly(ester urethane)s containing poly(e-caprolactone),
acceleration of degradation which showed convex-upward poly(ethylene glycol) and poly(propylene glycol). Biomaterials.
2008;29(22):3185–94. doi:10.1016/j.biomaterials.2008.04.015.
at final step (a [ 0.8) was related with char formation. The 8. Woodruff MA, Hutmacher DW. The return of a forgotten poly-
values of n verified the consecutive nature of degradation mer—polycaprolactone in the 21 st century. Prog Polym Sci.
of the elastomer. The wide variation of Ea is characteristic 2010;35(10):1217–56. doi:10.1016/j.progpolymsci.2010.04.002.
of complexity for continuous degradation mechanism and 9. Cherng JY, Hou TY, Shih MF, Talsma H, Hennink WE. Poly-
urethane-based drug delivery systems. Int J Pharm.
summarized in Table 5. The TG/DTG curves were com- 2013;450(1–2):145–62. doi:10.1016/j.ijpharm.2013.04.063.
parable to soybean, castor [51] and PCL [56]. 10. May-Hernández L, Hernández-Sánchez F, Gomez-Ribelles JL,
Sabater-i Serra R. Segmented poly(urethane-urea) elastomers
based on polycaprolactone: structure and properties. J Appl
Polym Sci. 2011;119(4):2093–104. doi:10.1002/app.32929.
Conclusions 11. Nair LS, Laurencin CT. Biodegradable polymers as biomaterials.
Prog Polym Sci. 2007;32(8–9):762–98. doi:10.1016/j.progpo
A new class of enhanced bio-polyurethane elastomer was lymsci.2007.05.017.
successfully developed from medium functionality non-ed- 12. Saalah S, Abdullah LC, Aung MM, Salleh MZ, Awang Biak DR,
Basri M, et al. Waterborne polyurethane dispersions synthesized
ible vegetable oil-based polyol which both chemical struc- from jatropha oil. Ind Crops Prod. 2015;64:194–200. doi:10.
ture and physical properties comparable to PCL-derived 1016/j.indcrop.2014.10.046.
polyurethane. Effects of polyol network structure toward 13. Petrović ZS. Polyurethanes from vegetable oils. Polym Rev.
polyurethane physical properties at different crosslink den- 2008;48(1):109–55. doi:10.1080/15583720701834224.
14. Petrović ZS, Guo A, Javni I, Cvetković I, Hong DP. Polyurethane
sities were characterized by thermal and mechanical analy- networks from polyols obtained by hydroformylation of soybean
sis. Application of medium functionality and high molecular oil. Polym Int. 2008;57(2):275–81. doi:10.1002/pi.2340.
weight polyol had beneficially revamped the polymer net- 15. Petrovic ZS, Javni I, Guo A, Zhang W, inventors; Method of
work formation and physical properties. Produced elas- making natural oil-based polyols and polyurethanes therefrom.
US Patent 6686435 patent 6686435. 2004.
tomers had highly elastic mechanical behavior and very low 16. Zlatanić A, Lava C, Zhang W, Petrović ZS. Effect of structure on
Tg which was vital to conserve resilience at low temperature properties of polyols and polyurethanes based on different veg-
or to withstand a high frequency environment. The thermal etable oils. J Polym Sci Part B Polym Phys. 2004;42(5):809–19.
and mechanical properties demonstrated tunable properties doi:10.1002/polb.10737.
17. Petrović ZS, Yang L, Zlatanić A, Zhang W, Javni I. Network
by resolving the crosslink densities. The produced veg- structure and properties of polyurethanes from soybean oil.
etable oil-based polyurethanes were thermally stable and J Appl Polym Sci. 2007;105(5):2717–27. doi:10.1002/app.26346.
had relatively high onset degradation temperature. 18. De Genova R, Malsam J, Zlatanic A, Wazirzada Y. Blo-based
polyols for the flexible slabstock foam industry. J Oil Palm Res.
Acknowledgements This work is supported by the Graduate 2008; Spec. Iss. October: 53–60.
Research Fellowship, Universiti Putra Malaysia. The authors 19. Badri KH, Ahmad SH, Zakaria S. Production of a high-func-
acknowledge personnel at Universiti Putra Malaysia, Universiti tionality RBD palm kernel oil-based polyester polyol. J Appl
Kebangsaan Malaysia, Malaysian Nuclear Agency and Malaysian Polym Sci. 2001;81(2):384–9. doi:10.1002/app.1449.
Palm Oil Board for their expert contribution and research facilities. 20. Gübitz GM, Mittelbach M, Trabi M. Exploitation of the tropical
The authors thank Director of Advanced Oleochemical Technology oil seed plant Jatropha curcas L. Biores Technol.
Division and Nurul Farhana Omar for kind proofread this work. 1999;67(1):73–82. doi:10.1016/s0960-8524(99)00069-3.
21. Ahmed WA, Salimon J. Phorbol ester as toxic constituents of
tropical Jatropha curcas seed oil. Eur J Sci Res. 2009;31(3):429–36.
22. Hazmi ASA, Aung MM, Abdullah LC, Salleh MZ, Mahmood
MH. Producing Jatropha oil-based polyol via epoxidation and
References ring opening. Ind Crops Prod. 2013;50:563–7. doi:10.1016/j.
indcrop.2013.08.003.
1. Modern Plastic International. Mod Plast Int. Lausanne; 23. Flory PJ, Rehner J. Statistical mechanics of cross-linked polymer
1990;20(1):31. networks I. Rubberlike Elast J Chem Phys. 1943;11(11):512.
2. Ionescu M. Chemistry and technology of polyols for poly- 24. Krevelen DWV, Nijenhuis KT. Properties of polymers: their
urethanes. Shawbury: Rapra Technology; 2005. p. 13–50. correlation with chemical structure; their numerical estimation
3. Desroches M, Escouvois M, Auvergne R, Caillol S, Boutevin B. and prediction from additive group contributions. 4th ed. Ams-
From vegetable oils to polyurethanes: synthetic routes to polyols terdam: Elsevier Science; 2009.
and main industrial products. Polym Rev. 2012;52(1):38–79. 25. Petrović ZS, Guo A, Zhang W. Structure and properties of
doi:10.1080/15583724.2011.640443. polyurethanes based on halogenated and nonhalogenated soy–
4. Lligadas G, Ronda JC, Galià M, Cádiz V. Plant oils as platform polyols. J Polym Sci Part A Polym Chem. 2000;38(22):4062–9.
chemicals for polyurethane synthesis: current state-of-the-art. doi:10.1002/1099-0518(20001115)38:22\4062:aid-pola60[3.0.
Biomacromol. 2010;11(11):2825–35. doi:10.1021/bm100839x. co;2-l.

123
Understanding intrinsic plasticizer in vegetable oil-based polyurethane elastomer as… 933

26. ASTM Standard D3418-03 I. Standard test method for transition 43. Clark AJ, Hoong SS. Copolymers of tetrahydrofuran and
temperatures and enthalpies of fusion and crystallization of epoxidized vegetable oils: application to elastomeric poly-
polymers by differential scanning calorimetry. West Con- urethanes. Polym Chem. 2014;5(9):3238–44. doi:10.1039/c3py
shohocken, PA; 2003. 01527k.
27. ASTM Standard E-1641-99 I. Standard test method for decom- 44. Menard KP. Dynamic mechanical analysis: a practical introduc-
position kinetics by thermogravimetry. West Conshohocken, PA; tion. Boca Raton: CRC Press; 2008.
1999. 45. Kevin M. Thermomechanical and dynamic mechanical analysis.
28. ASTM Standard D5026-01 I. Standard test method for plastics: In: Lobo H, Bonilla JV, editors. Handbook of plastics analysis.
dynamic mechanical properties: In: Tension. West Con- Boca Raton: CRC Press, 2003. p. 102–90.
shohocken, PA; 2001. 46. Zieleniewska M, Auguścik M, Prociak A, Rojek P, Ryszkowska
29. ASTM Standard D638-03 I. Standard test method for tensile J. Polyurethane-urea substrates from rapeseed oil-based polyol
properties of plastics. West Conshohocken, PA; 2003. for bone tissue cultures intended for application in tissue engi-
30. Nishikida K, Coates J. Infrared and Raman analysis of polymers. neering. Polym Degrad Stab. 2014;. doi:10.1016/j.poly
In: Lobo H, Bonilla JV, editors. Handbook of plastics analysis. mdegradstab.2014.03.010.
CRC Press; 2003. p. 650. doi:10.1201/9780203911983.ch7. 47. Puszka A, Kultys A. New thermoplastic polyurethane elastomers
31. Lapprand A, Boisson F, Delolme F, Méchin F, Pascault JP. based on aliphatic diisocyanate. J Therm Anal Calorim.
Reactivity of isocyanates with urethanes: conditions for allo- 2017;128(1):407–16. doi:10.1007/s10973-016-5923-7.
phanate formation. Polym Degrad Stab. 2005;90(2):363–73. 48. Guan J, Sacks MS, Beckman EJ, Wagner WR. Biodegradable
doi:10.1016/j.polymdegradstab.2005.01.045. poly(ether ester urethane)urea elastomers based on poly(ether
32. Delebecq E, Pascault J-P, Boutevin B, Ganachaud F. On the ester) triblock copolymers and putrescine: synthesis, characteri-
versatility of urethane/urea bonds: reversibility, blocked iso- zation and cytocompatibility. Biomaterials. 2004;25(1):85–96.
cyanate, and non-isocyanate polyurethane. Chem Rev. doi:10.1016/S0142-9612(03)00476-9.
2013;113(1):80–118. doi:10.1021/cr300195n. 49. Hong Y, Guan J, Fujimoto KL, Hashizume R, Pelinescu AL,
33. Wirpsza Z. Polyurethanes, chemistry, technology and applica- Wagner WR. Tailoring the degradation kinetics of poly(ester
tions. Polymer science and technology series. New York: Ellis carbonate urethane)urea thermoplastic elastomers for tissue
Horwood; 1993. engineering scaffolds. Biomaterials. 2010;31(15):4249–58.
34. Sonnenschein MF. Introduction to polyurethane chemistry. doi:10.1016/j.biomaterials.2010.02.005.
Polyurethanes. Wiley; 2014. p. 105–26. 50. Aung MM, Yaakob Z, Kamarudin S, Abdullah LC. Synthesis and
35. Dušek K, Špı́rková M, Ilavský M. Network formation in poly- characterization of Jatropha (Jatropha curcas L.) oil-based
urethanes due to allophanate and biuret formation: gel fraction polyurethane wood adhesive. Ind Crops Prod. 2014;60:177–85.
and equilibrium modulus. Makromol Chem Macromol Symp. doi:10.1016/j.indcrop.2014.05.038.
1991;45(1):87–95. doi:10.1002/masy.19910450112. 51. Javni I, Petrović ZS, Guo A, Fuller R. Thermal stability of
36. Semsarzadeh MA, Navarchian AH. Effects of NCO/OH ratio and polyurethanes based on vegetable oils. J Appl Polym Sci.
catalyst concentration on structure, thermal stability, and cross- 2000;77(8):1723–34. doi:10.1002/1097-4628(20000822)77:
link density of poly(urethane-isocyanurate). J Appl Polym Sci. 8\1723:aid-app9[3.0.co;2-k.
2003;90(4):963–72. 52. Beyler CL, Hirschler MM. Chapter 7, Thermal decomposition of
37. Redakcji O, Ryszkowska J. Supermolecular structure, morphol- polymers. In: SFPE handbook of fire protection engineering, vol.
ogy and physical properties of urea-urethane elastomers. Polim- 2. 2002. p. 111–31.
ery. 2012;57(11–12):777–85. 53. Hablot E, Zheng D, Bouquey M, Avérous L. Polyurethanes based
38. Dey J, Xu H, Shen J, Thevenot P, Gondi SR, Nguyen KT, et al. on castor oil: kinetics, chemical, mechanical and thermal prop-
Development of biodegradable crosslinked urethane-doped erties. Macromol Mater Eng. 2008;293(11):922–9. doi:10.1002/
polyester elastomers. Biomaterials. 2008;29(35):4637–49. doi:10. mame.200800185.
1016/j.biomaterials.2008.08.020. 54. Prime RB, Bair HE, Vyazovkin S, Gallagher PK, Riga A. In:
39. Thomson T, inventor Timothy Thomson, assignee. Hydrophilic Menczel JD, Prime RB, editors. Thermogravimetric analysis
polyurethane as a matrix for the delivery of skin care ingredients. (TGA). Thermal analysis of polymers. Wiley; 2009. p. 241–317.
US patent 20020182245. 2002. doi:10.1002/9780470423837.
40. Macosko CW, Miller DR. A new derivation of average molecular 55. Kissinger HE. Reaction kinetics in differential thermal analysis.
weights of nonlinear polymers. Macromolecules. Anal Chem. 1957;29(11):1702–6. doi:10.1021/ac60131a045.
1976;9(2):199–206. doi:10.1021/ma60050a003. 56. Bogdanov B, Toncheva V, Schacht E, Finelli L, Sarti B, Scandola
41. Petrović ZS, Zhang W, Javni I. Structure and properties M. Physical properties of poly(ester-urethanes) prepared from
of polyurethanes prepared from triglyceride polyols by different molar mass polycaprolactone-diols. Polymer.
ozonolysis. Biomacromol. 2005;6(2):713–9. doi:10.1021/bm049 1999;40(11):3171–82. doi:10.1016/S0032-3861(98)00552-7.
451s.
42. Fox TG, Loshaek S. Influence of molecular weight and degree of
crosslinking on the specific volume and glass temperature of
polymers. J Polym Sci. 1955;15(80):371–90.

123

You might also like