Sei sulla pagina 1di 9

European Polymer Journal 85 (2016) 53–61

Contents lists available at ScienceDirect

European Polymer Journal


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

Macromolecular Nanotechnology

Production and characterization of polyurethane foams from a


simple mixture of castor oil, crude glycerol and untreated lignin
as bio-based polyols
Camila S. Carriço, Thaís Fraga, Vânya M.D. Pasa ⇑
Laboratório de Produtos da Biomassa, Departamento de Química, Universidade Federal de Minas Gerais, Av. Antonio Carlos, 6627, 31270-901 Belo Horizonte,
Minas Gerais, Brazil

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

Article history: Polyurethane foams from bio-based polyols have been studied to produce sustainable and
Received 11 May 2016 eco-friendly materials from biomass industrial wastes. Semi-rigid polyurethane foams
Received in revised form 9 August 2016 were synthesized from physical mixtures of kraft lignin, castor oil and residual glycerol.
Accepted 6 October 2016
The influence of different lignin contents (10, 12.5, 15, 17.5, 20, 25, 30 and 40%) on the
Available online 8 October 2016
foam properties was evaluated. An increase in the density and a decrease in the thermal
stability were observed upon an increasing amount of lignin. The foam synthesized with
Keywords:
17.5% lignin showed better dimensional and thermal properties as well as the best cell
Polyurethane foams
Lignin
homogeneity and this formulation was chosen for further investigation of the effects of
Crude glycerol the castor oil content, isocyanate/hydroxyl molar ratio (NCO/OH) and the blowing agent
Castor oil type. Increased foam densities and compressive strengths were observed with increased
castor oil content in the foams. Similar results were observed upon increasing the NCO/
OH molar ratio due to the higher amount of crosslinking reactions. The evaluation of the
blowing agent effect showed that using cyclopentane and n-pentane, the foams showed
higher densities and lower compressive strengths in comparison to the foams synthesized
with the chemical blowing agent (water). This result can be explained by the faster
volatilization of these organic blowing agents, producing smaller cells in comparison to
the CO2 generated by the reaction of water with isocyanate. Green polyurethane foams
with good properties and industrial interest were produced using a mixture of residual
raw materials (glycerol and lignin) and castor oil as bio-based polyols.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Among all the products derived from polyurethane, foams are versatile polymeric materials with a broad market that can
be classified as flexible, semi-rigid, or rigid depending on their mechanical performance characteristics and densities. The
main applications of foams are in the furniture industry, including in upholstery, insulation, and packing, and in automotive
applications, including seat cushioning, bumpers, and sound insulation. This polymer is produced by the reaction of an iso-
cyanate and a polyol in addition to other additives used to adjust the properties of the final foam product. However, most of
these reagents are derived from petrochemicals, increasing the dependence on petroleum. Therefore, the environmental

⇑ Corresponding author.
E-mail addresses: vmdpasa@terra.com.br, vanya@ufmg.br (V.M.D. Pasa).

http://dx.doi.org/10.1016/j.eurpolymj.2016.10.012
0014-3057/Ó 2016 Elsevier Ltd. All rights reserved.
54 C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61

concern and need for sustainable technologies motivate the production of the polyurethane foams from renewable feedstock
[1,2].
Polyurethane foams from bio-based polyols have been studied for the production of sustainable and eco-friendly mate-
rials from industrial waste biomass such as lignocellulosic [3,4], bio-pitch [5], glycerol [6] and vegetable oils [7]. Among
these raw materials, lignin is a potential candidate for producing bio-based polyurethane foams due to its abundance and
complex chemical structure that allows its use in a wide range of applications. This oligomer exists naturally in plants
and trees, and comprises about 30% of wood. Lignin is an aromatic complex polymer derived from coniferyl, coumaryl
and sinapyl alcohol that contains high concentrations of phenolic and aliphatic hydroxyl groups. Lignin is obtained in large
quantities from pulp and paper mills or, more recently, as a residue of cellulosic ethanol production process [2,8].
Recently, many studies were carried out using lignins and other by-products in polyurethane foam synthesis, as reported
by Hatakeyama et al. [9,10], who used different lignins dissolved in molasses and polyglycols. Another interesting study was
performed by Li and Ragauskas [11,12], who prepared a polyol by reacting lignin with propylene oxide catalysed by potas-
sium hydroxide. Among other studies, Cinelli et al. [1] liquefied lignin using glycerol and PEG400 as the liquefying agent, and
Mahmood et al. [8] also produced foams using hydrolytically depolymerized Kraft lignin.
These studies indicate that lignin can be used directly or modified chemically by hydrolysis, epoxidation, hydroxylation,
ozonolysis, and hydrogenation. Recently, many methods were developed for the conversion of biomass into bio-based poly-
ols via liquefaction [13]. However, in addition to the use of some solvents, most of these bio-based polyols production meth-
ods require the use of some energy source, mainly heat or microwave, to homogenize the system phases and optimize the
physical properties of the polyols.
The physical mixing of raw materials is desirable for the use of by-products from bio-fuel production and industrial waste
without any type of chemical or physical pre-treatment and for direct use in polyurethane synthesis. In addition to the phys-
ical mixing being a simple method for the homogenization of miscible materials, the use of raw materials without pre-
treatment is of great interest to industry because it minimizes the number of steps in production, thereby accelerating
the execution of the overall production process and reducing costs.
The production of polyols from industrial residual sources, such as lignin (from paper industry and cellulosic ethanol pro-
duction) and crude glycerol (obtained as by-product of biodiesel production), add value to these residues, and also contribute
to the economical feasibility of important cellulose industry, biodiesel and bioethanol commodities. In this study, bio-based
polyols were synthesized without pre-treatment processes using lignin, crude glycerol and castor oil. The polyols mixtures
were used to synthesize different polyurethane foams. The foams’ physical chemical and mechanical properties were eval-
uated, and these results appear to be promising for several industrial applications.

2. Materials and methods

2.1. Materials

Eucalyptus Kraft pulp lignin (Suzano Industry of Paper and Cellulose, Limeira, São Paulo - Brazil) was obtained from a cel-
lulose pulping process (black liquor) and is therefore used in its alkaline form. Crude glycerol, obtained as a co-product of
biodiesel production, was kindly supplied by Petrobrás (Usina Darcy Ribeiro-Montes Claros, Minas Gerais - Brazil). The castor
oil was provided by the Polyurethane Company (Betim, Minas Gerais - Brazil). Desmodur 44 V 20 (Bayer), which is a mixture
of 4,40 -diphenylmethane-diisocyanates, was used as the isocyanate source. Tegostab 8460 (Evonik), a polyether-modified
polysiloxane, was used as the surfactant. DBTDL (dibutyltin dilaurate), an organometallic catalyst called Kosmos 19 manu-
factured by Evonik, was used as the catalyst in the syntheses of the polyurethane foams. Cyclopentane, n-pentane (Sigma
Aldrich) and distilled water were used as the blowing agents.

2.2. Foam syntheses

Polyurethane foams were synthesized using the batch process method, also known as the box process. Initially, the poly-
ols were produced by the physical mixture method, consisting of mixing two or more compounds to obtain a homogeneous
material. Lignin (L), crude glycerol (G) and castor oil (CO) were mixed with a mechanical stirrer, Fisatom model 713 D, until
the system became completely homogeneous. The surfactant, catalyst and blowing agent were added to the polyol, and this
mixture was kept under vigorous stirring for 1 min at 500 rpm. Finally, the isocyanate was added and remained under agi-
tation for 20 s. The formulation was poured into a wooden mould with dimensions of 7.0  7.0  20.0 cm for the growth of
the polymer foam, and the mould was kept closed for 24 h at room temperature to perform curing. After curing, the box was
opened and the foam was characterized.
For preliminary scanning tests varying the lignin content, 2% of water was used as the blowing agent and 2% of DBTDL
(dibutyltin dilaurate) was used as the catalyst. The foam formulations for the bio-based polyols are shown in Table 1, and
the isocyanate was added at the NCO/OH ratio of 1.1. The influence of the NCO/OH molar ratio, catalyst type and content,
and blowing agent type and content were investigated for the foams that showed the best dimensional and thermal stability.
C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61 55

Table 1
Foam formulations for bio-based polyols.

Formulation A B C D E F G H
Reagents Parts by weight
Lignin 10 12.5 15 17.5 20 25 30 40
Glycerol 40 37.5 35 32.5 30 25 20 10
Castor oil 50 50 50 50 50 50 50 50
Catalyst 2 2 2 2 2 2 2 2
Surfactant 2 2 2 2 2 2 2 2
Blowing agent 2 2 2 2 2 2 2 2

2.3. Characterizations

The viscosity of raw materials was measured using a viscometer (Herzog Ubbelohde Viscometer, model UVH 481) accord-
ing to the ASTM D 445 standard. The polyol hydroxyl number values were determined according to the ASTM D4274 stan-
dard in an automatic titrator (Kyoto Electronics, AT 500 model). The FTIR spectra of the raw materials, polyols and foams
were collected in a ABB Bomer spectrometer using 0.1% KBr pellets with a resolution of 4 cm1, in the 4000–400 cm1 range.
Thermal analyses by thermogravimetry (TG) were carried out in a TA equipment model TA-50 Q at temperatures ranging
from 30 to 800 °C with a heating rate of 10 °C min1 under a nitrogen flow of 40 mL min1 to verify the thermal stability
at different temperatures. The optical microscopy images were obtained using an Olympus Optical Microscope Model
BX41M coupled to a TecVoz camera model DNS 480. Scanning electronic microscopy (SEM) was performed with a JEOL
JSM - 6360LV scanning microscope to observe and evaluate the cellular structure of the foam. The apparent density was mea-
sured according to the ASTM D1622 standard, using 50  50  25 mm samples in triplicate. Compression tests were carried
out using an Autograph Precision Universal Testing Machine AG-Xplus Series (Shimadzu) according to the ASTM D1621 stan-
dard, with at least five samples tested to obtain average values, and the specimens’ dimensions were 50  50  25 mm.

3. Results and discussion

3.1. Raw material and polyol characterization

The results presented in Table 2 showed that crude glycerol has the highest reactivity and can be considered as a
crosslinking agent due to its low size and high functionality. Lignin has a hydroxyl number that is 6 times smaller than that
of the crude glycerol, implying a lower reactivity; however, the reactivity of lignin is still higher than that measured for cas-
tor oil. These values are in agreement with the values reported in the literature [8,15].
Castor oil showed a viscosity value 10 times higher than that obtained for crude glycerol, which can be related to its struc-
ture. Glycerol possesses only 3 carbons with 3 hydroxyl groups, but the effect on the viscosity of the hydrogen linkage caused
by this polar group was smaller than that of the large chain of the castor oil, which has 18 carbon atoms.
The infrared spectra of the raw materials used for the preparation of polyols are shown in Fig. 1. The bands at approxi-
mately 3400 cm1 correspond to the vibration of the hydroxyl groups, and the bands at approximately 2920 and 2870 cm1
were assigned to CH2 and CH3 stretches of aliphatic chains. The characteristic stretches of double bonds in the castor oil
C@CAH and C@CAC groups were observed at 3012 and 1621 cm1, respectively. An average band at 1465 cm1 was due
to the deformation of alkene CH2 groups, also present in the castor oil structure. The average intensity of the bands at
1710–1700 cm1, characteristic of carbonyl and carboxyl groups, were observed in the spectrum of lignin and most notably
in the castor oil spectrum. The bands at 1600 and 1515 cm1 correspond to the vibration of the aromatic rings, and the bands
at 1460–1420 cm1, due to the deformation of the CH on the aromatic ring, were also observed in the lignin spectrum; addi-
tionally, several absorption bands were observed below 1400 cm1, representing the contribution of various vibrational
modes [8,14,15].
The changes of the hydroxyl number of LGCO polyols upon variation of the lignin content are shown in Fig. 2. As the lignin
content increases, the hydroxyl group value decreased, which can be related to the reduction of the glycerol content, as
shown in Table 2. Glycerol has a higher functionality owing to its high hydroxyl number, so that result was expected; the
lower content of glycerol and the higher content of lignin reduced the polyol hydroxyl number. The castor oil content
was unaltered for all polyol samples.

Table 2
Physical chemical properties of the raw materials.

Raw materials Hydroxyl number (mg KOH g1) Viscosity (mPa s1 at 30 °C)
Lignin 305.6 –
Crude glycerol 1827.6 111.7
Castor oil 163.0 1101.7
56 C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61

LGCO polyol
250
Lignin
200

Transmitance (%)
150
Castor oil

100 Crude glycerol

50

0
4000 3500 3000 2500 2000 1500 1000
-1
Wavenumber (cm )

Fig. 1. Spectra of raw materials for polyol production.

1100

1000
Hydroxyl number (mg KOH/g)

900

800

700

600

500

400

300

200

10 15 20 25 30 35 40
Lignin content (%)

Fig. 2. Hydroxyl number versus lignin content of the polyols used to produce containing a mixture of lignin, crude glycerol and castor oil (LGCO polyol).

3.2. Foam characterization

The composition of polyurethane foams was evaluated by FTIR analysis. The spectra are presented in Fig. 3 and show the
characteristic bands for the main functional groups. The NH stretching and vibrations were observed between 3808 and
3308 cm1 and 1512 and 1510 cm1, respectively. The presence of unreacted isocyanate can be observed in a small band
at 2270 cm1, which can be explained due to the molar ratio used in these formulations (NCO/OH = 1.1). Other foams were
formulated with lower NCO/OH ratio, but these materials did not show dimensional stability. The two thin bands at 2900
and 2890 cm1, corresponding to deformation of CH2 bonds and other CH bond vibrational modes, are also observed at
1464, 1418, 1364 and 1294 cm1. The band between 1730 and 1720 cm1 corresponds to the CO bond stretch of the free
urethane, and at approximately 1700 cm1, the hydrogen bond between the carbonyl and hydrogen atoms (from OCONH
groups) from the urethane is also observed. Furthermore, a band related to the stretching of OCONH asymmetric links at
1380 cm1 was also revealed [1,16,17]. The band at 1726 cm1, related to the carbonyl group, becomes more extended with
increasing lignin content owing to the appearance of the band corresponding to the hydrogen bond with C@O at 1705 cm1;
this result supports the formation of hydrogen bonds between the polar groups of the lignin and urethane [18].
The results of the thermal analysis for the various lignin contents in the foams are shown in Fig. 4. The thermal degra-
dation of polyurethanes occurs as a result of various physical and chemical phenomena that are not dominated by a single
process and depend mainly on the balance between the polymerization and depolymerization of the functional groups or
bonds present in the polymer chains. Three weight loss regions were observed in the TG curves. A small event observed
in the TG curve at approximately 260 °C (approximately 5% weight loss) is probably due to the loss of some unreacted small
molecules. The second event, observed at approximately 300 °C, corresponds to the degradation of the urethane linkage,
which is related to the hard segments of polyurethane structure. This event corresponds to a weight loss of 35%. The third
degradation event is observed at 350 °C and is related to the lignin decomposition. The fourth event at 450 °C is associated to
the thermal degradation of the soft segments [1,5,17].
C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61 57

150 A

B
125
C

Transmitance (%)
100 D

E
75
F
50
G
25
H

0
4000 3500 3000 2500 2000 1500 1000
-1
Wavenumber (cm )

Fig. 3. FTIR spectra of foams produced with different lignin content (A) 10%, (B) 12.5%, (C) 15%, (D) 17.5%, (E) 20%, (F) 25%, (G) 30% and (H) 40% of lignin.

100 0,000
A
90
B
-0,002
80
C
70
-0,004 D
Weight (%)

60
A
50 E
D -0,006
40 B
F
30 F -0,008
C G
20
10 H -0,010 H
G E
0
100 200 300 400 500 600 700 200 250 300 350 400 450 500 550 600
Temperature (°C) Temperature (°C)
(a) (b)
Fig. 4. TG (a) and DTG (b) curves of foams with LGCO polyol containing different amounts of lignin (A) 10%, (B) 12.5%, (C) 15%, (D) 17.5%, (E) 20%, (F) 25%, (G)
30% and (H) 40% of lignin.

The data for the foam densities for samples with different lignin contents are shown in Fig. 5; and these density values are
in agreement with the literature data for foams containing lignin [6,8,9,11]. It was also observed that there is a trend of
increased density with higher lignin content in the foams. This probably occurs because, in addition to the enhanced system
viscosity, the addition of lignin produced polymer materials with a lower free volume and a higher density; as a conse-
quence, smaller and denser cells were obtained [19].
Relating the results presented in Figs. 2 and 5, it was observed that polyols with a higher hydroxyl number (lower lignin
and higher glycerol amount) yielded foams with lower densities. While higher densities were expected from higher OH num-
bers, due to the increase of crosslinking reactions, an opposite tendency was observed for our results. This behavior can be
explained by the short size of glycerol chain, which does not allow crosslinking reactions, causing an increase in the polymer
free volume and consequently a decrease in the density values.
The images obtained by the optical microscopy of the cross-section foams produced from the polyol LGCO are shown in
Fig. 6. The cells exhibited an elongated shape (elliptical), and it was observed that increasing the lignin content decreases the
cell size, confirming the density data (Fig. 5). The presence of smaller particles agglomerated in the foams synthesized with
higher lignin content was also observed. These observations indicate a low homogeneity for the foams with high concentra-
tions of lignin, especially for those with values higher than 17.5%. These results are consistent with the apparent density of
the foams, suggesting that the addition of lignin in these formulations increases their density values.
Based on this analysis, it was observed that the foam containing 17.5% of lignin, 32.5% of crude glycerol and 50% of castor
oil (D formulation) presented the best dimensional and thermal stability as well as the best cell homogeneity. For this reason,
some variations on this formulation were performed for evaluate the effect of castor oil content, molar ratio NCO/OH and
type of blowing agent.
58 C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61

120

110

100

Density (kg.m -3 )
90

80

70

60

50
10 15 20 25 30 35 40
Lignin content (%)

Fig. 5. Variation of foams density with the lignin content.

Fig. 6. Images of LGCO foams with different contents of lignin: (a) A, (b) B, (c) C, (d) D, (e) E, (f) F, (g) G and (h) H (magnification of 50).

Table 3 shows the apparent densities of the foams with LGCO polyol; it was observed that increasing the castor oil content
in the foam formulation led to an increase in the apparent density, suggesting that the polymeric chains were more packaged
with a lower free volume and smaller cells, as also confirmed by the increased stiffness of these materials. An increase of the
density was also observed with the addition of excess isocyanate (molar ratios NCO/OH = 1.1–2.0), which can be explained
by the increase of reactions between the isocyanate with the polyol and consequently, an increase of crosslinking and a
higher content of rigid segments, as expected.
The examination of the effect of the blowing agent on the density of the foams showed that the physical blowing agents
(cyclopentane and n-pentane) yielded foams with higher densities than the samples synthesized with a chemical blowing
agent (water); this result can be explained by the rapid volatilization during the highly exothermic foam growth steps, caus-
ing cells with less gas in comparison with the CO2 produced by the reaction of water with isocyanate. Similar results have
been reported in the literature [20,21].
The mechanical properties of the foams are proportional to the density, due to the increase in the cell wall thickness and/
or a decrease in the cell size; according to Euler-Bernoulli beam theory, this gives rise to an increase in the bending moment
of the cell wall [22]. The compressive strength results showed an increase with the greater castor oil amount in the formu-
lations, indicating that there is a greater content of crosslinking and rigid segments with higher castor oil content. The
increase of isocyanate content also increases the compressive strength of the foams, except for the NCO/OH = 1.5 foam,
which shows a decrease in this mechanical property, suggesting a dimensional destabilization of this structure.
The foams produced using physical blowing agents (cyclopentane and n-pentane) showed lower compressive strength in
comparison with those synthesized with the chemical blowing agent (water), unlike the results reported in the literature.
This behavior can be explained by the low accessibility of the hydroxyl groups from lignin and short chain of glycerol used
C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61 59

Table 3
Apparent densities and mechanical properties of foams produced with LGCO polyol (D formulation).

Sample Density (kg m3) Compressive strength (kPa) Compressive modulus (kPa)
25% CO 52.7 ± 1.7 14.4 ± 5.1 9.8 ± 3.5
50% CO 54.9 ± 4.0 35.0 ± 4.3 10.2 ± 2.5
75% CO 108.7 ± 2.2 48.0 ± 1.3 32.1 ± 13.6
NCO/OH = 1.1 54.9 ± 4.0 35.0 ± 4.3 10.2 ± 2.5
NCO/OH = 1.5 77.1 ± 9.8 24.4 ± 5.0 26.3 ± 4.7
NCO/OH = 2.0 130.2 ± 7.0 64.9 ± 8.9 56.7 ± 15.3
H2O 54.9 ± 4.0 35.0 ± 4.3 10.2 ± 2.5
NP 71.1 ± 0.5 28.1 ± 14.3 34.4 ± 17.0
CP 81.3 ± 13,9 24.7 ± 4.0 39.6 ± 9.8
H2O/CP 71.4 ± 6.9 22.6 ± 10.1 32.1 ± 13.6

NP = n-pentane and CP = cyclopentane.

in our formulation, which can not react completely with the isocyanate. As consequence, the unreacted isocyanate, evi-
denced in FTIR spectrum (Fig. 3), is available to form urea linkages by the reaction of the isocyanate and water. These urea
groups exhibits strong cohesion and provide more stability and toughness for foams using CO2 as the blowing agent, differ-
ently from the literature data [20,21].
The green foams obtained in this study, which showed density values of 50–120 kg m3 and compressive strength of 10–
65 kPa, which are similar to those conventional commercial foams that present densities in the 15–130 kg m3 range and
their hardness values are in the 1.3–31 kPa range. Based on these results the foams produced from lignin, crude glycerol
and castor oil can potentially be used on industrial scale in the construction and packaging industry [23].
The foam structure is confirmed by the morphological analysis performed by SEM as shown in Figs. 7–10. It can be
observed that increasing lignin content decreases cell sizes (Fig. 7), in accordance with the apparent density and optical

Fig. 7. SEM micrographs of LGCO foams varying lignin content. (a) A, (b) B, (c) C, (d) D, (e) E, (f) F, (g) G and (h) H.

Fig. 8. SEM micrographs of LGCO foams varying castor oil content in D formulation. (a) 25, (b) 50 and (c) 75% of castor oil.
60 C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61

Fig. 9. SEM micrographs of LGCO foams varying NCO/OH molar ratio in D formulation. (a) NCO/OH = 1.1, (b) NCO/OH = 1.5 and (c) NCO/OH = 2.0.

Fig. 10. SEM micrographs of LGCO foams varying the blowing agent in D formulation. (a) Water, (b) n-pentane, (c) cyclopentane and (d) a mixture of water
and cyclopentane.

microscopy results. Foams with a lower castor oil content (Fig. 8) showed thick and large cells. An increasing NCO/OH molar
ratio was observed to result in a greater heterogeneity of the cell size and shape, as shown in Fig. 9. The SEM micrographs of
the samples obtained with different blowing agents (Fig. 10) showed that physical blowing agents produced foams with
smaller cell sizes than the foams synthesized using water as the blowing agent, as suggested by the density data. The
SEM micrographs also confirmed that the heterogeneity of these formulations led to a higher standard deviation.

4. Conclusions

The properties of the foams produced from polyols obtained by the physical mixture of lignin, castor oil and crude glyc-
erol may be adjusted according to the amount of reactants in the formulations. The synthesis process is simple and inexpen-
sive and does not require biomass pre-treatment. The produced foams with higher lignin contents showed a decrease in
thermal stability and an increase in the density and compressive strength. Compared to the other tested foams, the foam
with 17.5% lignin in its formulation showed a higher thermal and dimensional stability as well as a greater cell homogeneity.
C.S. Carriço et al. / European Polymer Journal 85 (2016) 53–61 61

The effects observed with the increase in the amount of castor oil and the molar ratio NCO/OH showed that these reagents
promote an increase in the amount of urethane crosslinking, providing greater stiffness to the foams. The use of volatile
blowing agents produce smaller cells with higher density in comparison with formulations using water as the blowing agent.
However, the use of water produced more rigid foams due the increase of urea linkages as a by-product of CO2 formation.
The use of these polyols from industrial residual sources in addition to castor oil and water as the blowing agent provides a
high sustainability for these materials and can contribute to the economic feasibility of the important cellulosic ethanol and
biodiesel commodities. Additionally, these environmentally friendly foams showed good properties and a potential for use
on the industrial scale.

Acknowledgements

The authors would like to acknowledge the financial support of CNPq – National Council for Scientific and Technological
Development – Brazil.

References

[1] P. Cinelli, I. Anguillesi, A. Lazzeri, Green synthesis of flexible polyurethane foams from liquefied lignin, Eur. Polym. J. 49 (6) (2013) 1174–1184.
[2] J. Bernardini, P. Cinelli, I. Anguillesi, M.B. Coltelli, A. Lazzeri, Flexible polyurethane foams green production employing lignin or oxypropylated lignin,
Eur. Polym. J. 64 (2015) 147–156.
[3] H.R. Zhang, H. Pang, L. Zhang, X. Chen, B. Liao, Biodegradability of polyurethane foam from liquefied wood based polyols, J. Polym. Environ. 21 (2013)
329–334.
[4] R. Gu, M.M. Sain, Effects of wood fiber and microclay on the performance of soy based polyurethane foams, J. Polym. Environ. 21 (2013) 30–38.
[5] R.C.S. Araújo, V.M.D. Pasa, B.N. Melo, Effects of biopitch on the properties of flexible polyurethane, Eur. Polym. J. 41 (6) (2005) 1420–1428.
[6] X.L. Luo, S.J. Hu, X. Zhang, Y.B. Li, Thermochemical conversion of crude glycerol to biopolyols for the production of polyurethane foams, Bioresour.
Technol. 139 (2013) 323–329.
[7] Y.C. Tu, P. Kiatsimkul, G. Suppes, F.H. Hsieh, Physical properties of water-blown rigid polyurethane foams from vegetable oil-based polyols, J. Appl.
Polym. Sci. 105 (2007) 453–459.
[8] N. Mahmood, Z. Yuan, J. Schmidt, C. Xu, Preparation of bio-based rigid polyurethane foam using hydrolytically depolymerized Kraft lignin via direct
replacement or oxypropylation, Eur. Polym. J. 68 (2015) 1–9.
[9] H. Hatakeyama, R. Kosugi, T. Hatakeyama, Thermal properties of lignin- and molasses-based polyurethane foams, J. Therm. Anal. Calorim. 92 (2008)
419–424.
[10] T. Hatakeyama, Y. Matsumoto, Y. Asano, H. Hatakeyama, Glass transition of rigid polyurethane foams derived from sodium lignosulfonate mixed with
diethylene, triethylene and polyethylene glycols, Thermochim. Acta 416 (2014) 29–33.
[11] Y. Li, A.J. Ragauskas, Kraft lignin-based rigid polyurethane foam, J. Wood Chem. Technol. 32 (2012) 210–224.
[12] Y. Li, A.J. Ragauskas, Ethanol organosolv lignin-based rigid polyurethane foam reinforced with cellulose nanowhiskers, RSC Adv. 2 (2012) 3347–3351.
[13] N. Karak, Vegetable Oil-Based Polymers: Properties, Processing and Applications, Woodhead Publishing, 2012.
[14] D.L. Pavia, G.M. Lampman, G.S. Kriz, Introduction to Spectroscopy, third ed., Brooks Cole, 2000.
[15] U. Stirna, B. Lazdina, D. Vilsone, M.J. Lopez, M.C. Vargas-Garcia, F. Suárez-Estrella, J. Moreno, Structure and properties of the polyurethane and
polyurethane foam synthesized from castor oil polyols, J. Cell. Plast. 48 (2012) 476.
[16] F. Chen, Z. Lu, Liquefaction of wheat straw and preparation of rigid polyurethane foam from the liquefaction products, J. Appl. Polym. Sci. 111 (1)
(2009) 508–516.
[17] M.A. Corcuera, L. Rueda, B. Fernandez d’Arlas, A. Arbelaiz, C. Marieta, I. Mondragon, A. Eceiza, Microstructure and properties of polyurethanes derived
from castor oil, Polym. Degrad. Stabil. 95 (11) (2010) 2175–2184.
[18] X. Luo, A. Mohanty, M. Misra, Lignin as a reactive reinforcing filler for water-blown rigid biofoam composites from soy oil-based polyurethane, Ind.
Crop. Prod. 47 (2013) 13–19.
[19] W. Vilar, Química e Tecnologia dos Poliuretanos, Vilar Consultoria, São Paulo, 2002.
[20] H. Lim, E.Y. Kim, B.K. Kim, Polyurethane foams blown with various types of environmentally friendly blowing agents, Plast., Rubber Compos. 39 (2010)
364–369.
[21] M. Modesti, V. Adriani, F. Simioni, Chemical and physical blowing agents in structural polyurethane foams: simulation and characterization, Polym. Sci.
Eng. 40 (2000) 2046–2057.
[22] H. Jeong, J. Park, S. Kim, J. Lee, N. Ahn, Fiber Polym. 14 (2013) 1301.
[23] Eurofoam, Technical Purpose Foams, <http://www.eurofoam.hu/habtipusok-en/Comfort-purpose-foams/> (accessed June 20, 2006).

Potrebbero piacerti anche