You are on page 1of 10

European Polymer Journal 64 (2015) 147–156

Contents lists available at ScienceDirect

European Polymer Journal


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

Flexible polyurethane foams green production employing lignin


or oxypropylated lignin
Jacopo Bernardini a, Patrizia Cinelli b, Irene Anguillesi a, Maria-Beatrice Coltelli a,
Andrea Lazzeri a,⇑
a
Department of Civil and Industrial Engineering, University of Pisa, Largo L. Lazzarino 1, 56126 Pisa, Italy
b
National Interuniversity Consortium of Materials Science and Technology (INSTM), Via G. Giusti, 9, 50121 Firenze, Italy

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

Article history: An innovative and green chemistry synthetic approach was developed in order to employ
Received 16 July 2014 lignin for the production of flexible polyurethane foams. Soda lignin and oxypropylated
Received in revised form 24 November 2014 soda lignin were tested and compared. Glycerol and PEG 400 were used as polyol fractions
Accepted 26 November 2014
for lignin liquefaction by microwave irradiation, which represents a novel green processing
Available online 3 December 2014
technique. The samples were produced with the ‘‘one-shot’’ technique, using two types of
chain extenders in combination with liquefied lignin: castor oil and polypropylene glycol
Keywords:
triol. Water was used as a single natural blowing agent, and polymeric diphenylmethane
Polyurethane
Flexible foam
diisocyanate (PMDI) was employed as the isocyanate fraction. The work was carried out
Lignin keeping the NCO/OH less than one hundred, thus enhancing the flexibility due to a lower
Oxypropylation crosslinking degree; all the foams were produced in free and controlled expansion. Two of
Microwave the most efficient chain extenders were individuated thus introducing flexible chains into
Renewable sources the macromolecular structure that can reduce the glass transition temperature of the
materials and therefore generate foams with higher flexibility. The properties of the
produced foams were compatible with the technical requirements for applications in
packaging and for the production of the interior part of car seats.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction in many industries such as footwear, construction,


electronics and many others. Actually, common polyols
Polyurethanes (PUs) are an important class of polymers employed to produce polyurethane foams are petrochemi-
synthesized by means of a polyaddition reaction between cal in origin and, due to the escalation of crude oil prices,
polyols, such as polyhydric alcohols and polyisocyanate the cost of polyurethane reactants steadily increased, thus
that form urethanic linkages. These polymers are very ver- sparking interest in the quest for renewable sources.
satile because they present structures ranging from rigid to Flexible PU foams are blocks of copolymers whose elas-
flexible and are commonly used in various forms such as tic properties depend on the phase separation between
foams, elastomers, adhesives and many others, therefore hard and soft domains. Hard blocks are rigid structures
they have a wide range of applications, i.e. automotive that are physically cross-linked and give the polymer its
seating, furniture, and packaging and medical devices. firmness; soft blocks are stretchable chains that give the
Polyurethanes are also widely used as thermal insulators polymer its elasticity. Therefore, polyurethane foams can
be customized by adapting the composition and the ratio
of these blocks.
⇑ Corresponding author. Tel.: +39 050 2217807; fax: +39 050 2217866. Polyurethanes can be synthesized by using renewable
E-mail address: andrea.lazzeri@unipi.it (A. Lazzeri). sources as polyols, such as vegetable oils, by replacing

http://dx.doi.org/10.1016/j.eurpolymj.2014.11.039
0014-3057/Ó 2014 Elsevier Ltd. All rights reserved.
148 J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156

petrochemical polyols partially or totally. Since the 1960s,


a wide range of vegetable oils have been considered for the
preparation of polyurethane; the most important ones are
highly unsaturated, such as sunflower, palm, rapeseed and
mainly castor and soybean oils [1–3].
Several attempts were made in order to use lignin as a
novel and appropriate renewable feedstock, thus supersed-
ing petrochemical phenols and overcoming the idea that it
is a waste and a low value by-product of wood pulping
[4,5]. The use of lignin as such, or after appropriate treat- Fig. 1. Schematic pathway for the oxypropylation process of lignin.
ments, was extensively investigated to prepare polyesters
and polyurethanes exploiting its phenolic and aliphatic
hydroxyl groups [6,7]. process because it does not need further steps involving
Lignocellulosic biomass is composed mainly of three solvents and other by-product removal, nor is there the
chemicals: cellulose, hemicellulose and lignin, and its com- need for separation or purification steps in order to recover
position varies with the source and type of biomass. Lignin the final polyol mixture [19]. In this sense, the purpose of
is an aromatic complex polymer that is highly branched oxypropylation is to increase the lignin functionality for
and based on subunits of alcohol 4-hydroxycinnamyl (p- synthesizing polyurethane foams [20].
coumaryl alcohol) having methoxylated derivatives in 3- Many studies of oxypropylation were also carried out
(coniferyl alcohol) and 3,5-positions (sinapyl alcohol) [8]. on renewable sources in order to produce low-cost natural
Lignin has a molecular weight of about 3000–7000 Da, an polyols. Several biomasses such as cork [21], chitin and
OH number of about 1000–1500 mg KOH/g, and a methoxy chitosan [19], and sugar beet pulp [22] were considered.
group content of about 13–14%. Currently, commercial lignins are not used in the indus-
There are many different types of lignin, whose nature trial production of polyurethane foams due to the high vis-
depends on the different preliminary treatments that they cosity of lignin-based polyols combined with a high
have undergone and on the different types of wood [9]. Lig- temperature and very long reaction times; the extended
nin is mainly isolated from wood pulping and papermaking structure of the lignin macromolecules must be frag-
operations, but only a small amount (1–2%) is recovered mented before the employment in polyurethanes prepara-
and employed as feedstock to produce other chemicals tion. This process was defined liquefaction as the solubility
[10,11]. These pulping treatments provide fragmented lig- of lignin was modified thanks to the occurring of fragmen-
nin, thus enhancing its reactivity due to lower molecular tation. Usually the reported methods consist in the use of
weights and higher solubility in organic solvents. Soda lig- proper solvent and heating. Microwave treatment can be
nins are sulphur-free and for this reason are considered as an effective alternative because it is able to heat the core
being closer to natural lignins; these lignins are mainly of the system. The microwaves were carefully studied to
obtained from non-woody plants. degrade hazardous compounds [23] and subsequently this
Many studies were carried out over recent years in treatment was tried on lignocellulosic biomasses [24–26]
order to use various lignins in polymeric synthesis. In this in order to obtain liquid polyols with shorter reaction
sense, Hatakeyama’s group dissolved several lignins, such times. This approach was tried on commercial lignin and
as Kraft, organosolv and sulphonate, into oligoether diols the final mixture contained a higher amount of phenolic
to obtain the polyolic mixture. By using the macrodiol as OH groups and a lower amount of methoxy groups, thus
both the solvent and a comonomer, they improved the improving the system reactivity [27].
reactivity of lignin OH groups [12], and the introduction Microwaves are able to penetrate and heat the present
of flexible oligoether as spacers within the polymer chain reaction medium, thereby reducing reaction times from
reduced the network stiffness of the isocyanate. Also hours to minutes; furthermore, when heating with micro-
semi-rigid polyurethane foams, developed for housing waves there is a faster increase in temperature and a lower
insulation, were prepared by means of using lignin polyolic thermal degradation of reagents than traditional heating
molasses, by controlling the apparent density, and by systems, thus being compatible with a green chemistry
changing the mixing ratio between lignin and polyethylene approach.
glycol [13]. In previous research [28], the authors produced poly-
Oxypropylation of biomass was widely studied and is urethane soft foams from kraft lignin, achieving a promis-
presently accepted as a method for improving the func- ing quality of foams in terms of density and mechanical
tionality of lignin [14–16]. A schematic pathway is properties. It resulted interesting to investigate lignin pro-
reported in Fig. 1. duced using different processes and modified lignin in
Glasser and his co-workers have modified lignins by order to achieve optimized properties. Thus the present
using different conditions of high temperatures and pres- work was focused on the production of flexible polyure-
sures [17], allowing for further investigations on it in thane foams from commercial soda lignin and commercial
recent years [18]. oxypropylated soda lignin by adopting a green synthetic
The oxypropylation process allows to free the phenolic pathway. Two chain extenders were used to produce flex-
groups of lignin from steric hindrances and to obtain a lig- ible foams in combination with liquefied lignin: polypro-
nin-based liquid polyol with lower glass transition temper- pylene glycol triol (PPG triol) and castor oil; polymeric
ature compared to the starting powder. It is a green diphenylmethane diisocyanate (PMDI) was employed as
J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156 149

isocyanate fraction. Cellular structure as well as physical and DABCO BL11 (bisdimethylamminoethylether 70% by
and thermo-mechanical properties were characterized in weight, dipropylene glycol 30% by weight, blowing agent).
order to compare the properties of the foams produced The NE catalysts used were DABCO NE1070 (3-dimethyl-
with the requirements set for commercial standard poly- aminopropyl urea, gelling agent) and DABCO NE300
urethane foams. (This research activity was performed in (N-[2-[2(dimethylamino) ethoxy]ethyl]-N-methyl-1,3-pro-
the EC Project FORBIOPLAST Grant agreement No. panediamine, blowing agent). These catalysts can reduce
212239, which focused on the use and valorisation of for- emissions because the molecules that act as catalysts can
est resources that include lignin as a by-product of the chemically bind themselves to the polyurethane matrix.
wood industries and bioethanol production.) DABCO DC2525, a silicone surfactant made from 70% by
weight of polysiloxane, was used to ensure the uniformity
2. Materials and method of structure and enhance cell opening.

The lignin used was Protobind™ 1000, produced as dry 2.1. Liquefaction of lignin
powder by ALM India Pvt. Ltd. using a patented GreenValue
SA technology; this lignin comes from soda pulping of Lignin samples were previously dried for 24 h in an
non-woody biomass. Protobind™ 1000 has high degree of oven with air circulation at 80 °C. Then lignin was mixed
purity (about 98%) and is free of hemicellulose. Oxypropy- with liquefying polyols (glycerol and PEG 400) at a
lation reaction was performed in aqueous alkali at room predetermined weight ratio, placed in a Teflon vessel,
temperature at moderate pH (only the phenolic OH groups and subjected to liquefaction treatment in a multimode
are ionized) (Fig. 1). microwave oven CEM MDS 2100. A power of about
Polyethylene glycol 400 and glycerol from Aldrich were 200 W was fixed to reach 135 °C in 2 min.
used as solvents to liquefy lignin. Polypropylene glycol
triol (PPG triol) and Castor oil [44] (CO) (Fig. 2), were used 2.2. Foam preparation
as chain extenders because of their flexible molecular
structure, thus suitable for flexible foams. They were both The liquefied mixture was mixed with the other polyol-
purchased from Sigma Chemicals and used as received. ic components and with PMDI; then the resulting mixture
PPG triol from Sigma Chemicals has a Molecular weight was deposited into a mould (85  85  55 mm) and
of 4800, functionality of 2.6, and viscosity of 850 mPa s at allowed to rise in controlled expansion by closing the
25 °C. ISO 116/1 was used as an isocyanate, a polymeric mould thus forcing the foam to take on its shape. Foams
diphenylmethane diisocyanate (PMDI) with 25.7% part by were also produced in free expansion in order to compare
weight of NCO content, provided by BASF Poliuretani Italia their mechanical properties depending on the type of
Spa. The isocyanate was added in excess to the OH groups expansion. Compositions used to synthesize the foams
on the polyols in order to react with distilled water, thus are reported in Table 1.
forming carbon dioxide, which acted as the foam-blowing All the foam samples were allowed to cure at ambient
agent. The catalysts used for the production of foams can conditions for a minimum of 2 days. Cream times and rise
be divided into gelling and blowing catalysts. Both of these times are reported in Table 2.
must be kept at a proper ratio in order to obtain the desired
products. They were kindly provided by Air Products and 2.3. Methods
belong to two different series: classic catalysts and new
generation catalysts (NE), with a low emission of amine. The hydroxyl number in the liquefied mixture was
The classic catalysts consisted of non-reactive amines that determined by following the ASTM D2849-69 norm: 1 g
can be released from the foams once they are produced. mixture and 25 ml of phthalate reagent were heated for
They were DABCO 33LV (triethylenediamine 67% by 20 min at 110 °C. Then, 50 ml of pure 1.4-dioxane and
weight, dipropylene glycol 33% by weight, gelling agent) 25 ml of distilled water was added and the resulting

Fig. 2. Structure of chain extenders.


150 J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156

Table 1
Formulations (grams) used for synthesis of PUs foams with polypropylene glycol triol (PPG) or castor oil (CO) as chain extenders.

PPG-A PPG-B PPG-C PPG-D PPG-E PPG-F PPG-G CO-A CO-B


Lignin/glycerol/PEG400 1/0.4/2 1/0.4/2 1/0.2/1 1/0.2/1 1/0.2/1 1/0.2/1 1/0.2/1 1/0.4/2 1/0.4/2
Liquefied lignin 14.2 11.4 10.0 10.0 10.0 10.0 10.0 17.0 17.0
PPG triol 20.8 16.6 14.0 10.0 7.6 14.0 10.0 – –
Castor oil – – – – – – – 27.5 27.5
Water 0.6 0.7 0.66 0.66 0.66 0.65 0.65 0.81 0.5
DABCO33LV (gelling) 0.53 – 0.36 0.34 0.31 – – 0.67 –
DABCOBL11 (blowing) 0.53 – 0.35 0.31 0.31 – – 0.67 –
DABCONE1070 (gelling) – 0.9 – – – 0.63 0.6 – 1.24
DABCONE300 (blowing) – 0.35 – – – 0.2 0.22 – 0.25
Surfactant 0.35 0.56 0.26 0.21 0.17 0.21 0.2 0.45 0.66
Isocyanate 25.0 23.0 20.0 16.8 15.0 20.0 16.9 36.8 32.0
Lignin (% by weight) 6.7 6.7 10 11.9 13.3 10.0 11.8 6.0 6.3

Table 2
Cream time and rise time for foams prepared starting from lignin Protobind (P) and oxypropylated lignin (HP).

Sample Cream time (s) Rise time (s) Sample Cream time (s) Rise time (s)
P-PPG-A 10 15 HP-PPG-A 10 10
P-PPG-B 20 40 HP-PPG-B 20 50
P-PPG-C 30 80 HP-PPG-C 25 40
P-PPG-D 50 170 HP-PPG-D 30 50
P-PPG-E 40 180 HP-PPG-E 30 60
P-PPG-F 40 150 HP-PPG-F 40 110
P-PPG-G 50 180 HP-PPG-G 50 150
P-CO-A 30 45 HP-CO-A 25 40
P-CO-B 70 190 HP-CO-B 60 120

mixture was titrated with a 1 M sodium hydroxide solu- and the experiments were performed under a nitrogen
tion up to the equivalence point. The phthalate reagent gas atmosphere. The samples were ground into a fine pow-
consists of a mixture of 150 g phthalic anhydride, 24.2 g der prior to measurement and were heated from room
imidazole and 1000 g dioxane. The hydroxyl number in temperature to 900 °C at a rate of 10 °C/min.
mg KOH/g of the sample was calculated by the following For dynamic mechanical analysis (DMTA), foams were
Eq. (1): cut into 20  20  20 mm cubes and tested under com-
pression mode between two serrated parallel plates with
ðB  AÞ  N  56:1
Hydroxyl number ¼ ð1Þ a 25 mm diameter (Gabo Eplexor 100 N). Storage modulus
W
(E0 ) and tan d were recorded at a frequency of 1 Hz over the
where A is the volume in ml of the sodium hydroxide solu- temperature range from 150 to 120 °C. The temperature
tion required for titration of liquefied lignin sample, B is ramp rate was 2 °C/min.
the volume of blank solution (ml), N is the normality of The density and 50% compression force deflection
the sodium hydroxide solution, and W is the weight of liq- (CFDV) of flexible polyurethane foams were measured fol-
uefied lignin (g). lowing the ASTM D 3574-05 norm, which describes the
The isocyanate/hydroxyl molar ratio required for the standard test methods for flexible cellular materials
reaction is calculated using the following Eq. (2): including slab, bonded and moulded urethane foams. The
M NCO W NCO sample size was 30 mm  30 mm  30 mm and the result
NCO=OH ¼ h P3 2 i  100 ð2Þ of ratio between weight (M) and volume (V) of the speci-
M OH W OH þ i¼1 M Adi W Adi þ 18 W H2 O
mens was the apparent density, which should be in kg/
where M NCO is the number of isocyanate groups in one m3. The compression force deflection test was performed
gram of isocyanate, W NCO is the weight of isocyanate (g), using Instron 1185 equipment, according to the ASTM D
M OH is the number of hydroxyl groups contained in one 3574-05 norm: pre-flex the specimen twice up to 75%
gram of polyol (mixture of liquefied lignin and chain and 80% of its original thickness at 240 mm/min; rest for
extender, index 1), W OH is the weight of polyols (g), M Adi 6 min and measure sample thickness; bring the compres-
is the number of hydroxyl groups in one gram of additives sion plates into contact with the specimen and apply a con-
(blowing catalyst, index 1, gelling catalyst, index 2, and tact load of 140 Pa; compress the specimen up to 50% of its
surfactant, index 3), W Adi is the weight of additives (g) thickness at 50 mm/min and observe the final load after
and W H2 O is the weight of water. 60 s. At least three replicates were tested for each sample.
Thermogravimetric analysis (TGA) measurements were Compression strength was evaluated as the ratio of the
carried out following the ASTM standard procedure D final load at the end of the procedure (N) and the cross
3850-94. A Rheometric Scientific instrument was used, section area of the specimen (mm2).
J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156 151

The FT-IR characterization was performed with a to entirely expand. All data are reported in Table 2. These
Thermo Scientific NicoletTM 380; the foams were previ- two parameters are very important to control the kinetic
ously dried at 105 °C for 4 h. The samples for FT-IR mea- process because the reaction between water and isocya-
surement were taken from the centre of the foam buns, nate produces carbamic acid, which is unstable and
ground and dried at 105 °C for 2 h before analysis to ensure decomposes to carbon dioxide and primary amine. Carbon
that no water was adsorbed inside. dioxide then diffuses in the mixture thus expanding the
A Scanning Electron Microscope (Jeol JSM5600LV) was foam structure. Clearly, the nucleation sites in the reacting
used to study the microstructure of the foams and gold mixture are checked by controlling the amount of carbon
sputtering was applied using an Edwards Sputter Coater dioxide released. Foams produced using classic catalysts
prior to investigation. showed reaction times slightly lower than samples pro-
duced using NE catalysts; this was probably due to the
higher reactivity of classic catalysts. Nevertheless, the
3. Results and discussion properties of the produced foams showed no relevant dif-
ferences regardless of the kind of catalyst employed.
The synthesis of polyurethane foams requires polyolic
reagents with low viscosity and high functionality, there-
fore solid powder lignin has been liquefied with the aid 3.1. FT-IR spectroscopy
of suitable reagents, such as glycerol and PEG 400, as
reported in literature for Kraft lignin [28]. Liquefied lignin Polyurethane foams were qualitatively characterized by
is rich in free hydroxyl group, but it has high viscosity and the FT-IR technique in order to assess the presence of unre-
hydroxyl values. For these reasons, it was not suitable to acted isocyanate within the final material and to appraise
produce flexible foams. Accordingly, the addition of a chain the formation of typical urethanic bonds. Fig. 3 shows
extender was necessary in order to decrease the viscosity the FT-IR spectra of a representative sample of each type
of the polyolic mixture and at the same time to increase of produced foam.
the flexibility of the final material. Polypropylene glycol The above spectra qualitatively confirmed the presence
triol (PPG triol) and castor oil (CO) were used as chain of urethane linkages because they presented approxi-
extenders in this work. mately the same trend regardless of the different composi-
The liquefaction process was carried out by microwave tions, type of lignin and type of chain extender. It was
irradiation and was optimized by increasing the amount of possible to find the typical urethanic absorption bands:
solvents used to reduce the mixture viscosity. the ANH vibrational stretching (3500–3200 cm1), the
One of the most important parameters in polyurethane characteristic vibration of AC@O groups (1730–
production is the hydroxyl value, which is generally deter- 1640 cm1) [31] and, eventually, the very weak peak of
mined following the ASTM D 2849-69 norm; the procedure unreacted ANCO group at about 2270 cm1 [32].
is based on an esterification between polyol and an excess Particularly, two absorption bands can be observed, one
of acetic or phthalic anhydride, the hydrolysis of unreacted centred at 1710 cm1 and the other one at 1640 cm1. The
anhydride which forms dicarboxylic acid. The resulting first peak at about 1710 cm1 was caused by the presence
solution is then titrated with sodium hydroxide reducing of free urea, meanwhile the second one, at about
the reaction time by using imidazole as an esterification 1640 cm1, was characteristic of bidentate urea that inter-
catalyst [29]. acted through hydrogen bonds with the surrounding mol-
The NCO/OH was fixed once the OH number was deter- ecules. These two signals were indicative of a phase
mined, thus setting the weight of isocyanate to be added in
each formulation. A NCO/OH was fixed at less than 100 in
order to maintain a low cross-linking degree in the final
foam, thus giving higher flexibility to the material [30].
The polyurethane foams were produced by adopting the
one-shot technique, which provided a very efficient mixing
among all the materials involved (polyols, chain extender,
silicon emulsifier, blowing agent, catalysts, isocyanate) in a
single step and in a short time. It was possible to use only
two components: one was the polyolic containing all
material except for the isocyanate, which was the second
component. The resulting polyurethane was therefore a
consequence of the very efficient contact between the
two components.
The kinetics of foam formation was monitored by
means of measuring cream and rise times. Cream time is
defined as the time required by the polyurethane mixture
to change from a clear colour to a creamy one (although
Fig. 3. Comparison among FT-IR spectra on the polyurethane foams
liquefied lignin has a very dark colour, it was possible to obtained starting from Protobind 1000 lignin (P) or oxypropylated
recognize a change in colour when cream time was Protobind 1000 lignin (HP) and by using polypropylene glycol triol
reached), while the rise one is the time it takes the foam (PPG) or castor oil (CO) as a chain extender.
152 J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156

separation within the material [33]. The hydrogen-bonded thermogravimetric curves and the derivative curves of
urea, including both monodentate and bidentate, were an the studied foams.
indication of hard domain inside the foam structure. The Foams produced by using PPG triol as a chain extender
bands in the region below 1600 cm1 represented the fin- presented the beginning of irreversible degradation at
gerprints of the polyurethanes. around 240 °C and a first step at 270 °C, which were typi-
The above spectra did not present significant differ- cally characteristic of urethane bond; two degradation
ences among the several type of foams, thus allowing stat- steps can also be observed at 389 °C and at 480 °C, which
ing that the preliminary step of oxypropylation and the were correlated to the presence of two phases within the
type of chain extender did not influence the chemical material [28]. Indeed the first is related to soft segments
behaviour of the foams. Comparing this work with the pre- (PPGtriol chains), and the second is related to the degrada-
vious one based on Kraft lignin [28], the intensity of unre- tion of hard segments. Also the foams produced by using
acted isocyanate band was lower, probably due to a lower castor oil as a chain extender started thermal degradation
steric hindrance and to a simpler structure of the soda lig- at about 240 °C, thus indicating that the onset of thermal
nin with respect to the Kraft one. Despite the fixed NCO/OH degradation did not change regardless of the employed
was less than 100, a small amount of unreacted isocyanate chain extender. However the derivative trend is quite dif-
groups was found, which was caused by the steric hin- ferent comparing the systems obtained by using the two
drance of the lignin structure that prevented the complete chain extenders. The derivative trend of castor oil contain-
reaction between the isocyanate and hydroxyl groups. ing foams [35], as reported also by Narine [1], shows three
main typical peaks, whereas the trend observed for PPG
triol is a simpler trend, with a main peak at 389 °C. This dif-
3.2. Thermogravimetric analysis
ference can be ascribed to the different structure of the two
chain extenders. In fact in PPG triol the thermal degrada-
Thermogravimetric analysis is a very useful technique
tion occurs by the scission of CAO bonds of the ether
to analyze the degradation steps of polyurethane materials
chains, whereas the thermal degradation of CO is more
based on renewable sources [34]. Fig. 4 displays the
complex, involving both CAO and CAC linkages. Moreover
in the CO molecules the presence of double bonds make
the structure weaker with respect to thermal degradation
[43], thus resulting in a consistent mass loss at 323 °C.
The degradation of pure not modified lignin (not shown
data) occurs in the samples in two main steps at 285 ad
410 °C. Hence, as the content of lignin is, in the different
formulations, below 10% by weight, its mass loss cannot
be easily distinguished in the TGA trends. Foams produced
with PPG triol presented a higher amount of carbon resi-
dues, about 15% by weight, than those produced with cas-
tor oil (less than 10%); this is probably due to the different
cellular structure between the two types of foams. In this
sense, PPG triol-based foams presented more regular cellu-
lar structure, hindering the regular heat transfer and there-
fore resulting in a less efficient pyrolysis; conversely,
castor oil-based foams presented a less regular cellular
structure, and this allowed for a better pyrolysis process.
The results confirm that, as observed in FT-IR analysis,
the oxypropylation step did not change the thermo-chem-
ical behaviour of polyurethane foams produced from
lignin.

3.3. Dynamic mechanical thermal analysis

Fig. 5 shows the storage modulus profile (E0 ) and tan d,


as a function of temperature, of foams produced starting
from Protobind 1000 or oxypropylated Protobind 1000 by
using PPG triol or castor oil as a chain extender. The Tg
value of each sample was determined by the tan d peaks.
Two different transitions are shown in the modulus
profiles: the first is related to the flexible phase in the
foams, and the second to the hard one. Clearly both
domains have different glass transition temperatures and
Fig. 4. TG (top) and DTG (bottom) curves of the produced polyurethane
mechanical stiffness: the soft phases (or polyol-rich
foams. All thermograms were acquired at constant heating rate domains), which usually show low Tg ranging between
(10 °C min1) under air atmosphere. 70 °C and 50 °C, give viscoelastic properties to the
J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156 153

Fig. 5. DMTA results of the produced polyurethane foams starting from


lignin (P) or oxypropylated (HP) lignin by using PPG or castor oil (CO) as
Fig. 6. Compression force deflection value (CFDV) for samples produced
chain extenders.
with PPG triol as a chain extender.

foams thus allowing a higher energy absorption. Con- distribution of crosslinking density and, mainly, a lower
versely, the hard phases (or polyurea-rich domains) pres- structural homogeneity.
ent a glass transition temperature over 100 °C and impart A wide peak was observed at about 100 °C for the
to the polyurethane material a higher modulus and greater foams produced starting from oxypropylated lignin and
thermal stability [28]. The first Tg is related to the soft seg- castor oil, which was likely due to the motion of dangling
ments and similar values mean a similar phase separation chains [38]. Therefore, this transition was assigned to a
between hard and soft domains [36]. Fig. 5 shows no differ- segmental relaxation of a second soft phase, which was
ences between foams produced by using the same chain caused by the compatibility between castor oil and oxy-
extender, but it shows significant differences when the propylated lignin [39].
chain extender is changed. PPG-based foams presented a
plateau at about 0 °C, which is due to the hard segments, 3.4. Mechanical properties and morphology
precisely to the physical crosslinks inside the network;
their breaking caused the end of the plateau, thus decreas- Table 3 reports the apparent density of all the produced
ing the storage modulus [37]. foams.
Tan d curves, as well as modulus profiles, showed no dif- Foams of series 1 were produced in free expansion,
ferences among foams based on the same chain extender; while those of series 2 were made in controlled expansion.
samples produced by employing PPG triol presented a Tg at Table 3 shows that the densities of PPG-based samples
about 50 °C but a different intensity of the related peak. were much lower than those based on castor oil as a chain
In this sense, foams made starting from oxypropylated lig- extender, which has probably a less regular structure. A
nin showed a larger peak, which means a greater amount similar difference was observed by using Kraft lignin [28].
of soft domains in the chain network, thus imparting The compression force deflection test measures the
greater flexibility to the material. force necessary to produce a 50% compression over the
Different considerations were made about castor oil- entire top surface area of the foam specimen. It was deter-
based foams in comparison with PPG triol-based foams. mined by following the ASTM norm D 3850-94 and repre-
They showed a Tg at about 0 °C and presented peaks with sents a parameter that increases with the foam stiffness
broader shapes; this greater amplitude signifies a wider and depends on the lignin content in the final material.

Table 3
Apparent density of polyurethane foams prepared in free (1) and controlled expansion (2).

Sample Apparent density (kg/m3) Sample Apparent density (kg/m3)


1 2 1 2
P-PPG-A 117.8 113.4 HP-PPG-A 101.6 124.8
P-PPG-B 107.9 147.9 HP-PPG-B 86.6 132
P-PPG-C 87.1 95.2 HP-PPG-C 72.1 99.9
P-PPG-D 80.0 103.7 HP-PPG-D 76.6 100.2
P-PPG-E 83.7 102.5 HP-PPG-E 78.8 93.4
P-PPG-F 69.8 105.0 HP-PPG-F 77.5 105.4
P-PPG-G 81.9 100.5 HP-PPG-G 85.9 96.3
P-CO-A 129.7 147.0 HP-CO-A 127.4 180.9
P-CO-B 145.1 178.7 HP-CO-B 151.1 207.8
154 J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156

thickness. The value of compression force deflection


increased linearly as the apparent density increased.
Fig. 7 reports the values of compression force versus
apparent density for foams produced by using castor oil
as a chain extender.
As reported above, the castor oil-based foams presented
a higher density when compared to PPG-based ones, and in
this case, as reported in Fig. 7, the density and the com-
pression force values cannot be directly correlated because
the structure of the foam was less regular. Thinner cell
walls and larger foam cells caused the decrease of com-
pressive strength even if the apparent density was higher
than in the PPG-based samples.
Taking into account the nature of lignin, both of the
above figures showed similar behaviour for each kind of
starting material; this means that there was no influence
Fig. 7. Compression force deflection value (CFDV) for samples produced on the trends of compression strength, whatever the
with castor oil as a chain extender. employed chain extender. Comparing an equal formula-
tion, the foams produced starting from oxypropylated lig-
nin presented a lower compression value than those
Higher values of compression strength indicate that foam based on commercial lignin; this was due to the previous
is sturdier, which depends on both higher crosslinking reaction step that had degraded the structure of lignin
density and greater apparent density. Fig. 6 reports a direct making it more fragmented and therefore lighter.
correlation between CFDV and apparent density for a sam- Low functionality, low hydroxyl number and secondary
ple produced with PPG triol. hydroxyl groups of castor oil led to a low reactivity when
This linear increase was in agreement with our above- directly used to produce flexible polyurethane foams. By
cited previous work: with the aim to increase the apparent mixing CO with polyols such as glycerol and PEG 400, a
density of the foam, it was possible to raise the NCO/OH, polyol mixture with higher hydroxyl number was
thus giving the foam a higher reticulation degree and obtained, which leads to less flexible foams. In the present
therefore increasing the compression strength value. All work, replacing PPG triol with CO as a chain extender
the foams produced with PPG triol as a chain extender, resulted in foams with a higher apparent density and,
after the preflex load and recover time of 1 min, showed mainly, apparent density and compression values cannot
a small decrease of thickness, less than 2% of its initial be directly correlated, as reported above.

Fig. 8. SEM micrographs (enlargement 50) of all investigated foams, produced either by changing the lignin or varying the chain extender.
J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156 155

The presence of an open cell structure, which is typical vents than commercial soda lignin, and, in addition, the
of flexible foams, was assessed by Scanning Electron miscibility between liquefied lignin and chain extenders
Micrographs (SEM). This aspect is very important since (in particular with castor oil).
the mechanical properties of the foams are primarily influ- All the polyurethane foams showed a high content of
enced by the apparent density, thickness and sizes of the open cells and their thermo-mechanical properties were
cells [40]. suitable for fillings and packaging applications, such as fill-
Fig. 8 shows SEM micrographs of foams produced by ing for the interior parts of car seats and packaging of fur-
using PPG triol or Castor Oil as chain extender and lignin niture. In this sense, foams produced by using PPG triol as a
or oxypropylated lignin as fraction of polyol. chain extender had an apparent density ranging from
The cell shape of the foam was polyhedral and showed about 70 to 150 kg/m3 and a compression force deflection
preferential orientation associated with the polyurethane value from about 2  102 to 7  102 MPa, while foams
rise direction. As reported in literature, the precipitation based on castor oil had an apparent density ranging from
of the urea phase and the formation of the urea aggregates about 130 to 210 kg/m3, and a compression strength rang-
can be directly related to the cell-opening event in flexible ing from about 7  103 to 3.5  102 MPa.
PU foams [41]. Starting from soda lignin or soda oxypropylated lignin
Regarding the foams produced by using PPG triol, they could be considered a good alternative way to produce
showed more regular and thinner cells than those pro- flexible polyurethane foams based on renewable sources,
duced starting from CO; also the cell shapes were more in agreement with previous work based on Kraft lignin.
regular. To confirm these statements, CO-based foams pre-
sented also a higher number of untouched intercellular
Acknowledgements
membranes, which were not suitable for flexible behav-
iour. Foams based on castor oil showed coarse cells being
The authors wish to acknowledge the support from FP7
large in size, contributing to a less homogeneous structure;
– KBBE project n° 212239 Forbioplast (Forest Resource Sus-
this heterogeneity led to a lower compressive modulus and
tainability through Bio-Based Composite Development) to
higher apparent densities, compared to the more regular
carry out this research.
samples produced with PPG triol.
Considering the effects of lignin treatment, the foams
manufactured starting from oxypropylated lignin exhib- References
ited anisotropic cellular structures with elongated cells
along the blow direction, as expected [42,43]. Samples pro- [1] Narine SS, Kong X, Bouzidi L. Physical properties of polyurethanes
produced from polyols from seed oils. II. Foams. J Am Oil Chem Soc
duced by using oxypropylated lignin and CO presented a 2007;84(1):65–72.
regular structure, despite the cells were less regular than [2] Petrovic ZS. Polyurethanes from vegetable oils. Polym Rev
the samples based on PPG triol. 2008;48(1):109–55.
[3] Sharma C, Kumar S, Raman Unni A, Aswal VD, Rath SK, Harukrishnan
G. Foam stability and polymer phase morphology of flexible
4. Conclusions polyurethane foams synthesized from castor oil. J Appl Polym Sci
2014:40668–75.
[4] Stewart D. Lignin as base material for materials applications:
Starting from the our previous work, this article reports chemistry, application and economics. Ind Crops Prod
the successful green and innovative synthesis of flexible 2008;27(2):202–7.
polyurethane foams starting from Protobind lignin and [5] Serrano L, Briones R, Melus A, Herseczk Z, Labidi J. Polyols from the
lignocellulosic waste of biodiesel production process. Chem Eng
oxypropylated Protobind lignin, by following an environ-
Trans 2010;21(1):1339–44.
mentally-friendly process based on the use of microwave [6] Belgacem MN, Gandini A. Monomers, polymers and composites from
irradiation in order to liquefy the lignin, and on the use renewable resources. Oxford: Elsevier; 2008.
of water as a blowing agent. Glycerol and PEG 400 were [7] Li Y, Ragauskas AJ. Kraft lignin-based rigid polyurethane foam. J
Wood Chem Technol 2012;32(3):210–24.
used as liquefaction solvents, while polypropylene glycol [8] Crestini C, Melone F, Sette M, Saladino R. Milled wood lignin: a linear
triol or castor oil were used as chain extenders in order oligomer. Biomacromolecules 2011;12(11):3928–35.
to produce flexible foams in controlled expansion, thus tai- [9] Lange H, Decina S, Crestini C. Oxidative upgrade of lignin – recent
routes reviewed. Eur Polym J 2013;49(6).
loring the properties of the final foams. All samples were [10] Bonini C, D’Auria M, Emanuele L, Ferri R, Pucciariello R, Sabia AR.
prepared using a one-shot approach maintaining the Polyurethanes and polyesters from lignin. J App Polym Sci
NCO/OH ratio less than one hundred, thus reducing signif- 2005;98(3):1451–6.
[11] Cateto CA, Barreiro MF, Rodrigues AE, Brochier-Salon MC,
icantly the degree of crosslinking and giving higher flexi- Thielemans W, Belgacem MN. Lignins as macromonomers for
bility to the material. polyurethane synthesis: a comparative study on hydroxyl group
Considering the characterizations carried out, foams determination. J App Polym Sci 2008;109(5):3008–17.
[12] Hatakeyama H. Polyurethanes containing lignin, chemical
based on PPG triol showed a lower Tg of the soft phases,
modification, properties and usage of lignin. New York: Kluver;
clearly separated from the hard ones, while samples based 2002.
on castor oil as a chain extender presented both phases [13] Hatakeyama H, Kosugi R, Hatakeyama T. Thermal properties of
lignin- and molasses-based polyurethane foams. J Therm Anal
mixed. The amount of lignin in the formulation ranged from
Calorim 2008;92(2):419–24.
6% to 13% and considering castor oil as renewable source, [14] Glasser WG, Wu LCF, Selin JF. Synthesis, structure, and some
the amount of green raw materials can be more than 45%. properties of hydroxypropyl lignins. In: Wood and agricultural
The kind of lignin did not greatly influence the structure residues – research on use for feed, fuels and chemicals. New York:
Academic Press; 1983. p. 149–66.
and properties of the produced foams, although oxypropy- [15] Crestini C, Crucianelli M, Orlandi M, Saladino R. Oxidative strategies
lated lignin was dissolved more easily in liquefaction sol- in lignin chemistry: a new environmental friendly approach for the
156 J. Bernardini et al. / European Polymer Journal 64 (2015) 147–156

functionalisation of lignin and lignocellulosic fibers. Catal Today [31] Goddard RJ, Cooper SL. Polyurethane cationomers with pendant
2010;156(1–2):8–22. trimethylammonium groups. 1. Fourier Transform Infrared
[16] Aniceto JP, Portugal I, Silva CM. Biomass-based polyols through Temperature studies. Macromolecules 1995;28(5):1390–400.
oxypropylation reaction. ChemSusChem 2012;5(8):1358–68. [32] Cole KC, Van Gheluwe P. Flexible polyurethane foam I. FTIR analysis
[17] Wu LC, Glasser W. Engineering plastics from lignin. I. Synthesis of on residual isocyanate. J App Polym Sci 1987;34(1):395–407.
hydroxypropyl lignin. J Appl Polym Sci 1984;29(4):1111–23. [33] Zhang XD, Bertsch LM, Makosco CW, Turner RB, House DW, Scott RV.
[18] Cateto CA, Barreiro MF, Rodrigues AE, Belgacem MN. Optimization Effect of amine additives on flexible molded foams properties. Cell
study of lignin oxypropylation in view of the preparation of Polym 1998;17(5):327–49.
polyurethane rigid foams. Ind Eng Chem Res 2009;48(5):2583–9. [34] Javni I, Petrovic Z, Guo A, Fuller R. Thermal stability of polyurethanes
[19] Fernandes S, Freire CS, Pascoal Neto C, Gandini A. The bulk based on vegetable oils. J App Polym Sci 2000;77(8):1723–34.
oxypropylation of chitin and chitosan and the characterization of [35] Satheesh Kumar MN, Siddaramaiah. Thermogravimetric analysis and
the ensuing polyols. Green Chem 2008;10(1):93. morphological behavior of castor oil based polyurethanes polyester
[20] Laurichesse S, Avérous L. Chemical modification of lignins: towards nonwoven fabric composite. J App Polym Sci 2007;106(5):3521–8.
biobased polymers. Prog Polym Sci 2014;39:1266–90. [36] Aneja A, Wilkes GL. Exploring macro- and microlevel connectivity of
[21] Evtiouguina M, Barros AM, Cruz-Pinto JJ, Neto CP, Belgacem N, Pavier the urea phase in slabstock flexible polyurethane foam formulations
C, et al. The oxypropylation of cork residues: preliminary studies. using lithium chloride as a probe. Polymer 2002;43(20):5551–61.
Bioresour Technol 2000;73(2):187–9. [37] Eceiza A, Martin MD, de la Caba K, Kortaberria G, Gabilondo N,
[22] Pavier C, Gandini A. Oxypropylation of sugar beet pulp. 1. Corcuera MA, et al. Thermoplastic polyurethane elastomers based on
Optimization of the reaction. Ind Crops Prod 2000;12(1):1–8. polycarbonate diols with different soft segment molecular weight
[23] Zhang L, Guo X, Yan F, Su M, Li Y. Study of degradation behaviour of and chemical structure: mechanical and thermal properties. Polym
dimethoate under microwave irradiation. J Hazard Mater Eng Sci 2008;48(2):297–306.
2007;149(3):675–9. [38] Danch A, Ilisch S, Sulkowski WW, Moczynski M, Radon A, Radusch
[24] Krzan A, Kunaver M. Microwave heating in wood liquefaction. J App HJ. DMTA study on the urethane network in rubber waste-urethane
Polym Sci 2006;101(2):1051–6. composites. J Therm Anal Calorim 2005;79(3):623–30.
[25] Zheng Z, Pan H, Huang Y, Chung YH, Zhang X, Feng H. Rapid [39] Hernandez R, Weksler J, Padsalgikar A, Choi T, Angelo E, Lin JS, et al.
liquefaction of wood in polyhydric alcohols under microwave A comparison of phase organization of model segmented
heating and its liquefied products for preparation of rigid polyurethanes with different intersegment compatibilities.
polyurethane foam. Open Mater Sci J 2011;5(1):1–8. Macromolecules 2008;41(24):9767–76.
[26] Xie J, Qi J, Hse CY, Shupe TF. Effect of lignin derivatives in the bio- [40] Gibson LJ, Ashby MF. Cellular solids: structure and properties.
polyols from microwave liquefied bamboo on the properties of Cambridge University Press; 1997.
polyurethane foams. Bioresources 2014;9(1):578–88. [41] Rossmy GR, Kollmeier HJ, Lidy W, Schator H, Wiemann M.
[27] Ouyang X, Lin Z, Yang D, Qiu X. Chemical modification of lignin Mechanism of the stabilization of flexible polyether polyurethane
assisted by microwave irradiation. Holzforschung 2011;65(5): foams by silicone-based surfactants. J Cell Plast 1981;17(6):319–27.
697–701. [42] Moreland JC, Wilkes GL, Turner RB. Viscoelastic behavior of flexible
[28] Cinelli P, Anguillesi I, Lazzeri A. Green synthesis of flexible slabstock polyurethane foams: dependence on temperature and
polyurethane foams from liquefied lignin. Eur Polym J relative humidity. I. Tensile and compression stress (load)
2013;49(6):1174–84. relaxation. J App Polym Sci 1994;52(4):549–68.
[29] Kurimoto T, Koizumi A, Doi S, Tamura Y, Ono H. Wood species effects [43] Gouveia de Souza A, Oliveira Santos JC, Conceição MM, Dantas Silva1
on the characteristics of liquefied wood and the properties of MC, Prasad S. A thermoanalytic and kinetic study of sunflower oil.
polyurethane films prepared from the liquefied wood. Biomass Brazil J Chem Eng 2004;21(02):265–73.
Bioenergy 2001;21(5):381–90. [44] Ogunniyi DS. Castor oil: a vital industrial raw material. Biores
[30] Pawlik H, Prociak A. Influence of palm oil-based polyol on Technol 2006;97:1086–91.
the properties of flexible polyurethane foams. J Polym Environ
2012;20(2):438–45.

You might also like