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

Polymer Testing 68 (2018) 135–145

Contents lists available at ScienceDirect

Polymer Testing
journal homepage: www.elsevier.com/locate/polytest

Rigid polyurethane foams reinforced with industrial potato protein T


a,∗ b a
Sylwia Członka , Massimo F. Bertino , Krzysztof Strzelec
a
Institute of Polymer & Dye Technology, Lodz University of Technology, Lodz, Poland
b
Department of Physics, Virginia Commonwealth University, Richmond, VA, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Potato protein (PP), the main product of thermal and acidic coagulation of potato juice obtained from potatoes
Polyurethane during starch production, was selected as a reinforcement filler in polyurethane (PU) foams. Rigid polyurethane
Rigid polyurethane foam foams (RPUFs) were successfully obtained from a two-component system and modified with PP, in the range of
Cellular structure 0.1–5 wt% of total polymer mass content. The influence of PP on the RPUFs morphology and on physical and
Filler
mechanical properties of the obtained porous materials were studied. PU foams were characterized using
Potato protein
morphological, mechanical and thermal analysis. The results show that an addition of 0.1 wt% PP improves
compressive strength (the final strength up to 230 kPa) due to the strengthening effect of the filler, low thermal
conductivity (0.029 W/m·K) and low water absorption (14.3%). The results obtained in this study confirm that
the addition of PP over a certain optimal level has a negative effect on the cell morphology, leading to a de-
terioration of the physico-mechanical properties of the modified foams.

1. Introduction sources has been observed.


A promising solution to improve the properties of obtained mate-
Polyurethanes (PUs) are polymers with a wide range of applications rials, reduce demand for non-renewable petrochemicals and limit the
that can display quite different properties according to the chemical emission of carbon dioxide is incorporation of natural fillers into PU
reagents used in their synthesis. One of the most important commercial foams. Natural fillers can increase biodegradation efficiency and also
applications of made of PU materials is rigid polyurethane foams reduce cost. In addition, natural fillers improve the mechanical and
(RPUFs) [1]. A common approach for making PU foams is by reaction of thermal properties of RPUFs. Their effectiveness as a reinforcing filler is
a polyol and an isocyanate, which forms the backbone urethane group. dependent on the method of incorporation [6], compatibility with the
Typically, a blowing agent is used to create cells within the foam. Ad- component [7–9], quality of the dispersion [9–11] and loading.
ditional modulating agents, such as, catalysts and surfactants, can be Several studies have reported the results of foams reinforced with a
added during production to modulate the properties of the resulting different types of fillers. Prociak et al. [14] synthesized polyurethane-
foam [2]. polyisocyanurate foams based on rapeseed oil and modified with
PU foams have rapidly grown to be one of the most diverse and carbon fiber. With increasing carbon fiber content, RPUFs have a re-
widely used plastics with a continuously increasing global market [3]. duced heat rate release, reduced flammability and greater thermal
RPUFs are commonly used in a variety of industries including auto- stability. Zhou et al. [11] have reported the synthesis of a series of semi-
motive, furniture, construction, packaging and manufacturing [4]. In rigid biopolyurethane foams based on palm-oil polyol and reinforced
buildings, they are usually employed as insulating materials for pipe- with cellulose nanocrystals. Foams with an apparent density of 50 kg/
lines and lightweight construction components e.g. parts of door and m3 and an improved compressive strength (up to 117 kPa) were ob-
window frames, furniture or home and commercial refrigeration tained. Compared to neat foam, modified compositions were char-
equipment [5]. acterized by superior dimensional stability in freezing and heating
Because of the large market and numerous applications, the interest conditions and greater water uptake. RPUFs reinforced with wheat
in PUs has increased. Manufactures of PU are looking for new methods straw lignin in the amount of 0–6.3 wt% of the total foam mass were
for their modification, which would be advantageous from an economic prepared by Paberza et al. [12]. RPUFs with an apparent density in the
and ecological point of view. A question related to the production of range of 45–60 kg/m3 were obtained. It was noticed that increasing
RPUFs is the environmental impact. In this regard, a growing tendency concentration of lignin in RPUFs improved thermal insulation proper-
to modify RPUFs with additives obtained from renewable, bio-based ties. The maximum value of compressive strength (0.35 MPa parallel to


Corresponding author.
E-mail addresses: sylwia.czlonka@edu.p.lodz.pl, sczlonka@vcu.edu (S. Członka).

https://doi.org/10.1016/j.polymertesting.2018.04.006
Received 3 January 2018; Received in revised form 26 February 2018; Accepted 7 April 2018
Available online 09 April 2018
0142-9418/ © 2018 Elsevier Ltd. All rights reserved.
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 1. PP observed in magnification of a)


×50 b) ×250 and c) the end-products
containing 0.1, 1 and 5 wt% of PP (from the
left, respectively).

the foaming direction) was reached for the foam with 1.2 wt% of lignin. protein (producer: PEPEES S.A., Łomża, Polska) was used as a biode-
Potato is the fourth most important food crop in the world, after gradable filler (Fig. 1a,b). According to the supplier, the potato protein
rice, wheat and corn [13]. Apart from being eaten directly, a large is composed of dry matter (80–85 wt%), ashes (below 4.5 wt%) and
amount of potato is used to produce potato starch (PS) [14]. The PS moisture (below 11 wt%). Before adding to the polyol, the filler was
industry releases large quantities of by-product, known as potato fruit dried to a constant weight at a temperature of 70 °C and subjected to
juice (PFJ) [14]. It was reported that PFJ contains 2–5 wt% of dry multistage disintegration by passing through a sieve with a square mesh
matter, of which one third is potato protein (PP) [14,15]. However, at size of 0.2 mm.
present, most PS processing factories discharge PFJ directly, causing
environmental pollution and resource waste, which seriously restricts 2.2. Preparation of RPUFs
the development of the potato processing industry [14]. PP has often
been overlooked but their removal from the starch industry byproduct RPUFs were produced by a single step method from a two-compo-
is necessary to help overcome the economic impact coupled to its high nent (A and B) system. Following to the information provided by the
polluting capacity. Because of this, the problem of utilization of PP is supplier the ingredients were mixed in a ratio 100:160 (ratio of
still not fully resolved and requires search for new opportunities and Izopianol 30/10/C to Purocyn B). RPUFs were prepared by chemical
technologies of its use. Addition of PP as a filler in polymer composites foaming with CO2 as the foaming agent produced by water and iso-
seems to be a logical approach to reduce the amount of waste PP and an cyanate reaction. Composite samples were prepared with 0.1, 1 or 5 wt
obvious methodology to increase the green material in polymeric ma- % of PP (Fig. 1c). Unmodified RPUF was used as a reference. The raw
terials. materials and formulations used in the synthesis of RPUFs are presented
Not surprisingly, the structure and properties of protein-based bio- in Table 1.
polymers are being actively investigated [16–21]. For example, bio- First, Izopianol 30/10/C was weighed out in the required propor-
polymers with PP and glycerol as a plasticizer have been successfully tions and placed in a 1000 ml polypropylene beakers. PP was then
prepared by Du et al. [22]. The resulting materials had tensile strength added and the mixture (component A) was homogenized with an
in the range of 20–25 MPa, depending on plasticizer concentration. overhead stirrer at 600 rpm under ambient conditions for approxi-
Molecular weight and protein structure also strongly affected the me- mately 30 s. Next, Purocyn B (component B) was added to component A
chanical properties. Comparable results were also obtained by Newson and the mixture was stirred for 10 s at a speed rate of 1800 rpm. The
et al. [23] who examined the suitability of commercial PP for bio-based mixture was then poured into an open mould and allowed to expand
films. Bio-films with tensile strength up to 20 MPa were obtained. The freely in the vertical direction. Obtained RPUFs were conditioned at
shape of the tensile stress–strain curve in thermoformed potato-based room temperature for 24 h. After this time, samples were cut with a
films was controlled by glycerol level and was independent of proces- band saw into appropriate shapes (determined by testing standards)
sing temperature. Tensile testing revealed that elongation at break in- and their physico-mechanical properties were investigated.
creased with processing temperature while Young's modulus was un-
affected by processing temperature.
2.3. Characterization methods
Despite the use of PP in different polymer composites PP has not
been considered before as a possible filler for RPUFs. In this work, we
The average size of PP was measured using a Zetasizer NanoS90
investigate the use of PP as a filler improve mechanical, thermal and
instrument (Malvern Instruments Ltd, UK). The size of particles in water
physical properties of RPUFs. The main goal of the present work is to
and polyol dispersion (0.04 g/l) was determined with dynamic light
optimize the foam formulation, including PP as a filler in the reaction
scattering (DLS).
mixture as well as appropriate synchronization of the amount of filler
The oil absorption parameter (DBPA) of the filler was determined
with the reaction temperature and viscosity of reaction mixture. These
according to ASTM D2414 [24] using Absorptometer C (Brabender
are key aspects of the blowing process, leading to obtaining products
GmbH & Co KG, Germany). Dibutyl phthalate (DBP) was added with a
with desirable properties. This modification can result in significant
titration rate of 4.0 ml/min to 20 g of powder filler. The oil absorption
benefits, not only from the application but also economic and ecological
parameter was recorded and determined as DBP measured volume per
point of view.
gram of analyzed sample.
The absolute viscosities of polyol and isocyanate were determined
2. Material and methods corresponding to ASTM D2930 [25], equivalent to ISO 2555 [26], using

2.1. Raw materials Table 1


Foam formulations.
RPUFs were synthesized from a two-component system (Izopianol
Foam symbol Mass content [by weight]
30/10/C and Purocyn B) supplied by Purinova Sp. z o.o. Izopianol 30/
10/C used in the reaction is a fully formulated mixture containing the Izopianol 30/10/C Purocyn B Potato protein
mixture of polyester polyol (hydroxyl number ca. 230–250 mgKOH/g
and functionality of 2), catalyst (N,N-Dimethylcyclohexylamine), flame FR 100 160 0
FP0.1 100 160 0.1
retardant (Tris(2-chloro-1-methylethyl)phosphate) and chain extender
FP1 100 160 1
(1,2-propanediol). Purocyn B is a polymeric diphenylmethane-4,4′-dii- FP5 100 160 5
socyanate (pMDI) containing 31 wt% of free isocyanate groups. Potato

136
S. Członka et al. Polymer Testing 68 (2018) 135–145

a rotary Viscometer DVII+ (Brookfield, Germany). The torque of according to ASTM D7264 [31], equivalent to ISO 178 [32], using
samples was measured as a range of shear rate from 0.5 to 100 s−1 in Zwick Z100 Testing Machine (Zwick/Roell Group, Germany). Samples
ambient temperature. were bent with a testing speed 2 mm/min. Flexural stress at break (εfB)
The apparent density of foams was determined accordingly to ASTM results for each sample were expressed as an average over 5 measure-
D1622 [27], equivalent to ISO 845 [28]. The densities of 5 specimens ments per each type of composition.
per sample were measured and averaged. Dynamic mechanical analysis (DMA) was determined using ARES
The morphology and cell size distributions of RPUFs were examined Rheometer (TA Instruments, USA). Torsion geometry was used with
from the cellular structure images of foam cross-sections, which were samples of thickness 2 mm. Measurements were examined in the tem-
taken using an optical stereomicroscope Leica MZ6 (Leica perature range 20–250 °C at a heating rate of 10 °C/min, using a fre-
Microsystems, Germany) at 250 times magnification. All the RPUFs quency of 1 Hz and an applied deformation at 0.1%.
were scanned in the free-rising direction. The average pore diameters Water absorption of the RPUFs was measured according to ASTM
and pore size distribution were calculated using ImageJ software D2842 [33], equivalent to ISO 2896 [34]. Samples were dried for 1 h at
(Media Cybernetics Inc.). 80 °C and then weighed. The samples were immersed in distilled water
The chemical structure of RPUFs was determined by Fourier trans- to a depth of 1 cm for 24 h. Afterwards, the samples were removed from
form infrared spectroscopy (FTIR). FTIR spectra were collected for the the water, held vertically for 10 s, the pendant drop removed and then
wavelength range of 3500–400 cm−1 at a resolution of 4 cm−1 using a blotted between dry filter paper (Fisher Scientific, USA) at 10 s and
Nicolet 6700 FTIR Spectrometer (Thermo Fisher Scientific, Germany). weighed again. The average of 5 specimens was used.
FTIR was performed with a DGTS/KBr detector. Data processing was Changes in linear dimensions were determined in according to
performed using OMNIC 3.2 software developed by Thermo Scientific ASTM D2126 [35], equivalent to ISO 2796 [36]. Samples were condi-
Products (Thermo Fisher Scientific, Germany). Reported spectra re- tioned at 70 °C for 48 h. Length, width and thickness of the samples
present an average of 64 individual scans. were measured before and after conditioning using a vernier caliper
Thermal properties were evaluated by thermogravimetric analysis with an accuracy of 0.01 mm. The reported values refer to an average of
(TGA) performed using a STA 449 F1 Jupiter Analyzer (Netzsch Group, 5 measurements per composition.
Germany). About 10 mg of the sample was placed in the TG pan and
heated in argon atmosphere at a rate of 10 K min−1 up to 600 °C. The 3. Results and discussion
initial decomposition temperatures, T10%, T50% and T80% of mass loss
were determined. 3.1. Characteristics of the filler
The compressive strength of foams was determined according to
ASTM D1621 [29], equivalent to ISO 844 [30], using a Zwick Z100 The chemical structure of PP was determined by FTIR. The obtained
Testing Machine (Zwick/Roell Group, Germany) with a load cell of 2 kN spectrum (Fig. 2) confirms the presence of reactive groups in the filler
and a speed of 2 mm/min. Samples of specified sizes (10 × 10 × 10 cm) molecule. The intense bands in the range of 1490, 1510 and 1620 cm−1
were cut with a band saw in direction perpendicular to the foam growth are attributed to the vibration of –COO groups, N-H bending vibration
direction. Samples were placed between two plates and the compres- (of secondary amines) and N–H bending vibration (of primary amines)
sion strength was measured as a ratio of load causing 10% deformation of protein molecules, respectively [37]. A weak broad band between
of sample cross-section in the parallel and the perpendicular direction 3200 and 3500 cm−1 indicates the presence of -OH groups in side
to the square surface. The result was averaged over 5 measurements per chains and terminal groups [37]. The band at 1210 cm−1 is related to
each sample. the stretching acyl groups of proteins molecules [38].
Three point bending tests were carried out at room temperature, Since PP has polar groups and hydrophilic properties, strong

Fig. 2. FTIR spectra of RPUFs.

137
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 3. Particle size distribution of the filler measured after 5, 10, 15, 20 and 25 min of ultrasonic mixing.

interfacial interaction, such as hydrogen bonding, can be formed be- presented results the viscosity increase from 0.1 wt% PP to 1 wt% is
tween the PP molecules and isocyanate leading to the formation of a large, while from 1 wt% PP to 5 wt% PP is a comparatively small
cross-linked structure. Amine, hydroxyl and carboxyl groups present in change. It can be also noticed that each of the modified system is
PP molecules can react with isocyanates even in the absence of catalyst characterized by the same viscosity characteristics. The initial viscosity
[39–42]. It has been reported in previous work, that isocyanate has a drop is quick, whereas for the higher shear rates (about 6s−1), when the
higher reactivity with amines than with hydroxyl and carboxyl groups. particles of liquids reach the best possible arrangement, the viscosity
The scheme of the amine reaction with isocyanate is shown in equation drop is significantly slower and finally becomes linear. Because the
(1). The reactions of hydroxyl and carboxyl groups with isocyanate viscosity of analyzed liquids decreases under shear strain, these fluids
follow a similar mechanism at much slower rates forming urethane and can be described as shear-thinning liquids [43,45].
amide bonds as the end-products, respectively [39–42].

The results presented in Fig. 3 indicate that the filler particles ex- 3.2. Foams characteristic
hibit high tendency to agglomeration. The size of particles increases
(from 190 to 490 nm) with time, while the intensity of the peak de- 3.2.1. Foaming kinetics
creases (from 31 to 11%), pointing to a reduced particle concentration. The reaction of synthesis of PU is highly exothermic. A large amount
After 25 min the particle size distribution shows two bands. The first of heat is released in a short period of time during the gelling and
one covers diameters characteristic of primary aggregates, while the blowing reaction [46,47]. The rate of increase in temperature de-
second band evidences the presence of agglomerates. The results pre- termines the activity of reaction mixture, which in turn is associated
sented in Table 2 indicate that the average size of particles measured in with the reactivity of the components of the mixture. The results pre-
polyol is significantly higher than in water what can be connected with sented in Table 3 indicate that the addition of small amounts of PP leads
swelling due to absorption of the liquid. to a significant increase of the maximum temperature (Tmax) of reac-
The result of DBPA number, expressed as the volume of dibutyl tion. The maximum temperature reached by the modified RPUFs in-
phthalate (DBP) absorbed by specified amount of filler, is about 72.3 creases with the incorporation of PP in the formulation and is attained
[ml/100 g], indicating an unbranched structure of analyzed filler. For at longer times (Fig. 4b). The temperature increase is likely due to the
comparison, the DBPA value of carbon black is about 172 ml/100 g, higher concentration of reactive groups. These additional groups are
pointing to a large amount of free spaces between the aggregates and a amines and hydroxyl groups introduced by PP and probably to a lesser
branched filler structure [43]. extent, by residual water present in the filler [48].
The viscosity of polyol premixes is a crucial parameter determining The foaming process was characterized by measuring the char-
overall quality and uniformity of foam structure [43]. The results of acteristic processing times, cream, extension and gelation time. The
viscosity of compositions modified with PP are presented in Fig. 4a. The cream time was measured from the start of mixing of components to a
viscosity of the mixture rapidly increases when the concentration of PP visible start of foam growth, the extension time until the highest vo-
rises from 0.1 to 5 wt%. The viscosity increase is probably associated lume of the foam was reached and the gelation time was the time when
with the swelling of filler particles in polyol [44]. According to the the foam solidified completely and the surface was no longer tacky
[49]. The results presented in Table 3 indicate that addition of PP
changes the different characteristic times (as compared with the FR
Table 2 foam) and therefore the foaming kinetics. Compared to the FR for
The average size of the filler agglomerates formed in different media (water, which cream time is about 43 s, compositions modified with PP exhibit
polyol) and DBPA oil absorption values. a slight extension of cream time, that is 49 s, 57 s and 68 s for foams
Average size of filler Average size of filler Dibutyl Phthalate FP0.1, FP1 and FP5, respectively. According to Silva et al. [50] this
agglomerates in water agglomerates in polyol Absorption dependence is strongly connected with the fact that a well-dispersed
[nm] [nm] DBPA
filler can serve as a nucleating agent in the nucleation process leading
[ml/100 g]
to greater bubble formation and extend the cream time. Cream time is
1030 4280 72.3 also increases because the viscosity of the reaction mixture is increased

138
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 4. a) The viscosity curves of foam components and b) temperature rise during the foaming process.

Table 3
Parameters of RPUFs formation process and selected properties of RPUFs.
Foam Tmax Time [s] Apparent density [kg/m3] Thermal
[oC] conductivity [W/K∙m]
cream extension gelation upper middle bottom

FR 109 ± 2 43 ± 2 277 ± 11 341 ± 25 37.0 ± 1.1 34.2 ± 1.1 30.2 ± 0.6 0.028
FP0.1 133 ± 4 49 ± 1 339 ± 10 450 ± 22 39.8 ± 1.3 38.1 ± 0.8 36.5 ± 1.3 0.029
FP1 157 ± 4 57 ± 2 350 ± 12 442 ± 14 37.3 ± 0.9 35.4 ± 1.4 33.4 ± 1.2 0.030
FP5 160 ± 3 68 ± 2 382 ± 10 475 ± 10 36.4 ± 1.1 35.6 ± 0.7 30.2 ± 1.0 0.032

in PP-modified foams, causing more resistance to bubble expansion and 203 μm for the foams FP0.1, FP1 and FP5, respectively. This effect has
limited diffusion of gases into the bubbles. Higher viscosity can also be already been reported in previous works [61,62] and is attributed to the
the reason of longer extension time for foams modified with PP. The attachment of fillers to the polymer cell wall. This attachment weakens
higher reactivity of the system and higher temperature of reaction can the cell structure, and leads to ruptures in several points [61]. Com-
lead to a rapid increase in viscosity, decreasing the diffusion rate of pared to the FR foam, the modified foams are characterized by a wider
carbon dioxide and slowing the diffusion of gases into the air bubble cell size range and higher cell distribution frequency (Fig. 6). The
[51]. While, the characteristic times measured for PP-modified systems widest range of cell distributions is exhibited by composition FP5.
are higher than those measured for FR foam, the obtained results are These changes are likely due to the different temperatures which
still within the range of characteristic of RPUFs [52,53]. Similar results change in viscosity, the foam kinetics and hence the structure. Foams
of the filler's influence on the foaming process can be found in the lit- modified with PP are characterized by higher Tmax compared to the FR.
erature. Lee et al. [54] showed that addition of carbon nanotubes The temperature decrease viscosity and surface tension, leading to
(CNTs) into water-blown PU increases the reaction time by minutes. thinning of the cell walls. On the other hand, the diffusion rate of CO2
Prociak et al. [55] showed that modification of the reference system gas generated from inside the foam cells outward is faster than the
with ground walnut shells increased the start and extension times by diffusion rate of air into foam cells. If the expansion is too fast and cell
about 60%. An increase of the total time with increasing amounts of gas pressure becomes lower than the atmospheric pressure, the cell
lignin was observed also by Arshanitsa et al. [56]. walls and struts are too thin to withstand the pressure inside the cells
caused by contraction of gas (carbon dioxide) enclosed in the cells,
3.2.2. Cellular structure of RPUFs leading to the rupture of membrane and shrinking of the RPUFs [63].
One of the most important parameters of RPUFs is their cellular This phenomenon is observed especially in the case of foam FP5 which
structure. Morphology parameters such as cell size, wall thickness and is characterized by the highest temperature of reaction. Sung and Kim
anisotropy, have a prominent impact on the physical and mechanical [64] explained that the alteration in the cell morphology can be also
properties of PU foams [5,57]. The structure of RPUFs, especially the connected with the fact that filler particles can act as a gas nucleation
cell size and the kind of cell (open/closed) depends strongly on process sites during the foaming process and assist the formation of nucleation
parameters like viscosity of the reaction mixture, pressure and tem- centers for gaseous phase [64,65], thus affecting local rheology sur-
perature gradients during the foaming process [58,59]. According to rounding the growing bubbles [66]. The addition of powder filler can
Kang et el. [60] suitable synchronizing of the temperature of the reac- also change the nucleation mode from homogenous to heterogeneous
tion with the viscosity, increases a key aspect of optimizing the blowing and reduce the nucleation energy, which in turn promotes the forma-
process, leading to the formation of a foam with a well-developed tion of large numbers of small cells [64], increasing the tendency of cell
cellular structure. The intercellular structure of pure and PP-modified coalescence and leading to higher inhomogeneous cell size distribution.
foams is shown in Fig. 5. The structure of FR (Fig. 5a) is mostly regular The same tendency noticed in this work were observed in previous
and smooth with thick struts (intercellular membrane). With increasing studies. The influence of soy protein on morphology of RPUFs was in-
content of PP in the polyol mixture (Fig. 5b–d), the foams are char- vestigated by Zhang et al. [67]. Composite foams with 5, 10, 15 wt%
acterized by an increasingly more inhomogeneous structure with a filler had a cell size of 274, 239 and 202 μm, respectively compared to
higher content of open cells. Compared with FR, modified foams have a 391 μm of unmodified foam. A similar trend was also observed by other
reduced cell diameter which decreases from 316 μm to 275, 240 and authors [68,69].

139
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 5. Morphology of a) FR b) FP0.1 c) FP1 d) FP5 observed parallel to the blowing direction by using an optical microscope in magnification ×250.

Fig. 6. Cell size distributions of a) FR b) FP0.1 c) FP1 d) FP5.

140
S. Członka et al. Polymer Testing 68 (2018) 135–145

3.2.3. Apparent density reached 1 wt%, the compressive strength slightly decreased to 222 kPa,
Effects of filler addition on apparent density of RPUFs are presented but still remained higher than foam FR. With higher content of filler, in
in Table 3. Results show that as the filler content increases, the ap- the amount of 5 wt%, the normalized compressive strength begins to
parent density decreases. Changes in apparent density are closely re- decrease and reaches 173 kPa. Compressive strength results relate to
lated with the morphology changes reported in Fig. 5. Increasing PP the morphologies of Fig. 5. Foams with a large number of closed cells
concentration leads to less uniform and larger cells with thinner struts, with thick walls and struts can better withstand the compressive load
hence to a reduced apparent density. Prociak et al. [55] explained that [74]. Besides this, even dispersion of filler in the foam rise direction
the reduction of the apparent density may be due to the influence of the enhancing the struts and walls leads to the greater compressive
filler on the reactivity of the PU system, which leads to growth of the strength, especially parallel in the foam rise direction. As mentioned
foam and increased volume. The authors also argued that the moisture previously, with increasing content of PP, the structure of modified
present in the fillers can acting as a chemical blowing agent and further RPUFs becomes more inhomogeneous with a lower content of closed
reduce density. The results presented in Table 3 also indicate that ap- cells. Furthermore, the structure is characterized by smaller cells with
parent density is depending on the part of the sample. The density of thinner walls and struts. Because of this, the cell walls and struts are not
each foam series increases gradually from the bottom to the top part in strong enough to withstand the compressive load and more stress is
the PU foam rise direction. The bottom part of foam is characterized by generated on the cell struts, leading to deterioration of mechanical
large holes and greater porosity due to gravity effect. Because of this the properties of RPUFs.
cell growth in the bottom part of RPUFs is more than in the top part [1]. The mechanical properties of modified materials are also dependent
on the interaction between the filler and the PU matrix. A high inter-
3.2.4. Thermal conductivity action i.e., good adhesion between the filler and the polymer matrix
One of the most important parameter of RPUFs is thermal con- makes the stress transfer between filler and the PU matrix more effi-
ductivity, which determines their potential use as insulating materials cient and improves the mechanical properties [78–80]. In the case of
[70]. Thermal conductivity of foams is governed by several factors, like RPUFs modified with PP, high tendency to aggregate filler particles, can
wall thickness and apparent density [47,71]. The results presented in lead to a weakened interfacial adhesion between filler and effective
Table 3 indicate that introduction of PP increases the thermal con- active surface. In consequence, RPUFs are characterized by microphase
ductivity of foams. The thermal conductivity of the prepared RPUFs is separation of the structure, inhibiting the enhancement of mechanical
in the range of 0.029–0.032 W/mK which is slightly higher than idea- properties of obtained foams. The same tendency was observed by
lized RPUFs (0.020–0.024 W/mK) [72]. This result, however, is con- Kurańska et al. [81], who investigated the impact of flax fillers on the
sidered satisfactory if one takes into account that CO2 was used as physico-mechanical properties of obtained composites. They noticed
blowing agent (as opposed to the use the more thermally insulating that compressive strength can be decreased when a fraction of the
chlorofluorocarbon liquids, CFC, which have a harmful effect on the natural fibers is not incorporated in the struts between the cells, causing
environment). According to Seo et al. [73], deterioration of heat in- a disordered cell structure. Madaleno et al. [82] reported an increase in
sulating properties is closely related with the cellular morphology of compressive strength by 20% for PU foams modified with 0.25 wt% of
RPUFs. As the cell size is lower, more walls are present which inhibit montmorillonite-carbon nanotubes. With increase in filler amount, up
radiation. Increasing the PP content dropped the content of closed cell to 0.5 wt% and 1 wt%, the compressive strength decreases, which was
content, leading to deterioration of the thermal conductivity of the attributed to the poor dispersion of nanoparticles at higher concentra-
porous materials. The same tendency has already been reported by tions. The results obtained in the present study confirm that the addi-
Mosiwiecki et al. [71]. The authors showed that the thermal con- tion of filler beyond a certain optimum level has a negative effect
ductivity of RPUFs modified with castor oil increases as the wood flour morphology and leads to the reduction of fracture toughness and me-
concentration increases and explained that behavior considering the chanical strength.
increase of small and/or clustered-small cells of the reinforced foams. The compression behavior of RPUFs measured parallel to the foam
rise direction is presented in Fig. 7a. Each sample exhibits three re-
3.2.5. Compressive strength gimes: a linear elastic region, a plateau region and a densification re-
Mechanical properties of PU foams are closely dependent on the gion [61]. Compared to FR, compositions FP0.1 and FP1 exhibit more
apparent density and cellular structure. Cell morphology like cell size, sharp transition from the linear region (< 10% deformation) to the
shape and anisotropy as well as the thickness of cell walls and struts plateau (around 10–25%). Composition FP5 is characterized by a more
have a prominent effect on the compressive strength [74]. The influ- smooth transition, which can be ascribed to a greater ductility of the
ence of PP content on the mechanical properties of RPUF composites is modified foam [83].
presented in Table 4. Compression tests were carried out in the direc-
tions perpendicular and parallel to the foam rise direction. Compression
strength in the parallel direction is significantly larger than in the 3.2.6. Flexural strength
perpendicular direction, which can be ascribed to the high anisotropy An analysis of the results presented in Table 4 indicates that the
of the cells [75–77]. Compared to FR, the foams FP0.1 and FP1 exhibit flexural strength of PP-modified foams slightly decreases with the in-
higher mechanical strength in both directions. The FR possessed the creased filler in mass content due to their greater elasticity, ascribed to
compression strength of 187 kPa, while the sample FP0.1 exhibit the the cellular structure of foams. The results presented in Fig. 7b indicate
highest compression strength of 230 kPa. When the amount of PP that stress-strain curves for flexural loadings are very similar, pointing
at a comparable mechanical performance of the modified foams. With
Table 4 increasing content of PP, the foams exhibit an extended range of
Mechanical properties of RPUFs. elongation (measured as a region before fracture occurs) as a result of
greater flexibility of the PU matrix. A similar trend was observed by
Foam Compressive strength Compressive strength Flexural
(parallel) (perpendicular) strength
Ciecierska et al. [84], who investigated the impact of carbon nanotubes
[kPa] [kPa] [MPa] (CNTs) on the physico-mechanical properties of PU composites. They
argued that reduction of flexural strength is caused by weak adhesion
FR 187 ± 23 114 ± 13 0.81 ± 0.06 between the PU foam and filler particles as well as a nonuniform dis-
FP0.1 230 ± 10 120 ± 10 1.90 ± 0.06
tribution of CNTs. The same tendency was also presented by other
FP1 222 ± 14 119 ± 2 0.95 ± 0.03
FP5 173 ± 15 103 ± 11 0.88 ± 0.02 authors [61,83].

141
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 7. a) Compression behaviors of RPUFs measured parallel to the foam rise direction b) flexural stress-strain curves of RPUFs.

Table 5 Table 6
Analysis of signal displacements in FTIR spectroscopy of characterized mate- Thermal properties of RPUFs.
rials.
Foam T10 [oC] T50 [oC] T80 [oC] Tg [oC]
N-H C=O N-H C-N
FR 266 349 522 167
Wavenumber [cm−1] FP0.1 257 365 525 147
FP1 241 359 494 137
FR 3379 1716 1506 1411 FP5 231 347 476 135
intensity 0.0022 0.018 0.023 0.010

FP0.1 3375 1709 1508 1412 with the fact, that with increasing content of filler in reaction mixture,
intensity 0.0024 0.025 0.029 0.012
the amount of water increases. Some isocyanate group react with water,
FP1 3374 1709 1509 1412 instead of hydroxyl group of polyol, and consequently, smaller number
intensity 0.0019 0.016 0.020 0.009 of urethane species is formed. A band at 1411–1412 cm−1, attributed to
C-N vibrations, indicates that composition FP1 is characterized by the
FP5 3375 1716 1507 1412
largest content of isocyanurate rings [85,90,91]. No significant shift of
intensity 0.0019 0.016 0.022 0.008
the bands attributed to the N-H stretching of the urethane group (3200-
3600 cm−1) was noticed. This suggests that PP does not interact sig-
3.2.7. FTIR analysis nificantly with PU molecules. Thus, incorporation of PP does not affect
The FTIR spectra (Fig. 2) confirm the slight difference in the che- the separation between hard and soft segments, which could lead to
mical structure of analyzed foams. The urethane moieties of PU are greater microphase separation.
confirmed by the presence of the main absorptions bands, mainly C=O
stretch (1700–1770 cm−1) [85], N-H stretch (3200–3600 cm−1) [86] 3.2.8. Thermogravimetric analysis
and N-H bend (1500-1509 cm−1) [87–89]. Signal shift and band in- Thermal stability measurements are presented in Fig. 8 and in
tensities are presented in Table 5. The FTIR spectra are extremely si- Table 6. The temperatures at 10, 50, and 80% weight losses are taken as
milar in each case, except for only differences in the intensity of some the criteria for thermal stability. Each foam presents a similar de-
bands. With increasing content of PP, the modified foams are char- gradation pattern. As the filler content in polymer matrix increases,
acterized by a lower intensity of the band at 1700-1770 cm−1 attribu- modified samples show slightly worse thermal properties, which are
table to the carbonyl group of polyurethane group. It can be connected related to the changes in cellular structure. The first step of degradation
(corresponding to the temperature at 10% of total weight losses) is
connected with dissociation of urethane bond [92,93]. With addition of
the filler, the onset of thermal degradation is lower for each series of
modified RPUFs and decreases from 257 °C down to 231 °C. It can be
ascribed to the fact that RPUFs modified with higher amount of the PP
are characterized by more ductile structure compared with FR. The
second step of degradation of RPUFs is associated with decomposition
of polyol segments [92,94]. The results indicate that decomposition
temperature is significantly affected by the amount of PP in mass
content. Foams FP0.1 and FP1 are characterized by better thermal
properties and degrade at higher temperature than unmodified foams,
while the foam FP5 exhibits slightly worse properties compared to FR.
A similar trend is observed in the third step of degradation (loss of
weight by about 80%), which is ascribed to the degradation of the
fragments generated during the second step [92,94]. Increasing the
percentage of filler reduces the thickness of walls and cell size, thereby
samples are more liable to higher temperature and exhibit worse
thermal properties. Deterioration of the thermal properties may be also
due to the degradation of PP [23]. Above 170 °C protein fragmentation
begins, leading to the damage of protein network and increasing the
Fig. 8. TGA curve of the RPURFs. thermal degradation of resulting materials [42,95]. Similar results can

142
S. Członka et al. Polymer Testing 68 (2018) 135–145

Fig. 9. a) Tanδ and b) storage modulus as a function of the temperature of RPUFs.

Table 7
Thermal stability of RPUFs measured by linear dimension change, weight loss during temperature exposition and water absorption.
Foam Dimensional stability [%] Loss in mass Water
[g] absorption [%]
length width Thickness

FR 0.90 ± 0.03 1.31 ± 0.12 0.65 ± 0.01 0.65 ± 0.13 18.2 ± 1.2
FP0.1 1.39 ± 0.03 1.38 ± 0.03 1.12 ± 0.00 0.84 ± 0.08 14.3 ± 1.4
FP1 1.42 ± 0.04 1.48 ± 0.07 1.14 ± 0.05 0.82 ± 0.10 13.7 ± 2.1
FP5 0.76 ± 0.01 1.29 ± 0.01 1.08 ± 0.14 0.60 ± 0.09 13.6 ± 0.8

be also found in literature [67,96,97]. compared with the standard deviation, and thus they can be attributed
mostly to experimental errors. In each case, the dimensional stability is
3.2.9. Dynamic-mechanical properties still considered to be mild and within commercially acceptable limits
The dynamic-mechanical properties of foams are presented in Fig. 9. [98].
Glass transition temperatures (Tg) were determined by the peak max- Water absorption of cellular materials depends mainly on cellular
imum of tanδ as a function of temperature (Fig. 9a). With increasing PP morphology and on the hydrophobic nature of PU [99–101]. The open
concentration Tg decreases, from 147 °C to 135 °C (Table 6). The tanδ and closed cell content as well as apparent density are the main factors
peak also broadens with increasing PP concentration, which can be affecting foam water absorption [100]. It can be noticed that despite
attributed to the increased inhomogeneity of cellular structure with the the presence of higher content of open cells, the modified RPUFs absorb
PP incorporation noticed in Fig. 5 [1]. With increasing filler content, less water than foam FR (Table 7). The reduced water uptake can be
the storage modulus (E′) decreases (Fig. 9b), which can be ascribed to attributed to the hydrophobic character of PP [14]. PP itself is water-
the plasticizing effect of filler particles on the polymer matrix and in- soluble, however after the industrial coagulation process, the protein is
dicates weaker interactions between PU matrix and fillers. Storage damaged and aggregated, which limits its interaction with water [14].
modulus is closely related to interfacial adhesion between and the A similar trend of fillers, preventing the penetration by water can be
polymer matrix. Good adhesion between the filler and the polymer also found in previous work. Kuranska et al. [44] reported that in-
matrix leads to a high cross-linked structure of resulting materials, re- corporation of microcellulose in amount to 9% into the PU system
sulting in high storage modulus. Poor interfacial adhesion between caused a reduce of water uptake by up to 1%. In our case, water ab-
filler and PU matrix is an effect of unbranched structure of the filler and sorption in slightly higher, however, this result is still considered sa-
leads to a lower storage modulus. tisfactory for use of RPUFs as insulating materials and construction
A different tendency as presented in this study was observed by components [102–104].
Formela et al. [69], who investigated the impact of spent grain and
ground tire rubber waste fillers on physico-mechanical properties of PU 4. Conclusion
composites. Addition of hybrid filler significantly increased the storage
modulus (E′) of PU and consequently the stiffness of studied composites A series of RPUFs with a different content of PP were sucsessfully
was also enhanced. They attributed this result to the presence of fillers synthesized. Addition of filler in the range of 0.1–5 wt% affects sig-
in PU foam matrix, which imposes serious limits on the mobility of nificantly influence the properties of the obtained foams. The best
polymer chains, affecting their higher stiffness. The difference between properties are of compositions with 0.1 wt% of PP. The modified foam
the results presented in previous work and the results obtained in our shows higher apparent density (38.1 kg/m3), greater mechanical
research can be connected with the interactions at the PU/filler inter- properties (0.23 MPa), comparable thermal conductivity (0.029 W/
face. We attribute the difference to the fact that inorganic fillers are m·K) and less water uptake (14.3%) as compared to the neat foam. PP
characterized by smaller particles and higher specific surface area as concentrations of 1 and 5 wt% result in a lower number of closed cells
compared to organic fillers which facilitates better adhesion between and enhanced porosity. These foams, are characterized by lower ap-
particles and RPUFs [96]. parent density (35.4 and 35.6 kg/m3 respectively for FP1 and FP5),
lower compression strength (0.22 and 0.17 MPa) and higher thermal
3.2.10. Dimensional stability and water uptake conductivity (0.030 and 0.032 W/m∙K). The analysis of the FTIR spectra
Regarding dimensional stability, in all cases the variations in the suggests that increasing the content of the filler in polyurethane matrix
sample's dimensions after the thermal treatment do not follow a clearly results in smaller amount of urethane moieties. Compared to the FR,
identifiable trend (Table 7). Overall, the changes are negligible when modified compositions are characterized by a lower transition glass

143
S. Członka et al. Polymer Testing 68 (2018) 135–145

temperature, that is, 147 °C, 137 °C and 135 °C for foams FP0.1, FP1 and C5RA00662G.
FP5, respectively. [24] ASTM D2414-Standard Test Method for Carbon Black – Oil Absorption Number
(OAN).
Incorporation of low amounts of PP appears to be a promising to [25] ASTM D2930-Standard Test Method for Low-temperature Viscosity of Automatic
improvement mechanical and physical properties of RPUFs. The de- Transmission Fluids, (Hydraulic Fluids, and Lubricants using a Rotational
velopment of PU compositions with this natural modifier could be very Viscometer).
[26] ISO 2555-Plastics - Resins in the Liquid State or as Emulsions or Dispersions -
interesting from a technological, ecological as well as economic point of Determination of Apparent Viscosity by the Brookfield Test Method.
view, significantly increasing the range of foam applications in the in- [27] ASTM D1622-Standard Test Method for Apparent Density of Rigid Cellular
dustry. Plastics.
[28] ISO 845-Cellular Plastics and Rubbers - Determination of Apparent Density.
[29] ASTM D1621-Standard Test Method for Compressive Properties of Rigid Cellular
References Plastics.
[30] ISO 844-Preview Rigid Cellular Plastics - Determination of Compression
Properties.
[1] S. Tan, T. Abraham, D. Ference, C.W. MacOsko, Rigid polyurethane foams from a
[31] ASTM D7264-Standard Test Method for Flexural Properties of Polymer Matrix
soybean oil-based polyol, Polymer 52 (2011) 2840–2846, http://dx.doi.org/10.
Composite Materials.
1016/j.polymer.2011.04.040.
[32] ISO 178-Plastics - Determination of Flexural Properties.
[2] M.L. Pinto, Formulation, preparation and characterization of polyurethane foams
[33] ASTM D2842-Standard Test Method for Water Absorption of Rigid Cellular
lab documentation, J. Chem. Educ. 87 (2010) 212–215.
Plastics.
[3] M. Zieleniewska, M.K. Leszczyński, M. Kurańska, A. Prociak, L. Szczepkowski,
[34] ISO 2896-Rigid Cellular Plastics - Determination of Water Absorption.
M. Krzyzowska, J. Ryszkowska, Preparation and characterisation of rigid poly-
[35] ASTM D2126-Standard Test Method for Response of Rigid Cellular Plastics to
urethane foams using a rapeseed oil-based polyol, Ind. Crop. Prod. 74 (2015)
Thermal and Humid Aging.
887–897, http://dx.doi.org/10.1016/j.indcrop.2015.05.081.
[36] ISO 2796-Cellular Plastics, Rigid - Test for Dimensional Stability.
[4] M.M.A. Nikje, M. Noruzian, S.T. Moghaddam, Investigation of Fe3O4/AEAP su-
[37] M. Zhang, W. Liu, G. Li, Isolation and characterisation of collagens from the skin of
permagnetic nanoparticles on the morphological, thermal and magnetite behavior
largefin longbarbel catfish (Mystus macropterus), Food Chem. 115 (2009)
of polyurethane rigid foam nanocomposites, Polimery/Polymers 60 (2015) 26–32,
826–831, http://dx.doi.org/10.1016/j.foodchem.2009.01.006.
http://dx.doi.org/10.14314/polimery.2015.026.
[38] F.R. Voort, A.A. Ismail, J. Sedman, G. Emo, Monitoring the oxidation of edible oils
[5] D. Yan, L. Xu, C. Chen, J. Tang, X. Ji, Z. Li, Enhanced mechanical and thermal
by Fourier transform infrared spectroscopy, J. Am. Oil Chem. Soc. 71 (1994)
properties of rigid polyurethane foam composites containing graphene nanosheets
243–253, http://dx.doi.org/10.1007/BF02638049.
and carbon nanotubes, Polym. Int. 61 (2012) 1107–1114, http://dx.doi.org/10.
[39] J. Mráz, P. Šimek, D. Chvalová, H. Nohová, P. Šmigolová, Studies on the methyl
1002/pi.4188.
isocyanate adducts with globin, Chem. Biol. Interact. 148 (2004) 1–10, http://dx.
[6] Y.H. Kim, S.J. Choi, J.M. Kim, M.S. Han, W.N. Kim, K.T. Bang, Effects of orga-
doi.org/10.1016/j.cbi.2003.06.003.
noclay on the thermal insulating properties of rigid polyurethane poams blown by
[40] K. Schwetlick, R. Noack, F. Stebner, Three fundamental mechanisms of base-cat-
environmentally friendly blowing agents, Macromol. Res. 15 (2007) 676–681,
alysed reactions of lsocyanates with hydrogen-acidic compounds, J. Chem. Soc.
http://dx.doi.org/10.1007/BF03218949.
Perkin Trans. 2 (1994) 599–608.
[7] G. Harikrishnan, S.N. Singh, E. Kiesel, C.W. Macosko, Nanodispersions of carbon
[41] R.G. Arnold, J.A. Nelson, J.J. Verbanc, Recent advances in isocyanate chemistry,
nanofiber for polyurethane foaming, Polymer 51 (2010) 3349–3353, http://dx.
Chem. Rev. 57 (1957) 47–76, http://dx.doi.org/10.1021/cr50013a002.
doi.org/10.1016/j.polymer.2010.05.017.
[42] C. Guo, L. Zhou, J. Lv, Effects of expandable graphite and modified ammonium
[8] G. Harikrishnan, T. Umasankar Patro, D.V. Khakhar, Polyurethane Foam−Clay
polyphosphate on the flame-retardant and mechanical properties of wood flour-
Nanocomposites: Nanoclays as Cell Openers, (2006), pp. 7126–7134, http://dx.
polypropylene composites, Polym. Polym. Compos. 21 (2013) 449–456, http://dx.
doi.org/10.1021/IE0600994.
doi.org/10.1002/app.
[9] G. Harikrishnan, C.I. Lindsay, M.A. Arunagirinathan, C.W. MacOsko, Probing
[43] T. Pingot, M. Pingot, M. Zaborski, The effect of chemical modification on me-
nanodispersions of clays for reactive foaming, ACS Appl. Mater. Interfaces 1
chanical properties of carbon black filled elastomer, Trends Colloid Interface Sci.
(2009) 1913–1918, http://dx.doi.org/10.1021/am9003123.
XXIV, Springer, Berlin, Heidelberg, 2011, pp. 143–146 https://doi.org/10.1007/
[10] A. Prociak, M. Kurańska, U. Cabulis, M. Kirpluks, Rapeseed oil as main component
978-3-642-19038-4_25.
in synthesis of bio-polyurethane-polyisocyanurate porous materials modified with
[44] M. Kuranska, A. Prociak, S. Michalowski, U. Cabulis, M. Kirpluks, Microcellulose
carbon fibers, Polym. Test. 59 (2017) 478–486, http://dx.doi.org/10.1016/j.
as a natural filler in polyurethane foams based on the biopolyol from rapeseed oil,
polymertesting.2017.03.006.
Polimery/Polymers 61 (2016) 625–632, http://dx.doi.org/10.14314/polimery.
[11] X. Zhou, M.M. Sain, K. Oksman, Semi-rigid biopolyurethane foams based on palm-
2016.625.
oil polyol and reinforced with cellulose nanocrystals, Compos. Part A Appl. Sci.
[45] M. Safouane, A. Saint-Jalmes, V. Bergeron, D. Langevin, Viscosity effects in foam
Manuf 83 (2016) 56–62, http://dx.doi.org/10.1016/j.compositesa.2015.06.008.
drainage: Newtonian and non-Newtonian foaming fluids, Eur. Phys. J. E. 19
[12] A. Paberza, U. Cabulis, A. Arshanitsa, Wheat straw lignin as filler for rigid poly-
(2006) 195–202, http://dx.doi.org/10.1140/epje/e2006-00025-4.
urethane foams on the basis of tall oil amide, Polimery/Polymers 59 (2014)
[46] H. Al-Moameri, Y. Zhao, R. Ghoreishi, G.J. Suppes, Simulation blowing agent
477–481, http://dx.doi.org/10.14314/polimery.2014.477.
performance, cell morphology, and cell pressure in rigid polyurethane foams, Ind.
[13] FAO, Food and Agriculture Organization of the United Nations Statistical
Eng. Chem. Res. 55 (2016) 2336–2344, http://dx.doi.org/10.1021/acs.iecr.
Programme of Work, (2017) http://www.fao.org/3/a-i4045e.pdf.
5b04711.
[14] A. Waglay, S. Karboune, I. Alli, Potato protein isolates: recovery and character-
[47] H. Fan, A. Tekeei, G.J. Suppes, F.-H. Hsieh, Rigid polyurethane foams made from
ization of their properties, Food Chem. 142 (2014) 373–382, http://dx.doi.org/10.
high viscosity soy-polyols, J. Appl. Polym. Sci. 127 (2013) 1623–1629, http://dx.
1016/j.foodchem.2013.07.060.
doi.org/10.1002/app.37508.
[15] M. Ulbrich, I. Wiesner, E. Flöter, Molecular characterization of acid-thinned
[48] M. Kuranska, A. Prociak, Porous polyurethane composites with natural fibres,
wheat, potato and pea starches and correlation to gel properties, Starch - Stärke 67
Compos. Sci. Technol. 72 (2012) 299–304, http://dx.doi.org/10.1016/j.
(2015) 424–437, http://dx.doi.org/10.1002/star.201400233.
compscitech.2011.11.016.
[16] A. Parker, J.J. Marcinko, Patent No.: US8623931B2, PROTEIN-containing FOAMS,
[49] K. Formela, A. Hejna, Ł. Zedler, M. Przybysz, J. Ryl, M.R. Saeb, Ł. Piszczyk,
MANUFACTURE and USE THEREOF, (2014).
Structural, thermal and physico-mechanical properties of polyurethane/brewers'
[17] C. Reviews, F. Science, Thermoplastic Processing of Proteins for Film Formation —
spent grain composite foams modified with ground tire rubber, Ind. Crop. Prod.
a Review vol. 73, (2008), pp. 30–39, http://dx.doi.org/10.1111/j.1750-3841.
108 (2017) 844–852, http://dx.doi.org/10.1016/j.indcrop.2017.07.047.
2007.00636.x.
[50] M.C. Silva, J.A. Takahashi, D. Chaussy, M.N. Belgacem, G.G. Silva, Composites of
[18] N. Reddy, Y. Yang, Thermoplastic films from plant proteins, J. Appl. Polym. Sci.
rigid polyurethane foam and cellulose fiber residue, J. Appl. Polym. Sci. 117
130 (2013) 729–736, http://dx.doi.org/10.1002/app.39481.
(2010) 3665–3672, http://dx.doi.org/10.1002/app.32281.
[19] W. Shi, M.J. Dumont, Review: bio-based films from zein, keratin, pea, and rape-
[51] Z.L. Song, L.Q. Ma, Z.J. Wu, D.P. He, Effects of viscosity on cellular structure of
seed protein feedstocks, J. Mater. Sci. 49 (2014) 1915–1930, http://dx.doi.org/10.
foamed aluminum in foaming process, J. Mater. Sci. 35 (2000) 15–20, http://dx.
1007/s10853-013-7933-1.
doi.org/10.1023/A:1004715926692.
[20] Y. Song, Q. Zheng, Ecomaterials based on food proteins and polysaccharides,
[52] J. John, M. Bhattacharya, R.B. Turner, Characterization of polyurethane foams
Polym. Rev. 54 (2014) 514–571, http://dx.doi.org/10.1080/15583724.2014.
from soybean oil, J. Appl. Polym. Sci. 86 (2002) 3097–3107, http://dx.doi.org/10.
887097.
1002/app.11322.
[21] J. Taylor, J.O. Anyango, J.R.N. Taylor, Developments in the science of zein, ka-
[53] J.F. Jin, Y.L. Chen, D.N. Wang, C.P. Hu, S. Zhu, L. Vanoverloop, D. Randall,
firin, and gluten protein bioplastic materials, Cereal Chem. 90 (2013) 344–357,
Structures and physical properties of rigid polyurethane foam prepared with rosin-
http://dx.doi.org/10.1094/CCHEM-12-12-0165-IA.
based polyol, J. Appl. Polym. Sci. 84 (2002) 598–604, http://dx.doi.org/10.1002/
[22] Y. Du, F. Chen, Y. Zhang, C. Rempel, M.R. Thompson, Q. Liu, Potato protein
app.10312.
isolate-based biopolymers, J. Appl. Polym. Sci. 132 (2015) 1–7, http://dx.doi.org/
[54] L.J. Lee, C. Zeng, X. Cao, X. Han, J. Shen, G. Xu, Polymer nanocomposite foams,
10.1002/app.42723.
Compos. Sci. Technol. 65 (2005) 2344–2363, http://dx.doi.org/10.1016/j.
[23] W.R. Newson, F. Rasheed, R. Kuktaite, M.S. Hedenqvist, M. Gällstedt, T.S. Plivelic,
compscitech.2005.06.016.
E. Johansson, Commercial potato protein concentrate as a novel source for ther-
[55] A. Prociak, L. Szczepkowski, M. Zieleniewska, J. Ryszkowska, Biobased poly-
moformed bio-based plastic films with unusual polymerisation and tensile prop-
urethane foams modified with natural, Polimery 60 (2015) 592–599, http://dx.
erties, RSC Adv. 5 (2015) 32217–32226, http://dx.doi.org/10.1039/
doi.org/10.14314/polimery.2015.592.

144
S. Członka et al. Polymer Testing 68 (2018) 135–145

[56] A. Arshanitsa, A. Paberza, L. Vevere, U. Cabulis, G. Telysheva, Two approaches for org/10.1016/j.compscitech.2012.11.014.
introduction of wheat straw lignin into rigid polyurethane foams, AIP Conf. Proc. [82] L. Madaleno, R. Pyrz, A. Crosky, L.R. Jensen, J.C.M. Rauhe, V. Dolomanova,
1593 (2014) 388–391, http://dx.doi.org/10.1063/1.4873806. A.M.M.V. De Barros Timmons, J.J. Cruz Pinto, J. Norman, Processing and char-
[57] V. Dolomanova, C.M.R. Jens, L.R. Jensen, R. Pyrz, A.B. Timmons, Mechanical acterization of polyurethane nanocomposite foam reinforced with montmor-
properties and morphology of nano-reinforced rigid PU foam, J. Cell. Plast. 47 illonite-carbon nanotube hybrids, Compos. Part A Appl. Sci. Manuf 44 (2013) 1–7,
(2011) 81–93, http://dx.doi.org/10.1177/0021955X10392200. http://dx.doi.org/10.1016/j.compositesa.2012.08.015.
[58] D. Niyogi, R. Kumar, K.S. Gandhi, Water blown free rise polyurethane foams, [83] M.C. Saha, M.E. Kabir, S. Jeelani, Enhancement in thermal and mechanical
Polym. Eng. Sci. 39 (1999) 199–209, http://dx.doi.org/10.1002/pen.11408. properties of polyurethane foam infused with nanoparticles, Mater. Sci. Eng. 479
[59] A.A. Septevani, D.A.C. Evans, P.K. Annamalai, D.J. Martin, The use of cellulose (2008) 213–222, http://dx.doi.org/10.1016/j.msea.2007.06.060.
nanocrystals to enhance the thermal insulation properties and sustainability of [84] E. Ciecierska, M. Jurczyk-Kowalska, P. Bazarnik, M. Gloc, M. Kulesza,
rigid polyurethane foam, Ind, Crop Prod. 107 (2017) 114–121, http://dx.doi.org/ M. Kowalski, S. Krauze, M. Lewandowska, Flammability, mechanical properties
10.1016/j.indcrop.2017.05.039. and structure of rigid polyurethane foams with different types of carbon reinfor-
[60] J.W. Kang, J.M. Kim, M.S. Kim, Y.H. Kim, W.N. Kim, W. Jang, D.S. Shin, Effects of cing materials, Compos. Struct. 140 (2016) 67–76, http://dx.doi.org/10.1016/j.
nucleating agents on the morphological, mechanical and thermal insulating rro- compstruct.2015.12.022.
perties of rigid polyurethane foams, Macromol. Res. 17 (2009) 856–862. [85] S. Matsumura, A.R. Hlil, C. Lepiller, J. Gaudet, D. Guay, Z. Shi, S. Holdcroft,
[61] R. Gu, S. Konar, M. Sain, Preparation and characterization of sustainable poly- A.S. Hay, Ionomers for proton exchange membrane fuel cells with sulfonic acid
urethane foams from soybean oils, JAOCS, J. Am. Oil Chem. Soc. 89 (2012) groups on the end-groups: novel branched poly(ether-ketone)s, Am. Chem. Soc.
2103–2111, http://dx.doi.org/10.1007/s11746-012-2109-8. Polym. Prepr. Div. Polym. Chem. 49 (2008) 511–512, http://dx.doi.org/10.1002/
[62] M. Nar, C. Webber, N.A. D'Souza, Rigid polyurethane and kenaf core composite pola.
foams mangesh, polym, Eng. Sci. 55 (2015) 132–144, http://dx.doi.org/10.1002/ [86] S. Chuayjuljit, T. Sangpakdee, Processing and properties of palm oil-based rigid
pen.23868. polyurethane foam, J. Met. Mater. Miner 17 (2007) 17–23.
[63] M.C. Hawkins, B. O'Toole, D. Jackovich, Cell morphology and mechanical prop- [87] L. Ugarte, S. Gómez-Fernández, C. Peña-Rodríuez, A. Prociak, M.A. Corcuera,
erties of rigid polyurethane foam, J. Cell. Plast. 41 (2005) 267–285, http://dx.doi. A. Eceiza, Tailoring mechanical properties of rigid polyurethane foams by sorbitol
org/10.1177/0021955X05053525. and corn derived biopolyol mixtures, ACS Sustain. Chem. Eng. 3 (2015)
[64] G. Sung, J.H. Kim, Influence of filler surface characteristics on morphological, 3382–3387, http://dx.doi.org/10.1021/acssuschemeng.5b01094.
physical, acoustic properties of polyurethane composite foams filled with in- [88] S. Ibrahim, A. Ahmad, N.S. Mohamed, Characterization of novel castor oil-based
organic fillers, Compos. Sci. Technol. 146 (2017) 147–154, http://dx.doi.org/10. polyurethane polymer electrolytes, Polymers 7 (2015) 747–759, http://dx.doi.
1016/j.compscitech.2017.04.029. org/10.3390/polym7040747.
[65] A. Viksne, A.K. Bledzki*, L. Rence, R. Berzina, C. Materials, water uptake and [89] U. Stirna, I. Beverte, V. Yakushin, U. Cabulis, Mechanical properties of rigid
mechanical characteristics of wood fiber-polypropylene composites, Polymer 42 polyurethane foams at room and cryogenic temperatures, J. Cell. Plast. 47 (2011)
(2006) 73–82. 337–355, http://dx.doi.org/10.1177/0021955X11398381.
[66] T. Widya, C. Macosko, Nanoclay-modified rigid polyurethane foam, J. Macromol. [90] R. Samborska-Skowron, A. Balas, Jakościowa identyfikacja pielścieni izocyjanur-
Sci. Part B Phys. 44 (2005) 897–908, http://dx.doi.org/10.1080/ owych w elastomelach uretanowo-izocyianurowych i w ich hydrolizatach,
00222340500364809. Polimery 48 (2003) 371–374.
[67] S. Zhang, A. Xiang, H. Tian, A.V. Rajulu, Water-blown Castor oil-based poly- [91] D.H. Park, G.P. Park, S.H. Kim, W.N. Kim, Effects of isocyanate index and en-
urethane foams with soy protein as a reactive reinforcing filler, J. Polym. Environ. vironmentally-friendly blowing agents on the morphological, mechanical, and
0 (2016) 1–8, http://dx.doi.org/10.1007/s10924-016-0914-0. thermal insulating properties of polyisocyanurate-polyurethane foams, Macromol.
[68] A. Wolska, M. Goździkiewicz, J. Ryszkowska, Thermal and mechanical behaviour Res. 21 (2013) 852–859, http://dx.doi.org/10.1007/s13233-013-1106-6.
of flexible polyurethane foams modified with graphite and phosphorous fillers, J. [92] L. Jiao, H. Xiao, Q. Wang, J. Sun, Thermal degradation characteristics of rigid
Mater. Sci. 47 (2012) 5627–5634, http://dx.doi.org/10.1007/s10853-012-6433-z. polyurethane foam and the volatile products analysis with TG-FTIR-MS, Polym.
[69] K. Formela, A. Hejna, Ł. Zedler, M. Przybysz, J. Ryl, M.R. Saeb, Ł. Piszczyk, Degrad. Stabil. 98 (2013) 2687–2696, http://dx.doi.org/10.1016/j.
Structural, thermal and physico-mechanical properties of polyurethane/brewers' polymdegradstab.2013.09.032.
spent grain composite foams modified with ground tire rubber, Ind. Crop. Prod. [93] S.V. Levchik, E.D. Weil, Thermal decomposition, combustion and fire-retardancy
108 (2017) 844–852, http://dx.doi.org/10.1016/j.indcrop.2017.07.047. of polyurethanes - a review of the recent literature, Polym. Int. 53 (2004)
[70] S. Estravís, J. Tirado-Mediavilla, M. Santiago-Calvo, J.L. Ruiz-Herrero, 1585–1610, http://dx.doi.org/10.1002/pi.1314.
F. Villafañe, M.Á. Rodríguez-Pérez, Rigid polyurethane foams with infused na- [94] D.K. Chattopadhyay, D.C. Webster, Thermal stability and flame retardancy of
noclays: relationship between cellular structure and thermal conductivity, Eur. polyurethanes, Prog. Polym. Sci. 34 (2009) 1068–1133, http://dx.doi.org/10.
Polym. J. 80 (2016) 1–15, http://dx.doi.org/10.1016/j.eurpolymj.2016.04.026. 1016/j.progpolymsci.2009.06.002.
[71] M.A. Mosiewicki, G.A. Dell'Arciprete, M.I. Aranguren, N.E. Marcovich, [95] M. Pommet, M.H. Morel, A. Redl, S. Guilbert, Aggregation and degradation of
Polyurethane foams obtained from castor oil-based polyol and filled with wood plasticized wheat gluten during thermo-mechanical treatments, as monitored by
flour, J. Compos. Mater. 43 (2009) 3057–3072, http://dx.doi.org/10.1177/ rheological and biochemical changes, Polymer 45 (2004) 6853–6860, http://dx.
0021998309345342. doi.org/10.1016/j.polymer.2004.07.076.
[72] G. Woods, The ICI Polyurethanes Book, second ed., John Wiley & Sons, New York, [96] M. Hoseinabadi, M. Naderi, M. Najafi, S. Motahari, M. Shokri, A study of rigid
1990, http://dx.doi.org/10.1002/pola.1991.080291220. polyurethane foams: the effect of synthesized polyols and nanoporous graphene, J.
[73] S. Seo, S. Karboune, A. Archelas, Production and characterisation of potato pa- Appl. Polym. Sci. 134 (2017) 1–6, http://dx.doi.org/10.1002/app.45001.
tatin–galactose, galactooligosaccharides, and galactan conjugates of great poten- [97] A.A. Septevani, D.A.C. Evans, C. Chaleat, D.J. Martin, P.K. Annamalai, A sys-
tial as functional ingredients, Food Chem. 158 (2014) 480–489, http://dx.doi.org/ tematic study substituting polyether polyol with palm kernel oil based polyester
10.1016/j.foodchem.2014.02.141. polyol in rigid polyurethane foam, Ind. Crop. Prod. 66 (2015) 16–26, http://dx.
[74] A.R. Hamilton, O.T. Thomsen, L.A.O. Madaleno, L.R. Jensen, J.C.M. Rauhe, doi.org/10.1016/j.indcrop.2014.11.053.
R. Pyrz, Evaluation of the anisotropic mechanical properties of reinforced poly- [98] K.H. Badri, S.H. Ahmad, S. Zakaria, ProductionofaHigh-functionali-
urethane foams, Compos. Sci. Technol. 87 (2013) 210–217, http://dx.doi.org/10. tyrbdpalmkerneloil-basedpolyesterpolyol, J. Appl. Polym. Sci. 81 (2001) 384–389,
1016/j.compscitech.2013.08.013. http://dx.doi.org/10.1002/app.1449.
[75] M. Modesti, A. Lorenzetti, Improvement on fire behaviour of water blown PIR-PUR [99] A.M. Atta, W. Brostow, T. Datashvili, R.A. El-Ghazawy, H.E.H. Lobland,
foams: use of an halogen-free flame retardant, Eur. Polym. J. 39 (2003) 263–268, A.R.M. Hasan, J.M. Perez, Porous polyurethane foams based on recycled poly
http://dx.doi.org/10.1016/S0014-3057(02)00198-2. (ethylene terephthalate) for oil sorption, Polym. Int. 62 (2013) 116–126, http://
[76] G. Gaidukova, A. Ivdre, A. Fridrihsone, A. Verovkins, U. Cabulis, S. Gaidukovs, dx.doi.org/10.1002/pi.4325.
Polyurethane rigid foams obtained from polyols containing bio-based and recycled [100] A. Kairyte, S. Vejelis, Evaluation of forming mixture composition impact on
components and functional additives, Ind. Crop. Prod. 102 (2017) 133–143, properties of water blown rigid polyurethane (PUR) foam from rapeseed oil
http://dx.doi.org/10.1016/j.indcrop.2017.03.024. polyol, Ind. Crop. Prod. 66 (2015) 210–215, http://dx.doi.org/10.1016/j.indcrop.
[77] E. Linul, L. Marsavina, Assesment of sandwich beams with rigid polyurethane 2014.12.032.
foam core using failure-mode maps, Proc. Rom. Acad. 16 (2015) 522–530. [101] M. Thirumal, D. Khastgir, N.K. Singha, B.S. Manjunath, Y.P. Naik, Mechanical,
[78] R.R. Maharsia, H.D. Jerro, Enhancing tensile strength and toughness in syntactic morphological and thermal properties of rigid polyurethane foam: effect of the
foams through nanoclay reinforcement, Mater. Sci. Eng. 454–455 (2007) 416–422, fillers, Cell. Polym. 26 (2007) 245–259.
http://dx.doi.org/10.1016/j.msea.2006.11.121. [102] Insta-panelstm. Polyurethane Faced Board Insulation, Technical Product Data,
[79] S. Nam, A.N. Netravali, Characterization of ramie fiber/soy protein concentrate (2012) Retrieved September 2013 from, (n.d.) http://www.instapanels.ca/wp-
(SPC) resin interface, J. Adhes. Sci. Technol. 18 (2004) 1063–1076, http://dx.doi. content/uploads/site.
org/10.1163/1568561041257504. [103] ENERLAB. Polyisocyanurate Foam Boards 2012. Retrieved September 2013 from,
[80] J.-K. Kim, Y.-W. Mai, High strength, high fracture toughness fibre composites with (n.d.) http://www.enerlab.ca/vw/fd/CartierECOANG.pdf/$fil.
interface control - a review, Compos. Sci. Technol. 41 (1991) 333–378, http://dx. [104] BASF.Highlyefficient Thermalinsulationwithpolyurethane-therightchoice for
doi.org/10.1016/0266-3538(91)90072-W. Structural Insulation Panels 2010. Retrieved September 2013 from, (n.d.) www.
[81] M. Kurańska, P. Aleksander, K. Mikelis, C. Ugis, Porous polyurethane composites polyurethanes.basf.de/pu/solutions/en.
based on bio-components, Compos. Sci. Technol. 75 (2013) 70–76, http://dx.doi.

145

You might also like