You are on page 1of 6

Radiation Physics and Chemistry 142 (2018) 94–99

Contents lists available at ScienceDirect

Radiation Physics and Chemistry


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

Polyurethane acrylate networks including cellulose nanocrystals: a MARK


comparison between UV and EB- curing
K. Furtak-Wronaa, P. Kozik-Ostrówkaa, K. Jadwiszczaka, J.E. Maigretb, V. Aguié-Béghinb,

X. Coquereta,
a
CNRS UMR 7312, Université de Reims Champagne-Ardenne, Institut de Chimie Moléculaire de Reims, BP 1039, 51687 Reims, France
b
INRA, UMR FARE 614 Fractionnement des Agro-Ressources et Environnement INRA/URCA, BP 224, 51686 Reims, France

A R T I C L E I N F O A B S T R A C T

Keywords: A water-based polyurethane (PUR) acrylate water emulsion was selected as a radiation curable matrix for
Cellulose nanocrystals preparing nanocomposites including cellulose nanocrystals (CNC) prepared by controlled hydrolysis of Ramie
Nanocomposites fibers. Cross-linking polymerization of samples prepared in the form of films or of 1 mm-thick bars was either
Polyurethane networks initiated by exposure to the 395 nm light of a high intensity LED lamp or by treatment with low energy electron
UV curing, EB-curing
beam (EB). The conversion level of acrylate functions in samples submitted to increasing radiation doses was
Thermal and mechanical properties of PUR
networks
monitored by Fourier Transform Infrared Spectroscopy (FTIR). Differential Scanning Calorimetry (DSC) and
Dynamic Mechanical Analysis (DMA) were used to characterize changes in the glass transition temperature of
the PUR-CNC nanocomposites as a function of acrylate conversion and of CNC content. Micromechanical testing
indicates the positive effect of 1 wt% CNC on Young's modulus and on the tensile strength at break (σ) of cured
nanocomposites. The presence of CNC in the PUR acrylate matrix was shown to double the σ value of the
nanocomposite cured to an acrylate conversion level of 85% by treatment with a 25 kGy dose under EB, whereas
no increase of σ was observed in UV-cured samples exhibiting the same acrylate conversion level. The occurrence
of grafting reactions inducing covalent linkages between the polysaccharide nanofiller and the PUR acrylate
matrix during the EB treatment is advanced as an explanation to account for the improvement observed in
samples cured under ionizing radiation.

1. Introduction exploit the unique properties of cellulose nanocrystals (CNCs) (Veigel


et al., 2014) as reinforcing additives in polyurethane acrylate materials.
Nanocomposites form a new class of materials with enhanced The recent developments in water-based polyurethane acrylates
physical, mechanical or use properties that continues to receive (Kozakiewicz, 2015, 2016; Xu et al., 2012) offer an interesting
increasing interest (Ansari et al., 2015; Barari et al., 2016; Dufresne, opportunity to introduce CNC in the form of aqueous suspension into
2013; Lee et al., 2014; Samir et al., 2005; Siqueira et al., 2010). the emulsion of matrix precursors. Using water as a common phase for
Bionanocomposites can be defined as a subclass of nanocomposites polymer and CNC mixing is expected to mitigate the risks for
either comprising a bio-based matrix with inorganic or synthetic uncontrolled CNC aggregation. Such a procedure avoids mixing
nanofillers (e.g. plasticized starch blended with nanoclays (Prakobna freeze-dried CNCs with molten or dissolved polymers that is known
et al., 2015)) or a non bio-based matrix including nanoparticles to induce the formation of nanocrystal bundles in nanocomposites
obtained from biomass (e. g. synthetic thermoplastic with dispersed (Khoshkava and Kamal, 2014; Santamaria-Echart et al., 2016).
cellulose nanowhiskers) (Ozgur Seydibeyoglu and Oksman, 2008; Yao Controlled acid hydrolysis of purified cellulose yields CNCs in the
et al., 2014)), or a mixture of both bio-based matrix and nanoparticle form of rod-like nanoparticles with a section in the range of 3–25 nm
components (e. g. poly (lactic acid) with dispersed cellulose microfibrils and lengths of a few hundred nanometers to 1–2 µm. Their size,
and nanowhiskers) (Fujisawa et al., 2013; Miao and Hamad, 2016)). dimension, as well as their shape depend primarily on the origin of
As a part of our current research on radiation-cured materials with cellulose source (Miriam de Souza Lima and Borsali, 2004), but also on
improved physical and mechanical properties (Pichavant and Coqueret, the conditions of hydrolytic preparation, such as the nature and
2008; Mougharbel et al., 2009), we have examined the possibility to concentration of the acid, reaction time and temperature, purity of


Corresponding author.
E-mail address: xavier.coqueret@univ-reims.fr (X. Coqueret).

http://dx.doi.org/10.1016/j.radphyschem.2017.04.013
Received 10 November 2016; Accepted 21 April 2017
Available online 22 April 2017
0969-806X/ © 2017 Elsevier Ltd. All rights reserved.
K. Furtak-Wrona et al. Radiation Physics and Chemistry 142 (2018) 94–99

the starting material, and ultrasonic treatment. The relative degree of 395 nm, 8 W cm−2) or with a 150 kV EB laboratory accelerator
crystallinity and the geometrical aspect ratio (length/cross-section) are operated at a typical dose rate of 8 kGy s−1 (Advanced Electron
the main parameters that control the properties of CNC-based materials Beam, Wilmington, USA). Radiation curing of thicker samples (1 mm-
(Lu and Hsieh, 2010; Siqueira et al., 2009). thick bars) was performed at IONISOS (Chaumesnil, France) with a
A recent article (Poaty et al., 2014) reports on the use of CNCs 10 MeV Linac electron accelerator at room temperature and under air
chemically modified for UV-curable wood coating formulations. In this (25 kGy or 2×25 kGy doses, at an average dose rate of 15 kGy s−1).
work, CNCs were acrylated for inducing the formation of covalent
bonds between the nanofiller and the host PUR acrylate matrix upon
2.4.1. Characterization of radiation-cured samples
UV-initiated polymerization. Gloss and haze measurements have shown
Determination of acrylate conversion – Fourier Transform Infrared
that the modified nanocellulose fibers ensure maintaining of esthetic
(FTIR) analyses were performed with a Vertex 70 spectrometer (Bruker,
properties of coatings and impart enhanced abrasion resistance of the
Germany). Polymerization monitoring was performed by Mid-InfraRed
coating at CNC contents as low as 2 wt% in the cured matrix.
spectroscopy (MIR) for thin films (10–50 µm) coated on a Si wafer or as
Our work was aimed at comparing the influence of the irradiation
free-standing films, and by near-infrared (NIR) for free-standing
method on the properties of cured PUR acrylate networks comprising
samples with the thickness up to 1 mm. Spectra were recorded using
small amounts (typically 1–5 wt%) of CNCs. The potential benefits of
32 scans at a resolution of 4 cm−1 from 400 to 4000 cm−1 and from
electron beam radiation in comparison with UV–visible was a central
2000 to 7000 cm−1, respectively for MIR and NIR. The degree of
issue worth to be addressed. We have selected CNCs obtained from
conversion (π) was calculated by the following formula:
Ramie as bio-based reinforcing nanofillers with an aspect ratio about to
30:1 (Aguié-Béghin et al., 2016) for enhancing the mechanical proper- ⎛ Aacrylate × A 0 ⎞
π = ⎜⎜1− ⎟⎟ × 100%
ref
ties of radiation-cured materials. 0
⎝ Aref × Aacrylate ⎠ (1)
Because of the formation of sulfate ester groups during the sulfuric
acid hydrolysis of cellulose, the obtained CNCs are negatively charged where: Aacrylate is the absorbance for acrylate peak for irradiated sample
(Li and Ragauskas, 2011). Consequently, the nanoparticles are expected (at 810 cm−1 for MIR or at 6164 cm−1 for NIR), Aref 0
is the absorbance
to exhibit improved stability in aqueous solution that facilitates their for reference peak for irradiated sample, Aref corresponds to the
blending with the water-based PUR acrylate emulsions. absorbance for reference peak for non-irradiated sample and Aacrylate 0

to the absorbance for acrylate band for non-irradiated sample (at


2. Experimental 810 cm−1 for MIR analyses or at 6164 cm−1 in the case of NIR).
Thermo-physical properties – Differential Scanning Calorimetry
2.1. Materials (DSC) analysis was carried out with a DSC Q100 differential calorimeter
(TA Instruments, USA) fitted with a liquid nitrogen cooling system.
The polyurethane acrylate dispersion (Bayhydrol UV 2282, aqueous Aluminum pans filled with the water-based formulations were heated at
emulsion with 39 wt% of dry matter) was received from Bayer 55 °C on a heating plate during 8 h for allowing extensive water
(Material Science), France. The photoinitiator (PI) used for UV-curing evaporation. The samples were irradiated with dose 25 kGy or
experiments was 2-hydroxy-2-methyl-1-phenyl-propan-1-one (Darocur 2×25 kGy. The heating and cooling rates were 5 °C min−1 from
1173) obtained from BASF. −50 °C to 150 °C to examine the effect of CNC content on the thermal
properties (particularly the glass transition temperature, Tg) of the
2.2. CNC preparation nanocomposite films.
Dynamic Mechanical Analysis (DMA) – Rheometric characterization
CNC were prepared from Ramie (Boehmeria nivea). The washed of nanocomposite films was carried out with a DMA Q800 equipment
fibers were cut into small pieces and treated with 2 wt% NaOH at 20 °C (TA Instruments, USA) working in tensile mode. The measurements
for 48 h to remove hemicelluloses, traces of pectin and residual were performed at a frequency of 1 Hz, with a strain amplitude of
proteins. The obtained suspension was hydrolyzed overnight with 0.05%, in the temperature range from −80 °C to 200 °C and a heating
65 wt% H2SO4 at 35 °C with mechanical stirring. After washing with rate of 5 °C min−1. The samples were prepared by cutting strips from
water until neutrality and dialysis with 6000 Da cut-off membranes, the the composite films (5×25 mm2 samples of thickness 200–300 µm).
purified aqueous colloidal suspension contains 3% (w/w) dry matter of
CNC. The stock suspension was stored at 4 °C and slightly sonicated
2.4.2. Measurements of the tensile properties of cured materials
with a Sonics vibra-cell (750 W, Fisher-Bioblock) before use
The mechanical behavior for the unfilled matrix and the nanocom-
(Hambardzumyan et al., 2012).
posites was analyzed with a micromechanical testing instrument
(Deben Ltd, UK). Experiments were performed with the elongation
2.3. Preparation of films and bars
speed of 0.4 mm min−1 on a total displacement of 20 mm at room
temperature. Samples (width 5 mm, length 25 mm) were prepared by
CNC suspension (1 wt% with the respect to the dry content of the
cutting strips from the nanocomposite films. The results were the
PUR acrylate dispersion) was added to the PUR acrylate emulsion and
average of at least five measurements.
stirred for 16 h with magnetic stirrer. The photoinitiator (1 wt% with
the respect to the dry content of the PUR acrylate dispersion) was added
to the formulations to be cured under UV–visible light. Formulations 3. Results and discussion
were cast in silicone molds and allowed to dry for 2 days at room
temperature, and for 2 more days at 25 °C. A final treatment at 50 °C The comparison between the two modes of radiation-initiated
was applied for achieving the coalescence process yielding a uniform polymerization that were selected for the present investigation requires
and transparent samples (10–50 µm-thick films or of 1 mm-thick bars) careful preparation of the nanocomposites from water-based suspen-
in the form of a free-standing material. sions and of the unfilled PUR acrylate matrix. Particular attention was
devoted to the experimental preparation of dry films and dry bars by
2.4. Radiation-induced curing evaporation and coalescence of the PUR acrylate emulsions including
the desired additives (CNC, photoinitiator for UV–visible curing
The resulting dry samples (10–50 µm-thick films) were irradiated by experiments), and to the control of the degree of curing in the samples
high intensity UV–vis LED lamp (FireLine™, Phoseon Technology submitted to physical analyses.

95
K. Furtak-Wrona et al. Radiation Physics and Chemistry 142 (2018) 94–99

Scheme 1. Typical molecular structure of an emulsible PUR acrylate and schematic of the corresponding emulsified particle in water.

3.1. Preparation of CNCs and of radiation curable formulations Firstly, higher level of nanofillers (e.g. 5–10 wt% with respect to the
PUR acrylate matrix) require introducing excessive amounts of water
The base PUR acrylate material (Scheme 1) is available as milky that may not lead to well-defined and regular nanocomposites, the large
fluid with 39 wt% of dry matter having a slightly alkaline pH (about amount of fluid likely favoring the transport of some materials and
7.5). The blends for elaborating nanocomposite materials are thus possibly inducing macroscale phase separation between the ingredients
easily prepared by adding the desired amount of CNC suspension. For a present in the complex suspensions. Secondly, we considered that the
final CNC content of 1 wt% in the PUR acrylate matrix, the volume of saturation by CNC of the interfaces between the prepolymer particles
CNC suspension to be added to a prepolymer emulsion of volume V0 undergoing coalescence may lead to a very complex microstructure that
represents about 13% of that volume, with limited impact on the drying would not allow for a good control of the system uniformity and for a
process and coalescence phenomena for obtaining free-standing mate- simple interpretation of the effects on nanocomposites mechanical
rials ready for radiation curing. Obviously, when larger amounts of CNC properties.
are desired in the nanocomposite, much greater volumes of water will
be added and have to be evaporated.
3.2. Control of the degree of curing in PUR acrylate materials by FTIR
For UV–visible curing under the 395 nm LED light, the photoini-
tiator is added and mixed under stirring at concentration of 1 part for
The mechanical properties of the uniform materials obtained after
100 parts of PUR acrylate. The low absorbance of PI at the working
drying are essentially driven by their chemical composition and by the
wavelength allows for the penetration of the light in the depth of the
degree of conversion of the acrylate functions. A detailed investigation
samples without significant gradient of photon absorption, even in
of the reactivity of different blends was conducted by FTIR spectroscopy
1 mm thick bars obtained after evaporation of the water.
(Patacz et al., 2000). MIR analyses were performed for thin films
The drying of the formulated proceeds in a sequence of steps
(50–100 µm-thick) and NIR measurements for thicker samples prepared
eventually leading to clear and homogeneous materials after extensive
in the form of plates or bars (up to 1 mm thick). Typical spectra
coalescence of the soft PUR acrylate particles (Steward et al., 2000).
recorded are reproduced in Fig. 1a) and 1 b). The progress of acrylate
The initial stage consists in the evaporation of excess water leading to
conversion as a function of radiation dose can be measured accurately
the compaction of the particles that gradually adopt a hexagonal type of
by using invariant bands as internal standards.
packing. Interdiffusion of the polymer segments is facilitated by the low
The degree of conversion of acrylate in the unfilled and in the
minimum film forming (MFF) temperature of the selected emulsion
nanocomposite materials was quantified on the basis of the changes of
(TMFF=0 °C). This step is finished when a dry film is formed as a
IR absorbance according to Eq. (1), as detailed in the experimental part.
transparent and free-standing material.
The corresponding kinetic profiles of unfilled and nanocomposite
As CNC are present in the form of rigid nanorods in the water
samples irradiated with 395 nm LED array are shown on the plots of
suspension, with low ability to diffuse inside the prepolymer particles,
Fig. 2 that reveal almost no difference in polymerization rates at the
we can reasonably assume that they are located at the interfaces of
different stage of the curing process. High levels of conversion are
hexagons during concentration stage and that they can play a beneficial
obtained within less than 1 min of exposure to UV–visible light, in spite
role by reinforcing the interstitial domain between the initially isolated
of the low absorbance of the selected PI at 395 nm. The polymerization
prepolymer particles. A schematic representation of the coalesced
rate levels off when acrylate reaches 70–80% of conversion, as the
material including CNC is shown in Scheme 2.
network starts vitrifying.
After having tested different drying conditions, we have adopted the
Similar conclusions were drawn from MIR monitoring of the curing
protocol detailed in the experimental part which afforded materials
process in thin films submitted to the beam of the 150 keV laboratory
with uniform appearance. We also focused our attention to samples
accelerator, with kinetic profiles resembling those already reported
containing limited amount of 1 wt% of CNC for two main reasons.
earlier for non water-based PUR acrylates (Patacz et al., 2000). The
presence of 1 wt% CNC in the nanocomposite material was shown not
to induce any significant differences with respect to the kinetic profiles
recorded with the unfilled PUR acrylate.
For comparing the physical and mechanical properties of the
unfilled and the nanocomposite samples, we defined target values of
85 and 90 mol-% for acrylate conversion. These values correspond to
the application of doses of 25 and 2×25 kGy, respectively, for 1 mm-
thick samples exposed to the 10 MeV Linac beam.
The conditions for curing 1 mm-thick bar samples under UV–visible
light were adjusted to reach the desired acrylate conversion level that
Scheme 2. Simplified representation of the location of CNCs at the boundary between was controlled by NIR spectroscopy. The corresponding times of
PUR acrylate latex particles undergoing coalescence. exposure were typically between 40 and 180 s under the 395 nm LED

96
K. Furtak-Wrona et al. Radiation Physics and Chemistry 142 (2018) 94–99

Fig. 1. FTIR spectra for PUR acrylate including 1 wt% of CNC upon 395 nm LED irradiation (8 W cm−2) a) MIR spectra recorded from a 50 µm-thick film at t=0, 16, 50, 100, 170, 240,
310, 380, 450, 520, 610, 700 s); b) NIR spectra recorded from a 1 mm-thick bar at t=0, 5, 15, 30, 60, 105, 165, 225 s).

100 the maximum of the loss factor tan δ between unfilled and nanocom-
posite samples cured under the same conditions at a given level of
80 acrylate conversion. More visible changes were observed by comparing
Conversion (%)

the Tg of samples cured under different radiation sources, and/or at


different doses. The lowest Tg were obtained for samples irradiated with
60
LED lamp, with Tg values ranging from 70 °C to 72 °C. Higher Tg values
were observed for the EB-cured materials, in the range of 73–78 °C after
40 treatment at 25 kGy, and up to about 80 °C for a 2×25 kGy dose,
regardless of the presence of CNC in PUR acrylate material. We can
20 conclude to absence of noticeable effect of CNC's on the relaxation
behavior of the cured networks. Another series of experiments con-
0 ducted with higher levels of CNCs (up to 5 wt%) incorporated in PUR
0 50 100 150 200 250 emulsions dried in different conditions (in the presence of ethanol, film
Irradiation time (s) formation at higher temperatures) and processed under EB to doses as
high as 200 kGy confirms the very small influence of the presence of
Fig. 2. Comparison of the kinetic profiles of acrylate conversion in thin films, unfilled (◻) CNC in samples cured under the same conditions.
and in nanocomposite including 1 wt% of CNC (○) upon UV–visible curing with a 395 nm
LED (1 wt% of PI, 8 W cm−2).
3.4. Mechanical properties of radiation-cured materials
array.
Tensile measurements performed with a micromechanical testing
3.3. Thermal and mechanical behaviors of nanocomposites equipment allowed for the comparison of the cured materials in terms
of Young's modulus and tensile stress at break. Because of the know
The changes of thermo-physical properties of UV- and EB-cured PUR sensitivity of ultimate tensile properties to small defects associated with
acrylate materials, either unfilled or filled with 1 wt% of CNC, were sample preparation or with the measurement itself, a large number of
characterized by DSC and by DMA. specimens were tested so as to enable significant comparison between
To evaluate the impact of the presence of CNC on thermal properties samples of different composition or with different conditions of curing.
of nanocomposites, the glass transition temperature was determined With respect to Young's modulus, the data presented in Fig. 3 reveal
from the inflexion point of the variation of heat capacity in the DSC a significant enhancement of the rigidity of radiation cured nanocom-
thermograms and from the maximum of the loss factor (tan δ) posite materials.
associated with the α relaxation appearing in the thermomechanical This behavior was present in all types nanocomposite samples,
spectra (DMA). irrespective of the irradiation source that was used. However, the
The Tg measured for acrylate prepolymers was already shown to highest enhancement of such features was observed for samples
depend strongly on the radiation dose via the induced change of exposed to the moderate EB dose of 25 kGy. The enhancement of
conversion which is the relevant descriptor of the curing process
(Krzeminski et al., 2010a, 2010b). In the case of EB irradiation of
1 mm-thick samples at a dose of 25 kGy, the measured values of Tg were
around 60–65 °C, for the unfilled and filled material prepared under
soft drying conditions. The Tg of the unfilled polymer material reached
slightly higher value when the sample was irradiated with the dose of
2×25 kGy. This is the expectable consequence of the increase of
conversion which progresses from 85% to 91%. However, the Tg of
samples containing 1 wt% of CNC submitted to 2×25 kGy dose was
maintained at the same level (Tg=64 °C) as the value measured for
samples cured with the dose of 25 kGy, in spite of a small increase of
conversion. The uncertainty on the degree of conversion as well the
possible of variable amounts of moisture from one sample to the other
incite us to be cautious on the potential effect on the influence of the Fig. 3. Young's modulus of unfilled cured PUR acrylate samples (hatched bars) or filled
presence of CNCs on the glass transition of the nanocomposites. with 1 wt% CNC materials (uniform gray bars) and submitted to EB irradiation, as
DMA analysis did not reveal significant differences in the position of determined by mechanical microtesting.

97
K. Furtak-Wrona et al. Radiation Physics and Chemistry 142 (2018) 94–99

Table 1 the polysaccharide nanofiller and the PUR acrylate matrix during the
Influence of the presence of CNC on the tensile strength of cured PUR acrylate samples EB treatment is advanced as an explanation to account for the
submitted to EB or to UV–vis LED −395 nm (PI content of 1 wt%) irradiation.
improvement observed in samples cured under ionizing radiation.
Radiation type Sample composition Radiation dose Acrylate Stress σ The benefits observed on the mechanical properties of samples contain-
(in wt-parts) conversion (MPa) ing 1 wt% of CNC cured by exposure to a 25 kGy dose were not as high
(%) when the material was submitted to 2 consecutive 25 kGy passes under
PUR PI CNC the electron beam. The loss of in the intrinsic mechanical properties of
EB 100 – – 25 kGy 85 20 ± 2 the polysaccharide nanofiller as a consequence of radiation-induced
100 – – 2×25 kGy 91 28 ± 10 degradation is believed to reduce the reinforcing effect observed at
100 – 1 25 kGy 85 39 ± 8 25 kGy. Increasing of the content of CNC above 1 wt% did not show
100 – 1 2×25 kGy 92 35 ± 1 further increasing in the values of modulus. Further work is needed to
UV–vis LED 100 1 – 200 J cm−2 89 24 ± 7 gain a deeper understanding of this phenomenon related to the
100 1 1 200 J cm−2 92 23 ± 4 uniformity of CNC dispersion in the PUR acrylate matrix.

Acknowledgement
Young's modulus was in the range of about 500 MPa. The lowest
increase in modulus (about 300 MPa) was recorded for nanocomposite
This work was conducted in the frame of the IAEA Coordinated
samples cured with EB with the dose of 2×25 kGy.
Research Project “Radiation processing of (nano)composites for enhan-
It is worth to add here that more experiments were achieved for
cing the features and utility in health care and industry”. Financial
examining the influence of the CNC content at levels up to 10 wt% of
support by Conseil Regional Grand Est, MENESR and EU-FEDER
the formulation. However, no significant increase in Young's Modulus
Programme (CPER Project PlAneT) and IAEA is gratefully acknowl-
was observed upon increasing of the content of reinforcing cellulose
edged. The authors are grateful to Dr. C. Kowandy and to Dr. G. Tataru.
nanofibers. As already discussed, the conditions of elaboration of such
KFW and PKO (UMCS, Lublin, Poland) as well as KJ (UT Lodz, Poland)
samples and the possible saturation of interstitial domains during
acknowledge the support received from the European Union exchange
coalescence cast some doubts on the relevance of measurements
program Erasmus for their mobility to URCA, France.
performed on samples with high CNC contents.
The ultimate tensile properties for UV- and EB-cured samples are
References
gathered in Table 1. Comparison of the data interestingly reveals
contrasting effects.
Aguié-Béghin, V., Paës, G., Molinari, M., Chabbert, B., 2016. Films and coatings from
An efficient reinforcing effect associated with the presence of 1 wt% lignocellulosic polymers. In: Montero, M.P., Gomez-Guillen, M.C., Lopez-Caballero,
of CNC is observed in EB-cured samples. The samples cured by M.E., Barbosa-Canovas, G.V. (Eds.), Edible Films and Coatings: Fundamentals and
photopolymerization do not benefit from the presence of CNC. This Applications. CRC Press, Boca Raton.
Ansari, F., Salajkova, M., Zhou, Q., Berglund, L.A., 2015. Strong surface treatment effects
phenomenon can be explained by the limited radiolysis of the cellulose on reinforcement efficiency in biocomposites based on cellulose nanocrystals in poly
in EB-cured samples. Additionally, a covalent coupling between the (vinyl acetate) matrix. Biomacromolecules 16, 3916–3924.
nanofiller and the matrix could be caused by free radical grafting Barari, B., Omranib, E., Moghadamb, A.D., Menezesc, P.L., Pillaia, K.M., Rohatgi, P.K.,
2016. Mechanical, physical and tribological characterization of nano-cellulose fibers
reaction at the interface between the cellulosic particles and the PUR reinforced bio-epoxy composites: an attempt to fabricate and scale the ‘Green’
acrylate matrix. During the irradiation with electron beam with the composite. Carbohydr. Polym. 147, 282–293.
dose of 25 kGy the degradation of cellulose by chain scission was Dufresne, A., 2013. Nanocellulose: a new ageless bionanomaterial. Mater. Today 16,
220–227.
expected not to alter dramatically the polysaccharide structure.
Fujisawa, S., Saito, T., Kimura, S., Iwata, T., Isogai, A., 2013. Surface engineering of
Samples submitted to 2 passes of 25 kGy gained a smaller improvement ultrafine cellulose nanofibrils toward polymer nanocomposite materials.
of mechanical properties when comparing with samples irradiated with Biomacromolecules 14, 1541–1546.
Hambardzumyan, A., Foulon, L., Chabbert, B., Aguié-Béghin, V., 2012. Natural organic
dose of 25 kGy. This is probably caused by some degradation of the
UV-absorbent coatings based on cellulose and lignin: designed effects on
polysaccharide chains during the exposure on EB of 2×25 kGy. For spectroscopic properties. Biomacromolecules 13, 4081–4088.
samples cured by photopolymerization, no significant enhancement of Khoshkava, V., Kamal, M.R., 2014. Effect of drying conditions on cellulose nanocrystal
mechanical properties is observed. Indeed, the selectivity of the (CNC) agglomerate porosity and dispersibility in polymer nanocomposites. Powder
Technol. 261, 288–298.
activation that only induces the photolysis of the initiator only triggers Kozakiewicz, J., 2015. Developments in aqueous polyurethane and polyurethane-acrylic
acrylate polymerization without causing significant free radical grafting dispersion technology. Part I. Polyurethane dispersions. Polimery 60, 523–600.
onto the polysaccharide nanofiller. Kozakiewicz, J., 2016. Developments in aqueous polyurethane and polyurethane-acrylic
dispersion technology. Part II. Polyurethane-acrylic dispersions and modification of
polyurethane and polyurethane-acrylic dispersions. Polimery 61, 79–156.
Krzeminski, M., Molinari, M., Troyon, M., Coqueret, X., 2010a. Calorimetric
4. Conclusions characterization of the heterogeneities produced by the radiation-induced
crosslinking polymerization of aromatic diacrylates. Macromolecules 43, 3757–3763.
CNCs available as dilute aqueous solutions can be easily introduced Krzeminski, M., Molinari, M., Troyon, M., Coqueret, X., 2010b. Characterization by
atomic force microscopy of the nanoheterogeneities produced by the radiation-
into PUR acrylate emulsions. Appropriate conditions were defined for
induced crosslinking polymerization of aromatic diacrylates. Macromolecules 43,
curing samples in the form of thin films (50–100 µm) or 1 mm-thick 8121–8127.
bars with well-controlled conversion levels that can be precisely Lee, K.-Y., Aitomäki, Y., Berglund, L.A., Oksman, K., Bismarck, A., 2014. On the use of
nanocellulose as reinforcement in polymer matrix composites. Compos. Sci. Technol.
measured by MIR or NIR spectroscopy, depending on the thickness of
105, 15–27.
the samples. Treatment under 395 nm LED light or by EB processing Li, Y., Ragauskas, A.J., 2011. Cellulose nano whiskers as a reinforcing filler in
allowed for preparing by the two initiation methods samples with the polyurethanes. Algae 75, 10–15.
same curing level. At a given conversion degree for the acrylate Lu, P., Hsieh, Y.-L., 2010. Preparation and properties of cellulose nanocrystals: rods,
spheres, and network. Carbohydr. Polym. 82, 329–336.
functions, the Tg of nanocomposite films was not significantly affected Miao, Ch, Hamad, W.Y., 2016. In-situ polymerized cellulose nanocrystals (CNC)-poly(l-
by the presence of 1 wt% of CNC. The presence of CNC in the PUR lactide) (PLLA) nanomaterials and applications in nanocomposite processing.
acrylate matrix was shown to double the σ value of the nanocomposite Carbohydr. Polym. 153, 549–558.
Miriam de Souza Lima, M., Borsali, R., 2004. Rodlike cellulose microcrystals: properties,
cured to an acrylate conversion level of 85% by treatment with a and applications. Macromol. Rapid Commun. 25, 771–787.
25 kGy dose under EB, whereas no increase of σ was observed in UV- Mougharbel, A., Mallégol, J., Coqueret, X., 2009. Diffusion behavior of isobornyl acrylate
cured samples exhibiting the same acrylate conversion level. The into photopolymerized urethane acrylate films: influence of surface oxidation during
curing. Langmuir 17, 9831–9839.
occurrence of grafting reactions inducing covalent linkages between

98
K. Furtak-Wrona et al. Radiation Physics and Chemistry 142 (2018) 94–99

Ozgur Seydibeyoglu, M., Oksman, K., 2008. Novel nanocomposites based on polyurethane nanocomposites. Carbohydr. Polym. 151, 1203–1209.
polyurethane and micro fibrillated cellulose. Compos. Sci. Technol. 68, 908–914. Siqueira, G., Bras, J., Dufresne, A., 2009. Cellulose whiskers versus microfibrils: influence
Patacz, C., Defoort, B., Coqueret, X., 2000. Electron-beam initiated polymerization of of the nature of the nanoparticle and its surface Functionalization on the thermal and
acrylate compositions 1: ftir monitoring of incremental irradiation. Radiat. Phys. mechanical properties of nanocomposites. Biomacromolecules 10, 425–432.
Chem. 59, 329–337. Siqueira, G., Bras, J., Dufresne, A., 2010. Cellulosic bionanocomposites: a review of
Pichavant, L., Coqueret, X., 2008. Optimization of a UV-curable acrylate-based protective preparation, properties and application. Polymer 2, 728–765.
coating by experimental design. Prog. Org. Coat. 63, 55–62. Steward, P.A., Hearn, J., Wilkinson, M.C. 2000. An overview of polymer latex film
Poaty, B., Vardanyan, V., Wilczak, L., Chauve, G., Riedl, B., 2014. Modification of formation and properties. Adv. Coll. Interf. Sci., 86, 195–267.
cellulose nanocrystals as reinforcing derivatives for wood coatings. Prog. Org. Coat. Veigel, S., Grüll, G., Pinkl, S., Obersriebnig, M., Müller, U., Gindl-Altmutter, W., 2014.
77, 813–820. Improving the mechanical resistance of waterborne wood coatings by adding
Prakobna, K., Galland, S., Berglund, L.A., 2015. High-performance and moisture-stable cellulose nanofibers. React. Funct. Polym. 85, 214–220.
cellulose−starch nanocomposites based on bioinspired core−shell nanofibers. Xu, H., Qiu, F., Wang, Y., Wu, W., Yang, D., Guo, Q., 2012. UV-curable waterborne
Biomacromolecules 16, 904–912. polyurethane-acrylate: preparation, characterization and properties. Prog. Org. Coat.
Samir, A., Alloin, F., Dufresne, A., 2005. Review of recent research into cellulosic 73, 47–53.
whiskers, their properties and their application in nanocomposite field. Yao, X., Qi, X., He, Y., Tan, D., Chen, F., Fu, Q., 2014. Simultaneous reinforcing and
Biomacromolecules 6, 612–626. toughening of polyurethane via grafting on the surface of microfibrillated cellulose.
Santamaria-Echart, A., Ugarte, L., García-Astrain, C., Arbelaiz, A., Corcuera, M.A., Eceiza, ACS Appl. Mater. Interfaces 6, 2497–2507.
A., 2016. Cellulose nanocrystals reinforced environmentally-friendly waterborne

99

You might also like