You are on page 1of 23

Accepted Manuscript

Title: PREPARATION AND CHARACTERIZATION OF


TRANSPARENT PMMA-CELLULOSE-BASED
NANOCOMPOSITES

Author: Esra Erbas Kiziltas Alper Kiziltas Shannon C. Bollin


Douglas J. Gardner

PII: S0144-8617(15)00232-5
DOI: http://dx.doi.org/doi:10.1016/j.carbpol.2015.03.029
Reference: CARP 9771

To appear in:

Received date: 9-11-2014


Revised date: 5-3-2015
Accepted date: 7-3-2015

Please cite this article as: Kiziltas, E. E., Kiziltas, A., Bollin, S. C., and Gardner,
D. J.,PREPARATION AND CHARACTERIZATION OF TRANSPARENT PMMA-
CELLULOSE-BASED NANOCOMPOSITES, Carbohydrate Polymers (2015),
http://dx.doi.org/10.1016/j.carbpol.2015.03.029

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
1 Highlights

2  Transparent PMMA/cellulose nanocomposites were successfully prepared by a solution


3 casting method which included a combination of ultrasonication and mechanical stirring
4 processes using acetone as the solvent.
5  Thermogravimetric analysis (TGA) indicated retained thermal stability of the transparent
6 nanocomposites.

t
7  The transmittance of PMMA/cellulose nanocomposites was reduced with increased

ip
8 loading of cellulose nanofibrils.
9  No significant differences were observed between the FTIR spectra of pure PMMA and

cr
10 cellulose nanocomposites.
11  Optical micrograph of PMMA/cellulose nanocomposite that some cellulose
12 nanomaterials might have reaggregated among themselves.

us
13  CNF can serve as promising reinforced nanomaterials for the production of PMMA
14 nanocomposites in diverse applications such as: packaging, flexible screens, optically
15 transparent films and light-weight transparent materials.
16

an
M
d
p te
ce
Ac

1
Page 1 of 22
16 PREPARATION AND CHARACTERIZATION OF TRANSPARENT PMMA-

17 CELLULOSE-BASED NANOCOMPOSITES

18
19 Esra Erbas Kiziltas1, 2†*, Alper Kiziltas1, 3 , Shannon C. Bollin4 and Douglas J. Gardner 1

t
20

ip
1
21 Advanced Structures and Composites Center, University of Maine, Maine, 04469, USA

cr
22
2

us
23 The Scientific and Technological Research Council of Turkey (TUBİTAK),

24 Tunus Cad, Kavaklıdere, Ankara, 06100, TURKEY

an
25
3
26 Department of Forest Industry Engineering, Faculty of Forestry,
M
27 University of Bartin, Bartin, 74100 TURKEY

28
d

4
29 Research and Advanced Engineering, Ford Motor Co. Dearborn, Michigan 48124, USA
te

30
p

31 †: Corresponding author.
ce

32 E-mail: esra.erbas@maine.edu

33 Tel: +1-207- 249-9346 Fax: +1-207-581-2074


Ac

34

35 *: First author.

36 E-mail: esra.erbas@maine.edu

37 Tel: +1-207- 249-9346 Fax: +1-207-581-2074

38

39

2
Page 2 of 22
40 Abstract
41
42 Nanocomposites of polymethylmethacrylate (PMMA) and cellulose were made by a
43 solution casting method using acetone as the solvent. The nanofiber networks were prepared
44 using three different types of cellulose nanofibers: (i) nanofibrillated cellulose (NFC), (ii)
45 cellulose nanocrystals (CNC) and (iii) bacterial cellulose from nata de coca (NDC). The loading
46 of cellulose nanofibrils in the PMMA varied between 0.25 and 0.5 wt. %. The mechanical

t
47 properties of the composites were evaluated using a dynamic mechanical thermal analyzer

ip
48 (DMTA). The flexural modulus of the nanocomposites reinforced with NDC at the 0.5 wt. %
49 loading level increased 23% compared to that of pure PMMA. The NFC composite also
50 exhibited a slightly increased flexural strength around 60 MPa while PMMA had a flexural

cr
51 strength of 57 MPa. The addition of NDC increased the storage modulus (11%) compared to neat
52 PMMA at room temperature while the storage modulus of PPMA/CNC nanocomposite
53 containing 0.25 and 0.5 wt.% cellulose increased about 46% and 260% to that of the pure

us
54 PMMA at the glass transition temperature, respectively. Thermogravimetric analysis (TGA)
55 indicated that there was no significant change in thermal stability of the composites. The UV-vis
56 transmittance of the CNF nanocomposites decreased by 9% and 27% with the addition of 0.25

an
57 wt. % CNC and NDC, respectively. This work is intended to spur research and development
58 activity for application of CNF reinforced PMMA nanocomposites in applications such as:
59 packaging, flexible screens, optically transparent films and light-weight transparent materials for
60 ballistic protection.
M
61
62
63 Keywords: Nanocellulose; Nanocomposites; Thermal analysis; Transparency; PMMA
d

64
65
te

66
67
68
p

69
70
ce

71
72
73
Ac

74
75
76
77
78
79
80
81
82
83

3
Page 3 of 22
84 1. Introduction
85 It is well known that cellulose, a linear homopolymer composed of β-1, 4-linked glucose
86 molecules, is the most abundant biopolymer on earth (Pecoraro et al. 2008). Characteristic
87 behavior of cellulose mostly depends on its unique structural hierarchy and its biological origin,
88 which provides excellent properties including; high mechanical properties, such as elastic
89 modulus of the crystalline region of cellulose I is ~138 GPa, biodegradability, high strength, high
90 aspect ratio, high specific surface area, low thermal expansion, and low density (Tanpichai et al.

t
91 2012; Nishino et al., 1995). The current demand in materials research is to develop

ip
92 multifunctional materials which comprise excellent features such as enhanced mechanical
93 properties and thermal stability, biodegradability, being eco-friendly, and low-cost. Thus,
94 cellulose nanomaterials from different sources have been widely used as a reinforcing phase in

cr
95 nanocomposites to produce innovative products for science, and technology especially in
96 medicine, electronics or energy production (Vitta et al. 2012).
97

us
98 Nanoscale cellulose fibers can be obtained from four different methods: (1)
99 microfibrillated/nanofibrillated cellulose plant cell fibers, (2) cellulose whiskers or cellulose
100 nanocrystals, (3) bacterial cellulose nanofibers (BC) and (4) cellulose nanofibers by

an
101 electrospinning (Gardner et al. 2008). Nanofibrillated cellulose (NFC) is typically a fibrous
102 component of cellulose fibers that have nanoscale (less than 100 nm) diameter and lengths up to
103 several micrometers (Stelte et al. 2009). Using NFC as a reinforcement agent in polymer
104 composite materials has gained increasing attention because of its excellent properties such as
M
105 large surface area, water retention value, transparency, sustainability, and its unique features
106 such as high strength and stiffness, low weight and biodegradability (Turbak et al. 1983; Nair et
107 al. 2013). Cellulose nanocrystals (CNCs), rod-like or whisker shaped crystalline particles
d

108 isolated from cellulose, have gained wide attention over a range of research areas (Habibi et al.
109 2010 and Klemn et al. 2011) attributed to its good mechanical properties, unique optical and self
te

110 assembly properties (Wang et al., 2013). These excellent material properties play a significant
111 role in developing functionalized optically transparent materials and lightweight nanocomposites
112 for many applications. On the other hand, bacterial cellulose, a microbial polysaccharide, has
p

113 gained a great deal of attention owing to its impressive physico-mechanical properties being
114 obtained through a bottom up process which is biosynthesis of a class of acetic acid producing
ce

115 bacteria that includes that includes Acetobacter xylinus (Ha et al., 2011; Hestrin & Schramm,
116 1954; Shaha et al., 2013). Although bacterial cellulose (BC) has the same molecular formula as
117 plant cellulose, it exhibits a unique three-dimensional micro and nano-porous network structure
Ac

118 that provides high purity, a high degree of polymerization, high crystallinity (of 70–80%), high
119 water content to 99%, and high mechanical stability (Barud et al., 2011). In recent years there
120 has been considerable interest in using bacterial cellulose as a reinforcement nanomaterial in the
121 preparation of optically transparent materials (Yano et. al.2005 and Nogi et. al.2005). Its nano-
122 sized fiber structure, typically a width of 50-80 nm, and a thickness of 3-8 nm (Tabuchi, 2007),
123 enables the reduction of light scattering. BC also has a low coefficient of thermal expansion,
124 which is more important for reinforcement fillers in optoelectronic devices (Nogi et. al., 2008).
125 The dimensions and mechanical properties of nanoscale cellulose fibers compared to
126 microcrystalline cellulose are summarized in Table 1. (Edge et al., 2000; Chauhan and
127 Chakrabart 2012; Vitta and Thiruvengadam 2012; Yano 2010; Eichorn and Young 2001; Lee et
128 al., 2009). Although the diameters of cellulose nanomaterials are less than 100 nm in Table 1,
129 purity, surface energy (NFC: 41mN/m, BC: 57mN/m and CNC: 69mN/m) and crystal structure

4
Page 4 of 22
130 of the cellulose nanomaterials are different. In this regard, it is important to compare the
131 reinforcing efficiency of cellulose nanomaterials in polymer composites (Lee et al., 2012;
132 Yuwawech et al, 2015).
133
134 In the past years, many transparent polymers including poly (methyl methacrylate),
135 polystyrene, and polycarbonate have gained great interest because of their excellent optical
136 clarity and low density. Poly (methyl methacrylate) (PMMA), a glassy polymer with excellent

t
137 transparency and good processing ability is also used as a model polymer for making nanofiller-

ip
138 reinforced transparent nanocomposites. Despite its huge potential, there are some certain
139 drawbacks including mechanical-dynamical properties (low strength, impact resistance and
140 storage modulus etc.), which limit its efficient use in engineering applications (Liu et. al. 2010;

cr
141 Li et. al., 2013). PMMA are also often used in place of glass in certain applications in which both
142 high mechanical-dynamical properties and optical transparency are required. However,
143 mechanical strength of PMMA still does not have sufficient for many current applications (Day

us
144 et al., 1997). These drawbacks can be overcome via reinforcements with nano- and microfillers
145 (Nussbaumer et. al., 2003; Tang et.al. 2006; Chen et.al. 2009). In the Liu et al. study, transparent
146 polymethylmethacrylate (PMMA) composites were fabricated using freeze dried cellulose

an
147 nanocrystals (CNCs) through a solution casting method. Their research showed that transparent
148 composite sheets had better mechanical properties and their thermal stability seemed retained
149 with respect to the matrix polymer (Liu et al. 2010). A general need also exists for increasing the
150 mechanical strength and stiffness of PMMA composites while still retaining their good optical
M
151 transparency (Day et al., 1997).
152
153 In cellulose-based transparent composites, it is important to maintain the optical
d

154 properties. Because of this cellulose nanofibrils need to have a certain cross sectional size such
155 as less than half of the shortest wavelength of visible light (380–570 nm) to have reduced
te

156 interfacial light scattering (Althues et al. 2007). However, dispersion is an important factor in
157 terms of optical properties. Poor dispersion of nanofillers in polymeric matrices, lead to
158 composites will have inadequate transparency even at very low loadings (Xu et. al., 2013). With
p

159 the homogenous dispersion of nanofillers in polmer matrices, nanocomposites continue to


160 maintain optical properties (Ben Mabrouk et. al., 2014). Processing methods also play an
ce

161 important role on the fabrication of composite films because the ultimate properties depend on
162 morphological and cellulose-polymer interactions. Thus, the solvent casting technique is a
163 favorable processing method to synthesize cellulosic nanocomposites films because of its slow
Ac

164 solvent evaporation that provides better nanofiller arrangement and also gives enough time for
165 the cellulose nanofibrils to form a percolation network (Dufresne et. al., 2013). In the Hajji et al.
166 study, the influence of three processing methods was compared and the solvent evaporation
167 method had pronounced success in enhancing the mechanical properties of cellulose
168 nanowhisker-based nanocomposites. Nanofiber-network reinforced optically transparent
169 composites with improved properties have also been reported (Nogi et al. 2005 and Okahisa et
170 al., 2011). Hence, PMMA and cellulose nanofiber reinforced PMMA-based nanocomposites
171 were fabricated by the solvent casting technique in the present study.
172
173 It has been recently reported in the literarture (Jonoobi et al. 2013 and Wang et al.2013)
174 that mechanical, thermal and optical properties were improved with cellulose nanomaterials.
175 However, it has been also reported that cellulose nanomaterials loading may often deteriorate

5
Page 5 of 22
176 these properties. In this respect, the effects of cellulose nanomaterials in polymers are not yet
177 fully understood. Therefore, the objective of this work is to elucidate the differences or
178 similarities among cellulose nanomaterials (NFC, CNC and NDC from BC) in terms of
179 mechanical, thermal and optical properties in PMMA matrix. This study also compared their
180 origin (bacteria versus trees), morphologies and dispersion properties in PMMA matrix,
181 interactions with matrix, and the resulting reinforcing effects on the matrix polymer. Thermal,
182 mechanical and optical properties of nanocomposites were evaluated by differential scanning

t
183 calorimetry (DSC), dynamic mechanical thermal analysis (DMTA), thermogravimetric analysis

ip
184 (TGA), and UV–visible spectroscopy, respectively. These results should allow us to further
185 investigate the industrial application areas for certain types of cellulose nanofibers.
186

cr
187 2. Material and Methods
188 2.1. Materials
189 Nanofibrillated Cellulose (NFC) suspension at 3.5 wt. % was kindly obtained from the

us
190 University of Maine Chemical and Biological Engineering Department. Cellulose Nanocrystals
191 (CNC) were provided by the United States Department of Agriculture (USDA) Forest Service,
192 Forest Products Laboratory. Solids content of CNC was 6.5 wt. %. Bacterial Cellulose (NDC)

an
193 was extracted from Nata de coco which is a food product from the fermentation of coconut milk
194 using bacteria Acetobacter xylinum. PMMA (OPTIX ® CA – 75 CLEAR) was supplied from
195 Plaskolite and acetone, (Purity ≥ 98%) was purchased from Sigma Aldrich, respectively.
196
M
197 2.2. Bacterial Cellulose Production from Nata De Coco
198 BC was extracted from 10 jars (450gr) of Nata-de-coco. First, Nata-de-coco was rinsed
199 with de-ionized water (3×10 dm³) to wash away the sugar syrup. The washed Nata-de coco then
d

200 was blended for 1 min using a laboratory blender. This BC blend was then homogenized for 5
201 min and centrifuged at 14,000 g to remove the excess water. To further purify the BC, the
te

202 centrifuged material was redispersed in deionized water (10 dm³). Sodium hydroxide (40 g, 1
203 mol) was added into this mixture and heated to 80˚ C for 60 min, while stirring to remove any
204 soluble polysaccharides. The purified BC was then successively centrifuged and homogenized
p

205 back to neutral pH using deionized water (Juntaro 2009).


206
ce

207 2.3. Production of Transparent Composites


208 PMMA (20g) was dissolved in 100 ml of acetone at 50°C. The maximum temperature for
209 the curing process was selected as 50 °C because the boiling point of acetone is 56 °C (Kim et al.
Ac

210 2005). PMMA-cellulose nanomaterial blends were prepared by dissolving different amounts of
211 cellulose nanofibers (0.25% and 0.5% by weight) in the PMMA-acetone solution under
212 continuous stirring for 1 hour followed by ultra-sonication for several hours at 50°C. The content
213 of PMMA was kept constant for all samples. Composites were prepared by casting this solution
214 in a glass Petri Dish and allowed to evaporate the solution at 50°C in oven overnight and then in
215 vacuum oven for two days. After this, the films were peeled from the glass plate using a doctor
216 blade to obtain free-standing films (Tomar et al. 2011).
217
218 2.4. Characterization of the Composites
219 The optical transmittance of the PMMA and transparent nanocomposites were measured
220 at wavelengths from 400 to 1000 nm using an ultraviolet (UV)–visible spectrophotometer
221 (HP8453). Transmission spectra were measured using air as a reference. Thermogravimetric

6
Page 6 of 22
222 Analysis (TGA) measurements were carried out using a Mettler Toledo analyzer on samples of
223 about 8 mg. Each sample was scanned over a temperature range from ambient temperature to
224 600 °C at a heating rate of 10°C/min under nitrogen with a flow rate 20 ml/min to avoid sample
225 oxidation. The samples used for the TGA measurement were randomly picked 5 individual
226 samples from ground samples (Kiziltas et al., 2014; Ozen et al., 2013). Differential scanning
227 calorimetry (DSC) experiments were used to determine glass transition temperature (Tg) of the
228 neat PMMA and nanocomposite materials using a TA Q2000 differential scanning calorimeter

t
229 (TA Instrument Inc., New Castle, DE, USA). The samples were heated from room temperature

ip
230 to 165°C to remove any possible solvent residues, cooled to 0°C, and then heated to 200° C at
231 the same heating ramp of 10° C /min under a nitrogen atmosphere. The data were collected on
232 the second heating ramp. At least three specimens were tested for each composition, and the

cr
233 results are presented as an average for tested samples. The viscoelastic properties of the
234 composites were determined with a Rheometric Scientific DMTA IV. The experiments were
235 conducted in tensile mode under isochronal conditions at a frequency of 10 Hz. The strain

us
236 amplitude was fixed at 0.01% to be in the domain of the linear viscoelasticity of the composites.
237 The temperature range was from 0 to 100°C at a scanning rate of 5°C /min. The storage modulus
238 (E’), loss modulus (E’’) and loss factor (tan δ) of the samples were measured as a function of

an
239 temperature. At least three specimens were tested for each composition, and the results are
240 presented as an average for tested samples. DMTA was also used to investigate flexural modulus
241 of elasticity and flexural strength of neat PMMA and nanocomposites. All specimens were
242 subjected to a three point bending at a test temperature of 21.7±0.3°C with a constant strain rate
M
243 of 1x10-4 mm/s. At least three specimens were tested for each composition, and the results are
244 presented as an average for tested samples. Attenuated Total Reflectance Fourier-Transform
245 InfraRed (ATR-FTIR) spectroscopy analysis was carried out on the neat PMMA and cellulose
d

246 nanocomposite samples on a Perkin Elmer Spectrum One FTIR spectrometer (Wellesley, MA,
247 USA), equipped with a Universal ATR accessory, using 200 scans and a resolution of 4 cm−1,
te

248 over the range 4000–400 cm−1. Before the analysis, samples were dried at 90 °C under vacuum
249 for 24 h (Erbas Kiziltas et al., 2015). Optical microscopy samples were prepared with a Leica
250 Ultracut Microtome, using a diamond blade at room temperature. Sections were cut to a 5 micron
p

251 thickness and floated in oil between a glass slide and coverslip. The samples were viewed in
252 transmission on a Nikon FX upright microscope.
ce

253
254 3. Results and Discussion
255 Figure 1 shows images of the neat PMMA and nanocomposite sheets with 0.25 wt%
Ac

256 NFC, CNC, NDC and 0.5 wt. % NFC. The patterns and letters in the background indicate that
257 the nanocomposite sheets are transparent. It can be seen from Figure 1 that the CNC reinforced
258 nanocomposites exhibited better transparency compared to the NFC and NDC reinforced
259 nanocomposites at the 0.25 wt. % loading level. This result indicated the light transmittance was
260 less affected by the cellulose nanocrystals at low loading levels because of their relatively small
261 size and homogenous dispersion (Liu et al. 2010). Consequently, an increased cellulose
262 nanomaterials loading level (0.5wt. %) became increasingly opaque because of CNF
263 agglomeration in the polymer matrix (Liu et al. 2010). Similar results were also observed for
264 PMMA/NFC composites with a loading level of 0.5 wt. % (Littunen et al. 2013).
265
266 Figure 2 shows quantitatively measured profiles of light transmittance versus the
267 wavelength of visible light for the pure PMMA matrix and cellulose nanocomposite sheets

7
Page 7 of 22
268 examined by a UV–visible spectrometer at a visible wavelength range of 400– 800 nm. The
269 percent transmittances at 400, 600 and 800 nm are also tabulated in Table 2. The pure PMMA
270 sample transmitted about 90% of the incident light. Figure 2 also showed that the UV-vis spectra
271 of PMMA nanocomposites reinforced with 0.25 wt. % CNC, NDC and 0.5 wt. % CNC
272 reinforced transparent composites offered 72%, 58% and 55% light transmittance respectively at
273 a wavelength of 600 nm. However, because of the heterogeneous nature of the composites, the
274 transmittance of nanocomposite was reduced with the increase of the loading content of cellulose

t
275 nanofibers. It has been shown that the optical transmittance of the CNF composites is largely

ip
276 dependent on the dispersion of CNFs in the polymer matrix (Liu et al. 2010). The decrease in
277 light transmittance by adding NDC (81%) and NFC (94%) at a percentage of 0.5% was found to
278 be relatively greater compared to NDC (26.5%) and NFC (73%) with a percentage of 0.25%

cr
279 compared to pure PMMA at 600 nm. These results suggest that inadequate dispersion and
280 agglomeration of cellulose nanofibers in the nanocomposites lead to reduced transparency. On
281 the other hand, further addition of NFC generated a significant decrease of composite

us
282 transparency although NFC exhibited better dispersion in the composites. This could probably be
283 attributed to NFCs relatively large lengths. However, high transparency and uniform dispersion
284 can be achieved by filtering cellulose suspensions prior to solvent casting (Nogi et al. 2009).

an
285
286 Figure 3 shows the thermal degradation behavior of PMMA, cellulose nanomaterials, and
287 PMMA/cellulose nanocomposites. The initial weight loss below 200 °C observed for CNFs
288 attributed to removal of moisture from the cellulose. The thermal degradation that occurred
M
289 between temperatures of 200−380 °C is attributed to the depolymerization of hemicellulose (for
290 NFC) and cleavage of glycosidic linkages of cellulose for all cellulose nanofibers (Lee et al.,
291 2012). Figure 3 also shows that thermal degradation behavior of CNC, NFC and NDC were very
d

292 similar, whereas the onset degradation temperature of CNC was higher than those of NFC and
293 NDC. The degradation onset temperatures corresponding to CNC, NFC and NDC were 287°C,
te

294 273°C and 260°C, respectively. The PMMA and cellulose nanocomposites underwent thermal
295 degradation at a higher temperature than that of cellulose nanofibers. This could be explained via
296 hydrogen bonding interactions between CNC hydroxyl groups and carbonyl groups on the
p

297 PMMA matrix to facilitate miscibility of polymer blends (Kuo et al., 2008; Dong et al., 2012)
298 where hydrogen bonding had a significant effect on the thermal properties of polymer blends.
ce

299 However, there was no significant difference found in the degradation temperatures of PMMA
300 and cellulose nanocomposites with different loadings probably because of the low contents of
301 cellulose nanofibrils.
Ac

302
303 The PMMA began thermal degradation at 305 °C with a mass loss of 10% which
304 indicated its onset decomposition temperature whereas the cellulose nanocomposites degradation
305 peaks varied in the range of 292–325 °C as shown in Figure 4.a. The maximum decomposition
306 temperature of the cellulose nanocomposites was increased 5 to 20ºC with the variation of
307 cellulose nanofibers content whereas CNC (0.25 wt.) and NDC (0.5 wt. %) showed lower
308 decomposition temperature compared to other cellulose nanocomposites. Sample preparations
309 (size, morphology and homogeneity) may influence heat transfer within the samples and thus
310 influence the course of reaction and the thermogravimetric measurements (Bottom 2008). In
311 NDC production, sodium hydroxide was added into this mixture and heated to 80˚ C for 60 min,
312 while stirring to remove any soluble polysaccharides. Residues of the polysaccharides remaining
313 after processing of NDC could be the reason for the lower thermal stability. The thermal stability

8
Page 8 of 22
314 of cellulose nanocrystals which were prepared by sulfuric acid hydrolysis was lower than that of
315 the original cellulose because of the presence of acid sulfate groups which decreased the thermal
316 stability by a dehydration reaction (Roman and Winter 2004; Wang et al. 2007). The maximum
317 decomposition temperature increased 20ºC with the NFC (0.25 wt. %) in the nanocomposites.
318 Figure 4.b. shows mass loss at 600°C. The maximum residual mass at 600 °C was obtained from
319 the NDC (about 35 %). There were no residual mass obtained for neat PMMA and the cellulose
320 nanocomposites at 600 °C attributed to the degradation of PMMA starting slowly at 220°C and

t
321 completely degraded at a temperature higher than 305°C (Pielichowski et al. 2005).

ip
322
323 The Tg of the cellulose composites was determined as the inflection point of the specific

cr
324 heat increment at the glass-rubber transition in DSC experiments. Figure 5 shows the effect of
325 cellulose nanofibers on the Tg of the composites measured by DSC. It can be seen that
326 incorporation of low contents of cellulose nanomaterials (0.25% and 0.5%) slightly increased the

us
327 glass transition temperature of PMMA. (~ 80°C and nanocomposites ~ 84°C). The trend of the
328 glass transition shifting to a slightly higher temperature suggests a restriction of the mobility of
329 polymer chains (Kuo, 2008). This result was expected since cellulose nanomaterials’ hydroxyl
330 groups could interact with the ester functional group (COOCH3) of the PPMA polymeric matrix

an
331 in its side chain and therefore hinder the rotation of polymeric chains (Kuo, 2008; Jonoobi et al.,
332 2010). A Similar result was reported for cellulose nanoparticle-reinforced PMMA composites in
333 the literature (Han et al. 2014).
M
334
335 The mechanical properties of the PMMA and the nanocomposites were also studied to
336 examine the influence of cellulose nanofibers on the deformation and viscoelastic behavior of the
337 nanocomposites. Figure 6 shows representative stress–strain curves of neat PMMA and cellulose
d

338 nanocomposites with varying cellulose nanomaterial contents from DMTA. The nanocomposites
339 with different loading content of CNC and NDC possessed slightly decreased values of flexural
te

340 strength attributed to higher surface tension for CNC (69 mN/m), NDC (57 mN/m) compared to
341 NFC (41 mN/m). However, NFC had a slight increase in the flexural strength of around 60 MPa
342 while the neat PMMA had a flexural strength of 57 MPa as presented in Figure 6. This could be
p

343 attributed to PMMA (41 mN/m) has better compatibility with the reinforcing material with lower
ce

344 surface energy (Oporto et al., 2011). The flexural strength values decreased with respect to the
345 strength of the PMMA matrix between approximately 7-13%. Figure 6 indicates that the flexural
346 modulus of the nanocomposites reinforced with NDC at the 0.5 wt. % loading level was 2.56
347 GPa, 23% greater than to that of the pure PMMA (2.07 GPa). The values of the flexural
Ac

348 modulus of elasticity for PMMA, PMMA/NDC (0.25%), PMMA/CNC (0.25%), PMMA/NFC
349 (0.25%), PMMA/NDC (0.5%), PMMA/CNC (0.5%) and PMMA/NFC (0.5%) were 2.07, 2.15,
350 2.05, 2.13, 2.56, 2.19 and 2.35 GPa, respectively. It can be seen that the NDC reinforced
351 nanocomposites showed better flexural modulus of elasticity compared to NFC and CNC
352 reinforced composites. This increase in flexural modulus of elasticity indicates that the NDC
353 acted as reinforcement in the polymer matrix by transferring load from the polymer matrix to the
354 NDC. It was also observed that from testing, PMMA/NFC composites did not have greater strain
355 to failure than PMMA/CNC and PMMA/NDC nanocomposites.
356
357 The storage moduli (E′) at room temperature and glass transition temperatures of the
358 PMMA and nanocomposites with different cellulose nanofiber loadings are shown in Table 3.
359 The glass transition of PMMA was detected at around 75°C, the glass transition temperature

9
Page 9 of 22
360 values were not significantly changed as the cellulose nanomaterials content was increased since
361 the viscoelastic properties of the composite are strongly influenced by the matrix polymer. The
362 PMMA composites had glass transition temperatures between 73 and 76° C. The storage moduli
363 of the nanocomposites slightly increased with cellulose nanofibers at loading contents of 0.25 wt.
364 % and 0.5 wt. % at room temperature. At low temperatures, the PMMA matrix was very rigid,
365 and therefore no vigorous reinforcing effect from cellulose nanomaterials was detected at the
366 magnitude of the storage moduli of nanocomposites (Han et al. 2014). The storage modulus of

t
367 neat PMMA at room temperature was 2.73 GPa. However, the addition of 0.25 wt. % of NDC

ip
368 had a positive effect and the storage modulus of the composite (3.04 GPa) increased 12%
369 compared to neat PMMA. This result could be attributed to the higher crystallinity of 71% of the
370 NDC compared to NFC (41%) (Lee at al. 2012). The storage moduli (E′) at the glass transition

cr
371 temperature and the tan δ maximum peak values of the neat PMMA and nanocomposites with
372 different cellulose nanomaterial loadings are shown in Table 3. The storage modulus of the
373 nanocomposites was greatly increased with cellulose nanofibers at loading contents of 0.25 wt.

us
374 % and 0.5 wt. % at the glass transition temperature. The reinforcement effect of the cellulose
375 nanomaterials was clear at the glass transition temperature because the PMMA matrix became
376 softened and the CNFs restricted the motion of the PMMA chains which promoted the rigidity of

an
377 the PMMA (Han et al. 2014). The magnitude of the tan δ peak values of the CNC reinforced
378 nanocomposites were significantly decreased around the glass transition temperature. This can
379 probably be attributed to restriction in chain segment mobility in the amorphous region of the
380 polymer (Kiziltas et al. 2011). The highest moduli of the PMMA/CNC nanocomposites
M
381 containing 0.25 and 0.5 wt. % cellulose were about 46% and 260% of that of pure PMMA at the
382 glass transition temperature. This result confirmed the creation of enhanced reinforcement with
383 the addition of cellulose nanocrystals into the PMMA matrix.
d

384
385 Figure 7 depicts the FT–IR spectra of neat PMMA and cellulose nanocomposites. It is
te

386 evident that all the four spectra are similar except for a few changes in the spectra of the
387 nanocomposites. No significant differences were observed between the FTIR spectra of pure
388 PMMA and cellulose nanocomposites. The fingerprint characteristic vibration bands of PMMA
p

389 appear at 1728 cm-1 C=O stretching mode. The bands at 3100 and 2800 cm–1 correspond to the
390 methylene C–H stretching while the bands at 1350 and 1450 cm–1 are associated with C–H
ce

391 symmetric and asymmetric stretching modes, respectively. The 1240 cm–1 band is assigned to
392 antisymmetric C-C-O stretch and the 1150 cm–1 band corresponds to skeletal vibrations coupled
393 to C-H deformations, while C–C stretching bands are at 1000 and 800 cm–1. Absence of any
Ac

394 additional bands by the presence of cellulose nanomaterials other than those of PMMA in the
395 spectrum of PMMA and further they remaining unperturbed in all the four spectra indicate (1)
396 the purity of the polymer obtained, (2) formation of the nanocomposites and (3) no chemical
397 interaction or chemical bond formation between PMMA and cellulose nanomaterials (Sain et al.,
398 2011; Ahmad et al., 2007).
399
400 Figure 8 show the optical images of the of PMMA and cellulose nanocomposites. In neat
401 PMMA, very few partially dissolved granules were observed in Figure 8. It was apparent from
402 the optical micrograph of PMMA/cellulose nanocomposite that some cellulose nanomaterials
403 might have reaggregated among themselves, adhered to the polymer surface and distribution was
404 not uniform. A few potential scenarios may be envisaged. (1) During the reaction, the
405 hydrophilic cellulose nanomaterials migrated to the polymer water interface, which might be

10
Page 10 of 22
406 because of either Van der Waal’s attraction or strong hydrogen bonding interaction among
407 themselves. (2) Agglomeration occurred due to the incompatibility between the hydrophilic
408 cellulose surface and hydrophobic nature of the PMMA (Sain et al., 2011; Ahmad et al., 2007).
409
410 4. Conclusions
411 PMMA-based transparent nanocomposites were successfully prepared by solution casting
412 with the reinforcement of cellulose nanofibers. The transmittance of PMMA/cellulose

t
413 nanocomposites was reduced with increased loading of cellulose nanofibrils. The UV-visible

ip
414 light transmittance of the nanocomposites decreased by 9% and 27% with the addition of 0.25
415 wt. % CNC and NDC respectively at 600 nm. Thermogravimetric analysis indicated retained
416 thermal stability of the transparent composites. The maximum decomposition temperature was

cr
417 increased 20ºC with an NFC content of 0.25 wt. % in the nanocomposites. The maximum
418 residual mass of the NDC was about 35 % obtained at a maximum temperature of 600 °C. The
419 DSC measurements also indicated that the presence of cellulose nanofibers slightly changed the

us
420 Tg of the PMMA matrix. The storage modulus of the PMMA/CNC nanocomposite containing
421 0.25 and 0.5 wt. % cellulose increased about 46% and 260% compared to that of the pure
422 PMMA at the glass transition temperature, respectively. The flexural modulus of the

an
423 nanocomposites reinforced with NDC at the 0.5 wt.% loading level was 2.56 GPa, which was
424 23% greater than the pure PMMA (2.07 GPa). The NFC composites showed enhanced flexural
425 strength of around 60 MPa compared to those comprised of CNC and NDC. These results imply
426 that at even low cellulose nanomaterials content enhanced thermal and mechanical properties of
M
427 the rigid-thermoplastic PMMA can be obtained and these serve as promising reinforced
428 nanomaterials for the production of transparent nanocomposites.
429
d

430 Acknowledgements
431 The republic of Turkey, The Scientific and Technological Research Council of Turkey
te

432 (TUBITAK) is greatly acknowledged for support of the scholarship of the researcher Esra Erbas
433 Kiziltas to do this study at the University of Maine. The authors would like to acknowledge the
434 contributions of Justin Crouse, Dr. Jason Bolton, Alex Nash, Donald Gjeta, Dr. Sanjeev Kumar
p

435 Kandpal, Connie Young Johnson and Chris West whose hard work made this paper possible. The
ce

436 authors would also like to thank U.S. Army Corps of Engineers, Engineer Research and
437 Development Center project 912HZ-07-2-0013 and Maine Agricultural and Forest Experiment
438 Station (MAFES) project ME09615-08MS and the Wood Utilization Research Hatch 2007-2008
439 project.
Ac

440
441 5. References
442 Ahmad S., Ahmad S. & Agnihotry S.A. (2007) Synthesis and characterization of in situ prepared
443 poly(methyl methacrylate) nanocomposites. Bull. Mater. Sci., 30(1): 31–35.
444 Althues H., Henle J. & Kaskel S. (2007) Functional inorganic nanofillers for transparent
445 polymers. Chemical Society Reviews 36: 1454-1465.
446 Barud H. S., Regiani T., Marques R. F. C., Lustri W. R., Messaddeq Y. & Ribeiro S. J. L. (2011)
447 Antimicrobial bacterial cellulose-silver nanoparticles composite membranes. Journal of
448 Nanomaterials http://dx.doi.org/10.1155/2011/721631. 1-8pp.
449
450

11
Page 11 of 22
451 Ben Mabrouk A., Brochier Salon M.C., Magnin A., Belgacem M.N. & Boufi S. (2014)
452 Cellulose-based nanocomposites prepared via mini-emulsion polymerization: Understanding the
453 chemistry of the nanocellulose/matrix interface. Colloids and Surfaces A: Physicochemical and
454 Engineering Aspects 448:1-8.
455 Biyun L., Huihua Y. & Yanzhong Z. (2013) Transparent PMMA-based nanocomposite using
456 electrospun graphene-incorporated PA-6 nanofibers as the reinforcement Composites Science
457 and Technology 89:134-141.

t
458 Bottom, R. 2008. Chapter 3 - Thermogravimetric analysis. In Gabbott, P. (editor) Principles and

ip
459 Applications of Thermal Analysis, UK. Blackwell Publishing. pp 87-118.
460 Chauhan V.S. & Chakrabart S.K. (2012). Use of nanotechnology for high performance cellulosic
461 and papermaking products. Cellulose Chemistry and Technology, 46(5-6): 389-400.

cr
462 Chen L.S., Huang Z.M., Dong G.H., He C.L., Liu L., Hu Y.Y. & Li, Y. (2009) Development of a
463 transparent PMMA composite reinforced with nanofibers. Polymer Composites 30:239-247.
464 Day D.E., Stoffer J.O. & Barr J.M. (1997) Optically transparent composite material and process

us
465 for preparing same. US Patent 5,665,450.
466 Dong H, Strawhecker K.E., Snyder J.F., Orlicki J.A., Reiner R.S. & Rudie, A.W. (2012)
467 Cellulose nanocrystals as a reinforcing material for electrospun poly(methyl methacrylate) fibers:

an
468 formation, properties and nanomechanical characterization. Carbohydrate Polymers 87:2488-
469 2495.
470 Dufresne A.(2011) Polymer nanocomposites from biological sources. In Nalwa H.S.
471 Encyclopedia of Nanoscience and Nanotechnology. 2nd ed., American Scientific Publishers
M
472 21:219-250.
473 Edge S., Steele D.F., Chen A., Tobyn M.J. & Staniforth J.N. (2000). The mechanical properties
474 of compacts of microcrystalline cellulose and silicified microcrystalline cellulose. International
d

475 Journal of Pharmaceutics, 200(1):67-72.


476 Eichorn S.J. and Young R.J. (2001).The Young's modulus of a microcrystalline cellulose.
te

477 Cellulose, 8(3):197-207.


478 Gardner D. J., Oporto G. S., Mills R. & Samir M. A. S A. (2008). Adhesion and surface issues in
479 cellulose and nanocellulose. Journal of Adhesion Science and Technology 22:545-567.
p

480 Ha J. H., Shah N., Ul-Islam M., Khan T. & Park J. K. (2011) Bacterial cellulose production from
481 a single sugar α-linked glucuronic acid-based oligosaccharide. Process Biochemistry 46:1717-
ce

482 1723.
483 Habibi Y. Lucia L. A. & Rojas O. J. (2010) Cellulose nanocrystals: chemistry, self-assembly,
484 and applications. Chemical Reviews 110:3479-3500.
Ac

485 Hajji P., Cavaille J.Y., Favier V., Gauthier C. & Vigier G. (1996) Tensile behavior of
486 nanocomposites from latex and cellulose whiskers. Polymer Composites 17: 537-647.
487 Han G., Huan S., Han J., Zhang Z. & Wu, Q. (2014) Effect of Acid Hydrolysis Conditions on the
488 Properties of Cellulose Nanoparticle-reinforced polymethylmethacrylate composites. Materials
489 7:16-29.
490 Hestrin S. & Schramm M. (1954) Synthesis of cellulose by Acetobacter xylinum. 2.Preparation
491 of freeze-dried cells capable of polymerizing glucose to cellulose. Biochemical Journal 58, 345-
492 352.
493 Jonoobi M. Harun J. Mathew A.P. & Oksman K. (2010). Mechanical properties of cellulose
494 nanofiber (CNF) reinforced polylactic acid (PLA) prepared by twin screw extrusion. Composites
495 Science and Technology 70:1742-1747.

12
Page 12 of 22
496 Jonoobi M., Aitomäki Y., Mathew A.P., & Oksman K. (2013). Thermoplastic polymer
497 impregnation of cellulose nanofibre networks: Morphology, mechanical and optical properties.
498 Composites: Part A 58:30-35.
499 Juntaro J. (2009). Environmentally friendly hierarchical composites. PhD Thesis, London,
500 England: Department of Chemical Engineering and Chemical Technology. University of
501 London, London.
502 Kim H.C., Lee S.E., Kim C.G. & Lee J.J. (2005) Mechanical improvement of multi-walled

t
503 carbon nanotube /poly (methyl methacrylate) composites. Key Engineering Materials 297-300:

ip
504 2545-2550.
505 Kiziltas A., Gardner D.J., Han Y., Yang H-S. (2011) Dynamic mechanical behavior and thermal
506 properties of microcrystalline cellulose (MCC)-filled nylon 6 composites Thermochimica Acta

cr
507 519: 38-43.
508 Erbas Kiziltas E., Kiziltas A. & Gardner D.J. (2015) Synthesis of bacterial cellulose using hot
509 water extracted wood sugars. Carbohydrate Polymers, 124:131-138.

us
510 Kiziltas, A., Nazari, B., Gardner, D.J. & Bousfield, D.W. (2014). Polyamide 6–cellulose
511 composites: effect of cellulose composition on melt rheology and crystallization behavior.
512 Polymer Engineering & Science, 54, 739-746.

an
513 Klemm D., Kramer F., Moritz S., Lindstrom T., Ankerfors M., Gray D. & Dorris A. (2011)
514 Nanocelluloses: a new family of nature-based materials Angewandte Chemie International
515 Edition 50:5438-5466.
516 Kuo S. W. (2008). Hydrogen bonding in polymer blends. Journal of Polymer Research 15:459-
M
517 486.
518 Lee K.-Y., Blaker J.J. & Bismarck A. (2009). Improving the properties of nanocellulose /
519 polylactide composites by esterification of nanocellulose. Can it be done? In 17th International
d

520 Conference on Composite Materials, Edinburgh, UK.


521 Lee Koon-Yang., Tammelin Tekla., Schulfter K., Kiiskinen H., Samela J. & Bismarck A. (2012)
te

522 High performance cellulose nanocomposites: comparing the reinforcing ability of bacterial
523 cellulose and nanofibrillated cellulose. ACS Applied Materials & Interfaces 4:4078-4086.
524 Littunen K., Hippi U., Saarinen T. & Jukka Seppälä (2013). Network formation of
p

525 nanofibrillated cellulose in solution blended poly (methylmethacrylate) composites.


526 Carbohydrate Polymers 91:183-190.
ce

527 Liu H.Y., Liu D.G., Yao F. & Wu, Q.L. (2010) Fabrication and properties of transparent
528 polymethylmethacrylate/cellulose nanocrystals composites. Bioresource Technology 101:5685-
529 5692.
Ac

530 Nair S.S., Zhu J.Y., Deng Y. & Ragauskas A.J. (2014) Hydrogels prepared from cross-linked
531 nanofibrillated cellulose. ACS Sustainable Chemistry & Engineering DOI: 10.1021/sc400445t.
532 Nishino T., Takano K. & Nakamae K. (1995) Elastic modulus of the crystalline regions of
533 cellulose polymorphs. Journal of Polymer Science Part B- Polymer Physics 33:1647-1651.
534 Nogi M. & Yano H. (2008) Transparent nanocomposites based on cellulose produced by bacteria
535 offer potential innovation in the electronics device industry. Advanced Materials 20:1849-52.
536 Nogi M., Handa K., Nakagaito A.N. & Yano, H. (2005) Optically transparent bionanofiber
537 composites with low sensitivity to refractive index of the polymer matrix. Applied Physics
538 Letters 87:243110-243112.
539 Nogi M., Iwamoto S., Nakagaito A. N. & Yano H. (2009) Optically transparent nanofiber paper.
540 Advanced Materials 21:1595-1598.

13
Page 13 of 22
541 Nussbaumer R.J., Caseri W.R., Smith P. & Tervoort T. (2003) Polymer-TiO2 nanocomposites: a
542 route towards visually transparent broadband UV filters and high refractive index materials.
543 Macromolecular Materials and Engineering 288:44-49.
544 Okahisa Y., Abe K., Nogi M., Nakagaito A.N., Nakatani T. & Yano, H. (2011) Effects of
545 delignification in the production of plant-based cellulose nanofibers for optically transparent
546 nanocomposites. Composites Science and Technology 71:1342-1347.
547 Oporto G.S., Gardner D.J., Kiziltas A. & Neivandt D.J. (2011) Understanding the affinity

t
548 between components of wood–plastic composites from a surface energy perspective. Journal of

ip
549 Adhesion Science and Technology, 25(15):1785-1801.
550 Ozen, E., Kiziltas, A., Erbas Kiziltas, E. & Gardner, D.J. (2013). Natural fiber blend-nylon 6
551 composites. Polymer Composites, 34, 544-553.

cr
552 Pecoraro E., Manzani D., Messaddeq Y. & Ribeiro, S.J.L. (2008) Bacterial cellulose from
553 Glucanacetobacter xylinus: preparation, properties and applications. In: Monomers, Polymers
554 and Composites from Renewable Resources. Eds.

us
555 Pielichowski K., J. Njuguna. (2005). Thermal degradation polymeric material. United Kingdom:
556 Rapra Technology Limited.
557 Roman M. & Winter W.T. (2004) Effect of sulfate groups from sulfuric acid hydrolysis on the

an
558 thermal degradation behavior of bacterial cellulose. Biomacromolecules 5:1671-1677.
559 Sain S., Ray D., Mukhopadhyay A., Sengupta S., Kar T., Ennis C.J. & Rahman P.K.S. M. (2012)
560 Synthesis and characterization of PMMA-cellulose nanocomposites by in situ polymerization
561 technique. Journal of Applied Polymer Science, 126, E127–E134.
M
562 Shaha N, Ul-Islama M, Khattaka W.A. & Parka. J.K. (2013) Overview of bacterial cellulose
563 composites: A multipurpose advanced material. Carbohydrate Polymers 98:1585-1598.
564 Stelte W. & Sanadi A. R. (2009) Preparation and characterization of cellulose nanofibers from
d

565 two commercial hardwood and softwood pulps. Industrial & Engineering Chemistry Research
566 48:11211-11219.
te

567 Tabuchi M. (2007) Nanobiotech versus synthetic nanotech? Nature Biotechnology 25:389-390.
568 Tang E.J., Cheng G.X. & Ma X.L. (2006) Preparation of nano-ZnO/PMMA composite particles
569 via grafting of the copolymer onto the surface of zinc oxide nanoparticles. Powder Technology
p

570 161:209-214.
571 Tanpichai S., Quero F., Nogi M., Yano H., Young R. J., Lindstrom T., Sampson W.W. &
ce

572 Eichhorn S.J.(2012) Effective young’s modulus of bacterial and microfibrillated cellulose fibrils
573 in fibrous networks. Biomacromolecules 13:1340-1349.
574 Tomar A.K., Mahendia S. & Kumar S. (2011). Structural characterization of PMMA blended
Ac

575 with chemically synthesized PAni. Journal of Applied Sciences Research 2:65-71.
576 Turbak A. F., Snyder F. W. & Sandberg, K. R. J. (1983) Microfibrillated cellulose, a new
577 cellulose product: properties, uses, and commercial potential. Journal of applied polymer
578 science. Applied Polymer Symposium 37:815-827.
579 Vitta S. & Thiruvengadam V. (2012). Multifunctional bacterial cellulose and nanoparticle-
580 embedded composites. Current Science, 102(10):1398-1405.
581 Vitta S. & Thiruvengadam, V. (2012) Multifunctional bacterial cellulose and nanoparticle-
582 embedded composites Current Science, 102: 10-25.
583 Wang D., Yu J., Zhang J., He J. & Zhang J. (2013). Transparent bionanocomposites with
584 improved properties from poly(propylene carbonate) (PPC) and cellulose nanowhiskers (CNWs).
585 Composites Science and Technology 85:83-89.

14
Page 14 of 22
586 Wang N., Ding E. & Cheng R. (2007) Thermal degradation behavior of spherical cellulose
587 nanocrystals with sulfate groups. Polymer 48:3486-3493.
588 Wang Q., Zhu J.Y. & Considine J.M. (2013) Strong and optically transparent films prepared
589 using cellulosic solid residue recovered from cellulose nanocrystals production. Waste Stream.
590 ACS Applied Materials & Interfaces 5:2527-2534.
591 Xu W., Qin Z., Yu H., Liu Y., Liu N., Zhou Z. & Chen, L. (2013) Cellulose nanocrystals as
592 organic nanofillers for transparent polycarbonate films Journal of Nanoparticle Research

t
593 15:1562.

ip
594 Yano H. (2010) Potential of cellulose nanofiber-based materials. TAPPI International
595 Conference on Nanotechnology for Renewable Materials. Otaniemi, Espoo Finland.
596 Yano H., Sugiyama J., Nakagaito A.N., Nogi M., Matsuura T. & Hikita M. (2005) Optically

cr
597 transparent composites reinforced with networks of bacterial nanofibers. Advanced Materials
598 17:153-165
599 Yuwawech K., Wootthikanokkhan J. & Tanpichai S. (2015) Effects of two different cellulose

us
600 nanofiber types on properties of poly(vinyl alcohol) composite films. Journal of Nanomaterials,
601 2015, 1-10.
602

an
603
604 LIST OF TABLES:

605 Table 1. Dimensions and mechanical properties of nanoscale cellulose fibers compared to
M
606 microcrystalline cellulose.
607 Table 2. UV-visible transmittances of the neat PMMA and cellulose nanocomposites at 400 nm,
608 600 nm and 800 nm.
d

609 Table 3. Glass transition temperature (Tg), tan δ maximum peak values (tan δ max. peak), storage
610 modulus at room temperature (SM at RT) and glass transition temperature (SM at Tg) for neat
te

611 PMMA and cellulose nanocomposites.


612
613 Table 1. Dimensions and mechanical properties of nanoscale cellulose fibers compared to
p

614 microcrystalline cellulose.


ce

Material Length Diameter Aspect Tensile Modulus of Density


(nm) (nm) Ratio (l/d) Strength (GPa) Elasticity (GPa) (g/cm3)
BC >1000 24-84 15-35 114-300 1.25
Ac

∼50

CNC 100-600 2-20 10-100 7.5 120-143 1.5


CNF >1000 10-40 100-150 3 138-141 -
MCC >1000 >1000 - 25 1.45

∼1

615
616
617

15
Page 15 of 22
618
619
620
621
622
623 Table 2. UV-visible transmittances of the neat PMMA and CNF nanocomposites at 400 nm, 600
624 nm and 800 nm.

t
Sample Name Transmittances (%)

ip
400 nm 600 nm 800 nm
Neat PMMA 90 79 81

cr
0.25 %NFC 22 21 29
0.25 %CNC 74 72 86
0.25 %NDC 50 58 75

us
0.5 %NFC 8 4.5 7
0.5 %CNC 56 55 69
0.5 % NDC 15 15 20

an
625
626 Table 3. Glass transition temperature (Tg), tan δ maximum peak values (Tan δ max. peak),
627 storage modulus at room temperature (SM at RT) and glass transition temperature (SM at Tg) for
628 neat PMMA and cellulose nanocomposites.
M
Group Name Tg (°C) Tan δ max. peak SM at RT (GPa) SM at Tg (GPa)
PMMA 71.87 (0.76) 0.99 (0.02) 2.73 (0.20) 0.023 (0.004)
0.25% NFC 75.51 (1.89) 0.96 (0.01) 2.77 (0.07) 0.024 (0.001)
d

0.25% CNC 78.67 (1.40) 0.81 (0.05) 2.80 (0.08) 0.034 (0.002)
0.25%NDC 74.35 (1.90) 0.96 (0.02) 3.04 (0.09) 0.030 (0.001)
te

0.5%NFC 72.25 (1.26) 0.86 (0.02) 2.97 (0.04) 0.032 (0.001)


0.5%CNC 76.41 (0.68) 0.65 (0.04) 2.88 (0.15) 0.060 (0.006)
0.5%NDC 72.72 (1.91) 0.97 (0.03) 2.91 (0.11) 0.030 (0.002)
p

629 Parenthesis indicates standard deviation.


ce

630 LIST OF FIGURES:


Ac

631 Figure 1. Photographs of the pure PMMA and CNF nanocomposites placed on a background
632 paper for demonstrating transparency.
633 Figure 2. UV–visible transmittance spectra of the neat PMMA and CNF nanocomposites.
634 Figure 3. TGA curves of the PMMA, cellulose nanofibers, and CNF nanocomposites.
635 Figure 4. a) The temperature at 10% and 50% M.L. and b) mass loss at 600 °C for neat PMMA,
636 cellulose nanofibers and CNF nanocomposites from their TGA curves, respectively.
637 Figure 5. DSC thermograms of the neat PMMA and CNF nanocomposites.
638 Figure 6. Stress-strain behavior and flexural strength (bars) and flexural modulus of elasticity
639 (lines) of neat PMMA and CNF nanocomposites, respectively.
640 Figure 7. FT-IR spectra of the neat PMMA and cellulose nanocomposites.
641 Figure 8. Optical micrographs of the neat PMMA and cellulose nanocomposites.
642

16
Page 16 of 22
643

PMMA+0.25wt% NFC PMMA+0.5wt% NFC

t
ip
cr
PMMA Sheet

us
an
M
d
p te
ce

PMMA+0.25wt% CNC PMMA+0.25wt% NDC


644
645 Figure 1. Photographs of the pure PMMA and CNF nanocomposites placed on a background
Ac

646 paper for demonstrating transparency.


647

17
Page 17 of 22
t
ip
cr
us
648
649

an
650
651 Figure 2. UV–visible transmittance spectra of the neat PMMA and CNF nanocomposites.
M
100
d

80
te
Mass Loss (%)

60
p

Neat PMMA
NFC
ce

CNC
40
NDC
0.25% NFC
0.25% CNC
20 0.25% NDC
Ac

0.5% NFC
0.5% CNC
0.5% NDC
0
100 200 300 400 500 600

652 Temperature (°C)


653 Figure 3. TGA curves of the PMMA, cellullose nanofibers, and CNF nanocomposites.
654

18
Page 18 of 22
380

360

(a)
Temperature (°C)

340

t
ip
320

300

cr
280

us
260
Temp. at 50% M.L.
Temp. at 10% M.L.
240
NFC

CNC

NDC

0.25% NFC

0.25% CNC

0.255 NDC

0.5% CNC

0.5% NDC
PMMA

0.5% NFC

655 Group Name an


M
100
d
Mass Loss at 600°C (%)

te

90
(b)
80
p
ce

70

60
Ac

50
NFC

CNC

NDC

0.25% NFC

0.25% CNC

0.255 NDC

0.5% NFC

0.5% CNC

0.5% NDC
PMMA

656 Group Name


657
658 Figure 4. a) The temperature at 10% and 50% M.L. and b) mass loss at 600 °C for neat PMMA,
659 cellulose nanofibers and CNF nanocomposites from their TGA curves, respectively.

19
Page 19 of 22
t
ip
cr
us
660
661 an
Figure 5. DSC thermograms of the neat PMMA and CNF nanocomposites.
M
662
663
664
665
d

666
667
te

668
p
ce
Ac

20
Page 20 of 22
6e+7

5e+7

t
Stress (Pa)

4e+7

ip
3e+7

cr
Neat PMMA
2e+7
0.25% NFC

us
0.25% CNC
0.25% NDC
1e+7 0.5% NFC
0.5% CNC
0.5% NDC

an
0
0 2 4 6 8 10

669 Strain (%)


M
Flexural Modulus of Elasticity (GPa)
3.0
d

60
Flexural Strength (MPa)

2.5
te

2.0

40
p

1.5
ce

1.0
20
Ac

0.5

0 0.0
0.25% NFC

0.25% CNC

0.25% NDC

0.5% NFC

0.5% CNC

0.5% NDC
PMMA

670 Group Name


671 Figure 6. Stress-strain behavior and flexural strength (bars) and flexural modulus of elasticity
672 (lines) of neat PMMA and CNF nanocomposites, respectively.
673
674

21
Page 21 of 22
t
ip
cr
us
an
675
676 Figure 7. FT-IR spectra of the neat PMMA and cellulose nanocomposites.
677
M
678
d
p te
ce
Ac

679
680 Figure 8. Optical micrographs of the neat PMMA and cellulose nanocomposites.
681

22
Page 22 of 22

You might also like