You are on page 1of 12

Journal of Natural Fibers

ISSN: 1544-0478 (Print) 1544-046X (Online) Journal homepage: https://www.tandfonline.com/loi/wjnf20

A New Approach for Conversion of Eucalyptus


Lignocellulosic Biomass into Cellulose
Nanostructures: A Method that Can Be Applied in
Industry

Alana G. de Souza, Giovanni F. De Lima, Rita C. L. B. Rodrigues, Ivana


Cesarino, Alcides L. Leão & Derval S. Rosa

To cite this article: Alana G. de Souza, Giovanni F. De Lima, Rita C. L. B. Rodrigues, Ivana
Cesarino, Alcides L. Leão & Derval S. Rosa (2019): A New Approach for Conversion of Eucalyptus
Lignocellulosic Biomass into Cellulose Nanostructures: A Method that Can Be Applied in Industry,
Journal of Natural Fibers, DOI: 10.1080/15440478.2019.1691125

To link to this article: https://doi.org/10.1080/15440478.2019.1691125

Published online: 27 Nov 2019.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=wjnf20
JOURNAL OF NATURAL FIBERS
https://doi.org/10.1080/15440478.2019.1691125

A New Approach for Conversion of Eucalyptus Lignocellulosic


Biomass into Cellulose Nanostructures: A Method that Can Be
Applied in Industry
Alana G. de Souzaa, Giovanni F. De Limaa, Rita C. L. B. Rodriguesb, Ivana Cesarinoc,
Alcides L. Leãoc, and Derval S. Rosaa
a
Engineering, Modeling and Applied Social Sciences Center (CECS), Federal University of ABC, Santo André, Brazil;
b
Department of Biotechnology, Lorena Engineering School, University of São Paulo, Lorena, Brazil; cSchool of
Agriculture, São Paulo State University (UNESP), Botucatu, Brazil

ABSTRACT KEYWORDS
This study aimed to explore the use of eucalyptus sawdust for the extrac- Bio-residue; cellulose
tion of nanocellulose. This residue was purified with a three-steps pretreat- nanostructures; morphology;
ment and ball milled for the isolation of nanocellulose. The method chemical composition;
crystallinity
proposed can be industrially applied due to the use of reactor and ball
mill and can be used for a variety of cellulose sources. Cellulose was the 关键词
main component detected in the eucalyptus residues, as observed after the 生物残渣; 纤维素纳米结
characterizations. The nanostructures showed an average size of 192 nm 构; 形态学; 化学成分; 结
and high crystallinity and thermal stability. The current study demonstrated 晶度
that the nanocellulose could be obtained to commercial-scale applications.

摘要
本研究旨在探索利用桉树锯末提取纳米纤维素的方法. 该残渣经三步预处
理和球磨分离得到纳米纤维素. 由于反应器和球磨机的使用,该方法可以
工业化应用,并可用于多种纤维素来源. 纤维素是桉树渣中的主要成分.
纳米结构的平均尺寸为192nm,具有较高的结晶度和热稳定性. 目前的研
究表明,纳米纤维素可以获得商业规模的应用.

Introduction
Eucalyptus is one of the essential forest genera commercially exploited worldwide. The diversity of
this genre among the wide adaptability to soils and climates and fast grown rate have made
Eucalyptus a vital source of biomass for paper pulp, fiberboard, charcoal industries and bioenergy
production (Barradas et al. 2018). During cutting, residual sawdust is generated. These residues are
composed of pieces of bark, leaves, and serrated trunks. In 2016, 47,8 million tons of Eucalyptus
sawdust was produced in Brazil. Its most common destination is the incineration (IBÁ 2017).
However, this wood biomass is composed of carbohydrate polymers (40-60% of cellulose and 20-
40% of hemicellulose), 15-40% of lignin and a smaller amount of extractives (Oksman et al. 2011).
The interest in reusing these residues aiming the reduction of environmental impacts related to its
disposal is growing in recent years. Considering its composition, Eucalyptus is a potential source for
the isolation of cellulose nanostructures (CNS) (Parize et al. 2017; Tonoli et al. 2012; Syverud et al.
2011; Ruiz Cuilty et al. 2018; Oksman et al. 2011).
Cellulose is a promising feedstock from lignocellulosic biomass, because it is a renewable,
biodegradable and biocompatible material, and its nanostructures have a high surface area,

CONTACT Derval S. Rosa dervalrosa@yahoo.com.br Engineering, Modeling and Applied Social Sciences Center (CECS),
Federal University of ABC, Santo André, Brazil
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/wjnf.
© 2019 Taylor & Francis
2 A. G. D. SOUZA ET AL.

lightweight, high aspect ratio and exceptional mechanical properties, like high stiffness and Young’s
Modulus (Tan, Hamid, and Lai 2015). Recently, CNSs has been attracting a great deal of research
interest due to the significant applicability and environmental benefits (Lu et al. 2014b).
Knowing the excellent properties of this biomass, this work presents a new approach for
converting lignocellulosic biomass of eucalyptus wastes into cellulose nanostructures. The pretreat-
ments were conducted in a reactor, which can treat 5 kg of biomass at a time. Also, the neutralization
and water withdrawal processes were conducted in the reactor, minimizing the total time spent and
labor. The isolation was performed in a ball mill, equipment known industrially for its ease of
handling, low cost, and high efficiency. This method also minimizes environmental impacts by not
using toxic reagents.
As reported by Reid et al., currently, the primary commercially nanocelluloses are obtained by
acid hydrolysis, which is harmful to the environment (Reid, Villalobos, and Cranston 2017). The
present work stands out for presenting a green methodology for obtaining nanocellulose, which can
be deployed in industries.
The efficiency of the proposed pretreatments for lignin and hemicelluloses removal were inves-
tigated by compositional analysis and Fourier transformed infrared spectroscopy.
Thermogravimetric analysis (TGA) and X-ray diffraction (XDR) were performed to evaluate the
changes in physicochemical properties of the samples after each pretreatment. Scanning electron
microscopy was used to evaluate the effect of the pretreatments on the morphology of the samples.
The nanoparticles obtained after ball milling were characterized, regarding particle size and mor-
phology, by dynamic light scattering (DLS) and atomic force microscopy (AFM). The Zeta-potential
analysis revealed electrochemical ability.

Experimental
Materials
Eucalyptus residues (Eucalyptus citriodora) were obtained after harvesting and cutting in Mato Grosso
(Brazil). Sodium Chlorite was purchased from Sigma-Aldrich. Sulfuric Acid, Sodium Hydroxide, and
Potassium Hydroxide were purchased from Labsynth. All reagents were of analytical grade.

Isolation of cellulose nanostructures


The Eucalyptus in natura (E) was dried in an oven at 60ºC for 24 h. After drying, the samples
were submitted to three pretreatments. In the first, the Eucalyptus residue was pretreated with
a sulfuric acid solution 0.8 wt.% (diluted acid hydrolysis). The mixture was carried out for
20 minutes at 133ºC. After that, the pretreated sample was washed and dried at 60ºC for 12 h
(sample named E-P1). Then, sample E-P1 was submitted to a new pretreatment (pretreatment 2)
in a sodium chloride solution (0,9 wt.%). The mixture was carried out for 2 h at 80ºC. After that,
the pretreated sample was washed and dried at 60ºC for 12 h (this sample was named E-P2).
Then, the sample was submitted to a third pretreatment in an aqueous solution of sodium
hydroxide (4 wt.%) and potassium hydroxide (4 wt.%). This mixture was carried out during
30 minutes at 80ºC. After that, it was washed dried at 60ºC for 12h (this sample was named
E-P3). After the chemical pretreatments, the pretreated material (E-P3) was mixed with ethanol
solution (80 wt.%), in the proportion 1:1 ml:g. This mixture was subjected to ball milling for 6 h
in a MA500/ball mill (Marconi) using alumina balls with weight ratio (balls/sample) of 70:1. This
sample is named CNS. The yield of the CNSs obtention was ~50 wt.% relative to Eucalyptus in
natura mass.
JOURNAL OF NATURAL FIBERS 3

Characterizations
Chemical analysis
The chemical composition of Eucalyptus in natura and pretreated samples was investigated accord-
ing to ASTM methods: lignin (ASTM D 1106–56), ash content (ASTM 1102–84), holocellulose
(ASTM D 1104–56) and alpha-cellulose (ASTM D 1103–60) (Lu et al. 2014a).

Fourier transformed infrared spectroscopy


Fourier transform infrared spectroscopy (FTIR) was performed using a Frontier 94942 (PerkinElmer,
USA) in the range from 400 to 4000 cm−1 with 64 scans and a resolution of 4 cm−1. It was used an
attenuated total reflectance accessory. Each sample was pressed into a thin film.

Thermogravimetric analysis
The thermal stability of Eucalyptus in natura, pretreated samples, and CNS was investigated using an
STA 6000 instrument (PerkinElmer, USA). The samples were heated from 20 to 600°C with
a heating rate of 20 °C/min under a nitrogen flow. The initial decomposition temperatures,
measured at 10% weight loss, were used to compare the thermal stabilities. The amount of sample
for each measurement was about 5.0 mg.

Scanning electron microscopy


The morphology of the samples was analyzed by scanning electron microscopy (SEM). The sample,
which was sputter-coated with gold, was examined in a JEOL – JCM 6010, operating with an
accelerating voltage of 5 kV.

Dynamic light scattering and zeta potential


The particle size of CNS dispersion in water was recorded by dynamic light scattering (DLS) system
using ALV-CGS3. The sample is illuminated by a polarized HeNe laser beam (22mW, λ = 633 nm)
and followed by detection of the resultant fluctuation of scattered light at 90º by an APD pair of
detectors operating in pseudo-correlation.
Zeta potential measurements were conducted using a Zetasizer Nano-ZS (Malvern Instruments)
to characterize the surface charge property of nanoparticles. The measurement was based on the
electrophoretic mobility of the particles, which was converted to zeta potential using the Helmholtz-
Smoluchowski equation (Daltin 2011).

X-ray diffraction
X-ray analysis determined the crystallinity indexes of all samples. The experiments were performed
on a STADI-P (Stoe), equipped with CuKα (λ = 1,54056 Å), in the 2θ range 0-40º, and a counting
time of 100 s at each 1.05º. The crystallinity index was calculated according to the method proposed
by Segal (Segal et al. 1959).

Atomic force microscopy


The morphologies were analyzed by AFM. For images examination, a droplet of the aqueous
suspension was sprayed onto a freshly cleaned silicon wafer and then airdried during 24 h, at
room temperature. The AFM experiments were performed in tapping mode, using an Agilent
4 A. G. D. SOUZA ET AL.

equipment (model 5500), at ambient relative humidity and temperature. The scan was done at a rate
of 0.5 Hz, with an image resolution of 256 × 256 pixel, using a silicon cantilever of constant force
between 25 and 75 N.m−1 and a resonance frequency of 120 kHz.

Results and discussion


Chemical analysis
Chemical analysis was carried out to evaluate the efficiency of the pretreatments. Table 1 shows the
composition of the studied sample after each pretreatment. The results show that eucalyptus in
natura contains 7.4% of extractives, 15% of hemicellulose, 23% of lignin, 55% of cellulose, and 1% of
ashes. This suggests that cellulose, lignin, and hemicellulose are the main constituents of this
biomass, and cellulose is the predominant polysaccharide (Pereira et al. 2013)
The three pretreatments were applied to remove the non-cellulosic components by the oxidation
and solubilization of these materials. After these pretreatments, the sample E-T3 contained 4% of
extractives, 1% of hemicellulose, 5% of lignin, 85% of cellulose, and 5% of ashes. It was possible to
observe a significant removal of the non-cellulosic materials by the studied methodology (94% of
hemicellulose and 72% of the total lignin content, regarding mass fraction) (Lu et al. 2014a). The
third step of pretreatments, which is an alkali pretreatment, resulting in the huge reduction of the
hemicellulose and lignin content due to its solubilization. The pretreatments were useful to increase
the surface area of the eucalyptus to make this more vulnerable to the isolation by ball mill (Ilyas,
Sapuan, and Ishak 2017). The yield of the process of removal of lignocellulosic components is
approximately 50%. Figure 1 shows a scheme representing the effects of the pretreatments on the
Eucalyptus fiber.

Fourier transformed infrared spectroscopy


The FTIR spectroscopy monitors the chemical structure by identifying the functional groups present
in each sample. The FTIR spectra of eucalyptus in natura and after each pretreatment are shown in
Figure 2. The spectra showed the typical band pattern for cellulose, hemicellulose, and lignin moiety.

Table 1. Chemical composition (mass percentage).


Insoluble Soluble
Cellulose Hemicellulose Lignin Lignin Extracts Ashes
Samples (%) (%) (%) (%) (%) (%)
Ea 54.9 ± 1.4 15.3 ± 1.2 21.4 ± 1.8 1.9 ± 0.9 7.4 ± 1.1 0.7 ± 0.02
E-P1a 64.1 ± 1.9 7.0 ± 0.9 18.1 ± 1.2 2.7 ± 0.9 5.9 ± 2.0 0.2 ± 0.1
E-P2a 73.5 ± 0.8 3.0 ± 0.1 10.8 ± 0.9 1.9 ± 0.8 5.1 ± 1.2 0.2 ± 0.1
E-P3a 84.6 ± 0.5 15.3 ± 1.2 3.9 ± 1.1 1.1 ± 0.2 4.0 ± 1.0 0.20 ± 0.02
a = E (Eucalyptus in natura); E-P1 (sample after pretreatment 1); E-P2 (sample after pretreatment 2); E-P3 (sample after
pretreatment 3)

Figure 1. Morphology of the samples after each pretreatment.


JOURNAL OF NATURAL FIBERS 5

Figure 2. FTIR spectra of all samples: eucalyptus in natura, pretreated samples, and CNS.

The broad bands at 3440 and 2920 cm−1 correspond to the stretching of -OH groups and the
C-H stretching of methyl and methylene groups of lignin, cellulose, and hemicellulose (Maheswari
et al. 2012; Lu et al. 2014a; Oksman et al. 2011). The band at 1730 cm−1 corresponds to the carbonyl
groups (C = O) due to the presence of acetyl ester and carbonyl aldehyde groups of hemicellulose
and lignin. This band disappeared entirely in the pretreated samples because of the removal of the
most non-cellulosic components by applied chemical extraction (Maheswari et al. 2012; Lu et al.
2014a). The other bands related to the aromatic ring of lignin (guaiacyl and syringyl rings) vibrating
at 1602 and 1515 cm−1 decreased after the chemical treatments, as expected (Ferreira et al. 2018).
The band at 1240 cm−1 corresponds to the -COO vibration of acetyl groups in hemicellulose (Reddy
et al. 2013; Lu et al. 2014a; Oksman et al. 2011). This band disappears after the chemical pretreat-
ments and in the CNS samples due to the most removal of hemicellulose. All samples showed the
presence of a predominant peak at 1115 cm−1 that corresponds to C-O stretching vibration of
cellulose, indicating the presence of cellulose as an essential content in the isolated CNSs (Kalita et al.
2015). A sharp peal observed at 1054 cm−1 corresponds to the stretching vibration of the
C-O-C pyranose ring (antisymmetric in phase ring) of cellulose molecule (Tan, Hamid, and Lai
2015). Fingerprint region of the spectra which lies in the range of 1100 cm−1 to 600 cm−1 indicated
the vibration of wagging, deformation and twisting modes of anhydro-glucopyranose which is
characteristic of β-glucosidic linkages (Johar, Ahmad, and Dufresne 2012; Mariano, Kissi, and
Dufresne 2016; Tan, Hamid, and Lai 2015).

Thermogravimetric analysis
Thermal stability of the eucalyptus in natura and the samples after each pretreatment were
investigated by the thermogravimetric method and the thermograms are shown in Figure 3(a) and
the derivative curves are presented in Figure 3(b). For all samples, the thermal decomposition
occurred in two main steps. This phenomenon can be observed in other reports related to the
lignocellulosic materials (Tan, Hamid, and Lai 2015; Tonoli et al. 2012). The first stage, below 100°C,
corresponds to the evaporation of low mass components and water loosely bound to the surfaces of
cellulose (CS, George, and Narayanankutty 2016; Movva and Kommineni 2017). The second stage,
which occurs above 220°C until approximately 400°C, corresponds to the degradation of
6 A. G. D. SOUZA ET AL.

Figure 3. Thermograms of the samples after each pretreatment (a) TGA curves and (b) DTG curves.

hemicelluloses, the breakdown of glycosidic linkage of cellulose and the begin of lignin degradation
(CS, George, and Narayanankutty 2016). These results are in good accordance with the work done by
Tan and his coworkers (Tan, Hamid, and Lai 2015). It is possible to observe that ball mill did not
alter thermal stability from E-T3 sample to CNS sample, indicating that this isolation methodology is
not harmful to the nanostructure. The increase on the thermal stability can be associated with the
rise of the crystallinity of the fiber, which acted as the thermal degradation resistant at high
temperature (this result will be discussed ahead) (CS, George, and Narayanankutty 2016; Johar,
Ahmad, and Dufresne 2012; Phanthong et al. 2016).

Scanning electron microscopy


Figure 4 reveals the morphology of the samples E, E-P1, E-P2, and E-P3. It can be observed that
Eucalyptus fiber in natura (Figure 4(a)) is a dense and uneven material (Ruiz Cuilty et al. 2018). Due
to the presence of binding and protection materials, like hemicelluloses and lignin, fibrils cannot be
distinguished (CS, George, and Narayanankutty 2016; Maheswari et al. 2012). After each
JOURNAL OF NATURAL FIBERS 7

Figure 4. SEM images of (a) E; (b) E-P1; (c) E-P2, and (d) E-P3.

pretreatment (Figure 4(b-d)), it is clear to see that the samples become more exposed, and porous
and it is possible to observe oriented fibers. That variations in the morphologies occur due to the
removal of the lignin and hemicelluloses, which was further supported by the chemical analysis data
given in Table 1.

Dynamic light scattering and zeta potential


Dynamic light scattering (DLS) is popularly used for the determination of the size of particles in
suspension. The advantage of this analysis includes the measure of nano sizes and the fact that it is
an easy and quick measure (CS, George, and Narayanankutty 2016). The size distribution of the
cellulose nanoparticles is given in Figure 5. It is possible to observe that the majority of the isolated
structures are in the nano-range, with an average size of 192 nm. DLS analysis also showed
a minimum particle size of 61 nm and maximum particle size of 1602 nm. Sofla et al. isolated
nanocellulose from sugarcane bagasse and obtained an average particle size in the range of 100 to
1000 nm, which is similar to the results obtained in the present work (Rahimi Kord et al. 2016). The
yield of the CNSs obtained by the mechanical method was 45 wt.%.
Zeta potential (δ) analysis was carried out to evaluate the electrostatic stabilization of the CNS
particles in suspension (Alves et al. 2015). Higher absolute values for δ indicate a higher capacity of
dispersion in water (Lu et al. 2014a). According to Mohaiyiddin and his coworkers, the agglomera-
tion of cellulose nanoparticles will occur if zeta potential absolute value is lower than 15 mV and the
same authors reported that a nanocellulose suspension could be considered stable if its zeta potential
absolute value is higher than 30 mV (Mohaiyiddin et al. 2016). The δ result for CNS sample isolated
8 A. G. D. SOUZA ET AL.

Figure 5. Particle size distribution for CNS sample.

in this work is −20 ± 1 mV. These results are in agreement with the commercialized nanoparticles, as
observed by Reid et al (Reid, Villalobos, and Cranston 2017).

X-ray diffraction
XRD patterns for the sample after each pretreatment and for the cellulose nanostructures are shown
in Figure 6. It can be observed that all samples showed diffraction peaks at 16° and 22°, which are
associated to crystallographic plans (110) and (200), respectively (Cheng et al. 2015; Ferreira et al.
2018). These peaks, related to crystals of cellulose I, are well described in the literature (Ferreira et al.
2018). After the pretreatments and the CNSs isolation, the diffractograms presented a sharp peak at
22,5°, representing cellulose I structure (Oksman et al. 2011). The evolution of the crystalline peaks
at the end of the isolation pretreatments is expected and is similar to other such reported materials

Figure 6. XRD patterns for the sample after each pretreatment and for CNS.
JOURNAL OF NATURAL FIBERS 9

Figure 7. AFM images of the cellulose nanostructures.

(Kalita et al. 2015). These results prove the effectiveness of the proposed methodology to obtain
crystalline cellulose.
Crystallinity index (CI) was calculated for all samples (26,6% for E; 69,6% for E-P1; 51,1% for
E-P2; 68,9% for E-P3 and 63,2% for CNS). The decrease of the CI from E-P3 to CNS can be justified
by the break of the crystalline region of cellulose during ball milling (Rahimi Kord et al. 2016). The
purpose of the ball milling is to isolate cellulose nanostructures by breaking the amorphous region of
the cellulose fiber, but it can be observed that part of the crystalline region is also broken.

Atomic force microscopy


The morphology and the dimensions of the samples were studied by AFM analysis. Figure 7 shows
AFM images for the CNS obtained after ball milling. The sample exhibited a rod-like shape, just like
the nanowhiskers reported in other works and the CNSs produced by different commercial produ-
cers (Oksman et al. 2011; Reid, Villalobos, and Cranston 2017). The end of the nanoparticles
appeared as irregular and grooved, as a result of the irregular grinding during the ball mill. Other
authors observed this behavior due to the different isolations methods of nanocellulose (Domingues
et al. 2016).
Figure 8 shows diameter (D) and length (L) distributions evaluated over 50 particles obtained by
several AFM images. These values are very similar to the hydrodynamic radius estimated by DLS
analysis. Such variations have been observed previously when both techniques were used for CNSs
size estimations (Kalita et al. 2015). It can be found an average diameter size of 107 nm and average
length size of 920 nm. The L/D aspect ratio for the CNS was 8.6. This result is consistent with values

Figure 8. (a) Diameter and (b) length distribution histograms for CNS.
10 A. G. D. SOUZA ET AL.

encountered in literature, Tonoli and his co-works obtained nanocellulose from Eucalyptus kraft
pulp with an aspect ratio of 10 (Tonoli et al. 2012). The most significant values of nanosizes,
comparing with other methodologies, can be attributed to the non-addition of chemical reagents in
the isolation process (Mascheroni et al. 2016).

Conclusions
A residue from forest trees such as eucalyptus is an exciting alternative as cellulose source for several
applications. Also, this work presents an useful and practical process to reuse this type of material is
the isolation of cellulose nanostructures after the use of chemical pretreatments for the lignin and
hemicellulose removal. The presented methodology showed similarities with industrially produced
CNSs, such as thermal properties, morphology, and crystallinity. The ball mill is an excellent
alternative to isolate CNSs due to its low cost and environmental impacts. The isolated CNSs
revealed excellent thermal stability, which is unusual in terms of applicability. In terms of dimen-
sions, the results derived from DLS and zeta potential confirmed that the CNSs are in the nanoscale,
with an average of ~200 nm and zeta potential of – 20 mV. The crystallinity of the pretreated
samples was higher than that of the eucalyptus in natura due to the removal of amorphous
components (i.e., lignin and hemicellulose). The morphology of the CNSs was cellulose nanocrystals,
with an aspect ratio of 8.6. This work demonstrated that with proper characterization, the transition
from lab scale to industrial production should be met with optimism, and another isolation method
can be applied.

Acknowledgments
The authors thank the UFABC, REVALORES Strategic Unit, CNPq (nº 306401/2013-4) and Multiuser Central
Facilities (UFABC).

Funding
This work was supported by the Conselho Nacional de Desenvolvimento Científico e Tecnológico [306401/2013-4].

References
Alves, L., B. Medronho, F. E. Antunes, M. P. Fernández-García, J. Ventura, J. P. Araújo, A. Romano, and B. Lindman.
2015. Unusual extraction and characterization of nanocrystalline cellulose from cellulose derivatives. Journal of
Molecular Liquids 210:106–12. doi:10.1016/j.molliq.2014.12.010.
Barradas, C., G. Pinto, B. Correia, B. B. Castro, A. J. L. Phillips, and A. Alves. 2018. Drought X disease interaction in eucalyptus
globulus under neofusicoccum eucalyptorum infection. Plant Pathology 67:87–96. doi:10.1111/ppa.12703.
Cheng, D., Y. Wen, L. Wang, A. Xingye, X. Zhu, and N. Yonghao. 2015. Adsorption of polyethylene glycol (PEG) onto cellulose
nano-crystals to improve its dispersity. Carbohydrate Polymers 123:157–63. Elsevier Ltd. doi:10.1016/j.carbpol.2015.01.035.
CS, C. J., N. George, and S. K. Narayanankutty. 2016. Isolation and characterization of cellulose nanofibrils from
arecanut husk fibre. Carbohydrate Polymers 142:158–66. doi:10.1016/j.carbpol.2016.01.015.
Cuilty, R., L. Karla, E. Ballinas-Casarrubias, R. D. S. Miguel, J. de Gyves, J. C. Robles-Venzor, and G. González-
Sánchez. 2018. Cellulose recovery from Quercus Sp. sawdust using ethanosolv pretreatment. Biomass and Bioenergy
111 (September 2016):114–24. doi:10.1016/j.biombioe.2018.02.004.
Daltin, D. 2011. Tensoativos: Química, Propriedades E Aplicações. São Paulo: Blucher. doi:541.34514.
Domingues, A. A., F. V. Pereira, M. Rita, O. J. Rojas, and D. F. S. Petri. 2016. Interfacial properties of cellulose
nanoparticles obtained from acid and enzymatic hydrolysis of cellulose. Cellulose Springer Netherlands.
doi:10.1007/s10570-016-0965-3.
Ferreira, F. V., M. Mariano, S. C. Rabelo, R. F. Gouveia, and L. M. F. Lona. 2018. Isolation and surface modification of
cellulose nanocrystals from sugarcane bagasse waste: from a micro- to a nano-scale view. Applied Surface Science
436:1113–22. Elsevier B.V. doi:10.1016/j.apsusc.2017.12.137.
IBÁ. 2017. Anuário Estatístico Da Indústria Brasileira de Árvore - 2017. Indústria Brasileira De Árvores. doi:10.1017/
CBO9781107415324.004.
JOURNAL OF NATURAL FIBERS 11

Ilyas, R. A., S. M. Sapuan, and M. R. Ishak. 2017. Isolation and characterization of nanocrystalline cellulose from sugar
palm fibres (Arenga Pinnata). Carbohydrate Polymers Elsevier Ltd. doi:10.1016/j.carbpol.2017.11.045.
Johar, N., I. Ahmad, and A. Dufresne. 2012. Extraction, preparation and characterization of cellulose fibres and
nanocrystals from rice husk. Industrial Crops and Products 37 (1):93–99. Elsevier B.V. doi:10.1016/j.
indcrop.2011.12.016.
Kalita, E., B. K. Nath, F. Agan, V. More, and P. Deb. 2015. Isolation and characterization of crystalline, autofluor-
escent, cellulose nanocrystals from saw dust wastes. Industrial Crops and Products 65:550–55. doi:10.1016/j.
indcrop.2014.10.004.
Lu, Q., W. Lin, L. Tang, S. Wang, X. Chen, and B. Huang. 2014a. A mechanochemical approach to manufacturing
bamboo cellulose nanocrystals. Journal of Materials Science 50 (2):611–19. doi:10.1007/s10853-014-8620-6.
Lu,, Q. lin, L. R. Tang, S. Wang, B. Huang, Y. D. Chen, and X. R. Chen. 2014b. An investigation on the characteristics
of cellulose nanocrystals from pennisetum sinese. Biomass and Bioenergy 70:267–72. Elsevier Ltd. doi:10.1016/j.
biombioe.2014.09.012.
Maheswari, C. U., K. Obi Reddy, E. Muzenda, B. R. Guduri, and A. Varada Rajulu. 2012. Extraction and
Characterization of cellulose microfibrils from agricultural residue E cocos nucifera L . Biomass and Bioenergy
46:555–63. Elsevier Ltd. doi:10.1016/j.biombioe.2012.06.039.
Mariano, M., N. Kissi, and A. Dufresne. 2016. Structural reorganization of CNC in injection-molded CNC/PBAT
materials under thermal annealing. Langmuir 32 (39):10093–103. doi:10.1021/acs.langmuir.6b03220.
Mascheroni, E., R. Rampazzo, M. A. Ortenzi, G. Piva, S. Bonetti, and L. Piergiovanni. 2016. Comparison of cellulose
nanocrystals obtained by sulfuric acid hydrolysis and ammonium persulfate, to be used as coating on flexible
food-packaging materials. Cellulose 23 (1):779–93. Springer Netherlands. doi:10.1007/s10570-015-0853-2.
Mohaiyiddin, M. S., O. H. Lin, W. T. Owi, C. H. Chan, C. H. Chia, S. Zakaria, A. R. Villagracia, and H. M. Akil. 2016.
Characterization of nanocellulose recovery from elaeis guineensis frond for sustainable development. Clean
Technologies and Environmental Policy 18 (8):2503–12. Springer Berlin Heidelberg. doi:10.1007/s10098-016-1191-2.
Movva, M., and R. Kommineni. 2017. Extraction of cellulose from pistachio shell and physical and mechanical
characterisation of cellulose-based nanocomposites. Materials Research Express 4 (4):045014. IOP Publishing.
doi:10.1088/2053-1591/aa6863.
Oksman, K., J. A. Etang, A. P. Mathew, and M. Jonoobi. 2011. Cellulose nanowhiskers separated from a bio-residue
from wood bioethanol production. Biomass and Bioenergy 35 (1):146–52. Elsevier Ltd. doi:10.1016/j.
biombioe.2010.08.021.
Parize, D., D. da Silva, J. E. de Oliveira, T. Williams, D. Wood, R. D. J. Avena-Bustillos, A. P. Klamczynski,
G. M. Glenn, J. M. Marconcini, and L. H. C. Mattoso. 2017. Solution blow spun nanocomposites of poly(lactic
acid)/cellulose nanocrystals from eucalyptus kraft pulp. Carbohydrate Polymers 174:923–32. Elsevier Ltd.
doi:10.1016/j.carbpol.2017.07.019.
Pereira, B. L. C., A. D. C. O. Carneiro, A. M. M. L. Carvalho, J. L. Colodette, A. C. Oliveira, and M. P. F. Fontes. 2013.
Influence of chemical composition of eucalyptus wood on gravimetric yield and charcoal properties. BioResources 8
(3):4574–92. http://ojs.cnr.ncsu.edu/index.php/BioRes/article/view/BioRes_08_3_4574_Pereira_Eucalyptus_
Gravimetric_Yield.
Phanthong, P., G. Guan, M. Yufei, X. Hao, and A. Abudula. 2016. Effect of ball milling on the production of
nanocellulose using mild acid hydrolysis method. Journal of the Taiwan Institute of Chemical Engineers
60:617–22. doi:10.1016/j.jtice.2015.11.001.
Rahimi Kord, S. M., R. J. Brown, T. Tsuzuki, and T. J. Rainey. 2016. A comparison of cellulose nanocrystals and
cellulose nanofibres extracted from bagasse using acid and ball milling methods. Advances in Natural Sciences:
Nanoscience and Nanotechnology 7 (3). doi: 10.1088/2043-6262/7/3/035004.
Reddy, M. M., S. Vivekanandhan, S. K. B. Manjusri Misra, and A. K. Mohanty. 2013. Biobased plastics and
bionanocomposites: Current status and future opportunities. Progress in Polymer Science 38 (10–11):1653–89.
Elsevier Ltd. doi:10.1016/j.progpolymsci.2013.05.006.
Reid, M. S., M. Villalobos, and E. D. Cranston. 2017. Benchmarking cellulose nanocrystals: From the laboratory to
industrial production. Langmuir 33 (7):1583–98. doi:10.1021/acs.langmuir.6b03765.
Segal, L., J. J. Creely, A. E. Martin, and C. M. Conrad. 1959. An empirical method for estimating the degree of
crystallinity of native cellulose using the X-ray diffractometer. Textile Research Journal 29 (10):786–94. doi:10.1177/
004051755902901003.
Syverud, K., G. Chinga-Carrasco, J. Toledo, and P. G. Toledo. 2011. A comparative study of eucalyptus and pinus
radiata pulp fibres as raw materials for production of cellulose nanofibrils. Carbohydrate Polymers 84 (3):1033–38.
Elsevier Ltd. doi:10.1016/j.carbpol.2010.12.066.
Tan, X. Y., S. B. A. Hamid, and C. W. Lai. 2015. Preparation of high crystallinity cellulose nanocrystals (CNCs) by
ionic liquid solvolysis. Biomass and Bioenergy 81:584–91. Elsevier Ltd. doi:10.1016/j.biombioe.2015.08.016.
Tonoli, G. H. D., E. M. Teixeira, A. C. Corrêa, J. M. Marconcini, L. A. Caixeta, M. A. Pereira-Da-Silva, and
L. H. C. Mattoso. 2012. Cellulose micro/nanofibres from eucalyptus kraft pulp: Preparation and properties.
Carbohydrate Polymers 89 (1):80–88. Elsevier Ltd. doi:10.1016/j.carbpol.2012.02.052.

You might also like