Sei sulla pagina 1di 10

G Model

ARTICLE IN PRESS

INDCRO-8646; No. of Pages 10

Industrial Crops and Products xxx (2016) xxxxxx

Contents lists available at ScienceDirect

Industrial Crops and Products


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

A comprehensive approach for obtaining cellulose nanocrystal from


coconut ber. Part I: Proposition of technological pathways
Diego M. do Nascimento a, , Jessica S. Almeida b , Maria do S. Vale c , Renato C. Leito c ,
Celli R. Muniz c , Maria Clea B. de Figueirdo c , Joo Paulo S. Morais d ,
Morsyleide de F. Rosa c,
a

Federal University of Cear (UFC), Department of Chemical, Bloco 940, CEP 60455-760 Fortaleza, Cear, Brazil
State University of Cear (UECE), Campus do Itapery, CEP 60455-900 Fortaleza, Cear, Brazil
c
Embrapa Tropical Agroindustry, Rua Dra Sara Mesquita 2270, Planalto do Pici, CEP 60511-110 Fortaleza, Cear, Brazil
d
Embrapa Cotton, Rua Oswaldo Cruz 1143, Centenrio, CEP 58428-095 Campina Grande, Paraba, Brazil
b

a r t i c l e

i n f o

Article history:
Received 30 October 2015
Received in revised form
21 December 2015
Accepted 28 December 2015
Available online xxx
Keywords:
Cellulose nanocrystal
Nanotechnology
Lignin
Green chemistry
Biorenery

a b s t r a c t
The high lignin content in the unripe coconut ber limits the use of this biomass as a cellulose nanocrystal source compared to other cellulose-rich materials. The aim of this study was to obtain lignin and
biomethane, and evaluate different approaches for extracting cellulose nanocrystal from unripe coconut
coir ber. The environmental evaluation of these approaches is presented in the second part of this
paper. Lignin was extracted by acetosolv pulping and cellulose by alkaline hydrogen peroxide bleaching
respectively. Were evaluated the biochemical methane potential of the efuents resulting from acetosolv pulping as well as the lignin concentration. Cellulose nanocrystals were prepared from cellulose
pulp via four methods: acidic hydrolysis with high acid concentration, acidic hydrolysis with low acid
concentration, ammonium persulfate oxidation, and high-power ultrasound. The cellulose nanocrystals
were analyzed by FTIR spectroscopy, X-ray diffraction, transmission electron microscopy, and TG analysis. Using these methods, the whole coconut ber could be used to produce cellulose nanocrystals and
lignin. Among the proposed methods, high-power ultrasound showed the highest efciency in cellulose
nanocrystal extraction.
2016 Elsevier B.V. All rights reserved.

1. Introduction
Cellulose nanocrystal is an abundant potential nanomaterial
that can be extracted from many renewable sources (Dufresne,
2013). This nanostructure has attracted attention for application in
several different areas because of its extraordinary physical properties, biodegradability, biocompatibility, and low cytotoxicity (Jor
and Foster, 2015; Rojo et al., 2015). Cellulose nanocrystals can be
obtained from wood, non-wood bers, algae, tunicates, and agroindustrial biomass, among other sources (Li et al., 2015).
Among the sources, lignocellulosic agroindustrial byproducts
are the most promising because of their low cost and availability.
Examples include sugarcane bagasse (Li et al., 2012a; Mandal and
Chakrabarty, 2014), corn straw (Huntley et al., 2015), palm-pressed
mesocarp ber (Souza et al., 2015), sisal (Rodrigues et al., 2015),
pineapple leaves (Deepa et al., 2015), cotton linter (Morais et al.,

Corresponding author.
E-mail address: morsyleide@gmail.com (M.d.F. Rosa).

2013), banana pseudostem (Pereira et al., 2014), banana peel, and


unripe coconut husk (Fahma et al., 2011; Nascimento et al., 2014;
Rosa et al., 2010). However, the prot ability of new biobased industries highly depends on integrating biomass conversion processes
to produce a range of fuels, power, materials and chemicals.
Worldwide coconut production was estimated to be approximately 62 million tons in 2013, of which nearly 3 million tons
were harvested in Brazil (FAOStat, 2015). Coconut water is the main
product of coconut crops. It is usually consumed directly from the
fruit in coastal cities or is bottled for shipping to inland locations.
Both processes generate large amounts of unripe coconut coir that
if not correctly collected and disposed of, causes environmental
problems, reducing the useful lifetime of landlls or causing water
pollution.
Unripe coconut bers can be extracted and used to manufacture several products, such as reinforced polymeric composites,
reinforced cement and concrete, geotextile fabrics and screens,
and wood-replacement berboards. Other alternatives for adding
value and reducing the disposal issues of such biomass involve
the extraction the cellulose nanocrystals. However, there are

http://dx.doi.org/10.1016/j.indcrop.2015.12.078
0926-6690/ 2016 Elsevier B.V. All rights reserved.

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model
INDCRO-8646; No. of Pages 10

ARTICLE IN PRESS
D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

two main limitations to performing this fractionation: high


power and reagent consumption to remove lignin and low
yield of cellulose nanocrystals from the original raw material
(Figueirdo et al., 2012). In this sense, all feasible approaches
for obtaining cellulose nanocrystals must also add value to other
macromolecules, such as lignin and hemicellulose.
Lignin is traditionally considered to be a troublesome waste and
is typically burned in mill boilers to generate power (Cotana et al.,
2014). Despite this traditional use, lignin has gained attention for
industrial applications because of its amorphous nature and highly
aromatic molecular structure. Thus, lignin can be used as a raw
material for producing bulk and ne chemicals, such as vanillin,
gallic acid, oils, phenols, acetic acid, lms, polyurethanes, carbon
bers, and other materials (Norgren and Edlund, 2014). Strassberger et al. (2014) report that the low price of natural gas and
the diversity of alternatives for lignin conversion is an atractive scenario to use lignin as a chemical instead of a simple fuel to be burned
in the boilers. The versatility of ne chemicals that can be produced
from lignin makes this biomacromolecule as lower carbon footprint alternative to oil, reducing environmental impacts related to
climate change (McDevitt and Grigsby, 2014). In addition to lignin
and cellulose, the coconut ber contains high hemicellulose levels
that typically are hydrolyzed during the extraction processes. Such
compounds are present in the efuents of cellulose pulp and can
be used as carbon sources for anaerobic fermentation and methane
production. Methane has a caloric power (50 MJ/kg) higher than
lignin (21 MJ/kg) and hemicelluloses (16 MJ/kg). The gas can be
burned to generate power for the nanocrystal extraction process,
reducing biogas production and emission in the efuents.
The aim of the present study was to develop new approach for
unripe coconut, allowing for the sustainable extraction of cellulose
nanocrystal and to add value to the extraction byproducts. Lignin
was recovered and pulping efuents were fermented to produce
methane. Here, the properties of the extracted cellulose nanocrystal are reported. In the Part II, the environmental impacts of the
extraction methods are evaluated with respect to life cycle analysis.

2. Experimental
2.1. Materials
Unripe coconut ber was provided by Embrapa Agroindstria
Tropical (Fortaleza, CE, Brazil). All chemicals were of analytical
grade and were used without further purication: 97% (w/w) NaOH,
30% (w/w) H2 O2 , 99.7% (w/w) CH3 COOH, 80% (w/w) NaClO2 , and
98% (w/w) H2 SO4 (Vetec Qumica Fina Ltda/SigmaAldrichDuque
de Caxias, RJ, Brazil).

2.2. Coconut ber fractioning


The fractionation owchart is presented in Fig. 1. Coconut
ber was ground in a Willye knife mill (STAR FT-80; Fortinox,
Piracicaba/SP, Brazil) with a 1-mm-large sieve. Delignication was
performed as described by Nascimento et al. (2014) with slight
modications.
Briey, after grinding, the ber was added to acetosolv solution
containing 93% (w/w) acetic acid and 0.3% (w/w) HCl in the ratio of
1:20 (w/v), heated under continuous reux and stirring for 3 h. After
pulping, the brous material was ltered through a Whatman no
2 paper lter and rinsed with fresh acetosolv solution at 80 C. The
black liquor, this efuent from this step, was stored for later lignin
recovery. The delignied cellulose pulp was rinsed with water until
the pH was neutral.
The ber was bleached in the proportion of 1:20 (w/v). First, the
pulp was stirred for 90 min at 50 C with 5% (w/w) H2 O2 and 3.8
(w/w) NaOH. The bleached pulp was ltered through a Whatman no
2 paper lter and rinsed with distilled water. This bleaching process
was repeated once more. Finally, the bers were stirred with 5.7%
(w/w) KOH for 120 min at 90 C, ltered through a Whatman no 2
paper lter, and rinsed with distilled water to obtain the cellulose
bleached pulp (Fahma et al., 2011).
The pulping black liquor was concentrated in a rotary evaporator (R-210/215; Buchi, Flawil, Switzerland), diluted 10 in hot
water (80 C), ltered through Whatman no 2 paper lter, and
rinsed with distilled water to recover lignin. The recovered lignin
was stored in a silica-gel desiccator until characterization analyses
(Morandium-Gianetti et al., 2012). The yield of recovered lignin,
the percentage of Klason lignin in the ber that was successfully
recovered, was calculated using Eq. (1):
%Yrecovered lignin =

mrecovered lignin
mber %LKber

100

(1)

where mrecovered lignin is the weight of the recovered lignin, mber


is the weight of the coconut ber and%LKber is the lignin Klason
contents of the samples.
The pulping efuent of the lignin recovery (efuent A) (172 mL
of efuent/1 g of coconut ber) and the mixed efuent of bleaching and lignin recovery (efuent AB) were collected (397 mL
of efuent/1 g of coconut ber) and analyzed to determine the
biodegradability and biochemical methane potential (BMP) assays
via anaerobic digestion.
2.3. Cellulose nanocrystal extraction
Cellulose nanocrystal extraction was performed using four
different methods: two acidic hydrolysis-based extractions, one

Fig. 1. General owchart of coconut ber fractionation.

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model

ARTICLE IN PRESS

INDCRO-8646; No. of Pages 10

D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

ammonium persulfate (APS) oxidation-based extraction, and one


high-power ultrasound-based extraction (Fig. 1). The yield of the
extraction was estimated as the ratio of the cellulose nanocrystals
mass and the cellulose mass in the ber used in the extraction step,
using Eq. (2):
%Ynanocrystals =

mnanocrystals
mpulp

%Ccellulose
100

100

(2)

where mnanocrystals is the weight of the extracted nanocrystals,


mpulp is the weight of the cellulose pulp (bleached pulp for CNH1,
CNH2 and CNU methods, and delignied pulp for CNO method)
and%C-cellulose is the -cellulose contents of the samples.
2.3.1. Acidic hydrolysis
Acidic hydrolysis was performed in the proportion of 1:20
(w/v) under stirring in two different conditions of H2 SO4 : (i)
CNH1, with low H2 SO4 concentration (44% w/w), long reaction
time (360 min), and high temperature (60 C); (ii) CNH2, with high
H2 SO4 concentration (60% w/w), short reaction time (45 min), and
low temperature (45 C) (Nascimento et al., 2014; Pereira et al.,
2014).
Cellulose nanocrystal cellulose extracted by both processes
(CNH1 and CNH2) were diluted by 10-fold and centrifuged (CR
22GIII; Hitachi, Tokyo, Japan) with deionized water for 15 min at
13 krpm (26.4 kg) until the supernatant became turbid (three centrifugation steps). Thereafter, the suspension was dialyzed in still
distilled water for 72 h, ultra-sonicated (UP400S; Hielscher, Teltow,
Germany) for 5 min, and freeze-dried. The nal yield was calculated
as the percentage of cellulose nanocrystal mass compared to the
initial bleached pulp mass (Eq. (2)).
2.3.2. APS oxidation
Delignied pulp was added to a 18.5% (w/w) APS solution
[(NH4 )2 S2 O8 ] at 60 C at a ratio of 1:100 (w/v) under vigorous
stirring for 16 h (Leung et al., 2011). The carboxylated cellulose nanocrystal (CNO) colloidal suspension was centrifuged with
deionized water at a pH close to neutral, and the yield was calculated as the percentage of cellulose nanocrystal mass compared to
the initial delignied pulp mass (Eq. (2)).
2.3.3. High-power ultrasound
Bleached pulp was mixed with deionized water at a ratio of
1:200 (w/v) and kept for 24 h. The mixture was wet-milled (T50,
IKA Works, Inc., Wilmington, NC, USA) for 10 min and transferred
to a jacketed stainless-steel reactor with cold-water circulation.
The mixture was treated with a high-power ultrasound processor
(UIP1500hd; Hielscher) at 20 kHz frequency and 2.5-cm-diameter
titanium sonotrode, which was immersed 2/3 of the length in the
middle of the mixture for 20 min at 1200 W (Li et al., 2013). The colloidal suspension of cellulose nanocrystals (CNU) was centrifuged
and freeze-dried. The nal yield was calculated as the percentage
of cellulose nanocrystal mass compared to the initial bleached pulp
mass (Eq. (2)).
2.4. Fiber chemical composition
The hemicellulose and -cellulose contents of the samples
were determined in accordance with TAPPI T203cm-99 (2009) and
Yokoyama (2002). Klason lignin was measured as described by the
TAPPI T 222 om-22 (2002) standards with minor modications.
2.5. Biodegradability and MPP
Anaerobic biodegradability evaluation was based on specic
methanogenic activity (Soto et al., 1993). The initial inoculum was

home sewer sludge from CAGECE (Water and Sewerage Company


of Cear). Methane production was recorded using an anaerobic
respirometer (Micro-Oximax; Columbus Instruments, Columbus,
OH, USA). Biodegradability and BMP were calculated as described
by Costa et al. (2014).
2.6. FTIR
Cellulose nanocrystal samples were ground and pelleted using
KBr. Spectra were recorded from 4000 cm1 to 400 cm1 (Cary640;
Agilent Technologies, Santa Clara, CA, USA) using Fourier transform
infrared spectroscopy.
2.7. XRD
Cellulose nanocrystal crystallinity indexes (CI) were analyzed
using an X-ray diffractometer (Xpert Pro MPD; PANalytical, Almelo,
Netherlands), with the Co tube at 40 kV and 40 mA in the 2 range
from 3 to 50 . CI was calculated using two different methods: (i)
by determining the ratio between the integrated area of all crystalline peaks and the integrated total area using PeakFit software
from Systat Software, Inc. (San Jose, CA, USA) (Garvey et al., 2005),
and (ii) by determining the ratio between the peak intensity of
the total crystallinity and the total peak intensity, considering the
amorphous (2 = 18,5 ) and crystalline (2 = 22,5 ) contents (Segal
et al., 1959).
2.8. Electronic microscopy
Fibers were gold sputtered in a metalizer (K550; Emitech, Fall
River, MA, USA) and analyzed by scanning electron microscopy
(DSM 940A; Zeiss, Jena, Germany) at an acceleration voltage of
15 kV.
Cellulose nanostructures were dropped in a 300-mesh copper
carbon grid, draining the excess of water during sample preparation. The cellulose nanocrystals were contrasted with a 2% (w/v)
uranyl acetate solution and analyzed by transmission electron
microscopy (JEM-1011; Jeol, Tokyo, Japan). Nanocrystal sizes and
aspect ratios (length/width) were calculated using GNU Image
Manipulation Program (GIMP 2.8) software.
2.9. Thermal analyses
Approximately 1520 mg of samples in an alumina sample holder were evaluated in a thermal analyzer (STA 6000;
PerkinElmer, Waltham, MA, USA) under a nitrogen atmosphere
with 40 mL/min of gas ux, heating rate of 10 C/min, and temperature ranging from 50 to 800 C.
3. Results and discussion
3.1. Fiber chemical composition
The chemical composition of the unripe coconut coir ber, delignied cellulose pulp, bleached cellulose pulp, and recovered lignin
are presented in Table 1. The major component of coconut ber
was lignin (35.1%), with a similar content (35%) as that reported by
van Dam et al. (2006). Both values are between the lignin contents
reported by Brgida et al. (2010) (43.1%) and Bledzki et al. (2010)
(27%); these values are all related to coconut ber.
Cellulose was the second major component (31.6%) and the
value was similar to that reported by Rosa et al. (2010) (32%) and
Muensri et al. (2011), but below the values recorded by Brgida
et al. (2010) (45.9%) and Bledzki et al. (2010) (34%). Hemicellulose
content (25%) was similar to that reported by Muensri et al. (2011)
(23%) and Bledzki et al. (2010) (21%).

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model

ARTICLE IN PRESS

INDCRO-8646; No. of Pages 10

D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Table 1
Chemical composition of coconut ber, delignied pulp, bleached pulp, and recovered lignin.
Sample

Coconut ber
Delignied pulp
Bleached pulp
Recovered lignin

Chemical composition (%)


-Cellulose

Hemicelluloses

Klason lignin

31.6 0.4
41.8 4.9
70.0 0.6

25.5 0.4
42.2 5.3
28.0 0.5

35.1 2.2
19.1 2.1
0
94.5 0.8

The chemical composition of plant bers varies as a function


of several factors, including genotypes, environments, physiological conditions, and harvest techniques, among others (Kruse et al.,
2008; Pettigrew, 2001). Differences among the chemical composition were similar to expected values and support the previous
characterization of each ber batch.
High lignin and hemicellulose contents decrease the cellulose
per mass of ber and block the accessibility of reagents to cellulose.
Coconut ber is a complex and interesting plant ber, and it is not
feasible to use it for cellulose nanocrystal extraction. Approaches
to add value to this ber must consider adding value to lignin and
hemicellulose, since more than half of the components are noncellulosic biomaterials. Thus, the three macrocomponents must
be simultaneously used on an industrial scale, and the developed
method may be useful as a model for other low-cellulose agroindustrial biomasses.
After acetosolv delignication and alkaline peroxide bleaching,
the cellulose content increased to 70% and lignin content decreased
to below the detection limit. Brgida et al. (2010) reported 41.9% cellulose and 47.3% lignin after coconut ber bleaching with sodium
chlorite, 62.8% cellulose and 45.1% lignin with sodium chlorite and
NaOH, and 43.9% cellulose and 42.7% lignin using hydrogen peroxide.
3.2. Lignin fractionation, anaerobic biodegradability, and
biochemical methane potential
The yield of recovered lignin was 30.1% and the purity was
94.5%. The yield may be a consequence of the long reaction time
and high solubilization of the recovered lignin from the black liquor.
This result is related to the reaction conditions and how such conditions affect the macromolecule depolymerization. The higher the
cleavage of ether bonds in the lignin subunits, the higher the dissolution of lignin fragments. Despite the relatively small amount
of recovered lignin, it was important to separate the lignin rather
than destroying it during the pulping process. This makes the
whole fractionation process more cost-competitive and improves
the manufacture of products from this biomass.
Residual lignin from acetosolv pulping usually presents lower
molecular mass and higher reactivity than residual lignin from
other pulping processes. Such characteristics are responsible for the
residual lignin in the coconut ber to be easily removed, reducing
the amount of chemicals in the bleaching process (Xu et al., 2007).
Residual lignin in the bleaching efuents usually is so degraded and
modied that it can not be recovered as an useful chemical (Pandey
and Kim, 2011). Further researches aiming to increase the overall
yield of the lignin recovery in the pulping step and to nd innovative uses for the residual lignin in the bleaching efuents shall be
performed.
The efuents did not show signicant differences in biodegradability and BMP after 30 days of treatment (Table 2). However,
efuent AB required a longer time for bacterial digestion, indicating
the presence of possible bacterial growth inhibitory chemicals, such
as furfural, 5-hydroximethylfurfural, phenolic monomers from

Table 2
Anaerobic biodegradability and biochemical methane potential (BMP) of efuents
A and AB.
Sample

Biodegradability (%)Total

PPM (Lmethane /kgeuente )

Efuent A
Efuent AB

45.21 3.3
43.4 0.7

2.71 0.20
2.00 0.02

lignin and reactive oxygen species, including H2 O2 , HO , and O2 1


(Monlau et al., 2014, 2013).
The BMP values for both efuents were inferior to the results
of Costa et al. (2014) for autohydrolyzed sugarcane bagasse
(197.5 Lmethane /kgefuent ). This difference is due to the no use of the
hydrolyzed with the ber, presence of inhibitory chemicals, low
content of hemicelluloses removed from the coconut ber (Table 1),
and to the use of no concentrated efuents. In spite of the low
methane production, efuent AB can be used as a methane source,
and this prevents it to be dispose in water bodies, preventing pollution in rivers, lakes, and groundwater.
3.3. Cellulose nanocrystal extraction
The yields of the cellulose nanocrystal extraction methods
(CNH1, CNH2, CNO, and CNU) ranged from 33% to 88% as described
in Eq. (2). The low yield of CNH2 (32.8 0.2%) relative to CNH1
(59.8 2.1%) (both of which involved H2 SO4 ) could be explained
by the acid concentration. Higher acid content in the reaction is
associated with harsher hydrolysis conditions, and even crystalline
domains in the cellulose structure are broken in oligo and monosaccharides, decreasing cellulose nanocrystal levels and overall yield
(Chenampulli et al., 2013).
The yield of CNO (49.6 0.3%) can be attributed to APS reaction
with cellulose, promoting concomitantly hydrolysis and degradation due formation SO4 2 and HSO4 1 species. Hydrolysis yield
with APS depends strongly on the reaction time and temperature. Harsh conditions (high time and temperature) result in higher
degradation and low yield (Cheng et al., 2014). Pulp after acetosolv
pulping still presents a high lignin content, and this biomacromolecule is a hinder for the ionic attack of SO4 2 and HSO4 1 . APS
procedure presents an advantage in the challenge of nanocellulose production from lignin-rich bers, as agroindustrial biomass,
because APS procedure can bleach and hydrolyze such bers in one
step.
The yield of cellulose nanocrystal from CNU (88.1 1.5%)
(20 min and 1200 W) was higher than that CNH1, CNH2 and CNO
samples. The use of mechanical methods lead to the production
of cellulose nanocrystals with higher yield (Li et al., 2012b, 2013;
Oksman et al., 2011). However, this method can produce heterogeneous and lower colloidal stability. These results indicate that CNU
had the best yield among the three different nanocrystal extraction approaches. Moreover, this yield was higher than that obtained
using H2 SO4 , CNH2, and CNH1, the standard procedures for cellulose nanocrystal extraction.
3.4. Electron microscopy
Fiber morphology was affected by pretreatment, which
removed some macromolecular components (Fig. 2) (Johar et al.,
2012). Raw coconut ber had an irregular and compact surface
compared to the delignied and bleached pulps, which had a rough
surface and loosely bundled macrobrils.
Decompression of the whole ber occurred because of the
removal of lignin and hemi-celluloses, which generally hold cellulose bers together. After disassembly of the lignocellulosic
matrix, cellulose macrobrils were more accessible for the cellulose nanocrystal extraction process because of the removal of

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model
INDCRO-8646; No. of Pages 10

ARTICLE IN PRESS
D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Fig. 2. Scanning electron microscopy of (a) coconut ber; (b) delignied pulp; and (cd) bleached pulp. Scale bar = 50 m to (a), (b) e (c), and 20 m to (d).

non-cellulosic components as well as the increased surface area


(Rosa et al., 2010).
Cellulose nanocrystal micrographies (Fig. 3) and the calculated
size parameters length (L), width (W), and aspect ratio (L/W)
(Table 3) indicated that CNH1 and CNH2 were needle-like and welldispersed, with average (L) values of 128 nm and 208 nm, average
(W) values of 6.6 nm and 4.9 nm, and aspect ratio of 19 and 42.
Cellulose nanocrystals produced by acidic hydrolysis showed an
aspect ratio close to that of nanocrystals from banana pseudostem
(21) (Pereira et al., 2014), Agave tequilana bagasse (28) (Espino
et al., 2014), brewers spent grains (36) (Martnez-Sanz et al., 2014),
cotton linter (24) (Morais et al., 2013), and bleached white coir, prepared using other pulping methods (22) (Nascimento et al., 2014)
or (39) (Rosa et al., 2010).
CNO had a spherical shape and was well dispersed, with an average (W) of 116 nm. Liocel cellulose nanocrystals had an average
width of 96 nm, while microcrystalline cellulose nanocrystal had
widths ranging from 100 to 270 nm. The aspect ratio for spherical nanoparticles is 1, which means a smaller specic surface area
compared to other morphologies nanoparticles and therefore, more
susceptibility of surface modifying reactions (Hsieh, 2013; Zhao and
Winter, 2014).

Table 3
Length (L), width (W), and aspect ratio of cellulose nanocrystals extracted using
different methods.
Sample

L (nm)

W (nm)

L/W

CNH1
CNH2
CNO
CNU

128 52
208 34
116 27
307 165

6.6 1.5
4.9 0.5
116 27
14.6 2.5

19
42
1
21

The CNU structure shows needle-like and cluster formation


because of the absence of surface-charged groups produced in other
nanostructures, with average (L) of 307 nm, average (W) of 14.6 nm,
and an aspect ratio of 21. A minimum aspect ratio of 10 is required
for efcient tension transfer between the load and the matrix,
enabling the nanostructure to be used as a reinforcement (Azeredo
et al., 2009).
CNH2 showed the highest aspect ratio because it had the shortest reaction time. Although high-power ultrasound did not produce
nanocrystals with the highest aspect ratio, the CNU aspect ratio was
still well above the minimum threshold, showing a similar size to
CNH1 and CNH2.

3.5. FTIR
FTIR spectra of raw coconut ber, delignied cellulose pulp,
bleached cellulose pulp, and cellulose nanocrystal are shown in
Fig. 4 and the summary of the main absorption peaks are shown
in Table 4.
The absence of the peaks around 1730, 1612, 1516, and
1450 cm1 , as well as the signicant reduction around 1639 cm1 ,
in the delignied and bleached pulps indicate lignin removal. The
persistence of peaks around 1163 and 899 cm1 in the bleached
pulp indicates the presence of hemicelluloses resulting from partial removal of this set of macromolecules, which was in accordance
with the chemical composition analyses (Section 3.1).
Cellulose nanocrystal samples showed peaks at 1431, 1319,
1034, and 897 cm1 , which typical cellulose absorption peaks. In
CNO, a peak was observed at around 1730 cm1 , indicating the presence of C O groups after oxidation with APS (Habibi et al., 2010).
Cellulose nanocrystal carboxylation preferentially occurs at the C6

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model
INDCRO-8646; No. of Pages 10
6

ARTICLE IN PRESS
D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Fig. 3. Transmission electron microscopy for cellulose nanocrystals extracted by sulfuric acid hydrolysis (CNH1 and CNH2), ammonium persulfate oxidation (CNO), and
high-power ultrasound (CNU). Scale bar = 500 nm.

of the crystal surface and allows more reactive nanoparticles to be


obtained. Cellulose oxidation degree (OD) can be calculated by the
equation OD = 0.01 + 0.7(I1735 /I1060 ), where I1735 is the peak intensity of carbonyl group, and I1060 is the peak intensity of cellulose
skeleton vibration (Habibi et al., 2006). The calculated OD for CNO

was 0.197. Their surfaces can be easily modied to improve the


versatility of the nanoller and result in nanocomposites that can
be more easily processed (Cheng et al., 2014; Leung et al., 2011). In
both CNH1 and CNH2, a band at 810 cm1 was observed, which is
related to sulfate esters in the nanocrystals (Bellamy, 1980; Felcio

Fig. 4. FTIR spectra of coconut ber, delignied pulp, bleached pulp, recovered lignin, CNH1, CNH2, CNO, and CNU.

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model

ARTICLE IN PRESS

INDCRO-8646; No. of Pages 10

D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Table 4
Main absorption wavelengths (cm1 ) and relation to chemical groups and lignocellulosic molecules (Goncalves et al., 2014; Rehman et al., 2013; Rosa et al., 2010 Rosa et al.,
2010).
Wavenumber (cm1 )

Chemical signicance

3411
2993
2920
1735
1718
1639
1612
1516
1450
1431
1377
1319
1265
1250
1225
1163
1119
1107
1060
1034
958
917
899
810

Axial deformation of O H
Axial deformation of methyl and methylene C H
Axial deformation of C H
Axial deformation of C O in non-conjugated ketones and esters, usually carbohydrates
Axial deformation of C O at position of carboxyl groups
Angular deformation of H2 O; Axial deformation of lignin C O
Vibration C C of lignin aromatic skeleton
Vibration C C of lignin aromatic skeleton
Vibration C C of lignin aromatic skeleton; Asymmetric angular deformation of CH3 e CH2
Angular deformation of cellulose and lignin C H
Angular deformation of cellulose, hemicellulose, and lignin C H
Wagging vibration of cellulose CH2
Vibration of guaiacyl (G) aromatic ring
Axial deformation of hemicelluloses and guaiacyl C O; and Angular deformation of cellulose C H
Axial deformation of C O e C C on the guaiacyl (G) aromatic ring
Asymmetric axial deformation C O C of hemicelluloses
Plane angular deformation of C H in the syringuyl and guaiacyl (GS) aromatic ring
Glucosidic ring vibration in cellulose; Plane deformation of lignin C H
Glucosidic ring vibration of C O
Vibration of p-hidrofenilpropano (H) units
Out of plane angular deformation of aromatic C H
Out of the plane angular deformation of C H in the guaiacyl (G) aromatic ring C-2, C-5, and C-6
Vibration of glucosidic ring in amorphous domains
Vibration of sulfate groups

et al., 2008). None of the nanocrystals showed absorption peaks at


1612, 1516, or 1450 cm1 , which are related to lignin. Thus, this
macromolecule contains no lignin.
3.6. XRD
The crystallographic prole of all samples is typical of cellulose
I, with diffraction peaks around 15 (plan 101), 17 (plan10), 22
(plan 021), 23 (plan 002), and 35 (plan 040) (Garvey et al., 2005)
(Fig. 5). The chemical reactions affected the crystallinity index, but
did not modify the cellulose allomorph structure in any sample.
The increment of the crystallinity index was observed from the
coconut ber to the delignied ber, to the bleached ber, and was
the highest in the nanocrystals (Table 5). This increment is a consequence of the removal of amorphous components, mainly lignin
and hemicelluloses.
CNO showed a high crystallinity index, indicating the removal
of non-cellulosic components, such as lignin, by oxidation. Lignin
is destroyed by oxidative processes, and the APS reaction produces cellulose nanocrystals and oxidizes residual lignin, acting as
a bleaching reaction.
CNH2 showed the highest crystallinity index, regardless of the
method used to calculate the index (Segal or deconvolution). This
was expected because H2 SO4 nanocrystal extractions are harsh
and generate nanocrystals with high crystallinity indexes (Mariano
et al., 2014).
The crystallinity index calculated for CNU and CNO differed
depending on the method used. While the deconvolution approach
showed values of 82% for CNU and 81% for CNO, the Segal approach
showed a smaller crystallinity index (75%) for CNU and a higher
index (85%) for CNO. Different methodological approaches may
Table 5
Crystallinity indexes of coconut ber, delignied pulp, bleached pulp, CNH1, CNH2,
CNO, and CNU.
Sample

Coconut
ber
Crystallinity index (%)

Delignied Bleached CNH1 CNH2 CNO


pulp
pulp

CNU

Deconvolution
Segal

59
64

82
79

52
57

61
77

68
79

71
80

81
85

generate different values, suggesting that it is important to use


methods that best approximate the actual values. Segals methodology (Segal et al., 1959) is simpler and is used more frequently,
but it is based only on the height of the recorded peaks (Terinte
et al., 2011). Deconvolution methodology is not typically used for
crystallinity index calculations, but presents values closer to the
crystallinity index measured by RMN 13 C, which is considered to
be the gold standard and reveal true values (Park et al., 2010).
CNO had a higher crystallinity index because hydrolysis of amorphous regions occurred. Cheng et al. (2014) obtained similar results
for the hydrolysis and oxidation of liocel ber with 18.5% (w/w) APS
after a 20-h reaction, with 94% of crystallinity as measured using
Segals method.
The increase in the crystallinity index in the CNU validates that
some cellulose amorphous regions were removed after high-power
ultrasound treatment. The results are corroborated by the results of
other studies, in which ultrasound was suggested to create cavity
effects that preferentially fragment cellulose amorphous regions
(Li et al., 2012a,b).
Segals crystallinity indexes for all cellulose crystals were higher
than those reported by Rosa et al. (2010) (66%) and Fahma et al.
(2011) (57%) for coconut ber from cellulose pulps prepared using
different pretreatments. Acetosolv delignication and bleaching as
well as the cellulose nanocrystal extraction methodologies were
quite effective for increasing the hydrolyzing reagents to the cellulose content and reducing the overall amorphous content in the
crystals.

3.7. Thermal analyses


Loss weight values and initial degradation temperatures (Tonset) are presented in Table 6.
The thermogravimetric curve for the raw coconut ber was similar to the curves reported by Rosa et al. (2010) and Nascimento
et al. (2014). The thermal stability of the coconut ber increased
after the pulping treatments. All analyzed samples showed a small
loss in weight at temperatures below and around 100 C, which
was related to water evaporation. The bleached pulp showed a single loss mass event at approximately 328 C, which was related to

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model

ARTICLE IN PRESS

INDCRO-8646; No. of Pages 10

D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Fig. 5. X-ray diffractograms of coconut ber, delignied pulp, bleached pulp, CNH1, CNH2, CNO, and CNU.

Table 6
Thermal events of TG/dTG thermograms under inert atmosphere.
Samples

Coconut ber
Delignied pulp
Bleached pulp
CNH1
CNH2
CNO
CNU

1st event

2nd event

3rd event

m (%)

Tonset ( C)

m (%)

Tonset ( C)

m (%)

Tonset ( C)

23.9
81.9
82.2
80.7
8.5
33.3
86.1

260.2
328.1
314.5
300.6
227.8
210.3
324.3

52.6

335.3

31.1
33.5

279.7
324.8

31.6

343.7

the removal of hemicellulose and lignin and increase in cellulose


content, which has a higher thermal stability.
Cellulose nanocrystal samples showed a small loss in water at
temperatures below and around 100 C. CNH1 presented a sole
mass loss event at approximately 300 C, while CNH2 showed three
mass loss events at 227 C, 279 C, and 343 C. The small thermal
stability of both samples in comparison to the bleached pulp was
related to the presence of sulfate groups on the surface of the cellulose chains. This reduced the activation energy and eased the
depolymerization, dehydration, and thermal decomposition reactions (Teodoro et al., 2011). The lower stability and higher number
of thermal events of the CNH2 are related to the greater number of
sulfate esters in the cellulose chains.
Rosa et al. (2010) reported that residual lignin in the nanocellulose suspensions prepared by acidic hydrolysis may contribute
to increasing the thermal stability and improving the compatibility of the ller with hydrophobic matrices. A further approach for
reaching this aim is to incorporate small amounts of lignin in the
cellulose pulp, reducing the severity of the bleaching treatment,
saving reagents and power.
CNO involved two main weight reduction events, including
at approximately 210 and 325 C. Bamboo cellulose nanocrystals
produced by APS oxidation showed higher thermal stability than
the nanocrystals produced by acidic hydrolysis (Hu, 2014). CNO
coconut cellulose nanocrystal showed thermal stability similar to
that observed for TEMPO-oxidized cellulose nanobrils (Carlsson
et al., 2014). Further studies are necessary to optimize the reac-

tion times and reagent concentrations for coconut delignied pulp


oxidation.
CNU showed a single weight reduction event at around 324 C,
which was 86.12% of the original mass. The presence of sulfate groups in the surface of the sulfuric acid-hydrolyzed CNCs
decreases the thermal stability in comparison to the CNU, which
were produced in aqueous medium, with no sulfur esters (Oksman
et al., 2011). This behavior is determined because thermal degradation reactions begin in the cellulose amorphous domains (Ciolacu
et al., 2011). The higher thermal stability of the ultrasoundproduced nanocellulose was a consequence of the absence of
sulfate groups on the crystal surface, combined with a high crystallinity index. Since thermal degradation begins in the amorphous
domains, fewer amorphous regions would result in the degradation
of fewer regions (Ciolacu et al., 2011).
Since CNU showed the highest thermal stability of all produced
nanocrystals, this nanocellulose can be incorporated in polymers
over a wide temperature range. For example, in further studies, this
cellulose nanocrystal can be used in polymer extruders because
most of the typically used thermoplastic matrixes have melting
temperatures of around 200 C (Chen et al., 2011; Roman and
Winter, 2004).
4. Conclusions
Coconut ber contains a low cellulose content compared to
other lignocellulosic biomasses, but high lignin content and fair
hemicellulose content. Thus, all main macromolecules in this

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model
INDCRO-8646; No. of Pages 10

ARTICLE IN PRESS
D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

industrial ber must be used as an economic source of added-value


chemicals or power, increasing its technical competitiveness and
economic viability to overcome issues related to coir disposal. In
this study, we developed an approach that complies with principles
of green chemistry and integral use of biomass.
Coconut ber was successfully fractionated into carbohydrate
and lignin materials without using chlorine processes. Hemicelluloses were successfully used as a power source for methane
production.
This study described several methods for transforming raw
ber in cellulose nanocrystals with high yield, purity, aspect ratio,
crystallinity index, and thermal stability above 200 C. Among the
evaluated methods, high-power ultrasound showed results that
were superior to the standard procedure based on sulfuric acid
hydrolysis, and may be scaled up for white coconut ber nanocellulose manufacture.
References
Azeredo, H.M.C., Mattoso, L.H.C., Wood, D., Williams, T.G., Avena-Bustillos, R.J.,
McHugh, T.H., 2009. Nanocomposite edible lms from mango puree reinforced
with cellulose nanobers. J. Food Sci. 74, N31N35, http://dx.doi.org/10.1111/j.
1750-3841.2009.01186.x.
Bellamy, L.J., 1980. The infrared spectra of complex molecules. Advances in
Infrared Group Frequencies, vol. 2. Chapman and Hall.
Bledzki, A.K., Mamun, A.A., Volk, J., 2010. Barley husk and coconut shell reinforced
polypropylene composites: the effect of bre physical, chemical and surface
properties. Compos. Sci. Technol. 70, 840846, http://dx.doi.org/10.1016/j.
compscitech.2010.01.022.
Brgida, A.I.S., Calado, V.M.A., Goncalves, L.R.B., Coelho, M.A.Z., 2010. Effect of
chemical treatments on properties of green coconut ber. Carbohydr. Polym.
79, 832838, http://dx.doi.org/10.1016/j.carbpol.2009.10.005.
Carlsson, D.O., Hua, K., Forsgren, J., Mihranyan, A., 2014. Aspirin degradation in
surface-charged TEMPO-oxidized mesoporous crystalline nanocellulose. Int. J.
Pharm. 461, 7481, http://dx.doi.org/10.1016/j.ijpharm.2013.11.032.
Chen, W., Yu, H., Liu, Y., Hai, Y., Zhang, M., Chen, P., 2011. Isolation and
characterization of cellulose nanobers from four plant cellulose bers using a
chemical-ultrasonic process. Cellulose 18, 433442, http://dx.doi.org/10.1007/
s10570-011-9497-z.
Chenampulli, S., Unnikrishnan, G., Sujith, A., Thomas, S., Francis, T., 2013. Cellulose
nano-particles from Pandanus: viscometric and crystallographic studies.
Cellulose 20, 429438, http://dx.doi.org/10.1007/s10570-012-9831-0.
Cheng, M., Qin, Z., Liu, Y., Qin, Y., Li, T., Chen, L., Zhu, M., 2014. Efcient extraction
of carboxylated spherical cellulose nanocrystals with narrow distribution
through hydrolysis of lyocell bers by using ammonium persulfate as an
oxidant. J. Mater. Chem. A 2, 251258, http://dx.doi.org/10.1039/c3ta13653a.
Ciolacu, D., Ciolacu, F., Popa, V.I., 2011. Amorphous cellulose structure and
characterization. Cellul. Chem. Technol. 45, 1321.
Costa, A.G., Pinheiro, G.C., Pinheiro, F.G.C., Dos Santos, A.B., Santaella, S.T., Leito,
R.C., 2014. The use of thermochemical pretreatments to improve the anaerobic
biodegradability and biochemical methane potential of the sugarcane bagasse.
Chem. Eng. J. 248, 363372, http://dx.doi.org/10.1016/j.cej.2014.03.060.
Cotana, F., Cavalaglio, G., Nicolini, A., Gelosia, M., Coccia, V., Petrozzi, A., Brinchi, L.,
2014. Lignin as co-product of second generation bioethanol production from
ligno-cellulosic biomass. Energy Procedia 45, 5260, http://dx.doi.org/10.1016/
j.egypro.2014.01.007.
Deepa, B., Abraham, E., Cordeiro, N., Mozetic, M., Mathew, A.P., Oksman, K., Faria,
M., Thomas, S., Pothan, L.a., 2015. Utilization of various lignocellulosic biomass
for the production of nanocellulose: a comparative study. Cellulose 22,
10751090, http://dx.doi.org/10.1007/s10570-015-0554-x.
Dufresne, A., 2013. Nanocellulose: From Nature to High Performance Tailored
Materials. Walter de Gruyter GmbH & Co. KG .
Espino, E., Cakir, M., Domenek, S., Romn-Gutirrez, A.D., Belgacem, N., Bras, J.,
2014. Isolation and characterization of cellulose nanocrystals from industrial
by-products of Agave tequilana and barley. Ind. Crops Prod. 62, 552559, http://
dx.doi.org/10.1016/j.indcrop.2014.09.017.
Fahma, F., Iwamoto, S., Hori, N., Iwata, T., Takemura, A., 2011. Effect of
pre-acid-hydrolysis treatment on morphology and properties of cellulose
nanowhiskers from coconut husk. Cellulose 18, 443450, http://dx.doi.org/10.
1007/s10570-010-9480-0.
FAOStat, 2015. FAOstat [WWW Document]. URL http://faostat3.fao.org/download/
Q/QC/E.
Felcio, R., de Debonsi, H.M., Yokoya, N.S., 2008.
4-(hidroximetil)-benzenossulfonato de potssio: metablito indito isolado da
alga marinha Bostrychia tenella (Rhodomelaceae, ceramiales). Quim. Nova 31,
837839, http://dx.doi.org/10.1590/S0100-40422008000400025.
de Figueirdo, M.C.B., Rosa, M. de F., Ugaya, C.M.L., Souza Filho, M. de S.M. de, da
Silva Braid, A.C.C., de Melo, L.F.L., 2012. Life cycle assessment of cellulose
nanowhiskers. J. Clean. Prod. 35, 130139, http://dx.doi.org/10.1016/j.jclepro.
2012.05.033.

Garvey, C.J., Parker, I.H., Simon, G.P., 2005. On the interpretation of X-ray diffraction
powder patterns in terms of the nanostructure of cellulose I bres. Macromol.
Chem. Phys. 206, 15681575, http://dx.doi.org/10.1002/macp.200500008.
Habibi, Y., Chanzy, H., Vignon, M.R., 2006. TEMPO-mediated surface oxidation of
cellulose whiskers. Cellulose 13, 679687, http://dx.doi.org/10.1007/s10570006-9075-y.
Habibi, Y., Lucia, L., Rojas, O., 2010. Cellulose nanocrystals: chemistry,
self-assembly, and applications. Chem. Rev. d, 34793500.
Hsieh, Y., 2013. Cellulose nanocrystals and self-assembled nanostructures from
cotton, rice straw and grape skin: a source perspective. J. Mater. Sci.,
78377846, http://dx.doi.org/10.1007/s10853-013-7512-5.
Huntley, C.J., Crews, K.D., Curry, M.L., 2015. Chemical functionalization and
characterization of cellulose extracted from wheat straw using acid hydrolysis
methodologies. Int. J. Polym. Sci. 2015, 19, http://dx.doi.org/10.1155/2015/
293981.
Johar, N., Ahmad, I., Dufresne, A., 2012. Extraction, preparation and
characterization of cellulose bres and nanocrystals from rice husk. Ind. Crops
Prod. 37, 9399, http://dx.doi.org/10.1016/j.indcrop.2011.12.016.
Jor, M., Foster, E.J., 2015. Recent advances in nanocellulose for biomedical
applications. J. Appl. Polym. Sci. 132, 119, http://dx.doi.org/10.1002/app.
41719.
Kruse, S., Herrmann, A., Kornher, A., Taube, F., 2008. Evaluation of genotype and
environmental variation in bre content of silage maize using a model-assisted
approach. Eur. J. Agron. 28, 210223, http://dx.doi.org/10.1016/j.eja.2007.07.
007.
Leung, A.C.W., Hrapovic, S., Lam, E., Liu, Y., Male, K.B., Mahmoud, K.A., Luong, J.H.T.,
2011. Characteristics and properties of carboxylated cellulose nanocrystals
prepared from a novel one-step procedure. Small 7, 302305, http://dx.doi.
org/10.1002/smll.201001715.
Li, F., Mascheroni, E., Piergiovanni, L., 2015. The potential of nanocellulose in the
packaging eld: a review. Packag. Technol. Sci. 28, 475508, http://dx.doi.org/
10.1002/pts.2121.
Li, J., Wei, X., Wang, Q., Chen, J., Chang, G., Kong, L., Su, J., Liu, Y., 2012a.
Homogeneous isolation of nanocellulose from sugarcane bagasse by high
pressure homogenization. Carbohydr. Polym. 90, 16091613, http://dx.doi.org/
10.1016/j.carbpol.2012.07.038.
Li, W., Yue, J., Liu, S., 2012b. Preparation of nanocrystalline cellulose via ultrasound
and its reinforcement capability for poly(vinyl alcohol) composite. Ultrason.
Sonochem. 19, 479485, http://dx.doi.org/10.1016/j.ultsonch.2011.11.007.
Li, W., Zhao, X., Huang, Z., Liu, S., 2013. Nanocellulose brils isolated from BHKP
using ultrasonication and their reinforcing properties in transparent poly
(vinyl alcohol) lms. J. Polym. Res. 20, 210, http://dx.doi.org/10.1007/s10965013-0210-9.
Mandal, A., Chakrabarty, D., 2014. Studies on the mechanical, thermal,
morphological and barrier properties of nanocomposites based on poly(vinyl
alcohol) and nanocellulose from sugarcane bagasse. J. Ind. Eng. Chem. 20,
462473, http://dx.doi.org/10.1016/j.jiec.2013.05.003.
Mariano, M., El Kissi, N., Dufresne, A., 2014. Cellulose nanocrystals and related
nanocomposites: review of some properties and challenges. J. Polym. Sci. Part
B Polym. Phys. 52, 791806, http://dx.doi.org/10.1002/polb.23490.
Martnez-Sanz, M., Vicente, A.A., Gontard, N., Lopez-Rubio, A., Lagaron, J.M., 2014.
On the extraction of cellulose nanowhiskers from food by-products and their
comparative reinforcing effect on a polyhydroxybutyrate-co-valerate polymer.
Cellulose 22, 535551, http://dx.doi.org/10.1007/s10570-014-0509-7.
McDevitt, J.E., Grigsby, W.J., 2014. Life cycle assessment of bio- and petro-chemical
adhesives used in berboard production. J. Polym. Environ. 22, 537544,
http://dx.doi.org/10.1007/s10924-014-0677-4.
Monlau, F., Aemig, Q., Trably, E., Hamelin, J., Steyer, J.-P., Carrere, H., 2013. Specic
inhibition of biohydrogen-producing Clostridium sp. after dilute-acid
pretreatment of sunower stalks. Int. J. Hydrogen Energy 38, 1227312282,
http://dx.doi.org/10.1016/j.ijhydene.2013.07.018.
Monlau, F., Sambusiti, C., Barakat, A., Qumneur, M., Trably, E., Steyer, J.-P.,
Carrre, H., 2014. Do furanic and phenolic compounds of lignocellulosic and
algae biomass hydrolyzate inhibit anaerobic mixed cultures? A comprehensive
review. Biotechnol. Adv. 32, 934951, http://dx.doi.org/10.1016/j.biotechadv.
2014.04.007.
Morais, J.P.S., Rosa de, M.F., de Souza Filho de s, M.M., Nascimento, L.D., do
Nascimento, D.M., Cassales, A.R., 2013. Extraction and characterization of
nanocellulose structures from raw cotton linter. Carbohydr. Polym. 91,
229235, http://dx.doi.org/10.1016/j.carbpol.2012.08.010.
Muensri, P., Kunanopparat, T., Menut, P., Siriwattanayotin, S., 2011. Effect of lignin
removal on the properties of coconut coir ber/wheat gluten biocomposite.
Compos. Part A Appl. Sci. Manuf. 42, 173179, http://dx.doi.org/10.1016/j.
compositesa.2010.11.002.
Nascimento, D.M., Almeida, J.S., Dias, A.F., Figueirdo, M.C.B., Morais, J.P.S., Feitosa,
J.P.A., Rosa, M. de F., 2014. A novel green approach for the preparation of
cellulose nanowhiskers from white coir. Carbohydr. Polym. 110, 456463,
http://dx.doi.org/10.1016/j.carbpol.2014.04.053.
Norgren, M., Edlund, H., 2014. Lignin: recent advances and emerging applications.
Curr. Opin. Colloid Interface Sci. 19, 409416, http://dx.doi.org/10.1016/j.cocis.
2014.08.004.
Oksman, K., Etang, J.A., Mathew, A.P., Jonoobi, M., 2011. Cellulose nanowhiskers
separated from a bio-residue from wood bioethanol production. Biomass
Bioenergy 35, 146152, http://dx.doi.org/10.1016/j.biombioe.2010.08.021.

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

G Model
INDCRO-8646; No. of Pages 10
10

ARTICLE IN PRESS
D.M.d. Nascimento et al. / Industrial Crops and Products xxx (2016) xxxxxx

Pandey, M.P., Kim, C.S., 2011. Lignin depolymerization and conversion: a review of
thermochemical methods. Chem. Eng. Technol. 34, 2941, http://dx.doi.org/10.
1002/ceat.201000270.
Park, S., Baker, J.O., Himmel, M.E., Parilla, P., a Johnson, D.K., 2010. Cellulose
crystallinity index: measurement techniques and their impact on interpreting
cellulase performance. Biotechnol. Biofuels 3, 10.
Pereira, A.L.S., Nascimento, D.M. do, Filho, M. de s M.S., Morais, J.P.S., Vasconcelos,
N.F., Feitosa, J.P.A., Brgida, A.I.S., Rosa, Rosa, M. de F., 2014. Improvement of
polyvinyl alcohol properties by adding nanocrystalline cellulose isolated from
banana pseudostems. Carbohydr. Polym. 112, 165172, http://dx.doi.org/10.
1016/j.carbpol.2014.05.090.
Pettigrew, W.T., 2001. Environmental effects on cotton ber carbohydrate
concentration and quality. Crop Sci. 41, 1108, http://dx.doi.org/10.2135/
cropsci2001.4141108x.
Rodrigues, B.V.M., Ramires, E.C., Santos, R.P.O., Frollini, E., 2015. Ultrathin and
nanobers via room temperature electrospinning from triuoroacetic acid
solutions of untreated lignocellulosic sisal ber or sisal pulp. J. Appl. Polym. Sci.
132, n/an/a, http://dx.doi.org/10.1002/app.41826.
Rojo, E., Peresin, M.S., Sampson, W.W., Hoeger, I.C., Vartiainen, J., Laine, J., Rojas,
O.J., 2015. Comprehensive elucidation of the effect of residual lignin on the
physical, barrier, mechanical and surface properties of nanocellulose lms.
Green Chem. 17, 18531866, http://dx.doi.org/10.1039/c4gc02398f.
Roman, M., Winter, W.T., 2004. Effect of sulfate groups from sulfuric acid
hydrolysis on the thermal degradation behavior of bacterial cellulose.
Biomacromolecules 5, 16711677, http://dx.doi.org/10.1021/bm034519+.
Rosa, M.F., Medeiros, E.S., Malmonge, J.A., Gregorski, K.S., Wood, D.F., Mattoso,
L.H.C., Glenn, G., Orts, W.J., Imam, S.H., 2010. Cellulose nanowhiskers from
coconut husk bers: effect of preparation conditions on their thermal and
morphological behavior. Carbohydr. Polym. 81, 8392, http://dx.doi.org/10.
1016/j.carbpol.2010.01.059.
Segal, L., Creely, J.J., Martin, A.E., Conrad, C.M., 1959. An empirical method for
estimating the degree of crystallinity of native cellulose using the X-ray
diffractometer. Text. Res. J. 29, 786794, http://dx.doi.org/10.1177/
004051755902901003.

Soto, M., Mndez, R., Lema, J.M., 1993. Methanogenic and non-methanogenic
activity tests. Theoretical basis and experimental set up. Water Res. 27,
13611376, http://dx.doi.org/10.1016/0043-1354(93) 90224-6.
Souza, N.F., Pinheiro, J.A., Silva, P., Morais, J.P.S., de Souza Filho, M., de s, M.,
Brgida, A.I.S., Muniz, C.R., de Freitas Rosa, M., 2015. Development of
chlorine-free pulping method to extract cellulose nanocrystals from pressed
oil palm mesocarp bers. J. Biobased Mater. Bioenergy 9 (8), 372379, http://
dx.doi.org/10.1166/jbmb.2015.1525.
Teodoro, K.B.R., Teixeira de, E.M., Corra, A.C., Campos, A., de Marconcini, J.M.,
Mattoso, L.H.C., 2011. Whiskers de bra de sisal obtidos sob diferentes
condices de hidrlise cida: efeito do tempo e da temperatura de extrao.
Polmeros 21, 280285, http://dx.doi.org/10.1590/S010414282011005000048.
Terinte, N., Ibbett, R., Schuster, K.C., 2011. Overview on native cellulose and
microcrystalline cellulose i structure studied by X-ray diffraction (Waxd):
comparison between measurement techniques. Lenzinger Berichte 89,
118131.
van Dam, J.E.G., van den Oever, M.J.A., Keijsers, E.R.P., van der Putten, J.C., Anayron,
C., Josol, F., Peralta, A., 2006. Process for production of high density/high
performance binderless boards from whole coconut husk. Ind. Crops Prod. 24,
96104, http://dx.doi.org/10.1016/j.indcrop.2005.03.003.
Xu, Y., Li, K., Zhang, M., 2007. Lignin precipitation on the pulp bers in the
ethanol-based organosolv pulping. Colloids Surfaces A Physicochem. Eng. Asp.
301, 255263, http://dx.doi.org/10.1016/j.colsurfa.2006.12.078.
Zhao, X.F., Winter, W.T., 2014. Cellulose/cellulose-based nanospheres:
perspectives and prospective. Ind. Biotechnol., 3443, http://dx.doi.org/10.
1089/ind.2014.0030.

Please cite this article in press as: Nascimento, D.M.d., et al., A comprehensive approach for obtaining cellulose nanocrystal from coconut
ber. Part I: Proposition of technological pathways. Ind. Crops Prod. (2016), http://dx.doi.org/10.1016/j.indcrop.2015.12.078

Potrebbero piacerti anche