You are on page 1of 10

Resources, Conservation & Recycling 143 (2019) 133–142

Contents lists available at ScienceDirect

Resources, Conservation & Recycling


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

Full length article

Valorization of industrial paper waste by isolating cellulose nanostructures T


with different pretreatment methods

Alana Gabrieli de Souza, Daniel Belchior Rocha, Fabiany Sayuri Kano, Derval dos Santos Rosa
Laboratório de polímeros ambientalmente amigáveis, Universidade Federal do ABC – Santo André, Avenida dos Estados, 5001, CEP: 09210-580, SP, Brazil1

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

Keywords: Industrial paper wastes are an underestimated source of lignocellulosic materials. In this work cellulose na-
Cellulose nanostructures nostructures (CNS) were produced from this residue by acid hydrolysis and two chemical pretreatments con-
Industrial paper wastes sisting of an alkaline (CNS-I) or an acid (CNS-II) treatment. These pretreatments and the obtained CNSs were
Pretreatment influence characterized using compositional analysis, Fourier-transform infrared spectroscopy (FTIR), thermogravimetric
Controlled acid hydrolysis
analysis (TGA), scanning electron microscopy (SEM), dynamic light scattering (DLS), zeta potential, atomic force
Morphology
microscopy (AFM), and X-ray diffraction (XRD). Results reveal a reduction in the non-cellulosic compounds in
pre-treated samples, with different levels of components removal and lower lignin values for alkali treatment.
TGA reveals that CNSs present lower thermal stability than the residue. The SEM images revealed that pre-
treatment led to fiber breakage. The XDR diffractions show the co-existence of cellulose I and cellulose II, with
different crystallinity index for CNSs. The DLS and AFM revealed that the production of spherical nanoparticles
was following the dimensions of ˜195 for CNS-I and ˜350 nm for CNS-II. All the analysis allowed a more com-
prehensive understanding of the CNS properties and their possible applications. Finally, it is possible to conclude
that the conversion of paper wastes into CNS is an excellent opportunity to recovery this residue and aggregate
value to them.

1. Introduction quantities of generated residues, promoting the reuse and recycling of


these materials (Bayard et al., 2017).
Recently, the demand for sustainable and green materials has Research focusing on cellulosic waste has expanded quickly within
prompted the development of new materials using industrial residues as the last decade. Agricultural residues such as rice husks (Johar et al.,
raw material (Feng et al., 2015; Satari and Karimi, 2018). The paper 2012), eucalyptus fiber (Domingues et al., 2016), banana rachis
industry is a major consumer of natural resources (e.g., wood and (Hernandez and Rosa, 2016), cotton fibers (Sun et al., 2016) and algae
water) and energy and results in high emissions of pollutants into the marine biomass (Chen et al., 2016b) have been studied as sources for
environment (Bousios and Worrell, 2017; Corcelli et al., 2018; Jaria the production of nanocellulose fibers and crystals. Also, industrial
et al., 2017; Thompson et al., 2001). These industrial residues result residues have been studied, such as the coffee industry by-products and
from the mechanical and chemical process necessary to transform wood tobacco wastes (Janissen and Huynh, 2018; Tuzzin et al., 2016). Cel-
fiber into cellulose pulp (Mastrantonio et al., 2015). The process results lulose nanostructures (CNSs) are good options in the development of
in lignocellulosic residue with a significant amount of cellulose nanocomposites because of their high crystallinity and large superficial
wrapped by high amounts of hemicellulose and lignin (Mendes et al., area, essential properties for high-performance composites (Johar et al.,
2016). 2012).
Furthermore, the process also results in emissions containing heavy CNSs are typically produced from sources with a significative con-
metals, but at levels accepted by legislation. The pollutants are partially tent of cellulose and small contents of other lignocellulosic compounds,
dissolved in the residue during the paper production and are generally such as hemicellulose and lignin. These components are chemically
burned with the final waste (Chen et al., 2016a; Thompson et al., 2001). complex and highly resistant to the isolation of cellulose. The removal
Various management strategies have been studied to minimize the high of these lignocellulosic compounds through chemical pre-treatments


Corresponding author.
E-mail addresses: alana.gabrieli@ufabc.edu.br (A.G.d. Souza), daniel.rocha@ufabc.edu.br (D.B. Rocha), fkano@aluno.ufabc.edu.br (F.S. Kano),
derval.rosa@ufabc.edu.br (D.d.S. Rosa).
1
www.ufabc.edu.br

https://doi.org/10.1016/j.resconrec.2018.12.031
Received 14 September 2018; Received in revised form 26 December 2018; Accepted 27 December 2018
0921-3449/ © 2018 Published by Elsevier B.V.
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

results in greater efficiency in the isolation processes. Therefore, pre- Commercial nanocellulose was used as a reference for comparison. This
treatment is needed to increase the hydrolysis efficiency (Mohapatra work also presents an alternative to the revalorization of the industrial
et al., 2017). Different methods have been investigated for delignifi- paper wastes using the final sludge of paper production to isolate CNSs.
cation of the cellulosic wastes using alkalis, acids, high-pressure steam
explosion, and others (Li et al., 2017). The most common pretreatment 2. Materials and methods
process is the aal method, which makes use of different concentrations
of sodium hydroxide (e.g., 2, 4, 10, 20 and 30 wt%) (Chen et al., 2016a; 2.1. Materials
Sisti et al., 2016; Yue et al., 2015). This method possesses advantages
such as high lignin solubilization and limited formation of inhibitors via The paper residue (ETE sludge – according to the test report) used
fermentation. The removal mechanism involves the saponification of was kindly donated by Multiverde Papeis Especiais Ltda Company
intermolecular ester bonds between the hemicellulose and lignin (Lee (Mogi das Cruzes, SP, Brazil) (pH 9.33, solid content 24.3 wt%). Acetic
et al., 2014). However, other studies have used diluted acid solutions to acid (glacial, 99.85%), sulfuric acid (96%), hydrogen peroxide (34%)
remove lignocellulosic compounds. The most common acid pretreat- and sodium hydroxide (99.85%) were obtained from Sigma-Aldrich
ment are aqueous solutions of sulfuric and phosphoric acid. The ad- (Sao Paulo, Brazil).
vantages of these methods include higher reaction rates and increased
accessibility of cellulose (Lee et al., 2014). These pretreatments are a 2.2. Methods for the production of cellulose nanostructures
process to break the compact structure of lignocellulosic residues.
Bleaching pretreatments are conducted to ensure the breaking of lignin The cellulose nanostructures (CNSs) were obtained through two
(Pérez et al., 2002). Oxidative delignification is the most common different pretreatments. First, the paper residue (PR) was treated using
process, and this method uses peracetic acid or hydrogen peroxide and two different methodologies, an alkali (methodology I) and an acid
sodium hydroxide, which are efficient at removing lignin and in- (methodology II), as described below. The methods chosen for this work
creasing biomass digestibility (Lee et al., 2014). were applied after a study of different parameters present in the lit-
The production of CNSs via acid hydrolysis is the most common erature. Also, several tests were conducted to select the better condi-
process (Mascheroni et al., 2016). Acid hydrolysis typically uses sul- tions, aiming at lower reagent, energy and time expenditure, and en-
furic, hydrochloric, or phosphoric acid to remove less-stable structures suring a similar result to the conditions already reported in other
(e.g., the amorphous region of cellulose and the hemicellulose) after the studies. The selected parameters are present below.
lignin is removed. This process reduces the size and molar mass of the Methodology I: the alkali pretreatment was done adding the residue
cellulose, enabling the formation of 100–500 nm nanocrystals (Brinchi into a alkaline aqueous solution of 2% wt/v sodium hydroxide solution,
et al., 2013). An extensive array of approaches (Table 1) has been in the ratio 1:20 (residue (g): solution (ml)), at 50 °C for 1 h. Then, the
considered intending to decreasing the process severity of nanocellulose residue was washed to neutral pH, filtered and dried for 4 h in an oven
isolation and enhancing the yield/reaction efficiency. Besides that, with air circulation at 100 °C (sample PR-TI) (Berglund et al., 2016; Lee
there are other few methods to obtain nanostructures from different et al., 2014; Pinheiro et al., 2017; Yue et al., 2015).
materials, study their properties and ensure their stability. Methodology II: the acid pretreatment was done adding the residue
The methods described here are well developed and can be applied into an aqueous acid solution 92% v/v CH3COOH and 1.4% v/v H2SO4,
to raw materials in their natural state. However, industrial wastes are in the ratio 1:20 (residue (g): solution (ml)), at 80 °C for 3 h. Then, the
subject to harsh chemical processes and contain different residues in its residue was washed to neutral pH, filtered and dried for 3 h in an oven
composition. For paper residues, most of the cellulose has already been with air circulation at 50 °C (sample PR-TII) (Achaby et al., 2018; Lee
removed to produce the cellulose pulp. This situation leads to a material et al., 2014; Reddy et al., 2009; Souza et al., 2017).
with high hemicellulose and lignin content and creates a challenge in After the respective pretreatments, bleaching was done for both
defining a method to isolate the cellulose. Only a few studies have fo- methodologies, submitting the pre-treated residues in 34% v/v H2O2
cused on isolating CNSs from industrial wastes such as tobacco wastes and 4% wt/v NaOH solution, during 1 h at 50 °C (Ferreira et al., 2018;
(Tuzzin et al., 2016) and waste cotton fabrics (Xiong et al., 2012). Owolabi et al., 2016). The samples were washed until neutral pH, fil-
The conversion of waste materials into valued products reduces the tered and dried during 2 h at 80 °C. The bleached sample from PR-TI
environmental impacts of papermaking and adds value to the produc- was called PR-BI, and from PR-TII was called PR-BII.
tion chain. This study investigates isolation of CNS from an inexpensive The methodology of acid hydrolysis was adopted because already a
and underused raw material – paper residue. Also, the efficiency of two method used by industrial producers, and that can be applied in large
different pretreatments (alkali and acidic, in concentrations well es- scale. As reported by Reid et al., 2017, laboratory production leads to
tablished in the literature) on the removal of the lignocellulosic com- the same results that industrial production (Reid et al., 2017). The acid
pound from this industrial residue and pretreatments effect in the hydrolysis was done with 40% v/v aqueous H2SO4 solution for 50 °C
properties of the obtained CNSs by acid hydrolysis were evaluated. and 1 h, and the methodology was similar that the use by USDA Forest

Table 1
Different approaches for CNSs production.
Source Form Obtention Method Average size (nm) Yield (w %) Reference

Arecanut husk fiber CNF Homogenization 3–5 26 (Chandra et al., 2016)


Red algae marine biomass CNC H2SO4 hydrolysis 22 52 (Chen et al., 2016b)
Sugarcane bagasse CNC H2SO4 hydrolysis 30 – (Sofla et al., 2016)
CNF Ball Milling 50
Sago (Cycas circinalis) seed shells CNC H2SO4 hydrolysis 50 – (Naduparambath et al., 2018)
Tobacco stems CNF Supermass colloider 40 50 (Tuzzin et al., 2016)
wheat straw and kenaf fibers CNF Ball Milling 30 66 (Nuruddin et al., 2016)
Amorpha fruticosa Linn CNF High-pressure homogenization 10 – (Zhuo et al., 2017)
Fluff cellulose pulp CNF Dry grinding 118 67 (Lee and Mani, 2017)
Medium density fiberboards CNC H2SO4 hydrolysis 17 27 (Gu et al., 2017)
Garlic straw CNC H2SO4 hydrolysis 500 20 (Kallel et al., 2016)
Sawdust wastes CNC HCl hydrolysis and homogenization 100 – (Kalita et al., 2015)

134
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Products Lab (Houtman, 2016; Reid et al., 2017), where are used re- 3. Results and discussion
actors with 10 kg/day capacity. The solution was centrifuged at 6000
RPM during ten minutes and neutralized in a dialysis membrane for five 3.1. Compositional analysis
days. The hydrolyzed sample from PR-BI was called CNS-I, and from
PR-BII was called CNS-II. The chemical composition of the paper waste and pre-treated
The commercial nanocellulose (CNS-C) provided by Suzano Papel e samples are listed in Table 2. Based on the presented results, the in-
Celulose S.A. (Suzano, Brazil) (cellulose nanocrystals) was used to dustrial treatments have a significant effect on the chemical composi-
compare a commercial product with CNS-I and CNS-II. tion of these samples. The PR sample had a lower percentage of cellu-
lose (34.3%) and higher percentages of hemicellulose (23.3%) and
lignin (19.3%) compared to that of chemically treated samples. These
2.3. Characterization of the cellulose nanostructures values are similar to the lignocellulosic composition of softwood, the
feedstock for pulp production, whose lignin composition is between 20
The percentage composition of extractives and lignin was de- and 35 wt% (Hildebrandt et al., 2019; Lee et al., 2014).
termined by compositional analysis using an adapted method based on When applied the methodology I, the content of hemicellulose and
the neutral detergent fiber proposed by Van Soest (Van Soest et al., lignin decreased 22.7% and 35.2% respectively. When applied the
1991) of neutral and acid detergent and the TAPPI T 222om-22 (2002c) methodology II, the contents of hemicellulose and lignin decreased
methodology. 14.8% and 17.1%, respectively. So, it is possible to observe that the
Fourier transform infrared spectroscopy (FTIR) was conducted using methodology I was more efficient in the non-cellulosic components.
a Frontier 94,942 (PerkinElmer, USA) in the range from 400 to There is an enrichment of cellulose content because treatments partially
4000 cm−1 with 64 scans and a resolution of 4 cm−1. It was used an remove the non-cellulosic components from the surface of the samples
attenuated total reflectance accessory. (Chandra et al., 2016; Li et al., 2014).
The morphology of the samples was analyzed by scanning electron After the bleaching steps, lignin and hemicellulose were sig-
microscopy (SEM). After the samples were prepared and dried in an nificantly reduced, which resulted in enriched cellulose fibers con-
oven, as described in the experimental procedure, a little amount of the siderably. Although the samples show good results after the treatments,
dry powder of the samples was deposited with a spatula on carbon tape, the found contents are still above the ideal level described in the lit-
in an aluminum stub. The analyzed samples were: PR, PR-TI, PR-TII, erature (i.e., between 1 and 3%) (Delgado-Aguilar et al., 2016). Con-
PR-BI, and PR-BII. Subsequently, each sample was sputter-coated with sidering that the paper residue contains high concentrations of im-
gold. The samples were scanned in a JEOL - JCM 6010, operating with purities, the obtained result is positive, in this case, that aims at
an accelerating voltage of 5 KV. The Energy dispersive x-ray spectro- recycling of this previously discarded material. The above observations
scopy (EDS) spectra were obtained in the same equipment with samples indicate that the chemical treatments (alkali and acidic pretreatments
without gold coating. followed by bleaching) can damage the lignocellulosic structure, lead to
The size of the nanoparticles (cellulose nanostructures (CNSs)) was the cleavage of hemicellulose-lignin bonds and result into the removal
measured by the dynamic light scattering technique (DLS) (ALV-CGS3), of lignin (Li et al., 2014).
with 90° fixed scattering angle. A HeNe polarized laser (22 mW) with a The methodology I was more efficient at removing lignocellulosic
wavelength of 633 nm was used. Zeta potential values of the CNSs were components from industrial paper residues. However, both methodol-
measured using a zeta potential analyzer (Malvern Instruments). The ogies demonstrated a more significant decrease in hemicellulose and
suspensions were diluted to a concentration of 0.01 wt%. lignin compared with paper residue. Residual lignin can be a barrier
Atomic force microscopy (AFM) experiments were performed in a that reduces the effectiveness of the acid on the hydrolysis. Removal of
tapping mode in air using an AFM/SPM from Agilent, model 5500, at lignin could be improved by repeating the process or changing the
room temperature and contact operating mode. For images examina- temperature and time parameters to remove the remaining lignin
tion, a droplet of the aqueous suspension (1 wt% of CNS) was added (Delgado-Aguilar et al., 2016).
onto a cleaned silicon wafer and then airdried during 6 h, at room All of the samples presented significant ash in their composition,
temperature. The scan was done at a rate of 0.5 Hz, with an image re- which was expected considering the harsh chemical process involved to
solution of 256 × 256 pixels, with a silicon cantilever (nanosensor) of obtain cellulose pulp (Lynn et al., 2018; Thompson et al., 2001). The
constant force between 25 and 75 N.m−1 and resonance frequency of high ash concentration occurs due to the high concentration of che-
120 kHz. micals such as aluminum oxide, sodium hydroxide, titanium oxide and
Thermogravimetric analysis (TGA) was carried out using STA 6000 silicon oxide (Frías et al., 2015).
instrument (PerkinElmer, USA). The samples were heated from 20 to
600 °C with a heating rate of 20 °C/min under a nitrogen flow. 3.2. Fourier transform infrared spectroscopy
XRD results are presented in the Appendices. This analysis shows
the crystalline structure of the residue and its nanostructures. The FT-IR spectra allow us to evaluate the presence of lignocellulose
compounds functional groups and draw a correlation with composi-
tional analysis. The spectra of the PR, PR-TI, and PR-TII is presented in
Fig. 1(a), and the spectra of the PR, PR-BI, and PR-BII is presented in
Fig. 1(b). The chemically treated samples exhibit mutual spectral bands

Table 2
The chemical composition of the samples at each stage of treatment.
Sample Extractives (%) Hemicellulose (%) Insoluble Lignin (%) Soluble Lignin (%) Cellulose (%) Ashes and Residues (%)

PR 4.5 ± 0.5 20.3 ± 0.8 17.7 ± 2.8 1.6 ± 0.4 34.3 ± 4.8 21.6 ± 0.5
PR-TI 3.7 ± 0.3 15.7 ± 0.3 12.0 ± 3.4 0.5 ± 0.2 56.4 ± 5.3 11.7 ± 2.3
PR-TII 3.7 ± 0.6 17.3 ± 0.8 15.4 ± 3.3 0.6 ± 0.2 43.5 ± 6.1 19.5 ± 1.5
PR-BI 2.4 ± .0.1 11.7 ± 0.5 4.4 ± 3.2 0.1 ± 0.1 69.9 ± 4.3 11.5 ± 2.1
PR-BII 1.4 ± 0.4 12.1 ± 0.2 7.6 ± 3.8 0.3 ± 0.1 64.2 ± 5.6 18.4 ± 1.3
CNS-C 0.5 ± 0.1 12.8 ± 0.7 0.2 ± 0.1 – 85.0 ± 3.1 1.5 ± 0.9

135
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Fig. 1. The FT-IR spectrum of (a) PR, PR-TI, and PR-TII; (b) PR, PR-BI and PR-BII and (c) CNS-I, CNS-II, and CNS-C.

around 3340 and 1440 cm−1, that correspond to the stretching of −OH presence of cellulose structures after the acid hydrolysis (Henrique
groups and CH2 bending, respectively, which are associated with cel- et al., 2013; Jiang and Hsieh, 2013). Meanwhile, all CNSs exhibits
lulose, lignin and hemicellulose (Beltramino et al., 2016; Rosa et al., broadband in the region of 3500-3200 cm−1 due to the free OeH
2010; Soni et al., 2015; Tang et al., 2013). It is possible to identify peaks stretching vibration of hydroxyl groups in cellulose molecules (Cao
at 1250 and 1330 cm−1 corresponding to the lignin, and a peak at et al., 2012).
1040 cm−1 corresponding to C–OeC stretching; these vibration bands
are correlated with the lignin structure, which is probably due the
presence of residual lignin after the bleaching process (Chirayil et al., 3.3. Scanning electron microscopy
2014; Soucy et al., 2016; Xu et al., 2013; Zuluaga et al., 2009). Also, the
peak at 1640 cm−1 in all samples is relative to absorbed water in the The morphology of the fibers is believed to be affected by the
structure of cellulose. According to the literature, this band is asso- chemical treatments. Fig. 2(a) show the SEM images of the fibers at
ciated with the bending modes of water molecule due to a strong in- different stages of processing. The PR seems uneven due to the presence
teraction between cellulose and water (Haafiz et al., 2014). The pre- of binding materials like hemicellulose and lignin. These binding ma-
sence of residual lignin during the hydrolysis process shields the terials are removed, and thereby fibers get detached from the surface by
cellulose, protecting the amorphous fraction and negatively influencing the pretreatments. The surface changes after the treatments I and II
the likelihood of obtaining CNSs. indicate the removal of cementing materials, and the bleaching helps to
In Fig. 1(c), the spectra of CNS-I, CNS-II, and CNS-C are presented. eliminate the lignin (Chandra et al., 2016; Haafiz et al., 2014). The
The dominant peaks are at 1640 cm−1, 1120 cm−1, 1040 cm−1 and bleaching process results in saponification of the fibers, resulted by the
910 cm−1 that corresponds to absorbed water, C–O stretching vibra- alkalinity of the solution; and the addition of hydrogen peroxide causes
tion, C–H rock vibrations and C–H deformation vibration of cellulose an excess of groups OH− (Lee et al., 2014). After bleaching treatment,
(Li et al., 2012), indicating the presence of cellulose as a main com- fiber bundles become defibrillated (Chirayil et al., 2014). The sample
ponent in the CNSs (Haafiz et al., 2014; Rosa et al., 2010). There are PR-BI exhibits smoother and uniform fibrils surface, confirming the
also peaks of residual lignin at 1250 and 1330 cm−1 — also, the small removal of non-cellulosic components. However, it was not observed to
peaks in the 950–700 cm−1 region that represent the glycosidic -C1-O- PR-BII, which presented aggregated irregular shaped fibrils.
C4- deformation of the β-glycosidic link in cellulose, confirming the According to Frías et al. (2015), the paper sludge consists of two
fractions: the organic fraction (cellulose, lignin, and hemicellulose) and

136
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Fig. 2. a) representative scanning electron micrographs of PR, PR-TI, PR-TII, PR-BI, and PR-BII and b) chemical composition of the samples obtained by EDS analysis.

the inorganic fraction (kaolinite, quartz, and smaller proportion of 3.4. Dynamic light scattering and zeta potential
other constituents). The chemical composition obtained by the EDS
analysis of the samples is presented in Fig. 2(b) and shows the presence The DLS analysis considers the CNSs as spheres moving with
of natural components of lignocellulosic compounds, with significant Brownian motion (Mascheroni et al., 2016). By the measurement of the
content of carbon and oxygen. However, since this is a residue from diffusion coefficient of the nanostructures, it is possible to determine its
pulp production, there are present common elements of the industrial hydrodynamic radius (RH) (Brito et al., 2012). The size distribution of
bleaching process as titanium, aluminum and the silicon, derived from the particles is shown in Fig. 4(a).
their respective oxides (TiO2, Al2O3, and SiO2), which are the base The particle size distribution is remarkably monodispersed and ex-
composition of ashes obtained in compositional analysis (Frías et al., hibits a single sharp peak with sizes between 80 and 350 nm for CNS-I
2015; Segui et al., 2012; Yan and Sagoe-Crentsil, 2012). The samples and sizes between 187 and 1996 nm for CNS-II (Fig. 5(b)). The meth-
from methodology I present sodium from alkali treatment and odology I yielded smaller particles than methodology II. Fig. 4(c) shows
bleaching which is less present in methodology II, even after the a scheme that represents the morphology of the CNSs after acid hy-
bleaching process. All samples showed higher oxygen contents, which drolysis, considering the effect of the different pretreatments in the
probably occurs due to the presence of the oxides used in the paper final shape. The methodology I produced smaller nanospheres and rod-
industry on the wood treatments. Besides that, after each treatment, like structures; however, both methods result in nanostructures and
there was an increase in the metallic element contents, which can be aggregates. The yield of CNSs obtained concerning the initial amount of
justified by the removal of lignocellulosic components from the organic paper residue is around 31% and 24% for the CNS-I and CNS-II, re-
fraction. spectively. The smaller nanosizes of the CNS-I probably are due to the
The different mechanisms for the treatments are represented in higher access of the sulfuric acid to the interior of the fibers. These
Fig. 3. In treatment I, the alkaline solution acts via a saponification results were described before (FTIR and SEM analyses), and the sample
process, breaking ester bounds present on the lignin and hemicellulose PR-B1 has less lignin and hemicellulose in its composition, which makes
linkages and inducing the formation of surface roughness. In treatment the cellulose fibrils are more accessible to the acid. It results in higher
II, the acid solution hydrolyzes the bonds between the lignocellulosic efficiency in the scission of the cellulose structures and dissolution of
compounds and results in the degradation of superficial compounds to the amorphous phase of cellulose, which is between the crystalline
reach the internal fiber (Lee et al., 2014). The effectiveness of these phases. The amorphous cellulose removal facilitates the separation of
mechanisms has been confirmed based in the compositional analysis. the crystalline structures in nanocellulose (the changes in the crystal-
linity of these samples are showed in the Appendices).

137
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Fig. 3. Representation of mechanism of different treatments: alkaline and acid methodologies.

The use of DLS analysis presumes that the analyzed structure is a and the presence of residual lignin at the surface of obtained cellulose
spherical particle, which diverges from the majority of reports of CNS nanostructures. Lignin and hemicelluloses are not expected to be pre-
production (Chandra et al., 2016; Oksman et al., 2011). However sented at the surface of the CNSs, however, as seen at compositional
spherical CNSs have been reported for industrial residues of cotton fi- analysis, there is a residual content of these materials, which is prob-
bers and potato residue (Lu et al., 2013; Xiong et al., 2012). Atomic ably affecting the results (Tonoli et al., 2012). Thus, the SO3− groups
force microscopy analysis was conducted to confirm and support the are not presented in high enough quantity to change the zeta potential
majority formation of spherical CNSs. values, corroborating the EDS results (Yang et al., 2017).
The stability of nanocellulose suspension reflects the tendency of
CNSs to agglomerate. This characteristic can be studied by zeta po-
tential. The nanoparticles with similar electric charges repel each other 3.5. Atomic force microscopy
by the Coulomb force and thus affect the agglomeration and electro-
static stability of the suspension (Yang et al., 2017). The zeta potential The AFM topography images (Fig. 5(a), (b), (c) and (d)) confirm the
value for the CNS-I, CNS-II and CSN-C was −26.9 ± 7.0 and successful production of nanoparticles; individual nanoparticles and
−20.0 ± 2.71, −25.6 ± 0.4, respectively. The negative value in- aggregates were observed. The most abundant form of CNS was sphe-
dicates the presence of −OH groups on the cellulose surface rical (Fig. 5(a) and (b)). After treatment, individual crystals were ob-
(Naduparambath et al., 2018; Yang et al., 2017). From the data pre- tained showing a rod-like structure (Fig. 5(c) and (d)) (Haafiz et al.,
sented it may be inferred that the colloidal dispersions were not very 2014; Li et al., 2018; Nge et al., 2013; Xiong et al., 2012). The spheres
stable (Jiang and Hsieh, 2015). The zeta values of CNS-I are more (Fig. 5(a) and (b)) could be formed due to the self-assembled short
harmful than CNS-II, which means that CNS-I is more stable. Both cellulose rods via interfacial hydrogen bonds (Lu and Hsieh, 2010).
samples presented a similar zeta value comparing with CNS-C. The Also, the different forms of nanocrystals (spheres and rod-like struc-
dispersion difference could be related to the possibility of entanglement tures) cannot be separated by the conventional methods of filtration
and centrifugation (Lu and Hsieh, 2010). The spheres diameters were

Fig. 4. a) Scattering intensity curves as a function of the hydrodynamic radius (RH), (b) particles size of CNS-I, CNS-II, and CNS-C and (c) representation of CNSs
obtained after the hydrolysis.

138
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Fig. 5. 5 μm × 5 μm AFM topographical images of (a) CNS-I and (b) CNS-II; 2 μm × 2 μm AFM topographical images of (c) CNS-I and (d) CNS-II; (e) CNS-I and (f)
CNS-II diameter histogram.

estimated, and the data are presented in Fig. 5(e) and (f). The average compounds (Chandra et al., 2016). For the pre-treated samples
sizes are 200 nm for CNS-I and 305 nm for CNS-II, which corresponds (Fig. 6(a)), the initial decomposition temperature was 310, 295 and
well with the DLS results. These findings reveal that the different pre- 245 °C for the PR, PR-TI, and PR-TII, respectively. The PR sample
treatments result in differently sized nanospheres. The methodology I showed a higher T10% than the other samples due to the presence of the
yielded a smaller particle size. non-cellulosic constituents that act as barriers to initiate its thermal
degradation (Sá et al., 2015). Besides that, it is possible to observe that,
after the pretreatments, the samples tend to absorb higher humidity,
3.6. Thermogravimetric analysis which is verified by T10%. This behavior can be attributed to spaces left
vacant after the removal of hemicellulose and lignin. The open surfaces
The thermal stabilities of pre-treated, bleached and CNSs are pre- created in the pre-treated samples helps to absorb humidity (Sofla et al.,
sented in Fig. 6(a), (b) and (c), respectively. A first mass loss around 2016). Also, this thermal behavior is associated with the ashes and
100 °C is observed for both samples, and it is attributed to the eva- residues presence. The decrease in the thermal stability of the PR-TI and
poration of surface water, humidity, and low molecular mass

Fig. 6. Thermal curves obtained by TGA for (a) PR, PR-TI, and PR-TII; (b) PR, PR-BI, and PR-BII and (c) CNS-I, CNS-II, and CNS-C.

139
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

PR-TII is due to the greater exposition of the cellulose fibers due to the couple of years," which sounds, for now, a bit optimistic (Chauve and
efficiency of the pretreatments and bleaching (Du et al., 2016; Feng Bras, 2014; Reid et al., 2017). Nowadays, 1 pound (˜11.8 solids) can be
et al., 2015). purchased for $250 (Maine, 2018).
Fig. 6(c) shows the thermal behavior of the CNS-I, CNS-II, and CNS- Therefore, the developed CNSs have the great potential to be used as
C. The thermal degradation of cellulose and its nanostructures is critical reinforcement for the manufacture of high-performance nanocompo-
to assess the performance of the material and its possible applications sites, considering the thermal stability, dimensions, and morphology.
(Santmartí and Lee, 2018). For this reason, thermal stability has been Thus, this work demonstrates a method that can be applied in large-
extensively studied. In this study, the hydrolyzed samples exhibit a scale reactors, considering the adopted methodology, making the paper
lower T10% compared with the pre-treated samples. The observed va- waste commercially attractive. Overall, this study shows a scenario that
lues of the decomposition temperature are also lower than those of the can minimize the pulp consumption by the substitution of this raw
commercial nanocellulose, which is used as a comparison. The lower material to the CNS obtention by the paper residue and optimize the
degradation temperature could be due to the smaller particle dimen- capital investment by the high add value of this new subproduct. Some
sions (as shown in Fig. 3), which leads to higher specific surface area studies such as the life cycle assessment of this material could be con-
(Lu et al., 2013; Sofla et al., 2016; Soni et al., 2015). Also, smaller ducted to identify the potential risks associated with nanomaterials
particles result in many free end-chains where decomposition easily productions, and to estimate environmental impacts from this process
occurs, which justifies the lower thermal stability of the CNSs com- in lab scale, pilot scale and industrial scale.
paring with the pre-treated and bleached samples. Furthermore, CNS-C
has higher thermal stability due to the residual lignin present on the 4. Conclusions
CNS-I and CNS-II (Rambabu et al., 2016).
The differences in the thermal behavior of the hydrolyzed samples The present work showed that the cellulose nanostructures (CNSs)
and the CNS-C can be attributed to different factors, such as the degree could be successfully isolated from an abundant and low-cost residual
of crystallinity, presence of impurities and the type of cellulose sample source, paper residue, by acid hydrolysis. Before the isolation process,
(Santmartí and Lee, 2018). The difference in the first step of weight loss treatments with sodium hydroxide or acetic acid removed the non-
can also be attributed to a possible difference in the humidity of the cellulosic constituents resulting in fibers with high cellulose content.
CNS-C. However, the second step, the degradation, is a process linked to The alkaline treatment showed higher efficiency on the lignocellulosic
the scission of glycosidic bonds, leading to a reduction in the degree of compounds removal and yielded better results than acidic treatment, as
polymerization of cellulose. After the thermal activation, the next step it was confirmed via compositional analysis and FTIR concerning lig-
is the pyrolysis. This process is dependent on the cellulose structure and nocellulosic contents. So, the alkaline treatment was considered more
crystallinity, which can justify the difference between the hydrolyzed appropriate for CNSs isolation. After the treatments, CNSs were iso-
and commercial samples (Peng et al., 2013). It occurs because the hy- lated, and various characterizations showed the differences in the na-
drolyzed samples have cellulose I and II in its composition, and also nocellulose obtained from the pretreatments. CNSs revealed a reduction
some residues, which can act as a thermal catalyzer. Also, probably the in thermal stability. The nanocellulose present a combination of crys-
samples have a different degree of polymerization, which is influenced talline structures with cellulose I and cellulose II, with values of crys-
by the acid hydrolysis conditions. tallinity index for CNS-I, 75.9 and 68.6, and CNS-II, 66.9 and 57.6,
respectively (Appendices). The nanostructures presented spherical
3.7. Practical implications structures with average sizes varying according to the methodology:
195 nm for the alkaline method and 340 nm for the acid method, and
This work presents a nanotechnology applied in industrial residues, the yield was 31% and 24%, respectively; AFM and DLS analyses con-
and the developed material is based on three premises: (i) the resulting firmed these results. A possible correlation between the pretreatments
nanocellulose must be prepared from large-scale methodologies (i.e., and particles shape could be observed due to the residual lignin.
acid hydrolysis); (b) components must be available and abundant (such Comparing with the commercial CNS, the samples of this study showed
as the paper waste); (c) cost, environmental impact, and scalability excellent thermal stability, size distribution, and similar FTIR peaks.
must enable generalized use.
Concerning the acid hydrolysis, some aspects have been studied to Acknowledgments
minimize the environmental and economic impacts. Some options are
the recycling of reagents, the use of reactors and pipes, which must be The authors thank the UFABC and the Multiuser Central Facilities
highly resistant due to the extreme conditions of pH, temperature and/ (UFABC) for the experimental support. This work was supported by
or pressure required by the acid hydrolysis. Also, it is important a re- CNPq (grant numbers 447180/2014-2, 163593/2015-9).
cycling system of sulfuric acid to separate the acid of the digested su-
gars released during the reaction, which makes possible to produce Appendix A. Supplementary data
from 3 to 1000 kg/day of nanocellulose (Chauve and Bras, 2014).
The industries that produce CNS use cotton or softwood pulp as raw Supplementary material related to this article can be found, in the
material. Although these materials have high purity, properties like online version, at doi:https://doi.org/10.1016/j.resconrec.2018.12.
crystallinity, thermal and dimensional stability, size and morphology 031.
are similar to those presented in this work. Due to the extensive phy-
sicochemical characterizations carried out in this study, it is possible to References
compare the developed nanostructures with those currently commer-
cialized (Reid et al., 2017). Achaby, M.E., Kassab, Z., Barakat, A., Aboulkas, A., 2018. Alfa fibers as a viable, sus-
CNS are very expensive until now. The industrial plants were built tainable source for cellulose nanocrystals extraction: application for improving the
tensile properties of biopolymer nanocomposite films. Ind. Crop. Prod. 112, 499–510.
because the cost of CNS was estimated reasonable for the marketplace. https://doi.org/10.1016/j.indcrop.2017.12.049.
However, it needs to be cheaper to be commercially viable. For this, the Bayard, R., Benbelkacem, H., Gourdon, R., Buffière, P., 2017. Characterization of selected
current plants are constituted as a significant research laboratory that municipal solid waste components to estimate their biodegradability. J. Environ.
Manage. https://doi.org/10.1016/j.jenvman.2017.04.087.
understands and investigates the operational costs of a new process, to Beltramino, F., Roncero, M.B., Torres, A.L., Vidal, T., Valls, C., 2016. Optimization of
identify and evaluate new methods and process parameters, as well as sulfuric acid hydrolysis conditions for preparation of nanocrystalline cellulose from
new equipment for the production line. As anticipated by some players enzymatically pretreated fibers. Cellulose 23, 1777–1789. https://doi.org/10.1007/
s10570-016-0897-y.
in the field, CNSs will be sold "at just several dollars a kilogram within a

140
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Berglund, L., Noël, M., Aitomäki, Y., Öman, T., Oksman, K., 2016. Production potential of Jiang, F., Hsieh, Y.Lo, 2015. Cellulose nanocrystal isolation from tomato peels and as-
cellulose nanofibers from industrial residues: efficiency and nanofiber characteristics. sembled nanofibers. Carbohydr. Polym. 122, 60–68. https://doi.org/10.1016/j.
Ind. Crops Prod. 92, 84–92. https://doi.org/10.1016/j.indcrop.2016.08.003. carbpol.2014.12.064.
Bousios, S., Worrell, E., 2017. Towards a Multiple Input-Multiple Output paper mill: Johar, N., Ahmad, I., Dufresne, A., 2012. Extraction, preparation and characterization of
opportunities for alternative raw materials and sidestream valorisation in the paper cellulose fibres and nanocrystals from rice husk. Ind. Crops Prod. 37, 93–99. https://
and board industry. Resour. Conserv. Recycl. 125, 218–232. https://doi.org/10. doi.org/10.1016/j.indcrop.2011.12.016.
1016/j.resconrec.2017.06.020. Kalita, E., Nath, B.K., Agan, F., More, V., Deb, P., 2015. Isolation and characterization of
Brinchi, L., Cotana, F., Fortunati, E., Kenny, J.M., 2013. Production of nanocrystalline crystalline, autofluorescent, cellulose nanocrystals from saw dust wastes. Ind. Crops
cellulose from lignocellulosic biomass: technology and applications. Carbohydr. Prod. 65, 550–555. https://doi.org/10.1016/j.indcrop.2014.10.004.
Polym. 94, 154–169. https://doi.org/10.1016/j.carbpol.2013.01.033. Kallel, F., Bettaieb, F., Khiari, R., García, A., Bras, J., Chaabouni, S.E., 2016. Isolation and
Brito, B.S.L., Pereira, F.V., Putaux, J.L., Jean, B., 2012. Preparation, morphology and structural characterization of cellulose nanocrystals extracted from garlic straw re-
structure of cellulose nanocrystals from bamboo fibers. Cellulose 19, 1527–1536. sidues. Ind. Crops Prod. 87, 287–296. https://doi.org/10.1016/j.indcrop.2016.04.
https://doi.org/10.1007/s10570-012-9738-9. 060.
Cao, X., Ding, B., Yu, J., Al-Deyab, S.S., 2012. Cellulose nanowhiskers extracted from Lee, H., Mani, S., 2017. Mechanical pretreatment of cellulose pulp to produce cellulose
TEMPO-oxidized jute fibers. Carbohydr. Polym. 90, 1075–1080. https://doi.org/10. nanofibrils using a dry grinding method. Ind. Crops Prod. 104, 179–187. https://doi.
1016/j.carbpol.2012.06.046. org/10.1016/j.indcrop.2017.04.044.
Chandra, J.C., George, N., Narayanankutty, S.K., 2016. Isolation and characterization of Lee, H.V., Hamid, S.B.A., Zain, S.K., 2014. Conversion of lignocellulosic biomass to na-
cellulose nanofibrils from arecanut husk fibre. Carbohydr. Polym. 142, 158–166. nocellulose: structure and chemical process. Transfus. Apher. Sci. 2014. https://doi.
https://doi.org/10.1016/j.carbpol.2016.01.015. org/10.1155/2014/631013.
Chauve, G., Bras, J., 2014. Industrial Point of View of Nanocellulose Materials and Their Li, J., Wei, X., Wang, Q., Chen, J., Chang, G., Kong, L., Su, J., Liu, Y., 2012. Homogeneous
Possible Applications. [WWW Document]. https://doi.org/10.1142/ isolation of nanocellulose from sugarcane bagasse by high pressure homogenization.
9789814566469 0014. https://umaine.edu/pdc/cellulose-nanomaterials/order- Carbohydr. Polym. 90, 1609–1613. https://doi.org/10.1016/j.carbpol.2012.07.038.
nanocellulose/. Li, M., Wang, L., Li, D., Cheng, Y.-L., Adhikari, B., 2014. Preparation and characterization
Chen, N., Zhu, J.Y., Tong, Z., 2016a. Fabrication of microfibrillated cellulose gel from of cellulose nanofibers from de-pectinated sugar beet pulp. Carbohydr. Polym. 102,
waste pulp sludge via mild maceration combined with mechanical shearing. Cellulose 136–143. https://doi.org/10.1016/j.carbpol.2013.11.021.
23, 2573–2583. https://doi.org/10.1007/s10570-016-0959-1. Li, B., Ding, L., Xu, H., Mu, X., Wang, H., 2017. Multivariate data analysis applied in
Chen, Y.W., Lee, H.V., Juan, J.C., Phang, S.M., 2016b. Production of new cellulose na- alkali-based pretreatment of corn stover. Resour. Conserv. Recycl. 122, 307–318.
nomaterial from red algae marine biomass Gelidium elegans. Carbohydr. Polym. 151, https://doi.org/10.1016/j.resconrec.2016.12.007.
1210–1219. https://doi.org/10.1016/j.carbpol.2016.06.083. Li, X., Li, J., Gong, J., Kuang, Y., Mo, L., Song, T., 2018. Cellulose nanocrystals (CNCs)
Chirayil, C.J., Joy, J., Mathew, L., Mozetic, M., Koetz, J., Thomas, S., 2014. Isolation and with different crystalline allomorph for oil in water Pickering emulsions. Carbohydr.
characterization of cellulose nanofibrils from Helicteres isora plant. Ind. Crops Prod. Polym. 183, 303–310. https://doi.org/10.1016/j.carbpol.2017.12.085.
59, 27–34. https://doi.org/10.1016/j.indcrop.2014.04.020. Lu, P., Hsieh, Y.Lo, 2010. Preparation and properties of cellulose nanocrystals: rods,
Corcelli, F., Ripa, M., Ulgiati, S., 2018. Efficiency and sustainability indicators for pa- spheres, and network. Carbohydr. Polym. 82, 329–336. https://doi.org/10.1016/j.
permaking from virgin pulp—an emergy-based case study. Resour. Conserv. Recycl. carbpol.2010.04.073.
131, 313–328. https://doi.org/10.1016/j.resconrec.2017.11.028. Lu, H., Gui, Y., Zheng, L., Liu, X., 2013. Morphological, crystalline, thermal and physi-
Delgado-Aguilar, M., González, I., Tarrés, Q., Pèlach, M.À., Alcalà, M., Mutjé, P., 2016. cochemical properties of cellulose nanocrystals obtained from sweet potato residue.
The key role of lignin in the production of low-cost lignocellulosic nanofibres for Food Res. Int. J. 50, 121–128. https://doi.org/10.1016/j.foodres.2012.10.013.
papermaking applications. Ind. Crops Prod. 86, 295–300. https://doi.org/10.1016/j. Lynn, C.J., Dhir, R.K., Ghataora, G.S., 2018. Environmental impacts of sewage sludge ash
indcrop.2016.04.010. in construction: leaching assessment. Resour. Conserv. Recycl. 136, 306–314.
Domingues, A.A., Pereira, F.V., Rita, M., Rojas, O.J., Petri, D.F.S., 2016. Interfacial https://doi.org/10.1016/j.resconrec.2018.04.029.
properties of cellulose nanoparticles obtained from acid and enzymatic hydrolysis of Maine, The Process Development Center, 2018. Order Nanocellulose. [WWW Document]
cellulose. Cellulose. https://doi.org/10.1007/s10570-016-0965-3. https://umaine.edu/pdc/cellulose-nanomaterials/order-nanocellulose/ URL https://
Du, H., Liu, C., Mu, X., Gong, W., Lv, D., Hong, Y., Si, C., Li, B., 2016. Preparation and umaine.edu/pdc/cellulose-nano-crystals/. .
characterization of thermally stable cellulose nanocrystals via a sustainable approach Mascheroni, E., Rampazzo, R., Ortenzi, M.A., Piva, G., Bonetti, S., Piergiovanni, L., 2016.
of FeCl3-catalyzed formic acid hydrolysis. Cellulose 23, 2389–2407. https://doi.org/ Comparison of cellulose nanocrystals obtained by sulfuric acid hydrolysis and am-
10.1007/s10570-016-0963-5. monium persulfate, to be used as coating on flexible food-packaging materials.
Feng, X., Meng, X., Zhao, J., Miao, M., Shi, L., Zhang, S., Fang, J., 2015. Extraction and Cellulose 23, 779–793. https://doi.org/10.1007/s10570-015-0853-2.
preparation of cellulose nanocrystals from dealginate kelp residue: structures and Mastrantonio, G., Battaioto, L., Jones, C., Coustet, M., Chandi, H., Yamul, D.K., 2015.
morphological characterization. Cellulose 22, 1763–1772. https://doi.org/10.1007/ Chemical conversion of paper industry effluents into carboxymethylcellulose. Process
s10570-015-0617-z. Saf. Environ. Prot. 94, 315–321. https://doi.org/10.1016/j.psep.2014.07.005.
Ferreira, F.V., Mariano, M., Rabelo, S.C., Gouveia, R.F., Lona, L.M.F., 2018. Isolation and Mendes, C.V.T., Cruz, C.H.G., Reis, D.F.N., Carvalho, M.G.V.S., Rocha, J.M.S., 2016.
surface modification of cellulose nanocrystals from sugarcane bagasse waste: from a Integrated bioconversion of pulp and paper primary sludge to second generation
micro- to a nano-scale view. Appl. Surf. Sci. 436, 1113–1122. https://doi.org/10. bioethanol using Saccharomyces cerevisiae ATCC 26602. Bioresour. Technol. 220,
1016/j.apsusc.2017.12.137. 161–167. https://doi.org/10.1016/j.biortech.2016.07.140.
Frías, M., Rodríguez, O., Sánchez De Rojas, M.I., 2015. Paper sludge, an environmentally Mohapatra, S., Dandapat, S.J., Thatoi, H., 2017. Physicochemical characterization,
sound alternative source of MK-based cementitious materials. A review. Constr. modelling and optimization of ultrasono-assisted acid pretreatment of two
Build. Mater. 74, 37–48. https://doi.org/10.1016/j.conbuildmat.2014.10.007. Pennisetum sp. Using Taguchi and artificial neural networking for enhanced de-
Gu, J., Hu, C., Zhong, R., Tu, D., Yun, H., Zhang, W., Leu, S.-Y., 2017. Isolation of cel- lignification. J. Environ. Manage. 187, 537–549. https://doi.org/10.1016/j.jenvman.
lulose nanocrystals from medium density fiberboards. Carbohydr. Polym. 167, 70–78. 2016.09.060.
https://doi.org/10.1016/j.carbpol.2017.02.110. Naduparambath, S., J, T.V., Shaniba, V., S, M.P., Balan, A.K., Purushothaman, E., 2018.
Haafiz, M.K.M., Hassan, A., Zakaria, Z., Inuwa, I.M., 2014. Isolation and characterization Isolation and characterisation of cellulose nanocrystals from sago seed shells.
of cellulose nanowhiskers from oil palm biomass microcrystalline cellulose. Carbohydr. Polym. 180, 13–20. https://doi.org/10.1016/j.carbpol.2017.09.088.
Carbohydr. Polym. 103, 119–125. https://doi.org/10.1016/j.carbpol.2013.11.055. Nge, T.T., Lee, S.H., Endo, T., 2013. Preparation of nanoscale cellulose materials with
Henrique, M.A., Silvério, H.A., Flauzino Neto, W.P., Pasquini, D., 2013. Valorization of an different morphologies by mechanical treatments and their characterization.
agro-industrial waste, mango seed, by the extraction and characterization of its cel- Cellulose 20, 1841–1852. https://doi.org/10.1007/s10570-013-9962-y.
lulose nanocrystals. J. Environ. Manage. 121, 202–209. https://doi.org/10.1016/j. Nuruddin, M., Hosur, M., Uddin, M.J., Baah, D., Jeelani, S., 2016. A novel approach for
jenvman.2013.02.054. extracting cellulose nanofibers from lignocellulosic biomass by ball milling combined
Hernandez, C.C., Rosa, D.S., 2016. Extraction of cellulose nanowhiskers : natural fibers with chemical treatment. J. Appl. Polym. Sci. 133. https://doi.org/10.1002/app.
source, methodology and application. Polymer Science: Research Advances, Practical 42990.
Applications and Educational Aspects, 1st ed. Formatex Research Center. Oksman, K., Etang, J.A., Mathew, A.P., Jonoobi, M., 2011. Cellulose nanowhiskers se-
Hildebrandt, J., Budzinski, M., Nitzsche, R., Weber, A., Krombholz, A., Thrän, D., Bezama, parated from a bio-residue from wood bioethanol production. Biomass Bioenergy 35,
A., 2019. Assessing the technical and environmental performance of wood-based 146–152. https://doi.org/10.1016/j.biombioe.2010.08.021.
fiber laminates with lignin based phenolic resin systems. Resour. Conserv. Recycl. Owolabi, A.F., Haafiz, M.K.M., Hossain, S., Hussin, M.H., Fazita, M.R.N., 2016. Influence
141, 455–464. https://doi.org/10.1016/j.resconrec.2018.10.029. of alkaline hydrogen peroxide pre-hydrolysis on the isolation of microcrystalline
Houtman, C., 2016. Technoeconomic analysis of commercial-scale CNC production. cellulose from oil palm fronds. Int. J. Biol. Macromol. https://doi.org/10.1016/j.
TAPPI Nano. ijbiomac.2016.11.016.
Janissen, B., Huynh, T., 2018. Chemical composition and value-adding applications of Peng, Y., Gardner, D.J., Han, Y., Kiziltas, A., Cai, Z., Tshabalala, M.A., 2013. Influence of
coffee industry by-products: a review. Resour. Conserv. Recycl. 128, 110–117. drying method on the material properties of nanocellulose I: thermostability and
https://doi.org/10.1016/j.resconrec.2017.10.001. crystallinity. Cellulose 20, 2379–2392. https://doi.org/10.1007/s10570-013-0019-z.
Jaria, G., Silva, C.P., Ferreira, C.I.A., Otero, M., Calisto, V., 2017. Sludge from paper mill Pérez, J., Muñoz-Dorado, J., De La Rubia, T., Martínez, J., 2002. Biodegradation and
effluent treatment as raw material to produce carbon adsorbents: an alternative waste biological treatments of cellulose, hemicellulose and lignin: an overview. Int.
management strategy. J. Environ. Manage. 188, 203–211. https://doi.org/10.1016/j. Microbiol. 5, 53–63. https://doi.org/10.1007/s10123-002-0062-3.
jenvman.2016.12.004. Pinheiro, I.F., Ferreira, F.V., Souza, D.H.S., Gouveia, R.F., Lona, L.M.F., Morales, A.R.,
Jiang, F., Hsieh, Y.Lo, 2013. Chemically and mechanically isolated nanocellulose and Mei, L.H.I., 2017. Mechanical, rheological and degradation properties of PBAT na-
their self-assembled structures. Carbohydr. Polym. 95, 32–40. https://doi.org/10. nocomposites reinforced by functionalized cellulose nanocrystals. Eur. Polym. J. 97,
1016/j.carbpol.2013.02.022. 356–365. https://doi.org/10.1016/j.eurpolymj.2017.10.026.

141
A.G.d. Souza et al. Resources, Conservation & Recycling 143 (2019) 133–142

Rambabu, N., Panthapulakkal, S., Sain, M., Dalai, A.K., 2016. Production of nanocellulose Sun, B., Zhang, M., Hou, Q., Liu, R., Wu, T., Si, C., 2016. Further characterization of
fibers from pinecone biomass: evaluation and optimization of chemical and me- cellulose nanocrystal (CNC) preparation from sulfuric acid hydrolysis of cotton fibers.
chanical treatment conditions on mechanical properties of nanocellulose films. Ind. Cellulose 23, 439–450. https://doi.org/10.1007/s10570-015-0803-z.
Crops Prod. 83, 746–754. https://doi.org/10.1016/j.indcrop.2015.11.083. Tang, L., Huang, B., Lu, Q., Wang, S., Ou, W., Lin, W., Chen, X., 2013. Ultrasonication-
Reddy, K.O., Guduri, B.R., Rajulu, A.V., 2009. Structural characterization and tensile assisted manufacture of cellulose nanocrystals esterified with acetic acid. Bioresour.
properties of Borassus fruit fibers. J. Appl. Polym. Sci. 114, 603–611. https://doi.org/ Technol. 127, 100–105. https://doi.org/10.1016/j.biortech.2012.09.133.
10.1002/app. Thompson, G., Swain, J., Kay, M., Forster, C.F., 2001. The treatment of pulp and paper
Reid, M.S., Villalobos, M., Cranston, E.D., 2017. Benchmarking cellulose nanocrystals: mill effluent: a review. Bioresour. Technol. 77, 275–286. https://doi.org/10.1016/
from the laboratory to industrial production. Langmuir 33, 1583–1598. https://doi. S0960-8524(00)00060-2.
org/10.1021/acs.langmuir.6b03765. Tonoli, G.H.D., Teixeira, E.M., Corrêa, A.C., Marconcini, J.M., Caixeta, L.A., Pereira-Da-
Rosa, M.F., Medeiros, E.S., Malmonge, J.A., Gregorski, K.S., Wood, D.F., Mattoso, L.H.C., Silva, M.A., Mattoso, L.H.C., 2012. Cellulose micro/nanofibres from Eucalyptus kraft
Glenn, G., Orts, W.J., Imam, S.H., 2010. Cellulose nanowhiskers from coconut husk pulp: preparation and properties. Carbohydr. Polym. 89, 80–88. https://doi.org/10.
fibers: effect of preparation conditions on their thermal and morphological behavior. 1016/j.carbpol.2012.02.052.
Carbohydr. Polym. 81, 83–92. https://doi.org/10.1016/j.carbpol.2010.01.059. Tuzzin, G., Godinho, M., Dettmer, A., Zattera, A.J., 2016. Nanofibrillated cellulose from
Sá, R.Mde, Miranda, C.Sde, José, N.M., 2015. Preparation and characterization of na- tobacco industry wastes. Carbohydr. Polym. 148, 69–77. https://doi.org/10.1016/j.
nowhiskers cellulose from Fiber arrowroot (Maranta arundinacea). Mater. Res. 18, carbpol.2016.04.045.
225–229. https://doi.org/10.1590/1516-1439.366214. Van Soest, P.J., Robertson, J.B., Lewis, B.A., 1991. Methods for dietary Fiber, neutral
Santmartí, A., Lee, Koon-Yang, 2018. Crystallinity and Thermal Stability of detergent Fiber, and nonstarch polysaccharides in relation to animal nutrition. J.
Nanocellulose. Nanocellulose and Sustainability. pp. 68. https://doi.org/10.1002/ Dairy Sci. 74, 3583–3597. https://doi.org/10.3168/jds.S0022-0302(91)78551-2.
jhet.2254. Xiong, R., Zhang, X., Tian, D., Zhou, Z., Lu, C., 2012. Comparing microcrystalline with
Satari, B., Karimi, K., 2018. Citrus processing wastes: environmental impacts, recent spherical nanocrystalline cellulose from waste cotton fabrics. Cellulose 19,
advances, and future perspectives in total valorization. Resour. Conserv. Recycl. 129, 1189–1198. https://doi.org/10.1007/s10570-012-9730-4.
153–167. https://doi.org/10.1016/j.resconrec.2017.10.032. Xu, F., Yu, J., Tesso, T., Dowell, F., Wang, D., 2013. Qualitative and quantitative analysis
Segui, P., Aubert, J.E., Husson, B., Measson, M., 2012. Characterization of wastepaper of lignocellulosic biomass using infrared techniques: a mini-review. Appl. Energy
sludge ash for its valorization as a component of hydraulic binders. Appl. Clay Sci. 57, 104, 801–809. https://doi.org/10.1016/j.apenergy.2012.12.019.
79–85. https://doi.org/10.1016/j.clay.2012.01.007. Yan, S., Sagoe-Crentsil, K., 2012. Properties of wastepaper sludge in geopolymer mortars
Sisti, L., Totaro, G., Vannini, M., Fabbri, P., Kalia, S., Zatta, A., Celli, A., 2016. Evaluation for masonry applications. J. Environ. Manage. 112, 27–32. https://doi.org/10.1016/
of the retting process as a pre-treatment of vegetable fibers for the preparation of j.jenvman.2012.07.008.
high-performance polymer biocomposites. Ind. Crops Prod. 81, 56–65. https://doi. Yang, X., Han, F., Xu, C., Jiang, S., Huang, L., Liu, L., Xia, Z., 2017. Effects of preparation
org/10.1016/j.indcrop.2015.11.045. methods on the morphology and properties of nanocellulose (NC) extracted from corn
Sofla, M.R.K., Brown, R.J., Tsuzuki, T., Rainey, T.J., 2016. A comparison of cellulose husk. Ind. Crops Prod. 109, 241–247. https://doi.org/10.1016/j.indcrop.2017.08.
nanocrystals and cellulose nanofibres extracted from bagasse using acid and ball 032.
milling methods. Adv. Nat. Sci. Nanosci. Nanotechnol. 7. https://doi.org/10.1088/ Yue, Y., Han, J., Han, G., Aita, G.M., Wu, Q., 2015. Cellulose fibers isolated from en-
2043-6262/7/3/035004. ergycane bagasse using alkaline and sodium chlorite treatments: structural, chemical
Soni, B., Hassan, E.B., Mahmoud, B., 2015. Chemical isolation and characterization of and thermal properties. Ind. Crops Prod. 76, 355–363. https://doi.org/10.1016/j.
different cellulose nanofibers from cotton stalks. Carbohydr. Polym. 134, 581–589. indcrop.2015.07.006.
https://doi.org/10.1016/j.carbpol.2015.08.031. Zhuo, X., Liu, C., Pan, R., Dong, X., Li, Y., 2017. Nanocellulose mechanically isolated from
Soucy, J., Koubaa, A., Migneault, S., Riedl, B., 2016. Chemical composition and surface Amorpha fruticosa linn. ACS Sustain. Chem. Eng. 5, 4414–4420. https://doi.org/10.
properties of paper mill sludge and their impact on high density polyethylene (HDPE) 1021/acssuschemeng.7b00478.
composites. J. Wood Chem. Technol. 36, 77–93. https://doi.org/10.1080/02773813. Zuluaga, R., Putaux, J.L., Cruz, J., Vélez, J., Mondragon, I., Gañán, P., 2009. Cellulose
2015.1057647. microfibrils from banana rachis: effect of alkaline treatments on structural and
Souza, A.G.De, Kano, F.S., Bonvent, J.J., Rosa, D.S., 2017. Cellulose nanostructures ob- morphological features. Carbohydr. Polym. 76, 51–59. https://doi.org/10.1016/j.
tained from waste paper industry: a comparison of acid and mechanical isolation carbpol.2008.09.024.
methods. Mater. Res. 1–6.

142

You might also like