You are on page 1of 10

Research Article

Cite This: ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX pubs.acs.org/journal/ascecg

Water Hyacinth: A Sustainable Lignin-Poor Cellulose Source for the


Production of Cellulose Nanofibers
Supachok Tanpichai,*,†,‡,§ Subir Kumar Biswas,† Suteera Witayakran,⊥ and Hiroyuki Yano*,†

Research Institute for Sustainable Humanosphere, Kyoto University, Gokasho, Uji, Kyoto 611-0011, Japan

Learning Institute, King Mongkut’s University of Technology Thonburi, 126 Pracha Uthit Road, Bangkok 10140, Thailand
§
Cellulose and Bio-based Nanomaterials Research Group, King Mongkut’s University of Technology Thonburi, 126 Pracha Uthit
Road, Bangkok 10140, Thailand

Kasetsart Agricultural and Agro-Industrial Product Improvement Institute, Kasetsart University, 50 Paholyothin Road, Bangkok
Downloaded via UNIV OF CALIFORNIA SAN FRANCISCO on November 13, 2019 at 01:58:00 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

10900, Thailand
*
S Supporting Information

ABSTRACT: The extraction of cellulose nanofibers (CNFs) from a


lignocellulosic source containing less lignin would be an effective
way to avoid repetitious and energy-consuming chemical treatments.
In the present study, we used water hyacinth (Eichhornia
crassipes)a fast-growing, rapidly reproducing, sustainable, and
inexpensive raw material with a low lignin content (4.1%)to
successfully prepare CNFs with diameters of 10−30 nm and lengths
of several μm. We used three different chemical approaches:
chemical-free, alkaline, and combined sodium chlorite and alkaline
treatments. The results indicate that the alkaline treatment alone
was sufficient to eliminate most of the lignin and hemicellulose from
water hyacinth, providing CNFs with morphological, crystallinity,
and thermal characteristics similar to those of CNFs prepared using combined sodium chlorite and alkaline treatment. Also,
mechanical properties and thermal expansion of the nanopapers prepared from these chemically treated CNFs were
comparable. Water hyacinth has potential as a sustainable cellulose source for the large-scale production of CNFs for advanced
applications in tropical and subtropical countries in comparison with wood or other lignocellulosic sources due to a lower
requirement for chemical treatments. Moreover, water hyacinth has other positive aspects such as its rapid breeding rate,
availability, and economical price.
KEYWORDS: Cellulose nanofibers, Chemical treatment, Nanopaper, Mechanical properties, Thermal properties, Transparency

■ INTRODUCTION
Water hyacinth, Eichhornia crassipes, is a tropical monocotyled-
Each year, the Thai government spends a large amount of money
in an attempt to solve this problem, and the water hyacinth
onous aquatic species of the Pontederiaceae family (pick- population has become a national issue. After it is gathered from
erelweeds) that is native to the Amazon region of Brazil and areas of open water, water hyacinth typically ends up in landfill
Ecuador.1−3 It is a perennial plant that floats on water with thick or is used as a fertilizer or animal food.4,8 Small and medium-
leaves and purple flowers. The water hyacinth reproduces either sized enterprises extract fibers from water hyacinth for textile
by seeds or stolons. Vegetative propagation is the most common and paper applications. There have been several investigations
form of reproduction; it can double its population in 6−18 days into the possibility of using water hyacinth for the production of
and, in high-nutrient areas, in as little as 5 days. Within a month, ethanol and biogas.7,8
the number of water hyacinth plants can increase exponentially Water hyacinth is fast-growing, inexpensive, plentiful, and
from 1 to 1000 offspring.4 The yield of water hyacinth varies sustainable with a high rate of reproduction; furthermore, it
within the range of 60−150 tons per hectare.2,5 It is now contains less lignin and high cellulose contents.8,10,11 The lignin
widespread in the wetlands of tropical and subtropical countries content of water hyacinth is lower than 10 wt %,12−14 whereas
such as India, the U.S.A., Australia, and Thailand.1−3,6−9 Owing that of wood is 25−35 wt % and that of agricultural residues,
to its rapid rate of reproduction, large colonies of water hyacinth such as rice straw and sugar cane bagasse, is approximately 18 wt
form dense impermeable mats, clogging water bodies. This leads %.8,15−17 Fast-growing plants have a lower lignin content than
to environmental problems with fish populations, irrigation,
biodiversity, water quality, and water transportation.1,2,4,7,10,11 Received: July 19, 2019
The overall water hyacinth population floating on rivers in Revised: September 23, 2019
Thailand in January 2016 was reported to be 5.6 million tons.9

© XXXX American Chemical Society A DOI: 10.1021/acssuschemeng.9b04095


ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

wood or other slow-growing plants.8,18,19 Lignina complex


three-dimensional polymer that acts as an adhesive to bind
■ MATERIALS AND METHODS
Materials. We bought water hyacinth fibers from a local market in
microfibrils and cross-link wall componentsis essential for Thailand. Potassium hydroxide (KOH), acetic acid, and sodium
support, for upright growth in air, and for resisting gravity.8,20 As chlorite (NaClO2) were purchased from Wako Pure Chemical
in other aquatic plants, water hyacinth requires less lignin to Industries, Japan. All chemicals were used without any further
stiffen its body, and the less availability of lignin in its cell walls is purification.
the main characteristic of its evolution from a terrestrial to an Preparation of Cellulose Nanofibers. The preparation of CNFs
from water hyacinth is illustrated in Figure 1. We isolated the CNFs
aquatic existence.20,21 It is worth mentioning that the ancestors
of the aquatic family Pontederiaceae (water hyacinth) were
terrestrial before adaptation to life in an aquatic environment.22
Recently, cellulose nanofibers (CNFs) have received a great
deal of attention in a wide variety of applications such as
reinforcement in nanocomposites, biomaterials, paper, pack-
aging, electronic devices, and filtration membranes.23−27 This is
owing to their unique properties such as low density, low
thermal expansion, superior mechanical properties, biodegrad-
ability, biocompatibility, and a large surface-to-volume
ratio.16,28−31 Commonly, a series of chemical treatments is a Figure 1. Preparation of the cellulose nanofibers isolated from water
prerequisite for the removal of lignin and hemicellulose from hyacinth fibers.
lignocellulosic materials before disintegration.15,32,33 For
example, the well-known combination of acidified sodium
chlorite (NaClO2) and alkaline has long been used to treat
lignocellulosic materials such as wood,15 rice straw,15 potatoes,15 with the aid of alkaline treatment or a combination of NaClO2 and
alkaline treatments, and compared them to CNFs prepared using a
kenaf,33 and agricultural residues.32 With its long processing
chemical-free treatment. Before beginning the chemical treatment, we
time and high chemical and energy consumption, this chemical soaked the water hyacinth fibers in distilled water for 2 h at 90 °C to
purification process is very expensive.34 Moreover, an environ- remove water-soluble extractives such as carbohydrates.
mentally friendly enzymatic approach has been introduced to The acidified NaClO2 treatment was carried out according to a
facilitate nanofibrillation associated with the reduction of the method described by Abe and Yano,15 with modifications. Briefly, 25 g
energy consumption caused by mechanical treatment such as of the water hyacinth fibers were initially soaked in 1 L of distilled water
homogenization or grinding.35−37 But, this process requires an at 80 °C, and 10 g of NaClO2 and 1 mL of acetic acid were added to the
additional time-consuming step for the CNF preparation. fiber suspension. NaClO2 and acetic acid were introduced into the
Therefore, fibers with a limited lignin content would eliminate suspension every hour for 4 times. The treated fibers were filtered and
or reduce the need for such procedures and would provide a thoroughly rinsed with distilled water. The treated cellulose fibers were
promising alternative source for the preparation of CNFs to subsequently soaked in a 4 wt % KOH solution at 90 °C for 2 h, and this
step was conducted twice. After chemical treatment, fibrillation was
wood or plant-extracted cellulose sources, which contain a
initialized by a Vita-mix blender (ABSOLUTE 3; Osaka Chemical Co.,
higher lignin content.15,16 Ltd., Japan) at a speed of 37000 rpm for 2 min, and the slurry was then
In the present study, we selected the water hyacinthwhich passed twice through a grinder (MKCA6−2; Masuko Sangyo Co., Ltd.,
has a fast rate of growth, low lignin content, and a high rate of Japan) at 1500 rpm with a clearance gauge of −2.5 from the zero
reproduction and is sustainable and inexpensiveas a raw position (corresponding to 0.25 mm shift). The CNFs prepared using
material. To devise a greener approach using fewer chemicals, these treatments were designated as “S”.
we investigated three different chemical methods for preparing Alkaline-treated CNFs (referred to as “A”) were prepared as follows.
CNFs: no treatment, treatment with alkaline, and a NaClO2 and The cellulose fibers were soaked in a KOH solution (4 wt %) at 90 °C
alkaline treatment combination. The characteristics of these for 2 h, and this treatment was repeated twice. The treated fibers were
prepared CNFs were investigated using scanning electron primarily fibrillated with a Vita-mix blender and then treated in a
microscopy (SEM), chemical analysis, Fourier-transform infra- grinder under the same conditions described above. The CNFs
prepared without chemical treatment (referred to as “U”) were
red (FTIR) spectroscopy, X-ray diffraction (XRD), and
disintegrated by means of the blender and grinder. All the CNF
thermogravimetric analysis (TGA). We also produced nano- suspensions were kept at 4 °C before use.
papers from the CNFs and determined their transparency, Formation of Cellulose Nanofiber Nanopapers. We prepared a
wettability, mechanical properties, and thermal expansion. nanofiber suspension with a fiber content of 0.1% and vacuum-filtered
Although previous studies have reported the isolation of 250 g of the suspension on a glass filter funnel using a polytetrafluoro-
CNFs or cellulose nanocrystals from water hyacinth using a ethylene membrane filter with 0.1 μm pores to obtain a wet cellulose
combination of alkaline treatment and bleaching,10,13,38,39 as sheet. The wet sheet was placed between metal wire nets and filter
well as the use of water hyacinth-extracted CNFs and cellulose papers and dried at 110 °C for 30 min at a pressure of 0.1 MPa to
nanocrystals to reinforce polymer matrixes,40−42 none has taken produce a CNF nanopaper with a thickness of ∼40 μm and a diameter
advantage of the lower lignin content of water hyacinth to of ∼7.5 cm.
Chemical Component Analysis. The cellulose, hemicellulose,
prepare CNFs without the delignification step or with fewer
and lignin contents of the untreated and chemically treated water
chemical treatments for a greener approach. The use of water hyacinth fibers were determined according to the method proposed by
hyacinth would definitely be beneficial because it would reduce the Technical Association of the Pulp and Paper Industry (TAPPI).5
the production time, chemical cost, and solvent disposal expense The comparison of chemical components of water hyacinth and other
of CNF fabrication compared with similar considerations plants is presented in the Supporting Information.
involved in the use of wood or other lignocellulose sources Yield. The yield of the water hyacinth fibers after the chemical
with a high lignin content. treatment was calculated using the following equation

B DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

ij dried weight of the chemically treated fibers yz


Yield (%) = jjj
j
k dried weight of the raw fibers
zz × 100
zz
{
■ RESULTS AND DISCUSSION
Chemical Components. The water hyacinth used in the
(1) present study mainly comprised 57% cellulose, 25.6% hemi-
cellulose, and 4.1% lignin. We found that the contents of
Field-Emission Scanning Electron Microscopy. The morphol- cellulose and hemicellulose were higher than those previously
ogy of the raw water hyacinth fibers and CNFs was investigated using a reported owing to variations in the environment atmosphere
field-emission scanning electron microscope (JSM-7800F; JEOL, Ltd., such as water temperature, climate and density of water hyacinth
Japan) with an acceleration voltage of 1.5 kV and a working distance of
10 mm. All the samples were sputter-coated with platinum to avoid
in that area, maturity, and method of the chemical
charging by the ion sputter coater (JEC-3000FC; JEOL, Ltd., Japan) determination, whereas there was good agreement with previous
before investigation. We used the ImageJ program to determine the studies regarding the low amount of lignin.12−14 Owing to its
widths of the CNFs. flexibility and rapid rate of growth, water hyacinth has less lignin
Fourier-Transform Infrared Spectroscopy (FTIR). FTIR meas- and more cellulose than agricultural crops such as pineapple
urements of the CNF samples were obtained using a spectrometer leaves, bagasse, and switchgrass.12−14,45−49,17,50 Table S1 of the
(Nicolet IR200; Thermo Fisher Scientific Co., Ltd., U.S.A.) in Supporting Information compares the chemical components of
attenuated total reflectance (ATR) mode. The dried samples were water hyacinth with those of plants reported in previous studies.
scanned at wavenumbers in the range 400−4000 cm−1 and measured at Therefore, cellulose-rich water hyacinth with a low lignin
4 cm−1 resolution with 64 repeated scans. The background was content has potential as an alternative to wood and agricultural
recorded at the beginning of the measurement.
residues as a source of cellulose for CNF production.
X-ray Diffraction (XRD). We obtained the XRD profiles using an X-
ray diffractometer (D8 DISCOVERY; Bruker AXS, Germany)
The contents of cellulose, hemicellulose, and lignin of the
equipped with a Goebel mirror. The sample was mounted in a sample untreated and chemically treated water hyacinth fibers are
holder and scanned with Cu Kα radiation at a wavelength of 0.154 nm, presented in Figure 2. As is usual after chemical treatment, there
with the equipment operated at 40 kV and 40 mA. The scattered light
was detected in the 2θ range between 5 and 50°, and the sample was
scanned at a speed of 0.8 s per step and an angle step increment of 0.02°.
The degree of crystallinity (CI) was calculated, according to Segal’s
method,43 from the XRD profiles using the following equation:

ij I − IAM yz
CI (%) = jjj 200 zz × 100
j I200 zz
k { (2)

where I200 is the maximum intensity of the crystalline peak located at 2θ


between 22 and 23°, and the minimum intensity of the amorphous
region at 2θ between 18 and 19°.44
Thermogravimetric Analysis. We investigated the thermal
properties of the prepared CNFs using a Q50 thermogravimetric
analyzer (TA Instruments, U.S.A.). Approximately 5 mg of each dried
sample was placed on a platinum pan, stabilized at 110 °C for 10 min to
remove moisture, and analyzed from 110 to 600 °C at a heating rate of
10 °C min−1 under nitrogen supplied at a flow rate of 100 mL min−1. Figure 2. Chemical constituents of untreated and chemically treated
UV−vis Spectroscopy. The regular transmittances of the CNF water hyacinth.
nanopapers were measured using a UV−vis spectrophotometer (U-
4100; Hitachi High-Tech Corp., Japan) with a 60 mm diameter
integrating sphere. Each sample was placed close to the entrance port of is an increase in the proportion of cellulose and a reduction in
the integrating sphere and scanned at wavelengths from 200 to 800 nm. the proportions of lignin and hemicellulose in the treated fibers.
Contact Angle Measurement. To determine the surface Surprisingly, following treatment with alkaline only (treatment
hydrophilicity of the nanopapers, the contact angle was measured by A), the cellulose content increased to 89.5%, and the contents of
a sessile drop method using a static optical contact angle meter hemicellulose and lignin were markedly reduced to 5.3 and 0.4%,
(SL150E; USA KINO Industry Co., Ltd., U.S.A.) equipped with a
camera. A 1 μL water droplet was placed on a surface of the dried CNF
respectively. This large reduction in the contents of hemi-
nanopaper with a microsyringe at 23 ± 2 °C. cellulose and lignin may have been due to the porous structure of
Mechanical Testing. We determined the tensile properties of the the water hyacinth fibers, which facilitated the ingress of the
CNF nanopapers using an Instron 3365 Universal Testing Machine alkaline solution. During the alkaline treatment, the fibers
(Instron Corp., U.S.A.) equipped with a load cell of 5 kN. The 5 mm swelled, causing the solubilization of both hemicellulose and
wide samples were tested at a cross-head speed of 1 mm min−1 and a lignin and cleavage of the α-ether linkages and ester bonds
gauge length of 20 mm. Prior to testing, we kept the specimens in a between the lignin and hemicellulose.51 A longer treatment time
controlled room at 23 ± 2 °C and a relative humidity of 50 ± 2% for 24 and a higher alkaline concentration resulted in the elimination of
h. The Young’s modulus was calculated from the slope of the initial more lignin. In the present study, we observed a delignification
linear region of the stress-strain curve. At least five samples were tested efficiency of 91.1% following the treatment with 4% alkaline for
to obtain averages and standard deviations for the tensile strength, 120 min. After the fibers were treated with the combination of
Young’s modulus, and strain at break of each material.
Thermal Expansion. We used a thermomechanical analyzer
acidified NaClO2 and alkaline (treatment S), the contents of
(TMA/SS6100; Seiko Instruments, Japan) to determine the coefficient cellulose, hemicellulose, and lignin were 85.0, 11.7, and 0.2%,
of thermal expansion (CTE) values of the CNF nanopapers in tensile respectively. In comparison with the A fibers, the S fibers had a
mode over the temperature range 20−150 °C at a heating rate of 5 °C lower content of cellulose and a higher content of hemicellulose.
min−1 and a preload force of 29.4 mN under a nitrogen atmosphere. The increase of the hemicellulose content in the S sample could
The CTE values were calculated between 20 and 150 °C. be attributed to the degradation of cellulose caused by the
C DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

Figure 3. Scanning electron microscopy (SEM) images of a cross-section of the water hyacinth fiber (a) and cellulose nanofibers (CNFs) extracted
from water hyacinth (b) without chemical treatment, (c) by alkaline treatment, and (d) by treatment with a combination of sodium chlorite (NaClO2)
and alkaline. Arrows indicate the fiber aggregations.

Figure 4. (a) Fourier-transform infrared (FTIR) spectra, (b) X-ray diffraction (XRD) patterns, (c) thermogravimetric (TG) curves, and (d) derivative
thermogravimetric (DTG) curves of the cellulose nanofibers (CNFs) following the various chemical treatments.

D DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

reduction in the molecular weight and degree of polymerization, Chemical Structures. The chemical structures of the CNFs
resulting from the harsh NaClO2 and alkaline treatment. The prepared by the various chemical procedures are compared in
reduction in the degree of polymerization of cellulose caused by Figure 4a. After chemical treatment, the CNFs (A and S) had
acid hydrolysis and oxidative cleavage of the cellulose chains similar FTIR spectra because they had similar contents of
during the acidified NaClO2 treatment has already been cellulose, hemicellulose, and lignin. The absorption bands at
reported.52−54 Effect of the NaClO2 delignification on the 3,334 cm−1, attributable to the hydroxyl groups of cellulose,62
degree of polymerization of cellulose has been studied by and at 1,060 cm−1, attributable to C−O stretching of the
Hubbell et al.54 The degree of polymerization of cellulose cellulose structure,10 were more intense in the chemically
decreased to 303 and 294 when the lignin content in the treated CNFs (A and S) than in the U nanofibers. Furthermore,
cellulose samples were 1 and 0%, respectively, while untreated the A and S nanofibers, which contained less hemicellulose and
cellulose had a degree of the polymerization of 309. Similarly, lignin, produced hemicellulose- and lignin-related peaks with the
there was a significant reduction in the degree of polymerization weaker intensity, e.g., the 1,640 cm−1 peak, attributable to the
of cellulose (from 1500 to 410) when the filter paper was treated carbonyl groups of hemicelluloses,16,63 and the 1,240 cm−1
with acidified NaClO2 for 6 h.53 Moreover, when the NaClO2- peak,64,65 attributable to ester, ether, or phenol compounds.
delignified bamboo powders (degree of polymerization of 754 Peaks attributable to lignin in the A and S samples, such as 1,740
and molecular weight of 12 × 105) were treated with alkaline cm−1, assigned to either the acetyl and uronic ester groups of
treatment at room temperature for 15 h, the result was hemicelluloses or the ester linkage of the carboxylic groups of
reductions in the degree of polymerization to 556 and molecular lignin,10,16,61 and those at 1,512 and 858 cm−1, corresponding to
weight to 0.9 × 105.55 This might be the cause of the higher C=C and C−C stretching in the aromatic rings of the lignin
content of hemicellulose after treatment with NaClO2 and structure,6,16,61disappeared. These changes and the disappear-
alkaline treatments in this study. ance of the bands confirmed the removal of hemicellulose and
In contrast to a study by Nobuta et al.,33 treatment with lignin, and the incremental increase in the proportion of
alkaline alone slightly reduced the hemicellulose and lignin cellulose in the chemically treated CNFs, which agreed well with
contents of kenaf bast fibers owing to their high lignin content. the chemical component results.
The combined acidified NaClO2 and alkaline treatment is still Crystallinity and Thermal Stabilities. We carried out
required to cleave and dissolve lignin and hemicellulose in high- XRD measurements to investigate the effects of chemical
lignin content fibers and residues, whereas the alkaline treatment treatment on the crystalline structures of the nanofibers, because
is sufficient to purify cellulose fibers with a limited lignin content cellulose contains amorphous and crystalline regions. Figure 4b
extracted from the water hyacinth. shows the XRD patterns of the nanofibers prepared using the
Owing to the removal of lignin and hemicellulose and the various procedures. We found similar diffraction patterns at
partial degradation of cellulose, the fiber yield from the material 16.5, 22.5, and 34.5° corresponding to the (110), (200), and
treated with a combination of NaClO2 and alkaline (S) was (400) planes (the figures in parentheses represent the Miller
42.2% whereas the mild alkaline treatment yielded a higher fiber indices) for all three CNF samples, indicating typical cellulose I
content of 50.6%, and the yield from untreated hyacinth fibers structure.33,59 This suggests that the crystalline structure of the
was 83.1%. fibers could not be affected by the chemical treatments used in
Morphology. A cross-section of a water hyacinth fiber is the present study. A structural transition from cellulose I to
presented in Figure 3a. Unlike flax or wood, which contain a cellulose II arises when a high-concentration alkaline solution
central luminal cavity surrounded by a cell wall,56 water hyacinth (20 wt %) is used to treat the fibers.66 At this level of alkaline
fibers consist of porous subfibers that are similar to celery concentration, swollen fibers with less rigid interfibrillar
fibers.57 The large air cavities inside the fibers allow the water structures are formed, allowing the development of different
hyacinth to float on water.7,11,58 The morphologies of CNFs types of intra- and interhydrogen bonding. The degree of
produced using different chemical treatments are compared in crystallinity of the U CNFs was 67.7%, according to Segal’s
Figure 3b−d. The field emission scanning electron microscope method.43 Following chemical treatment, the degree of
(FE-SEM) images reveal the successful defibrillation of CNFs crystallinity of the CNFs increased (76.8% for A and 78.8%
produced from water hyacinth with or without chemical for S). This increase was unquestionably due to the removal of
treatment. Free-chemical treated CNFs (U) with diameters of hemicellulose and lignin. Similar increases in the degree of
23.7 ± 5.1 nm were successfully fibrillated from the water crystallinity after a series of chemical treatments have also been
hyacinth fibers, but groups of fiber aggregates were easily reported for other cellulosic materials such as wood, bamboo,
observed. The thicker widths of the U nanofibers and fiber rice straw, and flax.61 Furthermore, mechanical treatments, such
aggregates could be attributed to the less disintegration caused as grinding15,16 and high-intensity sonication,59 have a smaller
by the presence of lignin. With the aid of the alkaline treatment, effect on the cellulose structure but do improve its crystallinity.
the widths of the A nanofibers decreased slightly to 22.1 ± 5.2 It is worth noting that although the S nanofibers contained a
nm, and no fiber aggregation could be found for this sample, lower concentration of cellulose than the A sample, the
implying full disintegration. In the meantime, CNFs (S) with the combined sodium chlorite and alkaline treatment could destroy
widths of 18.4 ± 4.7 nm were disintegrated after the the amorphous region in cellulose, increasing the crystalline
combination of NaClO2 and alkaline treatment. This indicates region in cellulose.37 The dissolution of amorphous cellulose
that only the alkaline treatment is sufficient to isolate nanofibers could be found when cellulose was alkaline-treated for 60 min,
from water hyacinth with the aid of the grinding treatment. The increasing the crystallinity of the cellulose sample.67 Similar
nanofibers prepared from the water hyacinth in the present work increase of crystallinity has been found when the bamboo fibers
are comparable to those extracted from other plants such as were treated with the alkaline solution with a concentration of 4
banana,59 oil palm,60 wood,61 bamboo,61 and flax61 in terms of wt %.68
the width and length. It has been reported that the diameter We investigated the thermal stability of the CNFs produced
distribution of the CNFs depends on the cellulose source.15 by various chemical treatments using thermogravimetric analysis
E DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

Figure 5. (a) UV/visible transmittances of the cellulose nanofiber (CNF) nanopapers and (b) water contact angles of the nanopapers.

(TGA), and the thermogravimetric (TG) and derivative of hemicellulose and lignin.78,79 The onset degradation
thermogravimetric (DTG) curves of these CNF samples are temperature of U was 298.5 °C, which was lower than that of
shown in Figure 4c,d. A single thermal decomposition stage was the A and S CNFs. The lower thermal stability of the U CNFs
observed for the CNFs prepared by chemical treatments (A and was attributed to the presence of hemicellulose with the low
S) whereas the U CNF had several thermal decomposition thermal degradation temperature.78 The A and S CNFs began to
stages, implying multiple components in the nanofibers. The degrade at 313.7 and 312.6 °C, respectively. The similar thermal
initial decomposition of hemicelluloses begins in the range 220− stabilities of the A and S CNFs suggest that alkaline treatment is
315 °C, and cellulose subsequently decomposes at 315−400 °C sufficient to remove most lignin and hemicellulose from a
owing to the decarboxylation, depolymerization, and decom- cellulose source in comparison with the use of the NaClO2 and
position of glyosidic linkages.69,70 At temperatures higher than alkaline treatment. In addition, the higher residue content of the
400 °C, cellulose is almost pyrolyzed with a small amount of U nanofibers at 600 °C was attributed to the formation of a
solid residues.69 complex between cellulose and lignin and to the aromatic
The lower thermal stability of hemicellulose is caused by its structure of the lignin complexes in the fibers.69,74 A greater
structure of the random amorphous saccharides with rich of content of solid residues has been reported following lignin
branches (xylose, mannose, glucose, galactose, etc.), which is pyrolysis at 900 °C.69 Similarly, a lower thermal decomposition
easily converted to volatiles (CO, CO2, and hydrogen carbon) at temperature and higher amounts of carbonized residues have
low temperature. Cellulose is a linear polysaccharide polymer been reported for raw banana fibers in comparison with
with a high order of arrangement and higher molecular weight, bleached and acid-treated banana fibers.80
making it more stable than hemicellulose at high temper- Transparency and Wettability. The total light trans-
ature.69,71 Furthermore, lignin decomposes in a wider temper- mittances of the CNF nanopapers in the wavelength range 200−
ature range than hemicelluloses and cellulose (200−600 °C)72 800 nm are presented in Figure 5a. The U sample was brownish,
because its structure comprises various functional oxygen groups and the nanopapers became colorless following treatment with
with different thermal stabilities, such as aromatic carbons, acidified NaClO2 and alkaline, owing to the elimination of lignin
phenolics, and hydroxyphenolics.73,74 Thermal decomposition by the reactions of chlorine and the chlorite ions from
of lignin can be separated into two major states. The initial NaClO2.33 The S nanopaper had the highest light transmittance
decomposition of lignin is found between 120 and 300 °C, (49.8% at 600 nm) whereas the A and U nanopapers had
resulting in the fragmentation in the phenyl propane side chains transmittances of 33.6 and 16.7% at the same wavelength,
at the end positions which forms formic acid, formaldehyde, respectively. The high intensity of the brownish color observed
water, carbon dioxide, and sulfur dioxide.73 A small component in the U nanopapers was proportional to the availability of lignin,
of lignin is likely degraded in this state.71,75 The main thermal which contains chromophore groups such as phenolic units,
decomposition occurs between 300 and 480 °C with more than ketones, and other chromophores that absorb light81,82 while the
50% of the mass loss. The fragmentation and cleavage of the lower light transmittance of the U samples was due to the high
major chain linkages between monomeric phenol units in the lignin content, larger width CNF distribution, and fiber
lignin structure yield methanol, carbon dioxide, carbon aggregates. Furthermore, the lower light transmittance of the
monoxide, water, methane, and two phenolic compounds U samples can be attributed to less formation of hydrogen
(guaiacol and a 2-methoxy-4-akyl-substituted phenol).73,74,76 bonding between fibrils within the paper (small gaps formed
The difference of these two decomposition stages of lignin has between fibrils) interfered by lignin. The higher light trans-
been found to be dependent on the lignin sources.77 Zhang et mittance obtained from the S samples over the A samples could
al.75 studied the relationship between the activation energy and be attributed to the smaller widths of the S nanofibers and lesser
the thermal decomposition reaction of CNFs with various lignin content. A higher content of hemicellulose in the S CNFs
contents of lignin and found that higher energy was required to might be another key to provide a higher value of the light
thermally degrade CNFs which contained a higher content of transmittance. Hemicellulose has been reported to prevent the
lignin. coalescence of microfibrils (hornification) by physically
The thermal stability of the CNFs increased following hindering the direct contact between microfibils.83 The
chemical treatment. This increase was caused by the removal cohesion forces between microfibrils could be also deducted
F DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

Figure 6. (a) Stress−strain curves and (b) thermal expansion curves of the cellulose nanofiber (CNF) nanopapers.

by the moisture absorption of hemicellulose during rewet- owing to the reduced amount of the hydroxyl groups from
ting.83,84 cellulose. Therefore, the reduction of the surface energy of the
Moreover, the U nanopapers completely absorbed UV light fibrils could be caused by the appearance of lignin due to the
(200−380 nm) while around 7 and 35% of the UV light could larger percentage of C−C and C−H bonds and the lower
pass through the A and S nanopaper samples. The difference of composition of O to C in lignin compared with cellulose.85
the transmittance between the A and S nanopapers could be Mechanical Properties and Thermal Expansion. The
attributed to the lignin content. Although the lignin content in density and mechanical properties of the prepared nanopapers
the A and S samples was different by only 0.2 wt %, this are summarized in Table S2 of the Supporting Information, and
difference might be enough to block more light and exhibit the the stress−strain curves of the CNF nanopapers are shown in
lower light transmittance of the A nanopapers. This might be Figure 6a. The densities of the U, A, and S nanopapers were not
because during the nanopaper preparation, the nanopaper was significantly different with values close to ∼1.3 g cm−3. The
pressed at 110 °C for 30 min, and this process condition could filling of voids between nanofibrils with lignin during the process
be possibly enough for lignin as a binder to soften and fill cavities of hot-pressing and compression has been not related to the
between nanofibers.71,85 This random filling of lignin in pores densities of the nanopapers.85 The U nanopapers had the lowest
between nanopapers had an influence on transmittance, tensile strength (40.6 MPa) and Young’s modulus (4.3 GPa).
wettability, mechanical properties, and thermal expansion of These low values arose because lignin prevented the formation
the prepared nanopapers. Recent studies have shown that the of hydrogen bonds between the CNF fibrils and the fiber
light transmittance of CNF nanopapers decreases as the lignin aggregates; the mechanical properties of nanopapers are
content increases.82 It is worth mentioning that the U samples dominated by the properties of the fibers and by the bonding
exhibited a dip in transmittance at ∼670 nm (the red region) strength.82,87 With the removal of lignin, the values of the
whereas the other materials showed no such abnormal behavior. mechanical properties of the CNF nanopapers (strength, strain,
This might have been due to the presence of chromophores in and Young’s modulus) increased significantly. The S nanopapers
the lignin that are particularly light absorbent in that region. had a tensile strength of 73.3 MPa and a Young’s modulus of 6.3
Therefore, the lignocellulosic CNF nanopapers would be useful GPa whereas the A nanopapers had a tensile strength of 66.6
for UV protection. Moreover, surface light scattering plays an MPa and a Young’s modulus of 5.6 GPa. The slightly higher
important role in the transparency of nanopapers.86 mechanical properties of the S nanopapers might be ascribed to
The wettability values of the CNF nanopapers, evaluated by the smaller-width nanofibers and stronger interactions between
measuring the water droplet angles on their surfaces, are the individual CNF fibrils, but the presence of a small amount of
compared in Figure 5b. The U nanopapers had a higher contact lignin (0.4%) within the A nanopapers may have caused fibril
angle of >90°, suggesting that a greater proportion of lignin in slippage during tensile deformation, leading to a higher strain,
the CNFs makes the nanopaper less wettable. The A (60.2°) and and slightly lower values for strength and modulus. Also, the
S (48.1°) nanopapers, which contained less lignin, were more higher strength and modulus of the S nanopapers over the A
hydrophilic. Similar results have been reported for CNF nanopapers might also be due to the higher content of
nanopapers with different lignin contents.81,85 CNF nanopapers hemicellulose. Hemicellulose has been found to adhere
with a higher lignin content exhibited lower total water nanofibers and prevent hornification in the dried state, yielding
absorption capacity and less wettability. This was due to the improvement in the stiffness and strength of the nanopapers.83
higher hydrophobicity of lignin in comparison with cellulose and However, other works have revealed large differences in the
hemicellulose, resulting in the reduction of the accessibility of mechanical properties of nanopapers formed from CNFs
water molecules to interact with hydroxyls groups of cellulose produced by alkaline treatment and a combined acidified
and hemicellulose.81,82,85 Surface free energy has been found to NaClO2 and alkaline treatment.33,88 For example, in one study,
play a key role to control the interfacial properties such as the alkaline-treated CNF nanopapers had a tensile strength of
absorption and wettability.81,82,85 The total surface free energy 12.8 MPa and a modulus of 1.7 GPa, and the tensile strength and
has been dominated by the disperse component in the modulus increased to 68.2 MPa and 7.0 GPa, respectively, when
nanopapers which depends on the percentage and presence of the fibers were treated with an alkaline and bleaching solution.88
free hydroxyl groups on the surface. With increasing a higher Large improvements in the mechanical properties have also
lignin content, a value of the disperse component became lower been reported for nanopapers made from kenaf CNFs treated
G DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

with an alkaline solution and by bleaching, compared with those Notes


formed from alkaline-treated CNFs. The similar mechanical The authors declare no competing financial interest.
behaviors of the A and S nanopapers may be attributed to the
low lignin content of water hyacinth (4.1%), compared with
27.5% for parica88 and 18.8% for kenaf.33 Therefore, the lignin
■ ACKNOWLEDGMENTS
The authors are grateful for the financial support provided by
content of cellulosic raw materials could play an important role King Mongkut’s University of Technology Thonburi in the form
in CNF fibrillation and may also affect the mechanical properties of the KMUTT 55th Anniversary Commemorative Fund and
of nanopapers. Skill Development Grant (to Supachok Tanpichai) and by the
In general, materials tend to expand when the temperature Ministry of Education, Culture, Sports, Science, and Technol-
increases. Therefore, the thermal stability of materials at high ogy, Japan, for the Monbukagakusho Scholarship (No. 143492;
temperatures is important, especially with regard to composite to Subir Kumar Biswas).
and electronic applications. The coefficient of thermal expansion
(CTE) of each nanopaper material is reported in Table S2, and
the thermal expansion behavior of each nanopaper is presented
■ REFERENCES
(1) Malik, A. Environmental challenge vis a vis opportunity: The case
in Figure 6b. The CTE values of the nanopapers made from the of water hyacinth. Environ. Int. 2007, 33 (1), 122−138.
U, A, and S nanofibers were 15.6, 13.0, and 12.7 × 10−6 K −1, (2) Sullivan, P.; Wood, R. In Water hyacinth (Eichhornia crassipes
respectively. A reduced CTE was related to the crystallinity of (Mart.) Solms) seed longevity and the implications for management, 18th
chemically treated CNFs and the lignin content of the fibers, Australasian Weeds Conference, Melbourne, Australia; Weed Science
because lignin interferes with the formation of hydrogen bonds Society of Victoria Inc., Frankston, Australia, Melbourne, Australia,
between fibrils.33 All prepared CNF nanopapers exhibit 2012.
(3) Pintor-Ibarra, L. F.; Rivera-Prado, J. J.; Ngangyo-Heya, M.;
significantly limited expansion at higher temperature compared Rutiaga-Quiñones, J. G. Evaluation of the chemical components of
with engineering plastics such as polyethylene terephthalate, Eichhornia crassipes as an alternative raw material for pulp and paper.
polyether sulfones, polycarbonates, and nanopapers of CNFs BioResources 2018, 13 (2), 2800−2813.
extracted from kenaf and chitin.27,33,89 (4) Vanijajiva, O. Genetic characterization of invasive alien weed
Water hyacinth has a critical advantage over other Eichhornia crassipes in Thailand. Thammasat J. Sci. Technol. 2015, 22
lignocellulosic sources, which require tedious and time- (3), 486−496.
consuming treatment procedures for the preparation of CNFs (5) Anonymous TAPPI test method: The technical association of the pulp
with characteristics similar to those extracted from wood or and paper industry. TAPPI Press: Atlanta, GA, 2002-2003.
other lignocellulosic sources. The preparation of CNFs from (6) Chang, F.; Lee, S.-H.; Toba, K.; Nagatani, A.; Endo, T. Bamboo
lignin-poor water hyacinth would accelerate the production of nanofiber preparation by HCW and grinding treatment and its
application for nanocomposite. Wood Sci. Technol. 2012, 46 (1),
CNFs by the omission of some of the chemical treatment steps 393−403.
used for hemicellulose and lignin elimination. This would (7) Guna, V.; Ilangovan, M.; Anantha Prasad, M. G.; Reddy, N. Water
decrease the cost of CNF preparation because it would reduce hyacinth: A unique source for sustainable materials and products. ACS
the need for expensive chemicals and shorten the processing Sustainable Chem. Eng. 2017, 5 (6), 4478−4490.
time, while retaining the performance of the CNFs (i.e., close to (8) Kaur, M.; Kumar, M.; Sachdeva, S.; Puri, S. K. Aquatic weeds as the
that of fully chemically treated CNFs). Moreover, there were no next generation feedstock for sustainable bioenergy production.
significant differences in the mechanical properties and thermal Bioresour. Technol. 2018, 251, 390−402.
expansion behavior of nanopapers formed from CNFs prepared (9) Department of Public Works and Town & Country Planning
by the two different chemical approaches. Therefore, these Home Page. https://www.dpt.go.th (accessed July 17, 2019).
findings could facilitate the use of CNFs extracted from the (10) Sundari, M.; Ramesh, A. Isolation and characterization of
cellulose nanofibers from the aquatic weed water hyacinth Eichhornia
water hyacinth in a wide range of applications such as
crassipes. Carbohydr. Polym. 2012, 87, 1701−1705.
composites, electronics, packaging, and membrane applications (11) Abral, H.; Kadriadi, D.; Rodianus, A.; Mastariyanto, P.; Ilhamd,
in the near future.


S; Sapuan, S. M.; Ishak, M. R. Mechanical properties of water hyacinth
fibers - polyester composites before and after immersion in water.
ASSOCIATED CONTENT Mater. Eng. 2014, 58, 125−129.
*
S Supporting Information (12) Singh, A.; Bishnoi, N. R. Comparative study of various
The Supporting Information is available free of charge on the pretreatment techniques for ethanol production from water hyacinth.
ACS Publications website at DOI: 10.1021/acssusche- Ind. Crops Prod. 2013, 44, 283−289.
meng.9b04095. (13) Asrofi, M.; Abral, H.; Kasim, A.; Pratoto, A.; Mahardika, M.;
Park, J.-W.; Kim, H.-J. Isolation of nanocellulose from water hyacinth
Comparison of the chemical components of water fiber (WHF) produced via digester-sonication and its characterization.
hyacinth with those of aquatic weeds and agricultural Fibers Polym. 2018, 19 (8), 1618−1625.
crops, summary of the mechanical properties and (14) Sudiarto, S. I. A.; Renggaman, A.; Choi, H. L. Floating aquatic
coefficients of thermal expansion of the cellulose plants for total nitrogen and phosphorus removal from treated swine
nanofiber (CNF) nanopapers (PDF) wastewater and their biomass characteristics. J. Environ. Manage. 2019,


231, 763−769.
(15) Abe, K.; Yano, H. Comparison of the characteristics of cellulose
AUTHOR INFORMATION microfibril aggregates of wood, rice straw and potato tuber. Cellulose
Corresponding Authors 2009, 16 (6), 1017−1023.
*E-mail: supachok.tan@kmutt.ac.th. (16) Wang, B.; Li, D. Strong and optically transparent biocomposites
*E-mail: yano@rish.kyoto-u.ac.jp. reinforced with cellulose nanofibers isolated from peanut shell.
Composites, Part A 2015, 79, 1−7.
ORCID (17) Phinichka, N.; Kaenthong, S. Regenerated cellulose from high
Supachok Tanpichai: 0000-0002-6316-3080 alpha cellulose pulp of steam-exploded sugarcane bagasse. J. Mater. Res.
Subir Kumar Biswas: 0000-0002-9538-4726 Technol. 2018, 7 (1), 55−65.

H DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

(18) Ma, F.; Yang, N.; Xu, C.; Yu, H.; Wu, J.; Zhang, X. Combination (36) Nie, S.; Zhang, C.; Zhang, Q.; Zhang, K.; Zhang, Y.; Tao, P.;
of biological pretreatment with mild acid pretreatment for enzymatic Wang, S. Enzymatic and cold alkaline pretreatments of sugarcane
hydrolysis and ethanol production from water hyacinth. Bioresour. bagasse pulp to produce cellulose nanofibrils using a mechanical
Technol. 2010, 101 (24), 9600−9604. method. Ind. Crops Prod. 2018, 124, 435−441.
(19) Koyama, M.; Yamamoto, S.; Ishikawa, K.; Ban, S.; Toda, T. (37) Tao, P.; Zhang, Y.; Wu, Z.; Liao, X.; Nie, S. Enzymatic
Anaerobic digestion of submerged macrophytes: Chemical composi- pretreatment for cellulose nanofibrils isolation from bagasse pulp:
tion and anaerobic digestibility. Ecol. Eng. 2014, 69, 304−309. Transition of cellulose crystal structure. Carbohydr. Polym. 2019, 214,
(20) Martone, P. T.; Estevez, J. M.; Lu, F. C.; Ruel, K.; Denny, M. W.; 1−7.
Somerville, C.; Ralph, J. Discovery of lignin in seaweed reveals (38) Devi, R. R.; Dhar, P.; Kalamdhad, A.; Katiyar, V. Fabrication of
convergent evolution of cell-wall architecture. Curr. Biol. 2009, 19 (2), cellulose nanocrystals from agricultural compost. Compost Sci. Util.
169−175. 2015, 23 (2), 104−116.
(21) Kenrick, P.; Crane, P. R. The origin and early evolution of plants (39) Asrofi, M.; Abral, H.; Kasim, A.; Pratoto, A. XRD and FTIR
on land. Nature 1997, 389, 33−39. studies of nanocrystalline cellulose from water hyacinth (Eichornia
(22) Du, Z.-Y.; Wang, Q.-F.; Consortium, C. P. Phylogenetic tree of crassipes) fiber. J. Metastable. Nanocryst. Mater. 2017, 29, 9−16.
vascular plants reveals the origins of aquatic angiosperms. J. Syst. Evol. (40) Asrofi, M.; Abral, H.; Kasim, A.; Pratoto, A.; Mahardika, M.;
2016, 54 (4), 342−348. Hafizulhaq, F. Mechanical properties of a water hyacinth nanofiber
(23) Tanpichai, S.; Oksman, K. Crosslinked poly(vinyl alcohol) cellulose reinforced thermoplastic starch bionanocomposite: Effect of
composite films with cellulose nanocrystals: Mechanical and thermal ultrasonic vibration during processing. Fibers 2018, 6 (2), 40.
properties. J. Appl. Polym. Sci. 2018, 135 (3), 45710. (41) Asrofi, M.; Abral, H.; Kasim, A.; Pratoto, A.; Mahardika, M.;
(24) Li, K.; Skolrood, L. N.; Aytug, T.; Tekinalp, H.; Ozcan, S. Strong Hafizulhaq, F. Characterization of the sonicated yam bean starch
and tough cellulose nanofibrils composite films: Mechanism of bionanocomposites reinforced by nanocellulose water hyacinth fiber
synergetic effect of hydrogen bonds and ionic interactions. ACS (WHF): The effect of various fiber loading. J. Eng. Sci. Technol. 2018, 13
Sustainable Chem. Eng. 2019, 7 (17), 14341−14346. (9), 2700−2715.
(25) Gustafsson, O.; Manukyan, L.; Gustafsson, S.; Tummala, G. K.; (42) Yuwawech, K.; Wootthikanokkhan, J.; Tanpichai, S. Trans-
Zaman, S.; Begum, A.; Alfasane, M. A.; Siddique-E-Rabbani, K.; parency, moisture barrier property, and performance of the alternative
Mihranyan, A. Scalable and sustainable total pathogen removal filter solar cell encapsulants based on PU/PVDC blend reinforced with
paper for point-of-use drinking water purification in Bangladesh. ACS different types of cellulose nanocrystals. Mater. Renew. Sustain. Energy.
Sustainable Chem. Eng. 2019, 7 (17), 14373−14383. 2018, 7 (4), 21.
(26) Tanpichai, S.; Witayakran, S.; Srimarut, Y.; Woraprayote, W.; (43) Segal, L.; Creely, J. J.; Martin, A. E.; Conrad, C. M. An empirical
Malila, Y. Porosity, density and mechanical properties of the paper of method for estimating the degree of crystallinity of native cellulose
steam exploded bamboo microfibers controlled by nanofibrillated using the X-ray diffractometer. Text. Res. J. 1959, 29 (10), 786−794.
cellulose. J. Mater. Res. Technol. 2019, 8 (4), 3612−3622. (44) French, A. D. Idealized powder diffraction patterns for cellulose
(27) Biswas, S. K.; Tanpichai, S.; Witayakran, S.; Yang, X.; Shams, M. polymorphs. Cellulose 2014, 21 (2), 885−896.
I.; Yano, H. Thermally superstable cellulosic-nanorod-reinforced (45) Emmclan, L. S. H.; Zakaria, M. H.; Bujang, J. S. Utilization of
transparent substrates featuring microscale surface patterns. ACS aquatic weeds fibers for handmade papermaking. Bioresources 2018, 13
Nano 2019, 13 (2), 2015−2023. (3), 5684−5701.
(28) Abe, K.; Iwamoto, S.; Yano, H. Obtaining cellulose nanofibers (46) Ilangovan, M.; Guna, V.; Hu, C.; Nagananda, G. S.; Reddy, N.
with a uniform width of 15 nm from wood. Biomacromolecules 2007, 8, Curcuma longa L. plant residue as a source for natural cellulose fibers
3276−3278. with antimicrobial activity. Ind. Crops Prod. 2018, 112, 556−560.
(29) Tanpichai, S.; Quero, F.; Nogi, M.; Yano, H.; Young, R. J.; (47) Jung, S.-J.; Kim, S.-H.; Chung, I.-M. Comparison of lignin,
Lindstrom, T.; Sampson, W. W.; Eichhorn, S. J. Effective young’s cellulose, and hemicellulose contents for biofuels utilization among 4
modulus of bacterial and microfibrillated cellulose fibrils in fibrous types of lignocellulosic crops. Biomass Bioenergy 2015, 83, 322−327.
networks. Biomacromolecules 2012, 13 (5), 1340−1349. (48) Reddy, N.; Yang, Y. Characterizing natural cellulose fibers from
(30) Kang, X.; Sun, P.; Kuga, S.; Wang, C.; Zhao, Y.; Wu, M.; Huang, velvet leaf (Abutilon theophrasti) stems. Bioresour. Technol. 2008, 99
Y. Thin cellulose nanofiber from corncob cellulose and its performance (7), 2449−2454.
in transparent nanopaper. ACS Sustainable Chem. Eng. 2017, 5 (3), (49) Fareez, I. M.; Ibrahim, N. A.; Yaacob, W.; Razali, N. A. M.; Jasni,
2529−2534. A. H.; Aziz, F. A. Characteristics of cellulose extracted from Josapine
(31) Liu, Q.; Lu, Y.; Aguedo, M.; Jacquet, N.; Ouyang, C.; He, W.; pineapple leaf fibre after alkali treatment followed by extensive
Yan, C.; Bai, W.; Guo, R.; Goffin, D.; Song, J.; Richel, A. Isolation of bleaching. Cellulose 2018, 25 (8), 4407−4421.
high-purity cellulose nanofibers from wheat straw through the (50) Ovalle-Serrano, S. A.; Blanco-Tirado, C.; Combariza, M. Y.
combined environmentally friendly methods of steam explosion, Exploring the composition of raw and delignified Colombian fique
microwave-assisted hydrolysis, and microfluidization. ACS Sustainable fibers, tow and pulp. Cellulose 2018, 25 (1), 151−165.
Chem. Eng. 2017, 5 (7), 6183−6191. (51) Ciftci, D.; Flores, R. A.; Saldana, M. D. A. Cellulose fiber isolation
(32) Santos, R. M. d.; Flauzino Neto, W. P.; Silvério, H. A.; Martins, D. and characterization from sweet blue lupin hull and canola straw. J.
F.; Dantas, N. O.; Pasquini, D. Cellulose nanocrystals from pineapple Polym. Environ. 2018, 26 (7), 2773−2781.
leaf, a new approach for the reuse of this agro-waste. Ind. Crops Prod. (52) Kantouch, A.; Hebeish, A.; Elrafie, M. H. Action of sodium
2013, 50, 707−714. chlorite on cellulose and cellulose derivatives. Text. Res. J. 1970, 40 (2),
(33) Nobuta, K.; Teramura, H.; Ito, H.; Hongo, C.; Kawaguchi, H.; 178−184.
Ogino, C.; Kondo, A.; Nishino, T. Characterization of cellulose (53) Kumar, R.; Mago, G.; Balan, V.; Wyman, C. E. Physical and
nanofiber sheets from different refining processes. Cellulose 2016, 23 chemical characterizations of corn stover and poplar solids resulting
(1), 403−414. from leading pretreatment technologies. Bioresour. Technol. 2009, 100
(34) Tanpichai, S.; Witayakran, S.; Boonmahitthisud, A. Study on (17), 3948−3962.
structural and thermal properties of cellulose microfibers isolated from (54) Hubbell, C. A.; Ragauskas, A. J. Effect of acid-chlorite
pineapple leaves using steam explosion. J. Environ. Chem. Eng. 2019, 7 delignification on cellulose degree of polymerization. Bioresour. Technol.
(1), 102836. 2010, 101 (19), 7410−7415.
(35) Henriksson, M.; Henriksson, G.; Berglund, L. A.; Lindstrom, T. (55) Li, M.; Wang, J.; Yang, Y.; Xie, G. Alkali-based pretreatments
An environmentally friendly method for enzyme-assisted preparation of distinctively extract lignin and pectin for enhancing biomass
microfibrillated cellulose (MFC) nanofibers. Eur. Polym. J. 2007, 43 saccharification by altering cellulose features in sugar-rich Jerusalem
(8), 3434−3441. artichoke stem. Bioresour. Technol. 2016, 208, 31−41.

I DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX
ACS Sustainable Chemistry & Engineering Research Article

(56) Madsen, B.; Aslan, M.; Lilholt, H. Fractographic observations of (77) Shen, K. D.; Gu, S.; Luo, H. K.; Wang, R. S.; Fang, X. M. The
the microstructural characteristics of flax fibre composites. Compos. Sci. pyrolytic degradation of wood-derived lignin from pulping process.
Technol. 2016, 123, 151−162. Bioresour. Technol. 2010, 101, 6136−6146.
(57) Bakri, B.; Eichhorn, S. J. Elastic coils: Deformation micro- (78) Tao, P.; Wu, Z.; Xing, C.; Zhang, Q.; Wei, Z.; Nie, S. Effect of
mechanics of coir and celery fibres. Cellulose 2010, 17 (1), 1−11. enzymatic treatment on the thermal stability of cellulose nanofibrils.
(58) Zhang, W.; Fei, B.; Polle, A.; Euring, D.; Tian, G.; Yue, X.; Chang, Cellulose 2019, 26 (13), 7717−7725.
Y.; Jiang, Z.; Hu, T. Crystal and thermal response of cellulose isolation (79) Zhang, K.; Zhang, Y.; Yan, D.; Zhang, C.; Nie, S. Enzyme-assisted
from Bamboo by two different chemical treatments. Bioresources 2019, mechanical production of cellulose nanofibrils: thermal stability.
16 (2), 3471−3480. Cellulose 2018, 25 (9), 5049−5061.
(59) Khawas, P.; Deka, S. C. Isolation and characterization of cellulose (80) Deepa, B.; Abraham, E.; Cherian, B. M.; Bismarck, A.; Blaker, J.
nanofibers from culinary banana peel using high-intensity ultra- J.; Pothan, L. A.; Leao, A. L.; de Souza, S. F.; Kottaisamy, M. Structure,
sonication combined with chemical treatment. Carbohydr. Polym. morphology and thermal characteristics of banana nano fibers obtained
2016, 137, 608−616. by steam explosion. Bioresour. Technol. 2011, 102 (2), 1988−1997.
(60) Fahma, F.; Iwamoto, S.; Hori, N.; Iwata, T.; Takemura, A. (81) Jiang, Y.; Liu, X.; Yang, Q.; Song, X.; Qin, C.; Wang, S.; Li, K.
Isolation, preparation, and characterization of nanofibers from oil palm Effects of residual lignin on composition, structure and properties of
empty-fruit-bunch (OPEFB). Cellulose 2010, 17 (5), 977−985. mechanically defibrillated cellulose fibrils and films. Cellulose 2019, 26
(61) Chen, W. S.; Yu, H. P.; Liu, Y. X.; Hai, Y. F.; Zhang, M. X.; Chen, (3), 1−17.
P. Isolation and characterization of cellulose nanofibers from four plant (82) Bian, H.; Gao, Y.; Wang, R.; Liu, Z.; Wu, W.; Dai, H.
cellulose fibers using a chemical-ultrasonic process. Cellulose 2011, 18 Contribution of lignin to the surface structure and physical perform-
ance of cellulose nanofibrils film. Cellulose 2018, 25 (2), 1309−1318.
(2), 433−442.
(83) Iwamoto, S.; Abe, K.; Yano, H. The effect of hemicelluloses on
(62) Yue, Y.; Han, J.; Han, G.; Aita, G. M.; Wu, Q. Cellulose fibers
wood pulp nanofibrillation and nanofiber network characteristics.
isolated from energycane bagasse using alkaline and sodium chlorite
Biomacromolecules 2008, 9 (3), 1022−1026.
treatments: Structural, chemical and thermal properties. Ind. Crops
(84) Hult, E. L.; Larsson, P. T.; Iversen, T. Cellulose fibril aggregation
Prod. 2015, 76, 355−363.  an inherent property of kraft pulps. Polymer 2001, 42 (8), 3309−
(63) Pelissari, F. M.; Sobral, P. J. d. A.; Menegalli, F. C. Isolation and 3314.
characterization of cellulose nanofibers from banana peels. Cellulose (85) Rojo, E.; Peresin, M. S.; Sampson, W. W.; Hoeger, I. C.;
2014, 21 (1), 417−432. Vartiainen, J.; Laine, J.; Rojas, O. J. Comprehensive elucidation of the
(64) Tibolla, H.; Pelissari, F. M.; Martins, J. T.; Vicente, A. A.; effect of residual lignin on the physical, barrier, mechanical and surface
Menegalli, F. C. Cellulose nanofibers produced from banana peel by properties of nanocellulose films. Green Chem. 2015, 17 (3), 1853−
chemical and mechanical treatments: Characterization and cytotoxicity 1866.
assessment. Food Hydrocolloids 2018, 75, 192−201. (86) Nogi, M.; Iwamoto, S.; Nakagaito, A. N.; Yano, H. Optically
(65) Tanpichai, S.; Witayakran, S. All-cellulose composites from transparent nanofiber paper. Adv. Mater. 2009, 21 (16), 1595−1598.
pineapple leaf microfibers: Structural, thermal, and mechanical (87) Tanpichai, S.; Sampson, W. W.; Eichhorn, S. J. Stress-transfer in
properties. Polym. Compos. 2018, 39 (3), 895−903. microfibrillated cellulose reinforced poly(lactic acid) composites using
(66) Yue, Y.; Han, J.; Han, G.; Zhang, Q.; French, A. D.; Wu, Q. Raman spectroscopy. Composites, Part A 2012, 43 (7), 1145−1152.
Characterization of cellulose I/II hybrid fibers isolated from energycane (88) Scatolino, M. V.; Fonseca, C. S.; da Silva Gomes, M.; Rompa, V.
bagasse during the delignification process: Morphology, crystallinity D.; Martins, M. A.; Tonoli, G. H. D.; Mendes, L. M. How the surface
and percentage estimation. Carbohydr. Polym. 2015, 133, 438−447. wettability and modulus of elasticity of the Amazonian paricá
(67) Bali, G.; Meng, X. Z.; Deneff, J. I.; Sun, Q. N.; Ragauskas, A. J. nanofibrils films are affected by the chemical changes of the natural
The effect of alkaline pretreatment methods on cellulose structure and fibers. Euro. J. Wood. Wood. Prod. 2018, 76 (6), 1581−1594.
accessibility. ChemSusChem 2015, 8 (2), 275−279. (89) Biswas, S. K.; Sano, H.; Shams, M. I.; Yano, H. Three-
(68) Liu, Y.; Hu, H. X-ray diffraction study of bamboo fibers treated dimensional-moldable nanofiber-reinforced transparent composites
with NaOH. Fibers Polym. 2008, 9 (6), 735−739. with a hierarchically self-assembled “reverse” nacre-like architecture.
(69) Yang, H.; Yan, R.; Chen, H.; Lee, D. H.; Zheng, C. ACS Appl. Mater. Interfaces 2017, 9 (35), 30177−30184.
Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel
2007, 86 (12), 1781−1788.
(70) Naduparambath, S.; Purushothaman, E. Sago seed shell:
determination of the composition and isolation of microcrystalline
cellulose (MCC). Cellulose 2016, 23 (3), 1803−1812.
(71) Nair, S. S.; Yan, N. Effect of high residual lignin on the thermal
stability of nanofibrils and its enhanced mechanical performance in
aqueous environments. Cellulose 2015, 22 (5), 3137−3150.
(72) Jiao, Y.; Wan, C.; Qiang, T.; Li, J. Synthesis of superhydrophobic
ultralight aerogels from nanofibrillated cellulose isolated from natural
reed for high-performance adsorbents. Appl. Phys. A: Mater. Sci. Process.
2016, 122 (7), 686.
(73) Brebu, M.; Vasile, C. Thermal degradation of lignin - A review.
Cellulose Chem. Technol. 2010, 44 (9), 353−363.
(74) Abraham, E.; Deepa, B.; Pothen, L. A.; Cintil, J.; Thomas, S.;
John, M. J.; Anandjiwala, R.; Narine, S. S. Environmental friendly
method for the extraction of coir fibre and isolation of nanofibre.
Carbohydr. Polym. 2013, 92 (2), 1477−1483.
(75) Zhang, N.; Tao, P.; Lu, Y.; Nie, S. Effect of lignin on the thermal
stability of cellulose nanofibrils produced from bagasse pulp. Cellulose
2019, 26 (13), 7823−7835.
(76) Fenner, A. R.; Lephardt, O. J. Examination of the thermal
decomposition of Kraft pine lignin by Fourier Transform Infrared
evolved gas analysis. J. Agric. Food Chem. 1981, 29, 846−849.

J DOI: 10.1021/acssuschemeng.9b04095
ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

You might also like