You are on page 1of 9

International Journal of Biological Macromolecules 158 (2020) 1259–1267

Contents lists available at ScienceDirect

International Journal of Biological Macromolecules

journal homepage: http://www.elsevier.com/locate/ijbiomac

Preparation of lignin containing cellulose nanofibers and its application


in PVA nanocomposite films
Mingyan Yang a,b,⁎, Xiao Zhang a, Shuyi Guan a, Yan Dou a,b, Xiaofeng Gao a
a
School of Water and Environment, Chang'an University, Xi'an 710054, China
b
Shaanxi Key Laboratory of Exploration and Comprehensive Utilization of Mineral Resources, Xi'an 710054, China

a r t i c l e i n f o a b s t r a c t

Article history: Lignin containing cellulose nanofibers (LCNFs) were successfully prepared from wheat straw using an acid
Received 19 February 2020 hydrotrope of p-toluene sulfonic acid (p-TsOH) combined with ultrasonication. p-TsOH pretreatment was ap-
Received in revised form 19 April 2020 plied below 80 °C to selectively remove hemicellulose and lignin and generate purified cellulose fibers containing
Accepted 6 May 2020
approximately 15% lignin. Subsequently, high-intensity ultrasonication was used for b6 min to effectively defi-
Available online 11 May 2020
brillate the p-TsOH-pretreated cellulose fibers to nanoscale fibers. AFM and TEM analyses showed that the diam-
Keywords:
eter distribution of the resultant nanofibers decreased with the increase in ultrasonic intensity. The FTIR and XRD
p-Toluene sulfonic acid results indicated that the molecular structures and cellulose crystallinity were not changed during the ultrasonic
Ultrasonication process. An amount of 5 wt% of the obtained LCNFs was introduced into a polyvinyl alcohol (PVA) matrix. The
Lignin containing cellulose nanofibers (LCNFs) resulting nanocomposite products exhibited improved thermal performance and surface properties compared
with the pure PVA matrix. The mechanical properties, including the tensile stress and Young's modulus, were en-
hanced significantly, although the elongation at the break was slightly decreased. PVA composites with the addi-
tion of LCNFs are expected to be used in a variety of fields, such as biodegradable plastics, pharmaceutical carrier,
filtration media and packaging materials.
© 2020 Elsevier B.V. All rights reserved.

1. Introduction reinforcement and oxygen-barrier layers in composites, packaging ma-


terials, cosmetics, pharmaceuticals and biomedical devices [6–9].
The valorization of naturally abundant lignocellulosic resources, in- CNFs are conventionally produced using mechanical delamination
cluding agricultural residues, woods, energy crops and other industrial from microcrystalline cellulose, bleached chemical (lignin free) pulps
resources, can help solve environmental problems caused by the exces- or cotton litter. However, the process increases the product cost and
sive use of fossil resources [1]. Cellulose is the most important and abun- limits its commercial utilization [1,10,11]. In addition, the CNFs from
dant natural biopolymer resources, and it can be deconstructed, bleached pulps or with lower lignin content below 5% are highly hydro-
resulting in different types of nanocellulose, such as cellulose philic, leading to poor interaction and compatibility between CNFs and
nanocrystals (CNCs), cellulose nanowhiskers (CNWs), microfibrillated non-polar media, which are challenges that need to be addressed in
cellulose, microfibrile aggregates and cellulose nanofibers (CNFs) [2,3]. some applications [12]. The direct conversion of cellulose into nanofi-
Among these nanocelluloses, CNFs, which contain bundles of long, flex- bers from low-cost lignocellulosic biomass can generate a new kind of
ible and entangled fibers with a diameter in the range of 4–20 nm and a nanocellulose material: lignin containing cellulose nanofibrils (LCNFs)
length in micrometers, have been drawing great attention in recent [13]. The amount of lignin in the LCNFs depends on the degree of
years due to their unique structures and properties. In addition to biode- delignification of the biomass. Having the advantages of a higher yield,
gradability and biocompatibility, CNFs present markedly improved lower cost and less environmental impact, LCNFs have been demon-
properties, such as crystallinity, mechanical resistance, flexibility and strated to be a promising alternative to high-purity nanocellulose
barrier properties, due to their large specific surface area and surface ac- [14,15]. The residue lignin in the LCNFs is beneficial to improving the
tivity [4,5]. As a result, CNFs have been considered as a promising basic properties of LCNFs, including the reduction of the hydrophilicity and
building block for various applications, such as for additive polarity and the enhancement of thermal stability [16,17]. In addition,
lignin-based nanomaterial can be used as a delivery system or excipient
formulation to enhance drug delivery or improve drug properties
⁎ Corresponding author at: School of Water and Environment, Chang'an University,
[18–20]. This application opens the possibility of using lignin-
Xi'an 710054, China. containing nanomaterials for high-value applications in the pharmaceu-
E-mail address: yangmingyan67@163.com (M. Yang). tical industry. The major obstruction in separating cellulose from the

https://doi.org/10.1016/j.ijbiomac.2020.05.044
0141-8130/© 2020 Elsevier B.V. All rights reserved.
1260 M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267

complex polymeric matrix containing hemicellulose and lignin is the re- generator (GH-IID/GH-1000Y, China) at a concentration of 1 wt%.
calcitrance of the hierarchal structure of the plant wall [21]. A mechan- The ultrasonic process was performed in an ice bath to avoid excess
ical process combined with chemical pretreatment is believed to be a heat. Three samples were prepared by sonicating for 6 min at an out-
promising technology for resolving the recalcitrance problem of the put power of 800, 1200 or 1600 W, and they were labeled as LCNFs-
raw material. However, the existing chemical pretreatment process in- 800, LCNFs-1200 and LCNFs-1600, respectively. The suspensions of
volving strong concentrated acids, alkaline, (TEMPO)-mediated oxida- 0.3 wt% of the raw material, purified WIS, and cellulose nanofibers
tion and reducing agents is still limited by several factors, including after ultrasonication were allowed to rest for 0 and 30 min. The sta-
low solution recovery, negative environmental impacts and high inef- bility of the suspensions was studied through visual examination ac-
fective cost [11]. cording to other previous reports [27,28].
Recently, a recyclable acid hydrotrope was shown to be a promising
technology for biomass fractionation. Aqueous p-toluenesulfonic acid is 2.3. Chemical composition
capable of achieving high delignification efficiency in an aqueous solu-
tion at low temperature, resulting in cellulose-rich water insoluble The raw material and purified WIS were analyzed to determine their
solid (WIS) and spent liquor (SL) mainly composed of xylan and dis- chemical composition according to the procedures reported by the Na-
solved lignin [22,23]. The resultant WIS can be used for ethanol fermen- tional Renewable Energy Laboratory (NREL) as previously described
tation [24] or other high value-added production processes, including [29].
nanocellulosic materials based on a mechanical process, such as super
mass colloider (SMC) [10] and microfluidizer [23]. In addition, p-TsOH 2.4. Characterization of LCNFs
can be reused or recovered by crystallization technology [22–24].
Wheat is one of the most common crops in China at N120 million tons 2.4.1. FE-SEM observation
of production annually. Wheat residues account for 15.7% of all the post- The LCNFs samples were freeze dried using a freeze drier (Eyela
harvest crop residues, and they are traditionally buried in soil or burnt FDU-2110, Japan), coated with gold and analyzed via FE-SEM (Hitachi
in the field, which not only leads to the waste of biomass sources but su8020, Japan).
also negatively impacts the environment [25]. As a renewable, inexpen-
sive and underutilized resource, agricultural residues used for the pro-
2.4.2. AFM analysis
duction of high value-added products, such as nanomaterial, can
A mica substrate was prepared by depositing a drop of the LCNFs
introduce a new alternative resource and reduce the use of forest re-
suspension at approximately 0.01 wt% and drying for 24 h at room tem-
sources worldwide [26].
perature. The LCNFs image was obtained using AFM (Bruker, dimension
The p-TsOH pretreatment combined with ultrasonication to prepare
Icon) according to a previous study [20]. Gwyddion software (Depart-
LCNFs has not been reported previously. In addition, the role of LCNFs in
ment of Nanometrology, Czech Metrology Institute, Czech Republic,
improving the overall performance of biobased nanocomposite de-
64-bit) was used to analyze the AFM images.
pends on various factors, such as the biomass source, fractionation pro-
cess and extraction methods. In this study, p-TsOH pretreatment was
used in combination with ultrasonication to prepare LCNFs from 2.4.3. TEM
wheat straw. Field Emission Scanning Electron Microscope (FE-SEM), Samples were prepared by depositing a drop of the LCNFs suspen-
atomic force microscopy (AFM) and transmission electron microscopy sion at approximately 0.01 wt% on Cu grids and drying at room temper-
(TEM) were used to analyze the morphological and structural charac- ature. Then, the samples were staining with 2% phosphotungstic acid for
teristics of the resultant LCNFs. The changes of the functional groups 2–3 min and observed via TEM (Tecnai G2 F20, Japan). The diameter
and crystallization degree during the processing procedures were stud- and the size distribution of the LCNFs were analyzed using Nano Mea-
ied by Fourier transform infrared (FTIR) spectroscopy and X-ray diffrac- surer version 1.2.5 software (Department of Chemistry, Fudan Univer-
tion (XRD). The LCNFs were applied to PVA to prepare a bio- sity, China).
nanocomposite film. The thermal stability, mechanical performance
and surface hydrophobicity of the resulting films were explored. This 2.4.4. FTIR analysis
work will provide a novel, feasible, and environmentally friendly pro- The FTIR spectra of the raw wheat straw, purified WIS and freeze-
cessing technology to produce lignin containing cellulose nanomaterials dried LCNFs were performed using a FTIR instrument (Magna 560, Nico-
for a variety of applications. let, Thermo Electron Corp., USA). Spectra were recorded in the range of
400–4000 cm−1 with a resolution of 4 cm−1.
2. Material and methods
2.4.5. XRD analysis
2.1. Material The X-ray diffraction pattern spectra of the raw wheat straw, puri-
fied WIS and freeze-dried LCNFs samples were measured using an X-
Wheat straw collected from Chang'an County (Shaanxi Province, ray diffractometer (XRD, D8 Advance, Bruker) at a scan speed of 4°/
China) was cut into small pieces and milled into powder. The powder min. The crystallinity index (CrI) was calculated using the Segal method
sieved through a 20 mesh was collected for further use. (Eq. (1)).

2.2. Preparation of lignin containing cellulose nanofibrils (LCNFs)  


Iam
CrI ð%Þ ¼ 1−  100 ð1Þ
The production process of LCNFs consists of chemical pretreat- I 200
ment followed by an ultrasonic process. First, 10 g of the dry wheat
straw powder was mixed with 70 wt% p-TsOH solution at 80 °C for
10 min, and it had a liquor to mass ratio of 10:1. The reaction was Table 1
stopped by adding 133.3 mL DI water. The hydrolysate solution Chemical composition changes after p-TsOH fractionation.
was separated by filtration under vacuum. The obtained spent liquor
Solid yield (%) Glucan (%) Hemicellulose (%) Lignin (%)
(SL) was used for the recovery of acid and dissolved lignin. The ob-
tained water insoluble solids (WIS) were rinsed with DI water to Wheat straw 100 34.10 ± 0.28 24.96 ± 0.14 23.04 ± 0.21
WIS 47.55 ± 1.41 64.38 ± 0.18 12.87 ± 0.46 14.86 ± 0.44
neutral pH and subjected to ultrasonic processing in an ultrasonic
M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267 1261

Fig. 1. Dispersion states of 0.3 wt% suspensions of raw material (A), WIS (B), LCNFs-800 (C), LCNFs-1200 W (D), and LCNFs-1600 (E) at different times.

where I200 represents both crystalline and amorphous region material 2.5.1. Thermal stability analysis
and Iam represents amorphous material. The thermal stability of the LCNFs/PVA films obtained with different
ultrasonic output powers were tested in a thermogravimetric analyzer
(TGA Q50; TA Instruments-Waters LLC, USA) from 35 to 600 °C at a
2.5. Preparation of LCNFs/PVA nanocomposite films heating rate of 10 °C/min under nitrogen flow. The weight-loss rate, ini-
tial degradation temperature (Tonset) and maximum thermal degrada-
The solvent cast process was used to prepare nanocomposite films. tion temperature (Tmax) were obtained from the TGA and derivative
The obtained LCNFs were mixed with a 5 wt% PVA solution by stirring thermogravimetric (DTG) curves.
and then sonicated for 60 min to improve the dispersion of the
nanofillers in the matrix. Each mixture was degassed, poured onto a
polished glass plate and dried at 60 °C for approximately 6 h in an 2.5.2. Tensile properties
oven. A pure PVA film (without LCNFs) was prepared as a control The tensile properties, including the tensile strength (TS), tensile
using the above procedures. The resulting films were stored in a desic- modulus (TM), and elongation at break (EB%), of the studied films
cator before testing. were tested using a WDW-200 Universal Testing Machine (China) at a

Fig. 2. SEM images of the LCNFs obtained from wheat straw (A) LCNFs-800 (B) LCNFs-1200 and (C) LCNFs-1600.
1262 M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267

crosshead speed of 5 mm/min according to the standard of ASTM D1708 LCNFs-800 had a small portion of sediment, suggesting that
[30]. unfibrillated and partially fibrillated fibers still remained, while the fi-
bers in samples LCNFs-1200 and LCNFs-1600 were nearly dispersed in
2.5.3. Static water contact angles (WCAs) water, indicating that increasing the intensity of ultrasonication can
The WCAs of the nanocomposite film samples were tested using a promote the fibrillation process of the fibers.
contact angle analyzer (XG-CAMB, China). Images of the droplets were
recorded at different times after the drop touched the film surface. 3.2. FE-SEM of LCNFs

3. Result and discussions The FE-SEM morphologies of the LCNFs obtained from different in-
tensities of ultrasonic treatment are shown in Fig. 2 (A, B, C). Well-
3.1. Preparation and characterization of LCNFs delaminated nanocellulose fibers were observed in all three different ul-
trasonic treatments samples. Some dispersed globular shaped patches,
The wheat straw was fractionated into two parts by the p-TsOH pre- which are likely to be lignin nanoparticle (LNPs), were visible in all
treatment: cellulose-rich WIS and spent liquor (SL), which mainly three samples.
contained xylose and dissolved lignin. Table 1 shows the composition
changes after the p-TsOH pretreatment. The cellulose content in the 3.3. AFM analysis
WIS increased from 34% to 64%, with a recovery of approximately 90%.
Correspondingly, the hemicellulose and lignin content decreased to The nanoscale topography and the height distribution of the LCNFs
12.87% and 14.86%, respectively. The results indicated that the p-TsOH studied using AFM are plotted in Fig. 3. Long, flexible and filament-like
pretreatment can selectively remove lignin and hemicellulose from nanosized fibers and dispersed spherical nanoparticles were observed
the biomass under mild condition and produce more purified in all three samples. The average AFM height of the three different sam-
cellulose-rich WIS containing no N15% lignin content. Previous studies ples shown in Fig. 3 indicated that a higher intensity of ultrasonication
suggested that lignin content above 13% had a significant effect on im- can decrease the height of the fibers and thus promote the degree of
proving the properties of the resultant LCNFs [31,32]. The obtained nanofibrillation of the cellulose fibers. This result was consistent with
WIS were fibrillated via an ultrasonic treatment for 6 min with an out- the visual observations. In addition, one particular area in Fig. 3A1 re-
put power of 800, 1200 and 1600 W. The dispersion properties of the fi- corded the depolymerizing process from micron-sized cellulose
bers in water were substantially changed after the ultrasonic process. As branches to individual nanofibers with the impact of 800 W
shown in Fig. 1, the raw material and WIS fibers were precipitated ultrasonication. Small globular-shaped particles, which are likely lignin
completely at the bottom of the glass bottle, while all ultrasonic samples nanoparticles (LNPs), were dispersed around the crosslinked nanofi-
presented a good dispersion property in water. After 30 min of rest, bers. This fibril network and globular-shaped particles were similar

Fig. 3. Effects of the ultrasonic intensity on the morphologies and height distribution of the resulting LCNFs measured by AFM (First row-LCNFs 800; Middle row-LCNFs 1200; third row-
LCNFs 1600. Left two panels, scale bar = 10 μm; and right two panels, scale bar = 2 μm).
M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267 1263

Fig. 4. Effects of ultrasonic intensity on the morphologies and diameter distributions of the resulting LCNFs measured by TEM (left panel-LCNFs 800; middle panel-LCNFs 1200; right panel-
LCNFs 1600).

with the SEM result. Similar results for LCNFs derived from various 20.01, 13.32 and 10.12 nm, respectively. All these results indicated
sources were reported by several authors, such as Rojo et al. [15], Nair that increasing the output power of the ultrasonic treatment can posi-
and Yan [16], Herrera et al. [17] and Ewulonu et al. [33]. tively facilitate the degree of nanofibrillation and obtain more uniform
and slender nanofibers. In addition, lignin nanoparticles observed in
3.4. TEM the TEM image (Fig. 5) exhibited a spherical shape with the size range
of 20–50 nm, suggesting that the lignin nanoparticles were synthesized
The morphology and diameter distribution of the prepared LCNFs during the ultrasonic process as well [33].
were further studied by TEM. As shown in Fig. 4, the long nanofibers
are interconnected like a ‘cobweb’, and a similar result is observed in 3.5. FTIR analysis
the AFM images. Some strands agglomerated to each other to form a
large width. These agglomerations have been reported in the prepara- FTIR spectroscopy was performed to compare the changes of the
tion of nanofibers from different materials by previous studies chemical structure before and after each treatment. As depicted in
[33–35]. The TEM diameter distribution plots showed that N58% of the Fig. 6, the absorbance peaks at 3350, 2900, 1160, 1110, 1034 and
LCNFs-800 samples had a diameter of no N20 nm and exhibited a 897 cm−1 represent the typical features of cellulose appeared in all
wide range of diameter sizes. The diameter ranges of the nanofibers be- spectra [3,36]. The peaks at approximately 1737 cm−1 and 1252 cm−1
came more uniform and narrow as the ultrasonic intensity increased, in raw fiber represent ester and ether linkages formed between lignin
with 84.66% of the LCNFs-1200 and 93.33% of the LCNFs-1600 showing and hemicellulose, and they nearly disappeared in the WIS and LCNFs
a diameter range of no N20 nm. The average diameters of the resultant samples, indicating that most of the hemicellulose and lignin were re-
nanofibers of LCNFs-800, LCNFs-1200 and LCNFs-1600 samples were moved from the biomass after chemical pretreatment [24]. The removal

Fig. 5. TEM morphology of the lignin nanoparticles prepared by ultrasonic treatment.


1264 M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267

1.4
100 PVA
raw material LCNFs 800/PVA 1.2
80 LCNFs 1200/PVA
LCNFs 1600/PVA 1.0

DTG (mg/min)
Weight loss (%)
60
WIS 0.8
40
0.6
LNCFs 800 20
0.4
0
LCNFs 1200 0.2
-20
1503 1737 LCNFs 1600 0.0
897 2900 100 200 300 400 500
1252 3350
1160 1450 Temperature (℃)
1034 1110

1000 1500 2000 2500 3000 3500 4000 Fig. 8. Thermal stability of pure PVA and LCNFs/PVA nanocomposites.
-1
Wavenumber (cm )
both amorphous and crystalline cellulose. This result is in agreement
Fig. 6. FTIR spectra of the raw material, WIS and LCNFs with different ultrasonic
with previous reports [2,37].
treatments.
3.7. Thermal stability of the LCNFs/PVA nanocomposites

of lignin in the WIS and LCNFs samples can also be verified by eliminat- The thermal properties and degradation behavior are crucial in bio-
ing of the peaks at 1450 cm−1 and 1503 cm−1, which was attributed to nanocomposite processing. According to the TGA and DTG curves
the aromatic skeletal vibration of lignin. There was no difference be- shown in Fig. 8, three degradation stages were observed for all tested
tween the spectra of the WIS and LCNFs obtained from the different ul- samples. The first minor weight loss was observed below 100 °C due
trasonic treatment, suggesting that the ultrasonic treatment did not to evaporation of water or degradation of low molecular weight com-
destroy the molecular structure of cellulose. pounds. The second major weight loss occurred in the temperature
range from 200 to 400 °C because of the decomposition of the side
3.6. XRD analysis chains of PVA. The Tonset and Tmax of the second degradation process
(Table 2) for the LNCFs/PVA films were higher than those for pure
X-ray diffraction was carried out to compare the crystalline behavior PVA. With the addition of LCNFs, the Tonset and Tmax increased from
of the treated and untreated cellulose fibers. X-ray diffractograms are 201.1 °C/234.2 °C for pure PVA to 209.2 °C/246.8 °C for LCNFs-1600.
presented in Fig. 7. All samples exhibited similar diffractograms with The results indicated that the incorporation of LCNFs had a positive ef-
main characteristic peaks at 2θ angles of approximately 16° and 22° fect on the thermal stability of the nanocomposites. The improved ther-
and 34°, which represent typical peaks of cellulose I structure reflections mal stability of the LCNFs nanocomposites may have resulted from the
[36]. The results suggested that the inherent crystal structure of cellu- residual lignin content (approximately 15%) in the LCNFs, which are ex-
lose was not changed during the chemical and ultrasonic processes. pected to present a thermal barrier due to the stable aromatic units and
The CrI of the WIS was 49%, which represented an increase of 8.9% com- the covalent linkages between the cellulose and lignin [38,39]. In addi-
pared with the raw material due to the reduction of hemicellulose and tion, higher cellulose crystallinity of LCNFs is another factor that contrib-
lignin. No significant difference was observed between the WIS and utes to thermal stability of the composites [40]. Compared to LCNFs-
LCNFs, implying that the ultrasonic effect was non-selective in removing 1200 and LCNFs-1600, LCNFs-800 presented slower weight loss from
50 to 300 °C, indicating that the size of the fibers might affect the ther-
mal stability of the LCNFs. The mass loss started much sooner for more
Sample Crystallinity uniform nanocellulose due to the higher surface area when compared
(%) to the larger-sized fibers [16]. The third weight loss at 360 to 500 °C re-
Wheat straw 45 sulted from the decomposition of the PVA backbone.
WIS 49
Intensity / a.u.

LCNFs 800 50 3.8. Tensile properties


LCNFs 1200 50
LCNFs 1600 52 The mechanical properties of the nanocomposites are shown in
Fig. 9. All composite samples with LCNFs nanofillers exhibited excellent
mechanical properties compared to pure PVA. The tensile stress and
Young's modulus with a 5 wt% addition of LCNFs-1600/PVA were

Table 2
Degradation temperatures of main peaks and residues at 500 °C for PVA and LCNFs/PVA
nanocomposite films.

10 20 30 40 50 Sample Tonset (°C) Tmax (°C) Residues @ 500 °C (%)

2 Theta (degrees) PVA 201.1 234.2 12.4


LCNFs 800/PVA 209.1 244.1 12.9
LCNFs 1200/PVA 206.5 246.8 16.9
Fig. 7. X-ray diffraction patterns of the raw material, WIS and LCNFs with different
LCNFs 1600/PVA 209.2 246.8 22.8
ultrasonic treatment.
M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267 1265

Fig. 9. Mechanical properties of pure PVA and LCNFs/PVA nanocomposites.

76 MPa and 4.3 GPa, respectively, which were 1.75 and 3.3 times higher were expected [41]. The effect can result from the percolated network
than those of pure PVA film, indicating that LCNFs can be used as an ex- formed by strong intermolecular binding, including hydrogen bonds
cellent reinforcement agent to improve the mechanical performance of and mechanical interlocking between nanocellulose, which affect the
bio-based nanocomposites. In contrast, the elongation at break (EB%) stiffness and strength of composites [42,43].
decreased slightly with the presence of LCNFs. Generally, when the
nanocellulose was incorporated with the polymer matrix, increased 3.9. Surface hydrophobicity
tensile strength and Young's modulus as well as reduced elasticity
The water contact angle (WCA) of the pure PVA and PVA nanocom-
posites over a time course is shown in Fig. 10. The WCA of the pure PVA
50 films at 0 s was 43° and decreased to 21° after 80 s. Compared with the
PVA pure PVA film, the hydrophobicity of the LCNFs/PVA nanocomposite
LCNFs 800/PVA films over a time course of 80 s was improved due to the hydrophobicity
45 LCNFs 1200/PVA of residual lignin in the LCNFs. This result was consistent with previous
LCNFs 1600/PVA research [15]. Among the samples with different ultrasonic treatments,
40 the WCA deceased with the increase of ultrasonic intensity. According
to the AFM and TEM analyses, more uniform nanofibers with lower di-
WCA

35 ameters were obtained with the increase of ultrasonic intensity and re-
sulted in reduced surface roughness of the film, which contributed to
30 the lower contact angle [44].

4. Conclusion
25
LCNFs can lower the cost of production and reduce the polarity and
20 hydrophilicity of the nanocellulose, which represent factors that restrict
the large-scale utilization of polymer biocomposites. This study demon-
0 20 40 60 80 strated an easy, feasible, and environmentally friendly processing tech-
Time (s) nology to produce lignin containing cellulose nanomaterials (LCNFs)
from agricultural residue using a recyclable hydrotropic agent followed
Fig. 10. WCAs of pure PVA and LCNFs/PVA nanocomposites. by high intensity ultrasonication. Aqueous hydrotrope p-TsOH
1266 M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267

pretreatment was effective at removing hemicellulose and lignin below [16] S.S. Nair, N. Yan, Effect of high residual lignin on the thermal stability of nanofibrils
and its enhanced mechanical performance in aqueous environments, Cellulose 22
80 °C and resulted in cellulose-rich fibers with 15% lignin content. The (5) (2015) 3137–3150, https://doi.org/10.1007/s10570-015-0737-5.
AFM and TEM analyses showed that the extent of nanofibrillation of [17] M. Herrera, K. Thitiwutthisakul, X. Yang, P.-o. Rujitanaroj, R. Rojas, L. Berglund, Prep-
the obtained cellulose fibers was improved significantly as the intensity aration and evaluation of high-lignin content cellulose nanofibrils from eucalyptus
pulp, Cellulose 25 (5) (2018) 3121–3133, https://doi.org/10.1007/s10570-018-
of ultrasonication increased. The presence of these LCNFs in the PVA 1764-9.
composites films resulted in significant improvements in both tensile [18] M. Pishnamazi, H. Hafizi, S. Shirazian, M. Culebras, G.M. Walker, M.N. Collins, Design
strength and Young's modulus and slight decreases in elongation at of controlled release system for paracetamol based on modified lignin, Polymers 11
(6) (2019) 1059, https://doi.org/10.3390/polym11061059.
break. The thermal resistance and surface properties of the resultant [19] M. Pishnamazi, H.Y. Ismail, S. Shirazian, J. Iqbal, G.M. Walker, M.N. Collins, Applica-
nanocomposites were improved as well. Thus, LCNFs are expected to tion of lignin in controlled release: development of predictive model based on arti-
be a suitable candidate for various types of applications, including as ad- ficial neural network for API release, Cellulose 26 (10) (2019) 6165–6178, https://
doi.org/10.1007/s10570-019-02522-w.
ditives to composites and biomedical devices.
[20] M. Pishnamazi, S. Casilagan, C. Clancy, S. Shirazian, J. Iqbal, D. Egan, C. Edlin, D.M.
Croker, G.M. Walker, M.N. Collins, Microcrystalline cellulose, lactose and lignin
blends: process mapping of dry granulation via roll compaction, Powder Technol.
Credit author statement 341 (2019) 38–50, https://doi.org/10.1016/j.powtec.2018.07.003.
[21] R.J. Moon, A. Martini, J. Nairn, J. Simonsen, J. Youngblood, Cellulose nanomaterials
review: structure, properties and nanocomposites, Chem. Soc. Rev. 40 (7) (2011)
Mingyan Yang conceived the overall project and finalized the manu- 3941–3994, https://doi.org/10.1039/c0cs00108b.
script writing; Xiao Zhang performed the experiments and data analy- [22] L. Chen, J. Dou, Q. Ma, N. Li, R. Wu, H. Bian, D.J. Yelle, T. Vuorinen, S. Fu, X. Pan, Rapid
and near-complete dissolution of wood lignin at ≤80 °C by a recyclable acid
sis; Shuyi Guan and Xiaofeng Gao performed the experiments; Yan hydrotrope, Sci. Adv. 3 (9) (2017), e1701735. https://doi.org/10.1126/sciadv.
Dou supported the experiment and data analysis. 1701735.
[23] H. Bian, L. Chen, R. Gleisner, H. Dai, J. Zhu, Producing wood-based nanomaterials by
rapid fractionation of wood at 80 °C using a recyclable acid hydrotrope, Green Chem.
19 (14) (2017) 3370–3379, https://doi.org/10.1039/C7GC00669A.
Acknowledgments [24] M.Y. Yang, X.F. Gao, M. Lan, Y. Dou, X. Zhang, Rapid fractionation of lignocellulosic
biomass by p-TsOH pretreatment, Energy Fuel 33 (3) (2019) 2258–2264, https://
This work was supported by the Natural Science Foundation of doi.org/10.1021/acs.energyfuels.8b03770.
[25] Q. Liu, Y. Lu, M. Aguedo, N. Jacquet, C. Ouyang, W. He, C. Yan, W. Bai, R. Guo, D.E.
Shaanxi Province [grant numbers 2015SF263, 2018JQ4042]. Goffin, Isolation of high-purity cellulose nanofibers from wheat straw through the
combined environmentally friendly methods of steam explosion, microwave-
References assisted hydrolysis, and microfluidization, ACS Sustain. Chem. Eng. 5 (7) (2017)
6183–6191, https://doi.org/10.1021/acssuschemeng.7b01108.
[1] J.R. Pires, V.G. Souza, A.L. Fernando, Valorization of energy crops as a source for [26] T.S. Franco, D.C. Potulski, L.C. Viana, E. Forville, A.S. de Andrade, G.I.B. De Muniz,
nanocellulose production–current knowledge and future prospects, Ind. Crop. Nanocellulose obtained from residues of peach palm extraction (Bactris gasipaes),
Prod. 140 (2019), 111642. https://doi.org/10.1016/j.indcrop.2019.111642. Carbohydr. Polym. 218 (2019) 8–19, https://doi.org/10.1016/j.carbpol.2019.04.035.
[2] W. Li, J. Yue, S. Liu, Preparation of nanocrystalline cellulose via ultrasound and its re- [27] W. Chen, H. Yu, Y. Liu, P. Chen, M. Zhang, Y. Hai, Individualization of cellulose nano-
inforcement capability for poly (vinyl alcohol) composites, Ultrason. Sonochem. 19 fibers from wood using high-intensity ultrasonication combined with chemical pre-
(3) (2012) 479–485, https://doi.org/10.1016/j.ultsonch.2011.11.007. treatments, Carbohydr. Polym. 83 (4) (2011) 1804–1811, https://doi.org/10.1016/j.
[3] N. Amiralian, P.K. Annamalai, P. Memmott, D.J. Martin, Isolation of cellulose carbpol.2010.10.040.
nanofibrils from Triodia pungens via different mechanical methods, Cellulose 22 [28] H. Lee, S. Mani, Mechanical pretreatment of cellulose pulp to produce cellulose
(4) (2015) 2483–2498, https://doi.org/10.1007/s10570-015-0688-x. nanofibrils using a dry grinding method, Ind. Crop. Prod. 104 (2017) 179–187,
[4] H. Lee, S. Hamid, S. Zain, Conversion of lignocellulosic biomass to nanocellulose: https://doi.org/10.1016/j.indcrop.2017.04.044.
structure and chemical process, Sci. World J. (2014) 1–20, https://doi.org/10.1155/ [29] A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. Templeton, D. Crocker, Determi-
2014/631013. nation of structural carbohydrates and lignin in biomass, Laboratory Analytical Pro-
[5] J.R. Pires, V.G.L. De Souza, A.L. Fernando, Production of nanocellulose from lignocel- cedure, 1617, 2008, pp. 1–16.
lulosic biomass wastes: prospects and limitations, International Conference on Inno- [30] R. Megargle, ASTM (American Society for Testing and Materials) standards for med-
vation, Engineering and Entrepreneurship, Springer 2018, pp. 719–725, https://doi. ical computing, Computers in Healthcare, 11(2), 1990, pp. 25–26.
org/10.1007/978-3-319-91334-6_98. [31] A. Ferrer, E. Quintana, I. Filpponen, I. Solala, T. Vidal, A. Rodríguez, J. Laine, O.J. Rojas,
[6] K.-Y. Lee, Y. Aitomäki, L.A. Berglund, K. Oksman, A. Bismarck, On the use of Effect of residual lignin and heteropolysaccharides in nanofibrillar cellulose and
nanocellulose as reinforcement in polymer matrix composites, Compos. Sci. nanopaper from wood fibers, Cellulose 19 (6) (2012) 2179–2193, https://doi.org/
Technol. 105 (2014) 15–27, https://doi.org/10.1016/j.compscitech.2014.08.032. 10.1007/s10570-012-9788-z.
[7] Z. Wang, P. Tammela, M. Strømme, L. Nyholm, Nanocellulose coupled flexible poly- [32] K.L. Spence, R.A. Venditti, O.J. Rojas, Y. Habibi, J.J. Pawlak, The effect of chemical com-
pyrrole@ graphene oxide composite paper electrodes with high volumetric capaci- position on microfibrillar cellulose films from wood pulps: water interactions and
tance, Nanoscale 7 (8) (2015) 3418–3423, https://doi.org/10.1039/C4NR07251K. physical properties for packaging applications, Cellulose 17 (4) (2010) 835–848,
[8] P. Khawas, A.J. Das, S.C. Deka, Production of renewable cellulose nanopaper from cu- https://doi.org/10.1007/s10570-010-9424-8.
linary banana (Musa ABB) peel and its characterization, Ind. Crop. Prod. 86 (2016) [33] C.M. Ewulonu, X. Liu, M. Wu, Y. Huang, Ultrasound-assisted mild sulphuric acid ball
102–112, https://doi.org/10.1016/j.indcrop.2016.03.028. milling preparation of lignocellulose nanofibers (LCNFs) from sunflower stalks
[9] J.-H. Seo, T.-H. Chang, J. Lee, R. Sabo, W. Zhou, Z. Cai, S. Gong, Z. Ma, Microwave flex- (SFS), Cellulose 26 (7) (2019) 4371–4389, https://doi.org/10.1007/s10570-019-
ible transistors on cellulose nanofibrillated fiber substrates, Appl. Phys. Lett. 106 02382-4.
(26) (2015), 262101. https://doi.org/10.1063/1.4921077. [34] R. Zuluaga, J.L. Putaux, J. Cruz, J. Vélez, I. Mondragon, P. Gañán, Cellulose microfibrils
[10] J. Dou, H. Bian, D.J. Yelle, M. Ago, K. Vajanto, T. Vuorinen, J.J. Zhu, Lignin containing from banana rachis: effect of alkaline treatments on structural and morphological
cellulose nanofibril production from willow bark at 80 °C using a highly recyclable features, Carbohydr. Polym. 76 (1) (2009) 51–59, https://doi.org/10.1016/j.
acid hydrotrope, Ind. Crop. Prod. 129 (2019) 15–23, https://doi.org/10.1016/j. carbpol.2008.09.024.
indcrop.2018.11.033. [35] A. Kaushik, M. Singh, Isolation and characterization of cellulose nanofibrils from
[11] H. Bian, L. Chen, H. Dai, J. Zhu, Integrated production of lignin containing cellulose wheat straw using steam explosion coupled with high shear homogenization,
nanocrystals (LCNC) and nanofibrils (LCNF) using an easily recyclable di- Carbohydr. Res. 346 (1) (2011) 76–85, https://doi.org/10.1016/j.carres.2010.10.
carboxylic acid, Carbohydr. Polym. 167 (2017) 167–176, https://doi.org/10.1016/j. 020.
carbpol.2017.03.050. [36] M.I. Voronova, O.V. Surov, S.S. Guseinov, V.P. Barannikov, A.G. Zakharov, Thermal
[12] S. Herzele, S. Veigel, F. Liebner, T. Zimmermann, W. Gindl-Altmutter, Reinforcement stability of polyvinyl alcohol/nanocrystalline cellulose composites, Carbohydr.
of polycaprolactone with microfibrillated lignocellulose, Ind. Crop. Prod. 93 (2016) Polym. 130 (2015) 440–447, https://doi.org/10.1016/j.carbpol.2015.05.032.
302–308, https://doi.org/10.1016/j.indcrop.2015.12.051. [37] Y. Li, Y. Liu, W. Chen, Q. Wang, Y. Liu, J. Li, H. Yu, Facile extraction of cellulose
[13] C.M. Ewulonu, X. Liu, M. Wu, Y. Huang, Lignin-containing cellulose nanomaterials: a nanocrystals from wood using ethanol and peroxide solvothermal pretreatment
promising new nanomaterial for numerous applications, J. Bioresour. Bioprod. 4 (1) followed by ultrasonic nanofibrillation, Green Chem. 18 (4) (2016) 1010–1018,
(2019) 3–10, https://doi.org/10.21967/jbb.v4i1.186. https://doi.org/10.1039/C5GC02576A.
[14] M.A. Herrera, J.A. Sirviö, A.P. Mathew, K. Oksman, Environmental friendly and sus- [38] A. Kaushik, M. Singh, Isolation and characterization of cellulose nanofibrils
tainable gas barrier on porous materials: nanocellulose coatings prepared using from wheat straw using steam explosion coupeffectled with high shear ho-
spin-and dip-coating, Mater. Des. 93 (2016) 19–25, https://doi.org/10.1016/j. mogenization, Carbohydr. Res. 346 (1) (2011) 76–85, https://doi.org/10.
matdes.2015.12.127. 1016/j.carres.2010.10.020.
[15] E. Rojo, M.S. Peresin, W.W. Sampson, I.C. Hoeger, J. Vartiainen, J. Laine, O.J. Rojas, [39] A. Rangan, M.V. Manchiganti, R.M. Thilaividankan, S.G. Kestur, R. Menon, Novel
Comprehensive elucidation of the effect of residual lignin on the physical, barrier, method for the preparation of lignin-rich nanoparticles from lignocellulosic fibers,
mechanical and surface properties of nanocellulose films, Green Chem. 17 (3) Ind. Crop. Prod. 103 (2017) 152–160, https://doi.org/10.1016/j.indcrop.2017.03.
(2015) 1853–1866, https://doi.org/10.1039/C4GC02398F. 037.
M. Yang et al. / International Journal of Biological Macromolecules 158 (2020) 1259–1267 1267

[40] P. Yang, S. Kokot, Thermal analysis of different cellulosic fabrics, J. Appl. Polym. Sci. [43] A. Alemdar, M. Sain, Isolation and characterization of nanofibers from agricultural
60 (8) (1996) 1137–1146, https://doi.org/10.1002/(SICI)1097-4628(19960523)60: residues–wheat straw and soy hulls, Bioresour. Technol. 99 (6) (2008)
8b1137::AIDAPP6N3.0.CO;2-M. 1664–1671, https://doi.org/10.1016/j.biortech.2007.04.029.
[41] D. Liu, X. Sun, H. Tian, S. Maiti, Z. Ma, Effects of cellulose nanofibrils on the structure [44] A.G. Patil, M. Selvakumar, S. Anandhan, Characterization of composites based on
and properties on PVA nanocomposites, Cellulose 20 (6) (2013) 2981–2989, https:// biodegradable poly (vinyl alcohol) and nanostructured fly ash with an emphasis
doi.org/10.1007/s10570-013-0073-6. on polymer–filler interaction, J. Thermoplast. Compos. Mater. 29 (10) (2016)
[42] Q. Tarrés, N.V. Ehman, M.E. Vallejos, M.C. Area, M. Delgado-Aguilar, P. Mutjé, Ligno- 1392–1415, https://doi.org/10.1177/0892705714563130.
cellulosic nanofibers from triticale straw: the influence of hemicelluloses and lignin
in their production and properties, Carbohydr. Polym. 163 (2017) 20–27, https://
doi.org/10.1016/j.carbpol.2017.01.017.

You might also like