Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

The extraction of lignocelluloses and silica from rice husk using a single
biorefinery process and their characteristics
Ngoc Thuy Nguyen a,b, Nhat Thong Tran a,b, Tan Phat Phan a,b, Anh Thu Nguyen a,b, My Xuyen T. Nguyen a,b,
Nguyen Ngan Nguyen a,b,c,⇑, Young Ho Ko c,d, Dai Hai Nguyen e, Tran T.T. Van a,b, DongQuy Hoang a,b,⇑
a
Faculty of Materials Science and Technology, University of Science, Vietnam National University, Ho Chi Minh City 700000, Viet Nam
b
Vietnam National University, Ho Chi Minh City 700000, Viet Nam
c
Department of Chemical Engineering, Pohang University of Science and Technology, Pohang 37673, South Korea
d
Center for Self-assembly and Complexity, Institute for Basic Science, Pohang 37673, South Korea
e
Institute of Applied Materials Science, Vietnam Academy of Science and Technology, 01 TL29, District 12, Ho Chi Minh City 700000, Viet Nam

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

Article history: While the efficient usage of biomass waste can significantly help in addressing environmental issues,
Received 13 October 2021 there are only a few reports that discuss about processing such waste effectively at a low-cost. Such chal-
Revised 3 December 2021 lenge arises from the strong association between the components biomass. In this study, an abundant
Accepted 27 December 2021
agricultural byproduct, rice husk (RH), was used as the starting resource. A simple biorefining process
Available online 4 January 2022
of alkaline peroxide treatment followed by acid precipitation and ethanol extraction was performed on
RH to obtain cellulose, hemicellulose, lignin, and silica. The chemical structures, morphologies, and
Keywords:
physic-chemical properties of the separated components were identified through a wide range of char-
Rice husk
Recycling industry
acterization approaches. The final products obtained from of this process were (i) bundles of fiber-like
Cellulose cellulose with a fiber width of 6 mm and (ii) small particles of hemicellulose and lignin with non-
Hemicellulose uniform shapes. The lignocelluloses products had over 90 wt% carbon with 52.28% crystalline ratio.
Lignin Meanwhile, the other products comprising hemicelluloses, lignin, and silica were amorphous. The out-
Silica come of this study contributes to expanding and developing the simple and efficient conversion process
of biomass waste into sustainable value-added materials. It is crucial to reduce the environmental impact
by using renewable materials as the new building block resources for synthetic chemicals.
Ó 2022 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Introduction RH, a protective shell of rice grains, exists in four different lay-
ers: structural, fibrous, sponge-like, and cellular. The main con-
Rice is one of the most consumed and vital food items globally. stituents of RH are cellulose (35%), hemicellulose (25%), lignin
In 2017–18, the annual quantity of rice utilized globally was (20%), crude protein (3%), and inorganic compounds (17%) [4]. It
approximately 503 million tons, and the average per capita food should be noted that amorphous silica accounts for about 95–
consumption was approximately 53.7 kg per year [1]. It is esti- 98% weight, on average, of the inorganic fraction. While previous
mated that the demand for rice, including its production and con- studies have investigated the isolation of lignocellulosic biomass
sumption, will increase continuously. Roughly 20% w/w of paddy components [5–8], the majority of them have focused on sepa-
grain weight is rice husk (RH), which is a low-cost, abundant rately extracting one or two potentially high-value substances,
agro-industrial byproduct. However, rice waste, especially RH, is such as silica, cellulose nanocrystals, and lignin, for certain applica-
not reused and, thus, contributes to serious pollution. Therefore, tions or for even producing bioethanol from fermentable sugars
it is of immense benefit, both economically and environmentally, from lignocellulose biomass [9–11]. Tzong-Horng Liou [11] studied
to efficiently process RH into valuable materials [2,3]. the non-isothermal degradation of RH in an oxidizing atmosphere
to obtain uniformly sized ultrafine silica powder, with an average
⇑ Corresponding authors at: Faculty of Materials Science and Technology, particle size of 60 nm and average pore diameter of 5.4 nm. Nurain
University of Science, Vietnam National University, Ho Chi Minh City 700000, Viet Johar et al. [9] used classical chemical treatments methods, includ-
Nam. ing alkali and bleaching treatments, to remove non-cellulosic
E-mail addresses: ngannguyen@postech.ac.kr (N.N. Nguyen), htdquy@hcmus. material (hemicellulose and lignin) and obtain cellulose fibers with
edu.vn (D. Hoang).

https://doi.org/10.1016/j.jiec.2021.12.032
1226-086X/Ó 2022 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

diameters ranging from 7 to 170 lm and a crystallinity index of then poured into a three-fold volume of distilled water to obtain
59%. Subsequently, sulphuric acid (H2SO4) hydrolysis treatment hemicellulose 2. Hemicellulose 2 was also oven-dried at 45 °C.
was used to extract nanocrystals from these fibers. Claudia Fer- Alkaline/hydrogen peroxide treatment: The above alkaline-
nanda Lemonse Silva and et al. [10] produced second-generation treated RH was soaked in distilled water, in a flask overnight,
ethanol from lignocellulosic biomasses of giant reed (Arundo donax before adding 1 M NaOH until a pH 11 of 11 was obtained. H2O2
L.) through a simultaneous saccharification and fermentation pro- solution (30%), with a ratio between alkaline treated RH and
cess resulting in approximately 75 L of ethanol per ton of cellulose, H2O2 of 1:2 (w/w), was then slowly dropped into the solution. This
with an ethanol concentration of 39 g/L at 70 h. The concentration procedure was repeated twice, and cellulose was extracted from
upon extraction of one or two specific fractions and the removal of the mixture through a filtering and neutralization process. The
other components in biomass is considered wasting high-value pH of the rest of the solution was adjusted until it came down to
green materials. Only a few studies have discussed about extract- 3 by adding 1 M HCl to precipitate the solubilized lignin. The neu-
ing main lignocellulose fractions from RH and using it for potential tralized samples were oven-dried at 45 °C.
applications [12,13]. In these reports, a combination of acidic treat-
ment, isolation, and chemical activation with potassium hydroxide
was used to extract the products of hemicelluloses, activated car- Characterization and measurements
bon, and other fragments and activated. However, considering that The functional group and chemical structure analysis of cellu-
arabinose side groups and O-acetyl-galactoglucomannan within lose, hemicellulose, lignin, and silica were performed by using FTIR
hemicelluloses are prone to degradation in acidic conditions, the C1 (PerkinElmer Spectrum Version 10.5.2) and NMR spectrome-
reported reaction condition leads to the acid hydrolysis of these ters. The solid-state CP/MAS 13C NMR spectrum was recorded on
polysaccharides [14,15]. In this study, we specifically designed a digital Avance III HD 850 (Bruker AXS GmbH) spectrometer at
the extraction process in a way that would not cause structural 25 °C at a frequency of 125.78 MHz, with 10,240 scans at 5 s recy-
changes in the components. In fact, the clear presence of the pro- cle delay. The degree of crystallinity was measured by adopting the
duct’s arabinoxylans and glucomannan signals in the FTIR and equation used by Xiaolan Zhu et al. [16]:
NMR results in our study is a strong evidence that the final hemi-
cellulose was structural intact. It was found that the main chain SS
CrI% ¼  100 ð1Þ
and side groups were not degraded during the treatment process. Ss þ Si
Importantly, while our process simple and includes only two steps, where Ss is the area of the crystalline domain, and Si is the area
it could simultaneously extract every useful lignocelluloses and sil- of the amorphous domain.
ica from RH in purity. Moreover, through our high-yield biorefining Thermal analysis was conducted by using a differential scan-
process, we could conduct one of the first comprehensive investi- ning calorimetry (DSC) instrument (METTLER STARe SW 11.00).
gations to assess on the chemi-physical properties of RH-based lig- The sample was weighted to approximately 5 mg in a crucible
nocelluloses and silica. before placing it in a sample holder. It was heated from ambient
temperature to 350 °C, at a heating rate of 10 °C/min under a nitro-
gen atmosphere with a flow rate of 30 ml/min. The pyrolysis of bio-
Materials and methods mass components was performed in a thermogravimetric analysis
(TGA) instrument (LABSYS Evo STA 1600) under the following
Materials experimental conditions: temperature range from room tempera-
ture to 800 °C, heating rate of 10 °C/min, and N2 atmosphere with
RH was collected from local farms in the South of Vietnam. a purge rate of 30 mL/min.
Sodium hydroxide, hydrochloric acid, ethyl alcohol, and hydrogen The crystallinity of the samples was investigated using an X-ray
peroxide solution (30%) were purchased from JHD Chemical diffractometer (XRD) (Rigaku D/Max-2550 V) using Cu-Ka radia-
(Xilong, China). Distilled water was used in all preparations. tion (k = 0.15405 nm, 40 kV, 40 mA) in the scattering range (2h)
of 10-80° with a step rate of 0.25 °/min. The degree of crystallinity
was calculated as follows [17]:
Methods
Ið002Þ  Iam
Biorefinery recovery process CrI% ¼  100 ð2Þ
Ið002Þ
Silica, hemicellulose, cellulose, and lignin were obtained sepa-
rately through the biorefining process of RH. The materials were where I(002) is the maximum intensity at the (002) crystallo-
recovered simultaneously by following a two-step procedure graphic plane, and Iam is the minimum intensity between (110)

(Fig. 1). and (002) peaks with 2h  18 :
Alkaline treatment: Washed and dried RH was soaked in dis- The morphology and structure of the isolated biomass com-
tilled water for 24 h before being treated with 1.8% (w/v) NaOH pounds were analyzed through a scanning electron micro-scope
solution. It was then stirred for 3 h at 100 °C. The mixture was sub- (SEM). SEM investigations were performed under high-vacuum
sequently filtered, and the residue was washed with distilled water conditions using a S-4800 SEM instrument (Hitachi), with the fol-
repeatedly to acquire the neutralized sample for the next alkaline/ lowing parameter settings: accelerating voltage of 3 kV, current of
hydrogen peroxide treatment process while the liquor was 11 pA, working distance of 10 mm, and spot size of 250 nm. Energy
adjusted to attain a pH of 9 by adding 1 M HCl. The suspension dispersive spectroscopy (EDS) (EMAX, Horiba) mapping was used
from the liquor was filtered and washed with hot distilled water for the composition analysis in all the samples.
prior to being dried in the oven at 60 °C, followed by incineration
at 700 °C for 4 h inside a programmable furnace to obtain silica.
Hemicellulose 1 was isolated from the filtered solution by adding Results and discussion
1 M HCl until a pH 5 was obtained. The mixture was then poured
into a three-fold volume of ethanol. The recovered hemicellulose All the isolated lignocellulose fractions and silica (Fig. 2)
1 was oven-dried at 45 °C. After extracting hemicellulose 1, 1 M obtained through the biorefining process of RH were analyzed for
HCl was added to the solution until a pH of 5 was obtained and their identifications and chemi-physical characteristics.
151
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

Fig. 1. Brief of the biorefining process.

Fig. 2. Image of pristine rice husk (A), extracted cellulose (B), hemicellulose 1 (C), hemicellulose 2 (D), lignin (E), and silica (F).

152
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

FTIR analysis Table 1


Summary of the main functional groups of the three components in RH.

Organic fractions in RH account for 70–85% and the rest is inor- Band Assignment
ganic components, which depend on the type of species and differ- (cm1)
ent climatic and geographic conditions [4]. FTIR analysis was 3440–3400 O–H stretching
carried out to identify all the lignocellulosic components and silica 2920–2850 C–H stretching
through the special functional groups described in Fig. 3. These 1699 C@O stretching
1654–1599 Aromatic skeletal vibrations plus C@O stretching
details are summarized in Table 1. 1510 Aromatic ring, aromatic skeleton vibrations belonging to lignin
The FTIR spectra of RH show clear characteristic peaks of cellu- 1460 C–H deformations
lose, hemicellulose, and lignin. The absorbance bands at 3440– 1432 O-CH3
3000 cm1 and 2920–2850 cm1 for FTIR spectra arise due to O– 1373 Aliphatic C–H stretching in CH3 and phenolic OH
1234, 1167 C-O-C stretching vibration
H and C–H stretching, respectively. In the FTIR results of cellulose,
1032 S unit
the bands at 1644, 1511, 1432, and 1373 cm1 are associated with 897 b-glycosidic linkages
the stretching vibration of C = O link, aromatic skeleton vibrations,
the vibration of O-CH3 link, aliphatic C–H stretching in CH3, and
phenolic OH, respectively. Meanwhile, C-O-C stretching, OH associ-
mannan was observed at 803 cm1 [22], and the range at 661–
ation (C-OH), and the b-glycosidic linkages between glucose in cel-
400 cm1 was due to C–C stretching vibrations. However, the
lulose and C–C stretching vibrations were observed at 1234–1167,
absorbance zones at 1100–467 cm1 also indicate the presence of
1111, and 897 cm1, respectively [18,19].
silica residue in hemicellulose 1. There were different characteris-
For hemicellulose 1, the peaks at 1733, 1632, 1520, 1460–1413,
tic peaks in the FTIR spectrum of hemicellulose 2: the aromatic
1151–1099 cm1 seen in the spectrum were assigned to the C = O
compounds at 1696 cm1 [19], the characteristic band of uronic
stretching vibration of acetyl xylan, absorbed water, C = O stretch-
acid at 1599 cm1 [23], and the signals at 1461, 1423, 1360,
ing vibration in ketone and carboxyl groups, bending vibrations of
1329, and 1267 cm1 relate to the bending vibrations of C–H, C–
C–H, O–H, and CH2 in hemicelluloses, and C–OH ring vibrations
O, CH2, O–H, and C–C, respectively. The characteristic peak of ara-
overlapping with the stretching vibrations of the side groups and
binoxylans at 1129 cm1 was observed distinctively in the FTIR
C–O–C glycosidic bond vibrations, respectively [7,19,20]. The peak
spectrum of hemicellulose 2. Additionally, the occurrence of a peak
at 970 cm1 represents the appearance of arabinose substitution at
at 1510 cm1 could be correlated with a small amount of lignin in
C-3 of the xylose residues [6,21]. The characteristic band of gluco-
hemicellulose [6,21]. A characteristic signal of the substituted ara-

Fig. 3. (A) FTIR spectra of RH, lignocelluloses, and silica extracted from RH. (B) Chemical structures of lignin and the FTIR spectrum of lignin in the wavenumber range of 400–
2000 cm1.

153
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

binoxylans was also found at 1031 cm1. The absorbance range at 4.96 ppm and 5.10 ppm refer to the hydrogen H1 linking to the car-
700–400 cm1 corresponds to the characteristic of C–H stretching bon C1 of the glucose and mannose units, which are typical for glu-
vibrations [19]. comannans [31]. The region at d 5.72 ppm was caused by a-
The acid-extracted product at a pH of 3 produced peaks at 1699 configuration. b-configuration appeared at the region between d
and 1654 cm1 for lignin, demonstrating that carbonyl groups exist 4.1–4.5 ppm while the resonance signals appeared around 6.60–
in the b position and in the a and c positions of the phenylpropane 7.00 ppm, which originated from phenolic compounds due to the
unit and acetyl or uronic ester peaks, respectively. The appearance presence of small amounts of lignin in hemicellulose. Likewise,
of the characteristic regions at 1599, 1510, 1464, and 1423 cm1 in these signals can be seen in the spectra of hemicellulose 2 in
the spectrum of lignin relates to the aromatic skeletal vibrations Fig. 4(D). However, the signals at d 3.7 ppm (H-5 eq in b-D-
and C = O stretching, the aromatic rings and aromatic skeletal xylopyranose units) and around the region of d 1.0–1.1 ppm (to
vibrations, C–H deformations (especially asymmetric in –CH3 and the methyl protons of 4-O-methyl-D-glucuronic acid) were stron-
–CH2–), and the aromatic skeletal vibrations combined with C–H ger than in the spectra of hemicellulose 1.
in-plane deformations, respectively. In the band from 500 to
1350 cm1, there were clear existence of the S unit (1331, 1223, DSC and TGA analysis
and 1129 cm1), G unit (1267 cm1), and S + G unit, e.g., the aro-
matic C–H in-plane deformations at 1032 cm1, the C–H out-of- Thermal analysis methods provide information about the ther-
plane in positions 2, 5, and 6 (G + S unit), and the presence of aro- mal stability and decomposition reactions with respect to the char-
matic C = C and C–C groups at 835–616 cm1 [12,19]. acteristics of each biopolymer for identification purposes in some
The spectrum of silica isolated from RH, following after the cases. TGA and DSC are the methods available for studying the
alkali treatment and incineration at 700 °C in 4 h, shows that most thermal properties of samples.
of the organic components completely disappeared. The character- The thermalgravimetric curve of RH, with various thermal
istic peaks of silica at 1095, 808, and 465 cm1 correspond to the decomposition steps in the temperature range between 25–
O–Si–O bending and stretching vibrations. The shoulder peak at 800 °C, is shown in Fig. 5(A). The weight loss of absorbed water
1220 cm1 represents the asymmetric stretching mode of SiO4 in RH was 3.92 wt% at 128 °C, and the weight of RH began to signif-
coordination units [24]. icantly decrease above 200 °C. This continued to occur at even over
500 °C, with a total weight loss of 59.92 wt%. In this study, the first
NMR analysis and second step respectively occurred at 303–324 °C and at
358–380 °C, which are similar to earlier reports [32,33]. This
The functional groups and chemical structures of cellulose, lig- complex process was because of the superposition of pyrolysis of
nin, and hemicellulose fractions were analyzed by CP/MAS, 13C the three main constituents, including the degradation of hemicel-
NMR, and H-NMR and are illustrated in Fig. 4. The aromatic carbon lulose in the lower temperature range (150–350 °C), the decompo-
ring of cellulose appeared in the spectrum with six signals (Fig. 4 sition of cellulose in the higher temperature range (275–380 °C),
(A)). A single peak at 106.28 ppm relates to the C1 carbon atom, and lignin over the wider temperature range (250–550 °C)
and the peaks resonance at 73.78 and 76.22 ppm are assigned to [34,35]. The solid residue obtained after RH pyrolysis contained
C2, C3, and C5. The C4 glycosidic bond appeared at 85.36 ppm major amounts of silica and trace amounts of other elements.
(amorphous region) and 89.97 ppm (crystalline region). The peaks The thermal properties of the individual lignocellulosic compo-
at 64.09 and 66.21 ppm represent C6 carbons of amorphous cellu- nents extracted from RH were different from that of pristine RH. In
lose and crystalline cellulose. Pure cellulose exists in several allo- other words, the isolation process of all compounds in RH can
morphic forms. The result of 13C chemical shifting in the CP/MAS cause broken bonds or interactions among them, leading to non-
13
C NMR spectrum of cellulose indicates that the obtained cellulose overlapped pyrolysis. The TGA curves in Fig. 5(A) indicate that
from RH in this study was cellulose I [25]. Furthermore, as 13C NMR there were three phases during the biomass pyrolysis. The thermal
is particularly helpful to evaluate the crystallinity of cellulose [26], degradation of all lignocelluloses between 90–100 °C was caused
we estimated that the crystallinity index of cellulose was 52.28%. by moisture evaporation. The main degradation process of the cel-
In Fig. 4(B) of the CP/MAS 13C NMR spectrum, the peak that lulose fraction occurred between 278–508 °C, with two-stage
appeared at 173.40 ppm represents unconjugated aliphatic –COOH pyrolysis and a total weight loss of 94.7 wt%. The first stage began
groups of lignin [27] while the peak at 160.42 ppm is associated at 278–344 °C, and the maximum weight loss rate of 7.8 wt%/°C
with C4 in p-coumarate. The resonance regions at 153.48, 150.02, occurred at 318 °C. The second stage started at 363–508 °C, and
134.20, and 115.24 ppm were due to C3/C3 in etherified 5–5 units the maximum weight loss rate of 1.6 wt%/°C occurred at 426 °C.
in syringyl (S) and guaiacyl (G) residues, C3/C5 in etherified S units, The amount of solid residue was 0.4 wt%. There were two path-
C1 in etherified S units and C1 in etherified G units, and C5 in G ways in this major pyrolysis process of hemicellulose. However,
units, respectively. In addition, the signals of C2 and C6 in Ca- the main decomposition phase of hemicellulose took place rapidly
oxidized S units, Cb in G type b-O-4 units, C4 in b-D- when the temperature went from 256 to 709 °C in the case of
xylopyranoside, C3 in 3-O-acetyl-b-D-xylopyranoside, and Cc-in hemicellulose 1 and from 283 to 636 °C in the case of hemicellu-
cinnamyl acetate end-groups were observed in the CP/MAS 13C lose 2. In the first stage, the maximum weight loss rate of hemicel-
NMR spectrum of lignin at the resonances of 106.40, 84.00, lulose 1 and hemicellulose 2 occurred at 328 °C and 354 °C,
75.84, 73.96, and 64.03 ppm, respectively. The peak that appeared respectively. Hemicellulose 1 lost only 17.2 wt% at 434 °C while
at 56.81 ppm was caused by C in Ar–OCH3 while the resonances at hemicellulose 2 lost 40.6 wt% at 410 °C in the first stage. The sec-
47.97, 33.95, and 31.57 ppm correspond to CH3 groups, ketones, ond stage of hemicellulose 1 appeared in the temperature range
and the aliphatic chain, respectively [27–29]. of 485–709 °C and for hemicellulose 2, it occurred at 434–636 °C.
For hemicellulose 1, in Fig. 4(C), it can be seen that the DMSO-d6 There were 75.2 wt% residual char (hemicellulose 1) and 42.4 wt
solvent peak was at 3.35 ppm, and the strong signal at 2.50 ppm % residual char (hemicellulose 2) even at over 600 °C. The remnant
was caused by water in the solvent. The peaks at the range of weight following the hemicellulose 1 pyrolysis process was higher
1.0–1.1 ppm correspond to the methyl protons of 4-O-methyl-D- than that of hemicellulose 2 because of the existence of a small
glucuronic acid [30]. The weak signals at d 4.24/4.26 ppm relate amount of silica in hemicellulose 1. The weight of lignin sharply
to the anomeric protons of b-D-xylose grafted on C-3 residues declined from 255 to 700 °C, with a weight loss of 34.8 wt% at
while the appearance of the well-differentiated signals at d 255–361 °C in the first stage and a weight loss of 57.1 wt% at
154
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

13
Fig. 4. CP/MAS C NMR spectra of cellulose (A) and lignin (B). 1H NMR spectra of hemicellulose 1 (C) and hemicellulose 2 (D).

Fig. 5. (A) TGA curves of RH, lignocelluloses, and silica. (B) DSC curves of the extracted lignocelluloses and silica.

368–670 °C in the second stage. The maximum weight loss rates of The differences in pyrolysis of all the samples resulted from the
lignin were 6.8 wt%/°C at 326 °C and 3.5 wt%/°C at 520 °C. In the differences in the physical texture and the chemical structures of
case of silica, there was a negligible decrease in weight when the the biomass components. However, lignocellulosic pyrolysis
temperature went up from 25 to 800 °C. Thus, silica was stable at mainly released gas products (H2, CH4, CO, CO2) and some organic
this temperature range. compounds along with H2O [19,36]. At low temperatures, organic

155
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

compounds were released by reforming C@O and C–O–C bonds of


hemicellulose at below 400 °C and below 450 °C in the case of cel-
lulose while those from lignin were negligible. Among the incon-
densable gases, it was reported that the dominant gas products
were CO and CO2 while H2 and CH4 were the minor ones released
during pyrolysis. CO2 gas was released the most from cracking C–C
and C–O bonds connected with the main branch of hemicellulose,
at 280, 451, and 658 °C while the abscission of C@O groups in cel-
lulose created CO2 at 300 °C. Among the two temperature zones
of releasing CO2 from lignin at 340 and 700 °C, CO2 content at
700 °C was released more. The release of CO from cellulose degra-
dation was minor at  380 °C while most of the CO released from
hemicellulose pyrolysis occurred during the whole temperature
range while that from lignin occurred at high temperatures. CH4
gas, the main product during lignin pyrolysis, released at 400–
600 °C because of the highest methoxyl–O–CH3 content in lignin.
The release of CH4 from hemicellulose occurred at 280 and
520 °C. H2 released at a higher temperature (>400 °C).
Apart from the TGA curves indicating the pyrolysis of the com-
ponents, the DSC curves, showing the energy consumption charac-
teristics during the decomposition of samples, were also measured.
The results are plotted in Fig. 5(B). When the temperature range
Fig. 6. XRD patterns of cellulose, hemicellulose 1, hemicellulose 2, lignin, and silica.
was between 80–100 °C, the DSC curves of lignocelluloses pyroly-
sis indicated a similar tendency as the above TGA results. The SEM micrographs and EDS mapping analysis
endothermic peaks at 100 °C indicate the removal of waste mois-
ture. The glass transition temperatures (Tg) of hemicellulose 1, The SEM micrographs and EDS mapping in Fig. 7 reveal differ-
hemicellulose 2, and lignin observed in Fig. 5(B) were 136, 158, ences in the morphology and elemental distribution of the
and 94 °C, respectively. The Tg results of hemicelluloses in this extracted cellulose, hemicellulose 1, hemicellulose 2, and lignin
study are almost similar to those of hemicelluloses from grasses fractions. The smooth bundles of cellulose, the last product of the
[20]. With regard to lignin, it is difficult to determine its Tg because biomass extraction process from RH, were clearly seen in the
of the heterogeneity of lignin chemistry. However, the Tg of non- SEM images (Fig. 7(A)). The images indicate the successful removal
derivatized lignins has generally been found at temperatures of of waxes and encrusting substances, including hemicellulose and
90–180 °C depending on the molecular weight of lignins [37–39]. lignin, on the fiber surface. The width of the fiber was approxi-
Above 200 °C, the degradation process of cellulose, hemicellulose, mately 6.0 mm. Unlike cellulose, the lignin products were in the
and lignin began to occur, and the results were similar to those form of irregular micron particles with an average size of 2.6 mm.
obtained in the TGA method. These results are consistent with those obtained in previous
studies.
The SEM micrographs of the hemicelluloses fractions show that
the hemicellulose product comprised non-uniform small particles,
XRD analysis some of which agglomerated to create smooth and angular sur-
faces. Moreover, the surface of the hemicellulose 1 fraction was
X-ray diffraction patterns of lignocelluloses and silica were rougher than that of the hemicellulose 2 fraction due to silica.
extracted from RH and are shown in Fig. 6. Hemicellulose, being The results of EDS elemental mapping, showing the presence of
a group of several heteropolymers from different monosaccharides, elements in all the samples, can be seen in Fig. 7(B). Importantly,
has a random and amorphous structure that is present around the C was detected in the entire map area, corresponding to the ligno-
crystalline microfibrils of cellulose. This entire structure is embed- cellulose compounds. The content of C element in all the samples
ded in the matrix of lignin in almost all wood cell walls. The results was over 90 wt%, except for hemicellulose 1, which had C of
of XRD analysis indicate that the cellulose obtained following RH 48.80 wt% and Si of 27.94 wt% (Fig. 7(C)).
treatment still maintained its crystalline structure with two sharp
diffraction peaks at 2h = 16.1° of the lattice plane (110) and 22.8° of
(002) along with a small peak at 2h = 35.1°, corresponding to the Conclusion
lattice plane (004) [7,29]. These lattice planes were assigned to cel-
lulose Ib (PDF no.00-056-1718). The relative amount of crystalline The treatment of RH by following a two-step process—simulta-
material (CrI) in the cellulose fraction in this study was 84.63% neous alkaline and peroxide treatment—was effective in extracting
while the crystallinity index of cellulose analyzed through the high-purity cellulose, hemicellulose, lignin, and silica components.
NMR method and measured according to Eq. (1) was 52.28%. The The XRD patterns indicate that the high crystallinity of cellulose
crystallinity of cellulose measured by the XRD method was a little together with fully amorphous hemicelluloses, lignin, and silica
higher than this when using the NMR method, as reported earlier fractions were analyzed. The morphology of lignocelluloses was
[16]. illustrated through SEM images. A bundle-like structure of cellu-
A wide peak at 2h = 21° indicates the amorphous structure of lose fiber with a diameter of 6.0 mm and smaller particles of hemi-
hemicelluloses. Likewise, the less-ordering nature of lignin was celluloses and lignin with irregular shape were discovered. The
observed with a broad peak in its XRD pattern at 2h = 22.6°. The carbon content of over 90 wt% in all lignocelluloses corresponds
results were the same as those obtained by previous studies [40]. to hydrocarbon in the biomass fractions. The significant differences
The pattern of silica in this study was the same as obtained in in the thermal behavior among the lignocellulosic compounds
our earlier study [41]. A wide peak at 2h = 22° indicates that the were shown through TGA and DSC. Hemicelluloses from RH
obtained silica also had an amorphous structure. degraded in the temperature range of 256–603 °C. While the main
156
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

Fig. 7. (A) SEM micrographs, (B) elemental mapping, and (C) EDS analysis of lignocelluloses.

pyrolysis of cellulose rapidly occurred in the temperature range of characteristic of the main fractions of RH reported in this study
278–508 °C, the pyrolysis of lignin took place over a wider temper- lay a sound foundation and provide a fundamental understanding
ature range, from 255–700 °C. The Tg values of hemicelluloses and for carrying out further studies that aim to apply such bio-
lignin were determined at 136 °C (hemicellulose 1), 158 °C (hemi- materials in the fields of green agriculture, environment, and
cellulose 2), and 94 °C (lignin). The preparation method and the biomaterials.

157
Ngoc Thuy Nguyen, Nhat Thong Tran, Tan Phat Phan et al. Journal of Industrial and Engineering Chemistry 108 (2022) 150–158

Funding sources [17] L. Segal, J.J. Creely, A.E. Martin, C.M. Conrad, Textile Res. J. 29 (1959) 786–794,
https://doi.org/10.1177/004051755902901003.
[18] M. Krishaniaa, V. Kumar, R.S. Sangwan, Bioresource Technol. Rep. 1 (2018) 89–
This research is funded by Vietnam National University HoChi- 93, https://doi.org/10.1016/j.biteb.2018.01.001.
Minh City (VNU-HCM) under grant number C2020-18-23. [19] H. Yang, R. Yan, H. Chen, D.H. Lee, C. Zheng, Fuel 86 (2007) 1781–1788, https://
doi.org/10.1016/j.fuel.2006.12.013.
[20] W. Peng, L. Wang, M. Ohkoshi, M. Zhang, Cell. Chem. Technol. 49 (2015) 757–
Declaration of Competing Interest 764.
[21] F. Xu, J.X. Sun, C.F. Liu, R.C. Sun, Carbohyd. Res. 341 (2006) 253–261, https://
doi.org/10.1016/j.carres.2005.10.019.
The authors declare that they have no known competing finan-
[22] B.L. Xue, J.L. Wen, F. Xu, R.C. Sun, Carbohyd. Res. 352 (2012) 159–165, https://
cial interests or personal relationships that could have appeared doi.org/10.1016/j.carres.2012.02.004.
to influence the work reported in this paper. [23] T. Liang, L. Wang, Bioresources 12 (2017) 1128–1135, https://doi.org/
10.15376/biores.12.1.1128-1135.
[24] J.R. Martinez, F. Ruiz, Y. Vorobiev, F. Pérez-Robles, J. González-Hernández, J.
References Chem. Phys. 109 (1998) 7511–7514, https://doi.org/10.1063/1.477374.
[25] A. Isogai, M. Usuda, T. Kato, T. Uryu, R.H. Atalla, Macromolecules 22 (1989)
[1] FAO Rice Market Monitor, Food and Agriculture Organization of the United 3168–3172, https://doi.org/10.1021/ma00197a045.
Nations – FAO (Trade and Markets Division). http://www.fao.org/economic/ [26] S. Park, D.K. Johnson, C.I. Ishizawa, P.A. Parilla, M.F. Davis, Cellulose 16 (2009)
est/publications/rice-publications/rice-market-monitor-rmm/en/, 2018 641–647, https://doi.org/10.1007/s10570-009-9321-1.
(Accessed 27 April 2018). [27] Y. Lu, Y.C. Lu, H.Q. Hu, F.J. Xie, X.Y. Wei, X. Fan, J. Spectrosc. (2017) 8951658,
[2] J. Chun, Y.M. Gu, J. Hwang, K.K. Oh, J.H. Lee, J. Ind. Eng. Chem. 81 (2020) 135– https://doi.org/10.1155/2017/8951658.
143, https://doi.org/10.1016/j.jiec.2019.08.064. [28] J.L. Wen, S.L. Sun, B.L. Xue, R.C. Sun, Materials 6 (2013) 359–391, https://doi.
[3] Z. Chen, X. Wang, B. Xue, W. Li, Z. Ding, X. Yang, J. Qiu, Z. Wang, Carbon. 161 org/10.3390/ma6010359.
(2020) 432–444, https://doi.org/10.1016/j.carbon.2020.01.088. [29] S.S. Mohtar, T.N.Z.T.M. Busu, A.M.Md Noor, N. Shaari, H. Mat, Carbohyd. Polym.
[4] I.B. Ugheoke, O. Mamat, Maejo Int. J. Sci. Technol. 6 (2012) 430–448, https:// 166 (2017) 291–299, https://doi.org/10.1016/j.carbpol.2017.02.102.
doi.org/10.14456/mijst.2012.31. [30] F. Peng, J.L. Ren, F. Xu, J. Bian, P. Peng, R.C. Sun, J. Agric. Food Chem. 57 (2009)
[5] D.A. Cantero, C. Martínez, M.D. Bermejo, M.J. Cocero, Green. Chem. 17 (2015) 6305–6317, https://doi.org/10.1021/jf900986b.
610–618, https://doi.org/10.1039/C4GC01359J. [31] T.A. Nguyen, T.T. Do, T.D. Nguyen, L.D. Pham, V.D. Nguyen, Carbohyd. Polym.
[6] F.Y. Wang, H.Y. Li, H.M. Liu, Y.L. Liu, Bioresources 10 (2015) 5256–5266, 80 (2010) 308–311, https://doi.org/10.1016/j.carbpol.2009.11.043.
https://doi.org/10.15376/biores.10.3.5256-5266. [32] A. Sharma, T.R. Rao, Bioresource Technol. 67 (1999) 53–59, https://doi.org/
[7] L. Cong, Z. Li, Z. Guanqun, X. Jianguo, Z. Long, Pol. J. Chem. Technol. 17 (2015) 10.1016/S0960-8524(99)00073-5.
89–95, https://doi.org/10.1515/pjct-2015-0035. [33] K.G. Mansaray, A.E. Ghaly, Bioresource Technol. 65 (1998) 13–20, https://doi.
[8] J.H. Lee, J.H. Kwon, J.W. Lee, H. Lee, J.H. Chang, B.I. Sang, J. Ind. Eng. Chem. 50 org/10.1016/S0960-8524(98)00031-5.
(2017) 79–85, https://doi.org/10.1016/j.jiec.2017.01.033. [34] P.M. Stefani, D. Garcia, J. Lopez, A. Jimenez, J. Therm. Anal. Calorim. 81 (2005)
[9] N. Johar, I. Ahmad, A. Dufresne, Ind. Crop. Prod. 37 (2012) 93–99, https://doi. 315–320, https://doi.org/10.1007/s10973-005-0785-4.
org/10.1016/j.indcrop.2011.12.016. [35] K.G. Mansaray, A.E. Ghaly, Energ. Source. 20 (1998) 653–663, https://doi.org/
[10] C.F.L. Silva, M.A. Schirmer, R.N. Maeda, C.A. Barcelos, N. Jr Pereira, J. Biotechnol. 10.1080/00908319808970084.
18 (2015) 10–15, https://doi.org/10.1016/j.ejbt.2014.11.002. [36] N. Worasuwannarak, T. Sonobe, W. Tanthapanichakoon, J. Anal. Appl. Pyrolysis
[11] T. Liou, Mater. Sci. Eng. A 364 (2004) 313–323, https://doi.org/10.1016/j. 78 (2007) 265–271, https://doi.org/10.1016/j.jaap.2006.08.002.
msea.2003.08.045. [37] O. Gordobil, R. Moriana, L. Zhang, J. Labidi, O. Sevastyanova, Ind. Crop. Prod. 83
[12] D. Barana, A. Salanti, M. Orlandi, D.S. Ali, L. Zoia, Ind. Crop. Prod. 86 (2016) 31– (2016) 155–165, https://doi.org/10.1016/j.indcrop.2015.12.048.
39, https://doi.org/10.1016/j.indcrop.2016.03.029. [38] C. Crestini, F. Melone, M. Sette, R. Saladino, Biomacromolecules 12 (2011)
[13] Y.J. Heo, S.J. Park, J. Ind. Eng. Chem. 31 (2015) 330–334, https://doi.org/ 3928–3935, https://doi.org/10.1021/bm200948r.
10.1016/j.jiec.2015.07.006. [39] N. Cachet, S. Camy, B. Benjelloun-Mlayah, J.S. Condoret, M. Delmas, Ind. Crop.
[14] M. Mašura, Wood Sci. Technol. 21 (1987) 89–100, https://doi.org/10.1007/ Prod. 58 (2014) 287–297, https://doi.org/10.1016/j.indcrop.2014.03.039.
BF00349720. [40] S. Kumari, G.S. Chauhan, S. Monga, A. Kaushik, J.H. Ahn, RSC Adv. 6 (2016)
[15] F. Örså, B. Holmbom, J. Thornton, Wood Sci. Technol. 31 (1997) 279–290, 7768–77776, https://doi.org/10.1039/C6RA13308H.
https://doi.org/10.1007/BF00702615. [41] N.T. Nguyen, D.H. Nguyen, D.D. Pham, V.P. Dang, Q.H. Nguyen, D.Q. Hoang,
[16] X. Zhu, Y. Dai, C. Wang, L. Tan, Contributions to Tob. Res. 27 (2016) 126–135, Polym. J. 49 (2017) 861–869, https://doi.org/10.1038/pj.2017.58.
https://doi.org/10.1515/cttr-2016-0014.

158

You might also like