You are on page 1of 10

pubs.acs.

org/IECR Article

Optimization by Response Surface Methodology of Ethanosolv


Lignin Recovery from Coconut Fiber, Oil Palm Mesocarp Fiber, and
Sugarcane Bagasse
Francisco P. Marques, Aldo S. Colares, Maria N. Cavalcante, Jessica S. Almeida, Diego Lomonaco,
Lorena M. A. Silva, Morsyleide de Freitas Rosa, and Renato C. Leitão*
Cite This: Ind. Eng. Chem. Res. 2022, 61, 4058−4067 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The optimization of lignin recovery from coconut


fiber (CF), oil palm mesocarp fiber (OPMF), and sugarcane
Downloaded via CURTIN UNIV on March 24, 2022 at 12:08:12 (UTC).

bagasse (SCB) using ethanol organosolv was investigated as well as


the resulting lignin characteristics. Ethanol organosolv lignin
recovery was successfully optimized, and the lignin yield (YL)
was maximized. Ethanol concentration and temperature higher
than 74% and 187 °C maximized the YL, playing a significant role
in lignin yield. YL was higher for SCB than for other biomasses.
High-purity and high-quality lignins were obtained with slightly
different structural characteristics among the raw materials. The
obtained lignins were the p-hydroxyphenyl−guaiacyl−syringyl type
with a purity higher than 86%. SCB lignin seemed to be a better
choice than CF and OPMF lignins for application in aromatic polymerization, because of its higher lignin yield, higher purity, higher
thermal stability, and more reactive sites.

1. INTRODUCTION solution is the use of organosolv lignin extraction, which is a


The substitution of non-renewable resources, especially greener process for obtaining high-yield and high-quality
products derived from the petrochemical industry, has lignin, showing interesting characteristics such as a lower
increased the interest in the utilization of lignocellulosic carbohydrate content, a lower molecular weight, and higher
biomass. Lignocellulosic biomass is currently used as an energy reactive sites, among others.7−10 The organosolv process is
source in industrial ovens and boilers. This represents based on the solubilization of hemicellulose and lignin from
underutilization or a low rate of return on biomass use. This biomass using recoverable organic solvents, such as methanol,
is because these types of biomasses are rich in macromolecules ethanol, acetic acid, acetone, or mixtures of an organic solvent
with potential industrial interest, particularly lignin. Lignin is a and water in the temperature range of 100−250 °C for several
phenol-rich biomacromolecule that can be used in the minutes or hours.11 The literature reports that the organosolv
development of several technologies and high-value-added extraction efficiency is a function of solvent type and
products such as phenolic-based resins and adhesives, concentration, reaction temperature and time, catalyst, and
polyurethane foams, hydrogels, and anti-UV and antimicrobial biomass recalcitrance.10,12 The resulting lignin characteristics
additives, among others.1−3 depend on its chemical structure and properties, which in turn
Lignins are complex amorphous aromatic polymers with depend on the biomass source. 13 Therefore, process
three-dimensional cross-linked structures, mainly composed of optimization is required to achieve the highest lignin recovery
three phenylpropane units: p-coumaryl alcohol (H), coniferyl after precipitation.
alcohol (G), and sinapyl alcohol (S).4 Lignins can be classified Process optimization aims to determine the best reaction
based on their abundance of these basic units as type-G conditions for maximum lignin yield. In addition, it can give a
(softwood lignin), type-GS (hardwood lignin), type-H-G-S
(grass lignin), and type-H-G (compression wood lignin).5 It is
estimated that the annual production of lignin varies in the Received: November 4, 2021
range of 5−36 × 108 tons.6 Revised: February 7, 2022
Lignin is usually obtained as a byproduct of the pulp and Accepted: February 16, 2022
paper industry: kraft lignin, soda lignin, and lignosulfonate. Published: February 28, 2022
The process for obtaining these lignins often involves several
chemicals with potential environmental hazards. A possible

© 2022 American Chemical Society https://doi.org/10.1021/acs.iecr.1c04362


4058 Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

lignin extraction pattern behavior in relation to those variables, material was filtered using a Büchner funnel with a 28 μm filter
which makes it possible to manipulate the structure of the paper to separate the hydrolyzed material from black liquor
lignin depending on the conditions employed. Process (BL). The hydrolyzed solid material was washed with a
optimization can be conducted using the central composite twofold warm (60 °C) ethanol solution at the same
design (CCD), allied to the response surface methodology concentration used in the extraction process to extract the
(RSM), which allows modeling several variables in the same remaining lignin. To precipitate the lignin, the BL was diluted
experiment, saving experiment time and consumables with a to three volumes of distilled water and left to rest for 24 h. The
high statistical power. precipitated lignin was filtered using a paper filter with an 8 μm
In the past few years, our group has been working on the pore, washed with distilled water until constant pH, and dried
improvement of biopolymer extraction and characterization in an air-circulating oven at 50 °C until a constant weight was
from important worldwide biomasses such as coconut fiber achieved. The lignin yield was calculated using eq 1.

ij m yz
(CF), oil palm mesocarp fiber (OPMF), and sugarcane bagasse
YL = jjj L zzz × 100
j mIL z
(SCB), aiming at sustainable approaches and biomass
k {
valorization.10,14−17 These biomasses play a major role
(1)
worldwide, mostly in Brazil, Malaysia, India, China, Indonesia,
and the Philippines, due to their abundance, underutilization, where YL is the yield of lignin extracted on a dry basis (%), mL
and being a renewable source of biopolymers of industrial is the obtained recovered lignin after precipitation on a dry
interest, such as cellulose nanocrystals and lignin. There are no basis (g), and mIL is the initial mass (dry basis) of lignin in the
reports in the literature that aimed to evaluate the best fibers before the process (g). The mass of lignin was obtained
conditions for ethanosolv lignin extraction and the character- using the Klason lignin method (Section 2.4.1).
istics of lignin obtained from CF, OPMF, and SCB, which 2.3. Experimental Design and Statistical Analysis.
presents an opportunity to promote biomass valorization. Temperature (T), reaction time (t), and ethanol concentration
In this study, the organosolv process using ethanol as a (C) were optimized by CCD 23. The CCD comprised eight
solvent (hereafter referred to as ethanosolv) was optimized as factorial points, six axial points, and five center points, resulting
an extraction method for lignin from CF, OPMF, and SCB, in a total of 19 treatment points. RSM was applied to
three important Brazilian lignocellulosic biomasses. The lignin determine the effects of T, t, and C on Y for CF, OPMF, and
characteristics under each optimum condition were evaluated SCB.
by thermal behavior, chemical structure, and composition Analysis of variance (ANOVA) was used to determine the
using Klason lignin (purity), two-dimensional heteronuclear effect of independent variables (T, t, and C) on the dependent
single quantum coherence spectroscopy (2D-HSQC), Fourier variable Y (Tables S3−S5Supporting Information). The p-
transform infrared spectroscopy (FTIR), gel permeation values and lack of fit (LoF) as well as the coefficients of
chromatography (GPC), thermogravimetry analysis (TGA), determination (R2) were used to evaluate the suitability of the
and differential scanning calorimetry (DSC), in order to assess models. The models were validated by comparing the
their use as building blocks or additives for high-value-added experimental results obtained under optimal conditions and
products, such as resins, adhesives, low-molecular-weight the predicted values obtained by the models. Statistical analysis
phenolic compounds, and multifunctional hydrocarbons. of the data and response surface plots was performed using
Design Expert v. 11.
2. EXPERIMENTAL SECTION The conditions were as follows: temperature 156−224 °C;
2.1. Biomasses. In this study, CF, OPMF, and SCB were reaction time 26−94 min; and ethanol concentration, 35−85%
used as lignocellulosic biomasses. The CF, OPMF, and SCB (v/v). The catalyst concentration (H2SO4 0.5%, v/v) was
used were supplied by Embrapa Agroindústria Tropical constant, as was the ratio of fiber to ethanol solution (1:10, w/
(Fortaleza, Brazil), Embrapa Amazônia Oriental (Pará, Brazil), v) in all experiments. The complete experimental matrix is
and DIAGEO (Paraipaba, Brazil), respectively. The biomass presented in Table S2.
was dried in an air-circulating oven at 50 °C until constant 2.4. Lignin Characterization. 2.4.1. Lignin Structure and
weight and milled in a pilot mill (Fritsch Pulverisette 19 mill) Composition. The purity of the lignin samples was calculated
with a 0.5 mm sieve. All solvents and reagents were of as the sum of the acid-insoluble lignin (Klason lignin) and
analytical grade and were used as received without any acid-soluble lignin.19 A lignin sample (1.0 g) was mixed with
modification. The cellulose and lignin contents of the 17 mL of H2SO4 (72 wt %), ground in a mortar for 15 min at
biomasses were analyzed according to TAPPI standard room temperature, and kept at rest for 24 h. The solution was
procedures (T203 cm-99 and T222 om-2). The holocellulose diluted with deionized water until 4 wt % H2SO4 was added
and hemicellulose contents were determined by the procedure and refluxed for 4 h. Then, the system was cooled for 30 min at
described by Yokohama et al.18 The lignin concentrations in room temperature. The acid-insoluble lignin was obtained by
CF, OPMF, and SCB were 32.2, 28.4, and 24.1%, respectively. vacuum filtration using a #4 sintered glass funnel and washed
The main compositions of the studied biomasses are presented with deionized water until neutrality was achieved. The acid-
in Table S1 (Supporting Information). insoluble lignin was dried in an air-circulation oven at 50 °C
2.2. Ethanosolv Process. Lignin extraction was performed for 24 h. The acid-soluble lignin was calculated based on the
in a mini high-pressure reactor (Berghof Highpreactor BR- UV absorbance at 215 and 280 nm of the diluted filtrate (4 wt
300), with a total volume of 500 mL, without external % H2SO4, 1:10 by volume). All experiments were performed in
agitation. First, 20 g (dry weight) of fiber was added to 200 mL triplicate.
of ethanol solution (in the studied range concentrations, Table 2D-HSQC spectra were obtained using a 600 MHz DD2
S2Supporting Information) with sulfuric acid (0.5%, v/v) as NMR spectrometer (Agilent) equipped with a 5 mm One
the catalyst. After the designed reaction time, the reactor was Probe (2H−19F/15N−31P) with a z-axis field gradient.15 The
cooled to ∼25 °C and opened. The obtained hydrolyzed sample powder (50 mg) was dissolved in 500 μL of DMSO-d6
4059 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 1. Effects of process variables on the extraction yield of CFL, OPL, and SCL. T (temperature), t (reaction time), and C (ethanol
concentration).

with the aid of an ultrasonic bath for 24 h to promote better arithmetic average of 32 scans. FTIR data were processed and
solubilization. The DMSO solvent peak at δC/δH 39.5/2.49 analyzed using the PerkinElmer Spectrum IR software.
was utilized as an internal reference. The relative abundances GPC analyses were performed using a Shimadzu LC-20AD
of the lignin units were calculated based on the volume (Kyoto, Japan) at 40 °C using a setup comprising two
integration of the cross-peaks of each structure and expressed analytical GPC columns in series (Phenogel 5 μ 50 Å and
as a percentage of the total summation with p-hydroxyphenyl Phenogel 5 μ 103 Å, 4.6 mm × 300 mm, Phenomenex).21
(H), guaiacyl (G), and syringyl (S) units equal to 100%, as HPLC-grade THF was used as the mobile phase at a flow rate
previously reported.20 Lignin monomers in the oxidized forms, of 0.35 mL min−1. Lignin samples (2 mg) were dissolved in 2
such as ferulate (FA) and p-coumarate (pCA), were included mL of HPLC-grade THF and then filtered using a 0.22 μm
as part of each related monomer. Similarly, the relative PTFE filter. The filtered solution (20 μL) was injected into the
contents of the lignin linkages were obtained. Data analysis was GPC system at a flow rate of 0.35 mL min−1 and monitored
performed using MestReNova v. 12.0. using a UV−vis detector (Shimadzu SPD-M20A) at 280 nm.
FTIR spectra were obtained using a PerkinElmer Spectrum Standard calibration was performed using polystyrene stand-
Two FT-IR spectrometer using KBr pellets with a sample ards PSS (Mw range 162−1.3 × 105 g mol−1).
concentration of 5 wt %.15 The spectra were recorded between 2.4.2. Thermal Behavior of Lignins. TGA and derivative
4000 and 400 cm−1 with a resolution of 4 cm−1 using an thermogravimetry (dTG) analyses were performed using an
4060 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

STA 6000 thermal analyzer (PerkinElmer).21 All measure- form of heat to break the compact biomass structure and
ments were performed under a N2 atmosphere with a gas flow achieve maximum lignin recovery.
of 50 mL min−1 and a heating rate of 10 °C min−1 and heated Variable C showed significant positive linear and negative
from 30 to 900 °C. The mass of the sample was approximately quadratic effects for all biomass types. The variable C had the
10 mg. highest effect compared to the other variables, which indicates
DSC analyses were performed using a TA Instruments that the lignin yield is highly dependent on the ethanol
calorimetry model Q20.21 Lignin (5−7 mg) was first submitted concentration in the studied range. Moreover, the negative
to an annealing program (from 25 to 90 °C for 10 min and quadratic effect indicates that there is an optimum maximum
cooled from 90 to 0 °C for 3 min). Then, the sample was ethanol concentration within the studied range. Increasing the
heated from 0 to 200 °C under a N2 atmosphere at a flow rate ethanol concentration to a range of 60−80% led to an increase
of 50 mL min−1 and a heating rate of 10 °C min−1. Hermetic in lignin yield, as expected.9 High ethanol content solutions
closed aluminum pans were used in the experiment. favor the delignification due to the polar interactions that
ethanol provides, leading to a better lignin solubilization.24
3. RESULTS AND DISCUSSION Variable t had no significant effect on the response of all
3.1. Ethanosolv Process Optimization. The results of biomasses. The studied range of reaction times may have
the lignin extraction yield (YL) from CF, OPMF, and SCB hindered the detection of the reaction time effect, that is, the
according to the experimental matrix are summarized in Table applied reaction time range resulted in small changes in the
S2. ANOVA was used to examine the significance of the lignin yield.
mathematical models that were obtained from all biomasses The models showed similarities between themselves. First,
(Tables S3−S5). Table S6 (Supporting Information) summa- the ethanol concentration played a greater positive role than
rizes the effects of the variables on lignin yield. temperature and reaction time on the lignin yield because it
The coefficients of determination (R2) for CF, OPMF, and had a higher significance (lowest p-value) among the
SCB were higher than 0.95, with a confidence interval of 95% independent variables (data available in the Supporting
(α = 0.05), which is an indication of highly significant Information). This was also found in a previous study using
relationships between experimental values and values predicted corncob as a source of lignin.25 The second similarity is that
by the models. Model p-values were less than 0.0001, showing the reaction time did not affect the lignin yield within the
that the lignin yield models are significant at a confidence studied range. Hence, the shortest reaction time may be used
interval of 95% and satisfied by eqs 2−4. within this experimental domain. These behaviors were also
The LoF had no significant effect on the model. Therefore, observed by Pan et al. (2007),23 who found that increasing the
there was no evidence of a lack of adjustment in the obtained reaction temperature and ethanol concentration in lignin
models for all the fibers. Because the models were significant, it extraction from lodgepole pine caused higher lignin yields.
was possible to build response surfaces. Figure 1 shows the However, the reaction time did not affect the delignification
response surface plots of the lignin yields for the biomasses as a process. The same behavior was observed in a previous study
function of two process variables. The other variable was kept of hybrid poplar.22
constant at the central level. Delignification during ethanol organosolv is a combination
of the depolymerization and solubilization of lignin. Lower
YCF = −156.41162 + 1.22087T + 0.45469t + 2.31851C ethanol concentrations promote high acid-catalyzed cleavage of
− 0.00508T 2 − 0.00313t 2 − 0.02610C 2 ether bonds in the lignin structure due to higher hydrogen ion
concentrations, whereas high ethanol concentrations increase
+ 0.00030Tt + 0.00906TC − 0.00357tC (2)
lignin solubilization. Therefore, a balance between the
ethanol−water ratio may exist for lignin extraction, according
YOPMF = −321.02773 + 1.45026T + 0.49432t to Monteil-Rivera et al. (2012).26 Later, the authors reported
that the lignin yield increases with the ethanol concentration,
+ 5.88651C − 0.00304T 2 − 0.00318t 2 catalyst concentration, and temperature. These findings are
− 0.03738C 2 − 0.00044Tt − 0.00131TC similar to this study.
− 0.00007tC (3) Pinheiro et al. (2017)14 investigated lignin extraction from
SCB using acetic acid organosolv and found reaction time and
temperature effects similar to this work. However, because
YSCB = −289.77629 + 3.64991T − 0.37885t − 0.06210C lignin is more soluble in acetic acid than ethanol, a lower
− 0.01283T 2 − 0.00314t 2 − 0.01670C 2 temperature is necessary to maximize the lignin extraction. Xu
et al. (2006)27 studied several types of organic solvents and
+ 0.00310Tt + 0.01640TC + 0.00083tC (4)
their proportions in organosolv pulping at 85 °C for 4 h.
Variable T had a significant positive linear effect on Y for the Aqueous organic acid was more effective for delignification of
OPMF. This means that the lignin yield increased when the wheat straw than aqueous organic alcohol under the conditions
temperature increased within the studied range. In contrast, T used. In contrast, organic acids are more toxic and corrosive
had a significant negative linear effect on Y for CF and a than ethanol, making ethanol more suitable for this type of
negative quadratic effect for SCB. For CF and SCB, the lignin extraction. As demonstrated above, the three biomasses led to
yield increases at temperatures below 200 °C, which is in three different extraction conditions, indicating that the
accordance with previous optimization studies for ethanol delignification process occurs differently according to the
organosolv.22,23 This result may be related to the lower biomass. Nevertheless, it was demonstrated here that the
recalcitrance of CF and SCB owing to the lower lignin content. delignification process is strongly dependent on the ethanol
A low recalcitrance biomass may require less energy in the concentration and temperature.
4061 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 1. Optimal Condition, Predicted Yield (YP), Observed Yields (YO), and Purity for the Lignin Obtained from the
Biomassesa
condition
biomass T (°C) T (min) C (%) YP (%) YO (%)* P (%)*
CF 187 26 74 52.5 51.6 ± 3.2 86.3 ± 3.1
OPMF 218 26 75 68.7 73.9 ± 2.2 88.7 ± 1.5
SCB 200 26 85 85.7 83.4 ± 1.7 97.3 ± 4.9
a
Note: *Values are the mean of triplicate measurements ± standard deviation.

Figure 2. Aromatic region of the 2D-HSQC spectra of CFL, OPL, and SCL.

The models indicate a set of conditions that maximize the Table 2. Relative Content of Lignin Monomers and
response, also known as critical values (Table 1). In summary, Linkages Based on the Integration of 2D-HSQC Cross-
the observed values under the optimal conditions were very Peaks of CFL, OPL, and SCL
close to the predicted values, which shows the great capability
FA pCA Aγ A′γ A′α
of RSM to predict lignin yield in the studied range for all sample H (%) G (%) S (%) (%) (%) (%) (%) (%)
biomasses. The higher lignin yield obtained here indicates that CFL 22.0 59.5 18.5 3.1 11.9 11.7 64.8 23.4
ethanol organosolv leads to a greater cleavage of ether bonds OPL 20.1 55.9 24.0 3.0 12.3 18.1 70.1 11.8
compared to other processes. The higher the cleavage of ether SCL 25.9 61.1 12.9 31.5 57.1 11.4
bonds in the lignin substructures, the higher the dissolution of
lignin fragments.16 Accordingly, recovered optimized lignins
indicate the presence of carbohydrates linked to the lignin
from CF, OPMF, and SCB were designated as coconut fiber
structure. It was not possible to detect oxidized structures to
lignin (CFL), oil palm mesocarp fiber lignin (OPL), and SCL, indicating that SCL has a high purity among the samples
sugarcane bagasse lignin (SCL). according Table 2.
3.2. Lignin Characteristics. 3.2.1. Lignin Structure and Lignins extracted from CF, OPMF, and SCB were classified
Composition. The structural elucidation of lignins was as type p-hydroxyphenyl−guaiacyl−syringyl (HGS) owing to
evaluated by 2D-HSQC spectroscopy. Structural information the presence of the three basic units. The lignin monomers in
for the aromatic region (δC/δH 150−100/8.0−6.0) is provided this type of lignin usually vary in the range of 25−50% S units,
in Figure 2. Table 2 shows the relative contents of lignin 25−50% G units, and 10−25% H units.5 A high content of H
monomers (H, G, and S). The 13C−1H signal correlations of units and low content of S units are the best structural
the 2D-HSQC spectra were assigned according to previous characteristics for lignin application as building blocks in resins
studies.10,20 and adhesives. According to El Mansouri and Salvadó, H and
All extracted lignins were composed of similar monomer G units have unsubstituted C-3 and C-5 positions in the
units, composed of syringyl (S), ferulate (FA), guaiacyl (G), p- aromatic rings of phenolic aryl propane units, which are
hydroxyphenyl (H), and p-coumarate (pCA) units in different important reactive sites for the preparation of lignin-based
proportions, according to the biomass source (Table 2). The phenol−formaldehyde (e.g., adhesives, chelating resins, and
detection of oxidized monolignols, such as FA and pCA, could ion exchange resin) under alkaline conditions.28
4062 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 3. Oxygenated aliphatic regions of the 2D-HSQC spectra of CFL, OPL, and SCL.

Regarding the oxygenated aliphatic region of the 2D-HSQC the reference band at 1513 cm−1, which was attributed to the
(Figure 3), it was observed that all lignins were composed of aromatic skeleton of the lignin. Baselines were corrected at the
similar linkages, such as Aγ (non-acetylated β-aryl ether), A′γ same points in all spectra.
(acetylated β-aryl ether), and Aα (ethoxylated β-aryl ether). Stretching vibrations of the aliphatic and phenolic −OH
Table 2 shows the relative content of the linkages. The groups in hydrogen bonding were observed in the spectral
presence of Aα suggested that the ethanol organosolv may range of 3700−3000 cm−1. The aliphatic CH groups from the
insert ethoxyl groups in the aliphatic portion of the lignins. methyl and methylene stretches appeared at 2925 and 2853
This behavior was observed in other studies.29,30 Under acid cm−1, respectively. In the OPL, these bands were more
conditions and high temperatures, a benzylic cation prominent than CFL and SCL, which is related to the presence
intermediate forms, which could lead to oxidation (Hibbert of residual oil remaining during the industrial oil extraction
ketones) or a condensation reaction in the presence of elevated process.10
water contents. When the ethanol content is higher, The band at around 1710 cm−1 is attributed to the CO
ethoxylation could occur at the carbon-α of the lignin aliphatic stretch related to residual carbohydrates, which is indicative of
portion, suppressing degradation reactions.29 lignin−carbohydrate complexes that remain in the lignin linked
The functional group characteristics of the lignins and the by ester bonds.21 This band is higher in SCL, associated with
structural differences among the biomasses were assessed using the high amount of carbohydrates in SCB, which was not
FTIR spectroscopy. Figure 4 shows the FTIR spectra of CFL, hydrolyzed despite the high temperature of the extraction.
OPL, and SCL. The spectra were standardized according to However, SCL showed a higher purity (Table 1) among the
lignins, which could be associated with the presence of some
pseudo-lignin. Pseudo-lignin is a lignin-like material that is
formed under high severity low-pH biomass pretreatments,
being a combination of carbohydrate and lignin degradation
products.31 Pseudo-lignin differs mainly from real lignin due to
the presence of condensed structures derived from furfural and
5-hydroxymethylfurfural. In addition, a characteristic carbohy-
drate band was found at 1041 cm−1, assigned to C−O and C−
C stretching vibrations in polysaccharides.32
Strong absorptions related to aromatic skeletal vibrations
from lignin were observed at 1609, 1513, 1462, and 1426 cm−1.
Additionally, all three types of monomers were present in the
lignin spectra: p-hydroxyphenyl (H), guaiacyl (G), and syringyl
(S) units at 1116 cm−1 (typical GS lignin) combined with
1164 cm−1 (typical HGS lignin) and vibrations at 1270 and
1217 cm−1 attributed to guaiacyl units as well as C−C and C
O stretches. This suggests that all lignins belong to the HGS
type,13 which was confirmed by 2D-HSQC analysis.
GPC curves were used to evaluate the molecular weight
distributions of the lignins. Table 3 summarizes the molecular
Figure 4. FTIR spectra of CFL, OPL, and SCL. weight data obtained from Figure 5. All lignins presented a
4063 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 3. Apparent M̅ w (Weight Average), M̅ n (Number


Average), and Polydispersity (M̅ w/M̅ n) CFL, OPL, and SCL
sample M̅ n (g·mol−1) M̅ w (g·mol−1) PDI
CFL 879 1535 1.75
OPL 734 1233 1.68
SCL 771 1357 1.76

Figure 6. TGA curves and dTG of CFL, OPL, and SCL.

Figure 5. GPC curves of CFL, OPL, and SCL.

similar molecular weight and low polydispersity index, as


expected for organosolv lignins.11 Organosolv lignins usually
have lower molecular weights than other lignin types (i.e., kraft
or lignosulfonate), which could be related to lower steric
hindrance and a higher reactivity.
Previous studies using the acetosolv method reported higher
M̅ w for lignins from CF, OPMF, and SCB.14,15 This may Figure 7. DSC curves of CFL, OPL, and SCL.
indicate that ethanosolv is a better method for breaking the β-
O-4 linkages in the lignin structure. CFL showed the highest
molecular weight among the lignins, which could be related to Tonset and Tmax, probably related to the high content of
the lower severity of optimized ethanosolv extraction since impurities present in the OPMF lignins, such as residual oil. A
organosolv extensively cleaved some inter-unit bonds in lignin, previous study showed higher Tonset values obtained from
which is a function of the treatment time, temperature, and dewaxed organosolv lignin from OPMF.33 The absence of
acidity of the medium.8 residual oil can improve the thermal properties of the OPL.
3.2.2. Lignins’ Thermal Behavior. The thermal properties of However, SCL showed the best thermal properties with the
the lignins CFL, OPL, and SCL were evaluated through the highest Tonset, Tmax, and CY related to high purity, as
analysis of the thermal stability by TGA/dTG analyses and mentioned in Table 1.
state transition analysis by DSC. Table 4 summarizes the CY is related to the potential of a material for flame
retardancy. The relation is expressed as a percentage of the
limiting oxygen index (LOI) using the following equation:34
Table 4. Onset Degradation Temperature (Tonset),
LOI = 17.5 + 0.4CY. A CY higher than 26.3% (LOI > 28%)
Maximum Degradation Temperature (Tmax), CY, LOI, and
indicates that the polymer is self-extinguishing. All lignins
Glass Transition Temperatures (Tg) of the Studied Lignins
showed CY values higher than 26.3%, indicating their potential
sample Tonset (°C) Tmax (°C) CY (%) LOI (%) Tg (°C) to be applied in a material with a self-extinguishing
CFL 174.8 359.2 35.0 31.5 130.7 characteristic. In addition, the use of lignin as an additive in
OPL 172.9 227.3 36.9 32.3 136.9 polymeric materials can improve the thermal behavior, such as
SCL 234.4 383.0 45.3 35.6 140.1 self-extinguishing and fire-retardant properties.35−37
Amorphous polymers such as lignin undergo a state
transition called glass transition (Tg), which is a transition
obtained thermal data, such as the onset degradation from a glass-like rigid solid to a more flexible, rubbery
temperature (Tonset), maximum degradation temperature compound. Tg is associated with material processability and
(Tmax), char yield (CY), and glass transition temperature (Tg). indicates the start of molecular chain mobility when sufficient
Figures 6 and 7 show the TGA, dTG, and DSC curves, energy is supplied in the form of heat.14
respectively. As shown in Figure 6, all lignins presented similar Several factors affect the Tg of lignin factors, such as the
thermal profiles with some differences. OPL showed the lowest presence of low-molecular-weight contaminants (including
4064 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

water and solvents), extraction method, chemical modification, presence of more reactive sites for polymerization in the
molecular weight, thermal history, and crosslinking.25,38 Tg is aromatic ring. However, CF and OPMF cannot be set aside as
also related to the polymer molecular weight, which means that a source of this macromolecule because the lignins may require
a high molecular weight is usually associated with a high Tg. some purification.
In addition, a high glass transition temperature is associated
with strong bonding, probably due to hydrogen interactions,
which hinder the mobility of the lignin chains. In Table 4, the
■ ASSOCIATED CONTENT
* Supporting Information

lignins showed similar Tg values in a narrow temperature range The Supporting Information is available free of charge at
of 130−140 °C. This indicates that the obtained lignins have https://pubs.acs.org/doi/10.1021/acs.iecr.1c04362.
strong inter- and intramolecular bonds. The obtained Tg values
are in accordance with the literature, which shows a range from Main composition of the studied biomasses; independ-
100 to 150 °C for organosolv lignins.21,33,39,40 ent and dependent variables of the experimental design
3.2.3. Final Discussion. The process of lignin extraction and and yield of ethanol organosolv lignin for CF, OPMF,
characteristics, from the extraction process to compositional and SCB; ANOVA for CF, OPMF, and SCL
features and thermal behavior, have been elucidated in experimental designs; and the effect of variables on the
previous sections. Regarding the extraction process, SCB was increase of lignin yield from biomasses (PDF)


highlighted because of its higher lignin recovery (83%) and the
production of high-purity lignin (97%). Considering the AUTHOR INFORMATION
potential use of lignin for phenolic polymerization reactions,
the presence of abundant reactive sites in the aromatic ring is Corresponding Author
one of the most important features. Among the studied lignins, Renato C. Leitão − Embrapa Agroindústria Tropical, 60511-
SCL showed the highest content of reactive sites, H and G 110 Fortaleza-CE, Brazil; orcid.org/0000-0002-9969-
units, and the lowest content of unreactive sites, S unit. This 3059; Phone: +55 85 33917343; Email: renato.leitao@
makes lignin more suitable for application in resins and embrapa.br
adhesives. However, CFL and OPL also showed interesting Authors
monomeric composition, with over 75% of H or G units.
Francisco P. Marques − Departament of Organic and
Another important characteristic of aromatic polymerization
Inorganic Chemistry, Federal University of Ceará, 60440-900
is its molecular weight. All studied lignins showed lower
Fortaleza-CE, Brazil; orcid.org/0000-0002-5598-433X
molecular weights, which may indicate lower steric hindrance
Aldo S. Colares − Embrapa Agroindústria Tropical, 60511-
around the phenolic reactive sites compared to other technical
110 Fortaleza-CE, Brazil; orcid.org/0000-0002-5076-
lignins.11 The lower steric hindrance of the lignin chains
0548
improves the contact between the reactants and the reactive
Maria N. Cavalcante − Embrapa Agroindústria Tropical,
sites, which leads to a better reactivity, making them more
60511-110 Fortaleza-CE, Brazil; orcid.org/0000-0002-
suitable for polymerization reactions.28 These characteristics
7198-1497
are important attributes for polymerization in lignin-based
Jessica S. Almeida − Chemical Engineering Department,
phenol−formaldehyde, such as adhesives, chelating resins, and
Federal University of Ceará, 60455-760 Fortaleza-CE,
ion exchange resins. This was clearly better in the SCL than in
Brazil; orcid.org/0000-0002-1482-4452
the other lignin types. Moreover, all lignins showed high values
Diego Lomonaco − Departament of Organic and Inorganic
of Tonset and LOI, indicating their potential for application in
Chemistry, Federal University of Ceará, 60440-900
polymeric blends with high thermal degradation, which could
Fortaleza-CE, Brazil; orcid.org/0000-0001-5763-4336
also provide self-extinguishing and fire-retardant characteristics
Lorena M. A. Silva − Embrapa Agroindústria Tropical,
for these materials.
60511-110 Fortaleza-CE, Brazil; orcid.org/0000-0002-
6290-8933
4. CONCLUSIONS
Morsyleide de Freitas Rosa − Embrapa Agroindústria
The ethanol organosolv method was successfully optimized Tropical, 60511-110 Fortaleza-CE, Brazil; orcid.org/
using a CCD based on RSM. The ethanol concentration and 0000-0002-5224-9778
temperature played a significant role in the lignin yield.
However, reaction times within the range of 26−94 min did Complete contact information is available at:
not influence the lignin yield. A reaction time of 26 min was https://pubs.acs.org/10.1021/acs.iecr.1c04362
sufficient to maximize the lignin yield. In addition, temperature
Notes
and ethanol concentrations higher than 187 °C and 74%,
The authors declare no competing financial interest.


respectively, were required to maximize the lignin yield among
the biomasses. The lignin yield varied according to the type of
biomass: 51.6% for CF, 73.9% for OPMF, and 83.4% for SCB. ACKNOWLEDGMENTS
The obtained models were significant with p-values less than The authors thank the Ceará State Foundation for the Support
0.0001, and there was no evidence of a lack of adjustment in of Scientific and Technological Development (FUNCAP) for
the models. High-purity and high-quality lignins were obtained the scholarships granted to F.P.M., grant BMD-0008-
from optimized ethanol organosolv. These lignins had slightly 00622.01.12/16. This research was funded by the Brazilian
different structural characteristics, mainly because of the raw Agricultural Research Corporation (EMBRAPA) under grant
materials from which they were extracted. SCB lignin seems to 10.19.00.173.00.00 and the Brazilian National Council for
be a better choice than CF lignin and OPMF lignin for Scientific and Technological Development (CNPq) under
application in material science because of its higher lignin grants 443675/2014-7, 308807/2017-0, 409765/2018-0,
recovery, higher purity, higher thermal stability, and the 407291/2018-0.
4065 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research


pubs.acs.org/IECR Article

REFERENCES Morphology of Woody Plants. J. Agric. Food Chem. 2002, 50,


1040−1044.
(1) Meng, Y.; Lu, J.; Cheng, Y.; Li, Q.; Wang, H. Lignin-Based (19) TAPPI. T 222 Om-02: Acid-Insoluble Lignin in Wood and Pulp;
Hydrogels: A Review of Preparation, Properties, and Application. Int. TAPPI: Atlanta, GA, USA, 2002; pp 1−5.
J. Biol. Macromol. 2019, 135, 1006−1019. (20) Wen, J.-L.; Sun, S.-L.; Xue, B.-L.; Sun, R.-C. Recent Advances
(2) Jędrzejczak, P.; Collins, M. N.; Jesionowski, T.; Klapiszewski, Ł. in Characterization of Lignin Polymer by Solution-State Nuclear
The Role of Lignin and Lignin-Based Materials in Sustainable Magnetic Resonance (NMR) Methodology. Materials 2013, 6, 359−
Construction − A Comprehensive Review. Int. J. Biol. Macromol. 391.
2021, 187, 624−650. (21) Avelino, F.; Marques, F.; Soares, A. K. L.; Silva, K. T.; Leitão, R.
(3) de Sousa Nascimento, L.; da Mata Vieira, F. I. D.; Horácio, V.; C.; Mazzetto, S. E.; Lomonaco, D. Microwave-Assisted Organosolv
Marques, F. P.; Rosa, M. F.; Souza, S. A.; de Freitas, R. M.; Uchoa, D. Delignification: A Potential Eco-Designed Process for Scalable
E. A.; Mazzeto, S. E.; Lomonaco, D.; Avelino, F. Tailored Organosolv Valorization of Agroindustrial Wastes. Ind. Eng. Chem. Res. 2019,
Banana Peels Lignins: Improved Thermal, Antioxidant and Anti- 58, 10698−10706.
microbial Performances by Controlling Process Parameters. Int. J. (22) Pan, X.; Gilkes, N.; Kadla, J.; Pye, K.; Saka, S.; Gregg, D.;
Biol. Macromol. 2021, 181, 241−252. Ehara, K.; Xie, D.; Lam, D.; Saddler, J. Bioconversion of Hybrid
(4) Li, H.; Liang, Y.; Li, P.; He, C. Conversion of Biomass Lignin to Poplar to Ethanol and Co-Products Using an Organosolv Fractiona-
High-Value Polyurethane: A Review. J. Bioresour. Bioprod. 2020, 5, tion Process: Optimization of Process Yields. Biotechnol. Bioeng. 2006,
163−179. 94, 851−861.
(5) Li, C.; Zhao, X.; Wang, A.; Huber, G. W.; Zhang, T. Catalytic (23) Pan, X.; Xie, D.; Yu, R. W.; Lam, D.; Saddler, J. N.
Transformation of Lignin for the Production of Chemicals and Fuels. Pretreatment of Lodgepole Pine Killed by Mountain Pine Beetle
Chem. Rev. 2015, 115, 11559−11624. Using the Ethanol Organosolv Process: Fractionation and Process
(6) Bajwa, D. S.; Pourhashem, G.; Ullah, A. H.; Bajwa, S. G. A Optimization. Ind. Eng. Chem. Res. 2007, 46, 2609−2617.
Concise Review of Current Lignin Production, Applications, Products (24) Novo, L. P.; Curvelo, A. A. S. Hansen Solubility Parameters: A
and Their Environment Impact. Ind. Crops Prod. 2019, 139, 111526. Tool for Solvent Selection for Organosolv Delignification. Ind. Eng.
(7) Brosse, N.; Sannigrahi, P.; Ragauskas, A. Pretreatment of Chem. Res. 2019, 58, 14520−14527.
Miscanthus x Giganteus Using the Ethanol Organosolv Process for (25) Michelin, M.; Liebentritt, S.; Vicente, A. A.; Teixeira, J. A.
Ethanol Production. Ind. Eng. Chem. Res. 2009, 48, 8328−8334. Lignin from an Integrated Process Consisting of Liquid Hot Water
(8) El Hage, R.; Brosse, N.; Sannigrahi, P.; Ragauskas, A. Effects of and Ethanol Organosolv: Physicochemical and Antioxidant Proper-
Process Severity on the Chemical Structure of Miscanthus Ethanol ties. Int. J. Biol. Macromol. 2018, 120, 159−169.
Organosolv Lignin. Polym. Degrad. Stab. 2010, 95, 997−1003. (26) Monteil-Rivera, F.; Huang, G. H.; Paquet, L.; Deschamps, S.;
(9) Wei Kit Chin, D.; Lim, S.; Pang, Y. L.; Lam, M. K. Fundamental Beaulieu, C.; Hawari, J. Microwave-Assisted Extraction of Lignin from
Review of Organosolv Pretreatment and Its Challenges in Emerging Triticale Straw: Optimization and Microwave Effects. Bioresour.
Consolidated Bioprocessing. Biofuels, Bioprod. Biorefin. 2020, 14, Technol. 2012, 104, 775−782.
808−829. (27) Xu, F.; Sun, J.-X.; Sun, R.; Fowler, P.; Baird, M. S. Comparative
(10) Pereira Marques, F.; Lima Soares, A. K.; Lomonaco, D.; Study of Organosolv Lignins from Wheat Straw. Ind. Crops Prod.
Alexandre e Silva, L. M.; Tédde Santaella, S.; de Freitas Rosa, M.; 2006, 23, 180−193.
Carrhá Leitão, R. Steam Explosion Pretreatment Improves Acetic (28) Mansouri, N.-E. E.; Salvadó, J. Structural Characterization of
Acid Organosolv Delignification of Oil Palm Mesocarp Fibers and Technical Lignins for the Production of Adhesives: Application to
Sugarcane Bagasse. Int. J. Biol. Macromol. 2021, 175, 304−312. Lignosulfonate, Kraft, Soda-Anthraquinone, Organosolv and Ethanol
(11) Eraghi Kazzaz, A.; Fatehi, P. Technical Lignin and Its Potential Process Lignins. Ind. Crops Prod. 2006, 24, 8−16.
Modification Routes: A Mini-Review. Ind. Crops Prod. 2020, 154, (29) Zijlstra, D. S.; Lahive, C. W.; Analbers, C. A.; Figueirêdo, M. B.;
112732. Wang, Z.; Lancefield, C. S.; Deuss, P. J. Mild Organosolv Lignin
(12) Thoresen, P. P.; Matsakas, L.; Rova, U.; Christakopoulos, P. Extraction with Alcohols: The Importance of Benzylic Alkoxylation.
Recent Advances in Organosolv Fractionation: Towards Biomass ACS Sustainable Chem. Eng. 2020, 8, 5119−5131.
Fractionation Technology of the Future. Bioresour. Technol. 2020, (30) Bauer, S.; Sorek, H.; Mitchell, V. D.; Ibáñez, A. B.; Wemmer, D.
306, 123189. E. Characterization of Miscanthus Giganteus Lignin Isolated by
(13) Faix, O. Classification of Lignins from Different Botanical Ethanol Organosolv Process under Reflux Condition. J. Agric. Food
Origins by FT-IR Spectroscopy. Holzforschung 1991, 45, 21−28. Chem. 2012, 60, 8203−8212.
(14) Pinheiro, F. G. C.; Soares, A. K. L.; Santaella, S. T.; Silva, L. M. (31) Shinde, S. D.; Meng, X.; Kumar, R.; Ragauskas, A. J. Recent
A. e.; Canuto, K. M.; Cáceres, C. A.; Rosa, M. d. F.; Feitosa, J. P. d. A.; Advances in Understanding the Pseudo-Lignin Formation in a
Leitão, R. C. Optimization of the Acetosolv Extraction of Lignin from Lignocellulosic Biorefinery. Green Chem. 2018, 20, 2192−2205.
Sugarcane Bagasse for Phenolic Resin Production. Ind. Crops Prod. (32) Ratanasumarn, N.; Chitprasert, P. Cosmetic Potential of Lignin
2017, 96, 80−90. Extracts from Alkaline-Treated Sugarcane Bagasse: Optimization of
(15) Marques, F. P.; Silva, L. M. A.; Lomonaco, D.; Rosa, M. d. F.; Extraction Conditions Using Response Surface Methodology. Int. J.
Leitão, R. C. Steam Explosion Pretreatment to Obtain Eco-Friendly Biol. Macromol. 2020, 153, 138−145.
Building Blocks from Oil Palm Mesocarp Fiber. Ind. Crops Prod. 2020, (33) de Oliveira, D. R.; Nogueira, I. d. M.; Maia, F. J. N.; Rosa, M.
143, 111907. F.; Mazzetto, S. E.; Lomonaco, D. Ecofriendly Modification of
(16) Nascimento, D. M. d.; Almeida, J. S.; Vale, M. d. S.; Leitão, R. Acetosolv Lignin from Oil Palm Biomass for Improvement of PMMA
C.; Muniz, C. R.; Figueirêdo, M. C. B. d.; Morais, J. P. S.; Rosa, M. d. Thermo-Oxidative Properties. J. Appl. Polym. Sci. 2017, 134, 45498.
F. A Comprehensive Approach for Obtaining Cellulose Nanocrystal (34) van Krevelen, D. W. Some Basic Aspects of Flame Resistance of
from Coconut Fiber. Part I: Proposition of Technological Pathways. Polymeric Materials. Polymer 1975, 16, 615−620.
Ind. Crops Prod. 2016, 93, 66−75. (35) Cayla, A.; Rault, F.; Giraud, S.; Salaün, F.; Fierro, V.; Celzard,
(17) Rosa, M. F.; Medeiros, E. S.; Malmonge, J. A.; Gregorski, K. S.; A. PLA with Intumescent System Containing Lignin and Ammonium
Wood, D. F.; Mattoso, L. H. C.; Glenn, G.; Orts, W. J.; Imam, S. H. Polyphosphate for Flame Retardant Textile. Polymers 2016, 8, 331.
Cellulose Nanowhiskers from Coconut Husk Fibers: Effect of (36) Shukla, A.; Sharma, V.; Basak, S.; Ali, S. W. Sodium Lignin
Preparation Conditions on Their Thermal and Morphological Sulfonate: A Bio-Macromolecule for Making Fire Retardant Cotton
Behavior. Carbohydr. Polym. 2010, 81, 83−92. Fabric. Cellulose 2019, 26, 8191−8208.
(18) Yokoyama, T.; Kadla, J. F.; Chang, H.-m. Microanalytical (37) Dai, P.; Liang, M.; Ma, X.; Luo, Y.; He, M.; Gu, X.; Gu, Q.;
Method for the Characterization of Fiber Components and Hussain, I.; Luo, Z. Highly Efficient, Environmentally Friendly Lignin-

4066 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Based Flame Retardant Used in Epoxy Resin. ACS Omega 2020, 5,


32084−32093.
(38) Alzagameem, A.; Khaldi-Hansen, B.; Büchner, D.; Larkins, M.;
Kamm, B.; Witzleben, S.; Schulze, M. Lignocellulosic Biomass as
Source for Lignin-Based Environmentally Benign Antioxidants.
Molecules 2018, 23, 2664.
(39) Alriols, M. G.; Tejado, A.; Blanco, M.; Mondragon, I.; Labidi, J.
Agricultural Palm Oil Tree Residues as Raw Material for Cellulose,
Lignin and Hemicelluloses Production by Ethylene Glycol Pulping
Process. Chem. Eng. J. 2009, 148, 106−114.
(40) Ramezani, N.; Sain, M. Thermal and Physiochemical
Characterization of Lignin Extracted from Wheat Straw by Organo-
solv Process. J. Polym. Environ. 2018, 26, 3109−3116.

4067 https://doi.org/10.1021/acs.iecr.1c04362
Ind. Eng. Chem. Res. 2022, 61, 4058−4067

You might also like