You are on page 1of 24

Wood Science and Technology (2023) 57:173–196

https://doi.org/10.1007/s00226-022-01443-5

ORIGINAL

Birch wood biorefinery into microcrystalline,


microfibrillated, and nanocrystalline celluloses, xylose,
and adsorbents

B. N. Kuznetsov1,2   · I. G. Sudakova1 · N. V. Garyntseva1 · A. M. Skripnikov1,2 ·


A. V. Pestunov2 · E. V. Gnidan1,2

Received: 16 December 2021 / Accepted: 8 December 2022 / Published online: 22 December 2022
© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2022

Abstract
For the first time, it is proposed to combine the environmentally friendly heterogene-
ous catalytic processes of wood hemicelluloses hydrolysis and peroxide delignifica-
tion of «hemicellulose-free» wood for the biorefinery of birch wood into microcrys-
talline, microfibrillated, and nanocrystalline celluloses, xylose, and sorbents. The
use of the solid acid catalyst ­ZrO2/SO42− for the hydrolysis of birch-wood hemicel-
luloses at a temperature of 150 °C makes it possible to obtain xylose with a yield of
72.5% from the weight of hemicelluloses. The optimal conditions for peroxide del-
ignification of «hemicellulose-free» birch wood in a «formic acid–water» medium
over ­TiO2 catalyst, which ensure a high yield of microcrystalline cellulose (41.2%
from weight of wood), were established. By sulfuric acid hydrolysis and ultrasonic
treatment of microcrystalline cellulose, samples of microfibrillated and nanocrystal-
line celluloses were produced. Adsorbents with high sorption activity were obtained
from organosolv lignin formed as a side product of peroxide delignification of birch
lignocellulose. The birch biorefinery products were characterized by Fourier trans-
form infrared spectroscopy, X-ray diffractometry, scanning electron microscopy, gel
permeation chromatography, 31P nuclear magnetic resonance, dynamic light scatter-
ing, as well as chemical and elemental analysis.

* B. N. Kuznetsov
bnk@icct.ru
1
Institute of Chemistry and Chemical Technology, Krasnoyarsk Scientific Center, Siberian
Branch Russian Academy of Sciences, Akademgorodok 50, Bld. 24, Krasnoyarsk,
Russia 660036
2
Siberian Federal University, Pr. Svobodny 79, Krasnoyarsk, Russia 660041

13
Vol.:(0123456789)
174 Wood Science and Technology (2023) 57:173–196

Introduction

The huge woody biomass resources are renewable raw materials, which can
become a worthy alternative to fossil fuels (Cherubini 2010). Wood consists of
65–70% of carbohydrates (cellulose and hemicelluloses) and 25–30% of aromatic
polymer lignin. The carbohydrate component of wood is used to obtain the plat-
form chemicals: glucose, furfural, hydroxymethyl  furfural, levulinic acid, etha-
nol, sorbitol, xylitol, etc. (Takkellapati et al. 2018). Lignin also has a wide range
of potential applications, from binders and enterosorbents to monomeric phenolic
and aromatic compounds (Cao et al. 2019).
In contrast to softwood, hardwood contains a great amount of hemicellu-
loses. Birch wood hemicelluloses are represented mainly by glucuronoxylans.
The products of xylan hydrolysis (xylooligosaccharides and xylose) are used in
medicine, cosmetology, food industry, and other fields (Chemin et al. 2015). To
ensure the selective depolymerization of hemicelluloses, it is necessary to estab-
lish the wood hydrolysis conditions that would minimize the conversion of other
wood components. The selection of effective non-toxic catalysts is a key direction
in optimizing the processes of hydrolysis of polysaccharides (Degirmenci et  al.
2011; Wu et  al. 2016). The use of solid acid catalysts instead of mineral acids
will improve the environmental safety of hydrolysis, prevent the corrosive effect
of a reaction medium, and exclude additional costs for neutralizing reaction solu-
tions (Takagaki et al. 2011; Vicocq et al. 2014).
The main trend in the chemical processing of wood is the production of cel-
lulose. In recent years, there has been an increased interest in the synthesis of
cellulosic nanomaterials based on microfibrillated and nanocrystalline celluloses
(Moon et al. 2011; Charreau et al. 2013). Due to its unique properties, including
the low weight, large specific surface area, high porosity, non-toxicity, biocom-
patibility, and biodegradability, the microfibrillated cellulose (MFC) is in demand
for the production of aerogels (Zhou et  al. 2016), biocomposites, biodegradable
materials, medical implants (Du et al. 2019), and polymer-based reinforced com-
posites (Lepetit et al. 2017). Due to its good barrier properties, MFC is used in
films and paper coatings (Lavoine et al. 2012).
To obtain MFC and NCC from wood, multistage processes are used. First,
technical cellulose is obtained by sulfate and sulfide delignification of wood.
Then, residual lignin and amorphous cellulose are removed from technical cel-
lulose by bleaching and hydrolysis.
In the conventional industrial pulping and bleaching processes, toxic sulfur and
chlorine-containing delignifying agents are used, which damage the environment.
The developed environmentally friendly cellulose production methods are based
on the use of non-toxic organic and water-organic solvents (Shen and Sun 2021).
In the available processes of organosolvent delignification of lignocellulosic
biomass, concentrated acetic or formic acid solutions (60–90 wt.%) and hydrogen
peroxide solutions are used at elevated temperatures (120–180 °C) and pressures.
The advantages of formic acid in terms of its use in organosolvent delignifica-
tion are its easy regeneration, ability to dissolve lignin degradation products, and

13
Wood Science and Technology (2023) 57:173–196 175

possibility of obtaining formic acid production from biomass polysaccharides


(Gromov et al. 2018).
Earlier, Garyntseva et  al. (2019) demonstrated the possibility to obtain a high
yield of pure cellulose by peroxide delignification of abies wood in the «formic
acid–water» medium in the presence of catalyst ­TiO2 at a temperature of 100  °C,
ambient pressure, and low concentrations of the ­H2O2 (6 wt.%) and CHOOH (30
wt.%) reagents. The produced cellulose has a low residual lignin content (2.3 wt.%)
and can be used for processing into nanocellulose. The obtained organosolv lignin
differs from conventional technical lignins by the absence of sulfur, less condensed
structure, and high reactivity. These advantages of organosolv lignins, along with
their good solubility in organic solvents, make them attractive for further catalytic
processing into valuable chemical products (Kuznetsov et al. 2016) and for the pro-
duction of enterosorbents (Garyntseva et al. 2011).
The catalytic peroxide delignification in the «acetic acid–water» medium was
used as a key process for larch wood biorefinery into microcrystalline, microfilril-
lated, and nanocrystalline celluloses (Kuznetsov et al. 2021a).
In this work, it is proposed to perform birch wood biorefinery by integrating
the processes of catalytic hydrolysis of woody hemicelluloses with the formation
of xylose and lignocellulose, catalytic peroxide fractionation of lignocellulose into
microcrystalline cellulose (MCC) and organosolv lignin, and conversion of MCC
into microfibrillated cellulose (MFC) and nanocrystalline cellulose (NCC).
The birch wood biorefinery products were characterized by FTIR, XRD, SEM,
GPC, GC NMR, DLS and chemical methods.

Experimental

Materials and reagents

In the experiments, air-dried sawdust of birch wood harvested in the forest area near
Krasnoyarsk City (Russia) was used. The contents of cellulose, lignin, and hemi-
cellulose in the wood were determined by conventional wood chemistry methods
(Sjöström and Alern 1999). The birch wood chemical composition included 46.8
wt.% of cellulose, 21.7 wt.% of lignin, 27.3 wt.% of hemicelluloses, 3.2 wt.% of
extractive substances, and 0.6 wt.% of ash.
The birch sawdust (fraction 2–5  min) was mechanically activated in an AGO-2
planetary mill for 30 min at a centrifugal acceleration of 60 g using steel grinding
balls 3–8 mm in diameter.
The reagents used were zirconyl chloride (Merck 108,917), ammonium hydroxide
of reagent grade (GOST 24,147–80, Russia), sulfuric acid of reagent grade (GOST
4204–77, Russia), Ku-2–8 acid cation exchange resin (Thermax), Amberlyst®15
dry solid acid catalyst (Across Organics), a 3-mercaptopropyltrimethoxysilan
­C6H16O3SSi modifying agent (Aldrich, Cat.: 175,617), high-purity hydrogen perox-
ide (GOST 177–88, Russia), high-purity formic acid (GOST 5848–73, Russia), and
­TiO2 catalyst of analytical grade (DuPont, US). The solutions were prepared using
distilled water (GOST 6709–72). The monosaccharides used as chromatography

13
176 Wood Science and Technology (2023) 57:173–196

standards were crystalline hydrate glucose (GOST 975–88), Panreac D-xylose


142,080.1208, D-mannose 373,195.1208, and D-sorbitol.

Catalysts for hemicelluloses hydrolysis

The catalysts used for hydrolysis of birch wood hemicelluloses were an acid-modi-
fied SBA-15 catalyst with–SO3H groups and sulfated ­ZrO2 with –SO4−2 groups.
The acid modification of the SBA-15 included chemical fixing of 3-mercaptopro-
pyltrimethoxysilan on the catalyst surface and further oxidation of SH groups to sul-
fone groups –SO3H by treating with ­H2O2 (30%) at room temperature according to
a procedure described in Melero et al. (2006). Sulfated Z ­ rO2 was obtained by a con-
ventional two-stage method. At the first stage, zirconium hydroxide was precipitated
from the zirconyl chloride (­ ZrOCl2) solution with the ammonium hydroxide solution
at constant pH = 9.5 ± 0.5. At the second stage, sulfation with the 0.5 M ­H2SO4 solu-
tion with the subsequent calcination at 550–650 °C was performed.
The main characteristics of the modified catalysts used for hydrolysis of birch
wood hemicelluloses are given in Table 1.

Hydrolysis of activated birch wood

Hydrolysis of activated birch wood was carried out upon temperature variation from
110 to 170  °C in a steel rotary autoclave with a 35ml internal fluoroplastic tube
placed in an air cooling metal bath. The autoclave rotated at a rate of 11 rpm. The
wood and solid catalyst mixture was placed in a test tube and poured over with dis-
tilled water. The wood weight content in water was 40 g/l. The solid catalyst/wood
weight ratio was 1: 1.
After a specified time, the reactor was cooled and the non-hydrolyzed wood and
the catalyst were filtered on a Büchner funnel with a paper filter under vacuum,
washed with water, and dried in an oven at a temperature of 103 °C.
Due to the higher density of the catalysts (5.80 g/cm3 for Z
­ rO2/SO4 and 2.56 g/
cm3 for SBA –15) compared to the lignocellulose product (1.52 g/cm3), it is possible
to isolate more than 90 wt.% of the catalyst by combining intensive mixing and mul-
tiple decantation of the reaction mixture. It has been experimentally shown that the
residual impurities of a solid hydrolysis catalyst do not have a noticeable effect on
the process of subsequent peroxide fraction of a lignocellulose product in the pres-
ence of T­ iO2 catalyst.

Table 1  Textural characteristics of the solid acid catalysts


Catalyst Density, g/cm3 SBET, ­m2/g Vpores, ­cm3/g  < d > pores, nm

Acid-modified SBA- 2.56 417 0.59 5.6


15 with − ­SO3H
groups
Sulphated ­ZrO2 with 5.80 110 0.09 4.5
­SO4− 2 groups

13
Wood Science and Technology (2023) 57:173–196 177

Catalytic peroxide delignification of «hemicellulose‑free» birch wood

Catalytic delignification of hydrolyzed birch wood using ­H2O2 was carried out in
250 ­cm3 glass reactor equipped with a mechanical stirrer, reflux condenser, and ther-
mometer, according to the procedure described in Kuznetsov et  al. (2017). Wood
sawdust (10  g) was placed into a glass reactor. Then, a mixture of formic acid,
hydrogen peroxide, distilled water, and the catalyst was added. The reaction mixture
was vigorously stirred (700 rpm) at a temperature of 100 °C for 4 h. The contents
of the reaction mixture components ranged within 4–10 wt.% of hydrogen peroxide,
30–50 wt.% of formic acid and the liquid/wood ratio (LWR) 10–20. After comple-
tion of the reaction, the solid product was separated on a Büchner funnel under vac-
uum, washed with distilled water, and dried at 105 °C until constant weight.
The spent cooking solution was evaporated (formic acid was distilled off) and
lignin was precipitated by adding a fivefold excess of water.
The cellulose product yield was estimated by the gravimetric method and calcu-
lated as Y = (m/mo) × 100, where Y is cellulose product yield (wt.%), m is the abso-
lutely dry cellulose product weight (g), and m ­ 0 is the absolutely dry wood weight
(g).
Commercial ­TiO2 was used as a catalyst of peroxide fractionation of wood. The
amount of the catalyst in all the experiments was 1% of the wood weight. The T ­ iO2
catalyst phase composition included 92% of rutile and 8% of anatase with an aver-
age particle size of ~ 10 µm and a Brunauer–Emmett–Teller (BET) surface area of 3
­m2/g.

Production of nanocelluloses

The nanocellulose materials were obtained by controlled acid hydrolysis of cellulose


with a fractional composition of 0.1–0.25 mm. 1.0 g cellulose was added to 25 ml
of the H
­ 2SO4 solution (55 wt.%) and hydrolysis was carried out at a temperature of
25 °C for 2 h under continuous stirring. The hydrolysis process was stopped by add-
ing a tenfold volume of cold (∼10 °C) distilled water and stirring was continued for
further 15 min. The colloidal suspension was centrifuged in an OHAUS Multi-Pro
FC5718 centrifuge (Germany) at a rate of 12,000 rpm for 10 min and a precipitate
was separated from a centrifugate by decantation.
The hydrolyzed cellulose precipitate was washed from sulfuric acid by dialysis
against distilled water in an MF-503–46 MFPI cellophane dialysis bag (US) with a
pore size of 3.5 kDa (0.1 μm). Dialysis was carried out until neutral dialyzed solu-
tion; water in the bag was changed every hour. The neutral precipitate was added
to 25 ml of distilled water and the ultrasonic (US) treatment in a Grad ultrasound
bath (35 kHz, 55 W, Russia) was carried out for 15 min.
The obtained suspension was separated into MFC by centrifugation at 8000 rpm
for 10 min. A precipitate was separated from a centrifugate and the latter was also
subjected to the US treatment for 15 min and NCC was isolated by centrifugation at
12,000 rpm.

13
178 Wood Science and Technology (2023) 57:173–196

The obtained MCC, MFC, and NCC samples were kept at –18 °C for 24 h and
dried in a Iney-6 freeze dryer (Russia) until constant weight.

Production of adsorbents

To obtain adsorbents, 5g of organosolv lignin was treated with the 0.4% ­NaHCO3
solution at room temperature or with water at 95 °C for 30 min (hydromodule 20).
After the treatment, a solid precipitate was separated from the solution by filtration
on a Büchner funnel and the obtained adsorbents were neutralized and purified from
low-molecular substances using MF-505–46 MFPI cellophane dialysis bags (US)
with a pore size of 0.1  μm. Dialysis was carried out for 8  h with changing water
every half an hour. After dialysis, adsorbents were separated from the solution by
centrifugation at 8000 rpm for 8 min, kept at a temperature of + 5 °C for 24 h, and
brought to constant weight in a freeze dryer.

Methods of analysis

The cellulose, hemicellulose, and lignin contents in the initial wood and cellulosic
products were determined by conventional wood chemistry methods (Sjöström and
Alern 1999).
The cellulose content in wood was determined by the Kürschner method. The
lignin content was determined by hydrolysis of the sample with 72 wt.% of sulfuric
acid at 20 °C for 2.5 h followed by dilution of the solution with water and boiling for
1 h. The hemicellulose content was determined by the McKein and Shoorly method
using hydrolysis with 2 wt.% of HCl at 100 °C for 3 h.
Fourier transform infrared (FTIR) spectra were recorded on a Bruker Tensor–27
FTIR spectrometer in the wavelength range of 4000–400 ­cm−1. Samples for analysis
(4 mg for each) were prepared in the form of tablets with the KBr matrix. The spec-
tral data were processed in the OPUS/YR program (version 2.2).
The X-ray diffraction (XRD) analysis was carried out on a PANalytical X’Pert
Pro diffractometer using a Cu-Kα source (A = 0.154  nm) in the 2θ angle range of
5–70° at a scanning step width of 0.01°/scan. The samples were analyzed by the
powder method in a cuvette 2.5 cm in diameter.
The crystallinity index (CI) was calculated from the height ratio between the
intensity of the crystalline peak and the total intensity after subtraction of the back-
ground signal (Park et al. 2010).
Scanning electron microscopy (SEM) images were obtained on a HITACHI
TM-3000 microscope (Japan) with an acceleration potential of 15  kV. The speci-
mens were placed onto a carbon support.
The molecular weight of lignin was examined using an Agilent 1260 Infinity gel
permeation chromatograph (US) with a refractive index detector and an Agilent
PLgel Mixed-C column. The eluent used was chloroform and the eluent flow rate
was 1.0 ml/min at 40 °C. Typical sample volumes were 50 ml at a polymer concen-
tration of 2 mg/ml. Agilent narrow polydispersity polystyrene standards (US) were

13
Wood Science and Technology (2023) 57:173–196 179

used to generate a universal calibration curve for determining the weight-average


(Mw) and number-average (Mn) molecular weights and polydispersity.
The monosugar contents were determined using a VARIAN-450 GC gas chroma-
tograph with a flame ionization detector and a VF-624 ms capillary column with a
length of 30 m and an inner diameter of 0.32 mm.
The degree of cellulose polymerization was calculated according to ASTM 1795
standard test (ASTM D1795 2013) using the equation

DP0.905 = 0.805 × [𝜂],


where DP is the degree of polymerization and [η] is the intrinsic viscosity.
The nanoparticle hydrodynamic diameter was measured by the dynamic light
scattering (DLS) technique on a Zetasizer Nano ZS spectrometer (Malvern Instru-
ments, UK).
The zeta potentials of suspended particles were measured on a Zetasizer Nano ZS
device using the electrophoretic mobility in polycarbonate cells with Pd electrodes
at 20 °C without adding a background electrolyte or pH correction.

Results and discussion

In this study, it was proposed for the first time to carry out a biorefinery of birch
wood by integrating the processes of catalytic hydrolysis of wood hemicelluloses to
obtain xylose, catalytic peroxide fractionation of hemicellulose-free wood into MCC
and organosolv lignin, and conversion of MCC into MFC and NCC.
For the processes of hydrolysis of hemicelluloses and peroxide fractionation of
wood, active solid catalysts were applied.

Hydrolysis of birch wood hemicelluloses with solid acid catalysts

The catalytic properties of solid acid catalysts in the process of hydrolysis of birch
wood hemicelluloses to xylose were compared at temperatures of 110–170  °C.
The solid acid catalysts used were SBA-15 containing S ­ O3H groups and sulfated
­ZrO2 containing S ­ O4−2 groups. For comparison, a commercial Ku-2–8 acid cation
exchanger and an Amberlyst®15 sulfonic cation exchanger (copolymer of styrene
and divinylbenzene) were taken.
As is known (Kuznetsov et al. 2014), during the heterogeneous catalytic hydroly-
sis of plant raw materials, it is difficult to ensure the effective contact of a solid cata-
lyst with a solid reagent. To increase the contact area, the wood sawdust (a fraction
of 2–5 mm) was mechanically treated in an AGO-2 planetary mill. After mechanical
grinding, the sawdust turned into a finely dispersed homogeneous mass. Accord-
ing to the SEM data, the average wood particle size decreased from 2–5  mm to
0.1–0.25 mm after grinding (Fig. 1).
It was established by the gas chromatography method that the hydrolysates
obtained at temperatures of 110–130 °C, regardless of the catalyst used, contain low
amount of xylose. (Table 2).

13
180 Wood Science and Technology (2023) 57:173–196

Fig. 1  SEM images of birch sawdust (a) before and (b) after grinding for 30 min

At hydrolysis temperature of 150 °C, the ­ZrO2/SO4−2 and Amberlyst®15 catalysts


exhibited high catalytic activity. The xylose contents in hydrolysates were of 8.08
and 7.17 g/L (72.1% and 63.5% from woody hemicelluloses) respectively (Table 2).
These yields are comparable with xylose yields (of 72.4 and 77.1% from the hemi-
cellulose content in wood) obtained previously during birch wood hydrolysis with
2% and 3% H ­ 2SO4, at a temperature of 100 °C for 5 h (Yatsenkova et al. 2015).
An increase in the process temperature to 170 °C intensifies the hydrolysis of the
cellulose component of wood. In this case, the glucose yield obtained in the pres-
ence of the acid-modified Z­ rO2(–SO4−2) and SBA-15(− SO3H) catalysts increases to
3.91 and 4.91 g/L, respectively (Table 2).
In the absence of catalysts, hydrolysis of birch wood polysaccharides at tempera-
tures of 110 and 130 °C does not occur and, at a temperature of 150 °C, the xylose
and glucose yields are no higher than 0.07–0.06 g/L.
In the hydrolysates obtained at temperatures of 110–130 °C, no furfural, 5-HMF,
or levulinic acid were detected. As the temperature increased, the process tempera-
ture to 170 °C intensifies the hydrolysis of the cellulose component of wood. In this
case, the glucose yield obtained in the presence of the acid-modified ­ZrO2(–SO4−2)
and SBA-15(–SO3H) catalysts increases to 3.91 and 4.91 g/L, respectively (Table 2).
In the absence of catalysts, hydrolysis of birch wood polysaccharides at tempera-
tures of 110 and 130 °C does not occur and, at a temperature of 150 °C, the xylose
and glucose yields are no higher than 0.07–0.06 g/L.
The furfural and 5-HMF concentrations in the hydrolysates, obtained at the tem-
perature of 150 °C, were found to be 0.35 and 0.07 g/L, respectively, and levulinic
acid was almost absent.

Peroxide fractionation of birch lignocellulose in the «formic acid− water» medium


in the presence of the ­TiO2 catalyst

As previously shown (Garyntseva et  al. 2019), abies wood can be fraction-
ated into MCC and low-molecular-weight organosolv lignin in the «hydrogen

13
Table 2  Effect of the solid acid catalysts on the contents of xylose and glucose in the hydrolysates of birch wood (time 1 h, LWR 24, and wood/catalyst ratio 1: 1)
Catalyst Hydrolysis temperature °C

110 130 150 170


Wood Science and Technology (2023) 57:173–196

Xylose, g/L Glucose, g/L Xylose, g/L Glucose, g/L Xylose, g/L Glucose, g/L Xylose, g/L Glucose, g/L

Without catalyst – – – – 0.07 ± 0.001 0.06 ± 0.001 1.04 ± 0.015 0.08 ± 0.001


ZrO2(− ­SO4− 2) 0.07 ± 0.001 1.54 ± 0.023 1.36 ± 0.020 1.90 ± 0.029 8.08 ± 0.121 0.97 ± 0.014 1.78 ± 0.027 3.91 ± 0.059
SBA-15(− ­SO3H) 0.06 ± 0.0009 1.23 ± 0.018 0.30 ± 0.005 1.49 ± 0.022 4.00 ± 0.060 0.31 ± 0.005 7.21 ± 0.108 4.91 ± 0.074
Amberlyst®15 0.04 ± 0.0006 0.31 ± 0.005 0.78 ± 0.012 0.61 ± 0.009 7.17 ± 0.108 0.81 ± 0.012 4.62 ± 0.069 1.55 ± 0.023
Ku-2–8 0.05 ± 0.0006 0.12 ± 0.002 0.15 ± 0.002 0.14 ± 0.002 1.07 ± 0.016 0.19 ± 0.002 5.73 ± 0.086 0.74 ± 0.011
181

13
182 Wood Science and Technology (2023) 57:173–196

peroxide–formic» acid–water medium in the presence of the T ­ iO2 catalyst at a tem-


perature of 100 °C for 4 h.
Similar temperature and time were chosen for the fractionation of the lignocellu-
lose product formed during the hydrolysis of birch wood hemicelluloses in the pres-
ence of solid acid catalysts. The effect of the hydrogen peroxide and formic acid
concentrations and the liquid-to-wood ratio (LWR) on the yields and composition of
the products of birch lignocellulose peroxide fractionation was studied.
The lignocellulose product was obtained by hydrolysis of birch wood hemicel-
luloses in the presence of sulfated Z­ rO2(–SO4−2). The chemical composition of the
product (wt.%) was: 61.2 of cellulose, 27.3 of lignin, and 9.5 of hemicelluloses.
With an increase in the concentration of ­H2O2 from 4 to 7 wt.%, the cellulose
content in the cellulosic product and the soluble organosolv lignin yield increase
(Table  3). With a further increase in the hydrogen peroxide concentration to 10
wt.%, the organosolv lignin yield decreases due to the intensification of secondary
oxidation reactions. In this case, the molecular weight of lignin decreases and low-
molecular-weight products of its oxidation (formic, oxalic acids, etc.) are formed.
A decrease in the yield of the cellulosic product at the high ­H2O2 concentration is
apparently due to the reaction of destructive oxidation of the amorphous part of
cellulose.
The growth of the formic acid concentration in the reaction mixture from 30 to
50 wt.% leads to a decrease in the content of residual lignin in the cellulosic product
from 3.2 to 0.5 wt.%. The cellulosic product formed at a formic acid concentration
of 50 wt.% has the highest cellulose content: 95 wt.% (Table 3).
The growth of the LWR value facilitates diffusion of the reagents into the inter-
cellular space of wood and removal of delignification products into the solution.
However, using a high LWR value is economically impractical. It was found that, at
LWR 10, a cellulosic product with a cellulose content of 85.0 wt.% is formed. The
growth of the LWR to 15 and 20 leads to an increase in the cellulose content in cel-
lulosic product to 92.8 and 93.0 wt.%, respectively (Table 3).
A higher yield of cellulose product (67.2 wt.%) containing 91.0 wt.% of cellu-
lose was obtained at an HCOOH concentration of 30%. To obtain a cellulose prod-
uct with higher cellulose content (92.8 wt.%) the process of peroxide fractionation
of birch lignocellulose should be carried out under the following conditions: ­H2O2
concentration of 7 wt.%, HCOOH concentration of 40 wt.%, LWR 15, temperature
100 °C, time 4 h. Since the majority of formic acid will be returned to the techno-
logical cycle, its concentration (30 − 40 wt.%) should have only a little effect on the
cost of the produced cellulose product.

Characterization of cellulose and lignin obtained by the catalytic peroxide


fractionation of birch lignocellulose

Some characteristics of cellulose obtained under the optimal conditions of peroxide


fractionation of birch lignocellulose were investigated by the FTIR, XRD, and SEM
methods.

13
Table 3  Effect of the ­H2O2 and HCOOH concentration and LWR on the yield and composition of the products obtained by peroxide fractionation of birch lignocellulose at
100 °C for 4 h
H2O2 concentration, wt.%a HCOOH concentration, wt.%b LWRc
4 7 10 30 40 50 10 15 20
Wood Science and Technology (2023) 57:173–196

Cellulose product yield, wt.% ⃰ 68.3 ± 2.0 65.9 ± 1.9 60.4 ± 1.8 67.2 ± 2.0 65.9 ± 1.9 64.4 ± 1.9 68.7 ± 2.0 65.9 ± 1.9 65.8 ± 1.9
Cellulose product composition, wt.%**:
Cellulose 89.6 ± 1.3 92.8 ± 1.2 90.5 ± 1.4 91.0 ± 1.4 92.8 ± 1.2 95.0 ± 1.4 85.0 ± 1.3 92.8 ± 1.2 93.0 ± 1.4
Lignin 6.9 ± 0.2 2.0 ± 0.03 0.5 ± 0.01 3.2 ± 0.06 2.0 ± 0.03 0.5 ± 0.01 4.5 ± 0.07 2.0 ± 0.03 1.8 ± 0.03
Hemicelluloses 4.2 ± 0.06 4.2 ± 0.06 4.0 ± 0.05 4.8 ± 0.07 4.2 ± 0.06 3.5 ± 0.05 5.5 ± 0.08 4.2 ± 0.06 4.2 ± 0.06
Organosolv lignin yield, wt.% ⃰ ⃰ 7.5 ± 0.2 11.2 ± 0.3 5.9 ± 0.2 9.5 ± 0.3 11.2 ± 0.3 11.5 ± 0.4 11.0 ± 0.3 11.2 ± 0.3 8.5 ± 0.2
* a b c
 On abs. dry «lignocellulose product», **on abs. dry cellulose product, at 40 wt.% of HCOOH and LWR 15, at 7 wt.% of ­H2O2 and LWR 15, at 7 wt.% of ­H2O2 and
40 wt.% of HCOOH
183

13
184 Wood Science and Technology (2023) 57:173–196

The FTIR spectrum (Fig. 2) of the birch wood cellulose sample includes absorp-
tion bands at 3000–3600, 2700–300, 1500–1300, and 1000–1200  ­ cm−1, which
correspond to stretching vibrations of –OH and –CH2–CH2 groups and deforma-
tion and stretching vibrations of the C= O and C–O–C bonds of the cellulose ring,
respectively (Fan et al. 2012). The absorption band at 1727 ­cm− 1 indicates the pres-
ence of hemicelluloses in the birch cellulose. The absence of peaks in the range
of 1509–1609  ­cm−1, which corresponds to aromatic ring vibrations, evidences the
complete removal of lignin from cellulose.
According to the XRD data (Fig. 3), the unit cell of cellulose obtained from birch
wood is identical to the monoclinic unit cell of cellulose I (Nishiyama et al. 2002).
The crystallinity index (CI) of birch cellulose calculated as in Park et al. (2010) is
0.72. The results obtained show that birch cellulose has the same composition and
structure as MCC.
The morphologies of the samples of birch wood and birch cellulose were com-
pared. The SEM image, clearly demonstrates the microstructure of birch wood
(Fig. 4a). The wood particle surface is heterogeneous and contains bordered pores
in the libriform fiber walls. Cellulose obtained from birch wood (Fig. 4b) consists
of microfibrils, some of which are bundled. The microfibril surface is smooth and
homogeneous due to the removal of extractives, lignin, and hemicelluloses.
According to the GPC data, (Fig.  5) the low-molecular-weight lignin obtained
under the optimal conditions of the catalytic peroxide fractionation of birch lig-
nocellulose has number-average molecular weight 806  g/mol and weight-average
weight 1702  g/mol. A rather low value of polydispersity (2.112) indicates a fairly
high homogeneity of birch organosolv lignin.
The FTIR spectrum of lignin (Fig.  6) contains characteristic absorption bands
that point out the presence of aromatic structures. The bands at 1510, 1465, and
1424  ­cm−1 correspond to skeletal vibrations of the aromatic ring and the bands

Fig. 2  FTIR spectra of (1) the cellulose sample produced from birch lignocellulose and (2) commercial
Vivapur MCC

13
Wood Science and Technology (2023) 57:173–196 185

Fig. 3  X-ray diffraction patterns of (1) the cellulose sample produced from birch lignocellulose and (2)
commercial Vivapur MCC

Fig. 4  SEM images of birch wood (a) and cellulose obtained from birch wood (b) and commercial
Vivapur MCC (c)

Fig. 5  Molecular weight dis-


tribution of organosolv lignin
isolated under the optimal con-
ditions of peroxide fractionation
of birch lignocellulose

13
186 Wood Science and Technology (2023) 57:173–196

Fig. 6  FTIR spectrum of organosolv lignin isolated under the optimal conditions of peroxide fractiona-
tion of birch lignocellulose

at 1128 and 1330  ­cm−1 indicate the presence of S-type aromatic structures in the
lignin. A feature of the spectrum of lignin is the high intensity of the absorption
band at 1728  ­cm−1 corresponding to vibrations of a C = O bond in the carboxyl
group. This indicates a high content of oxygen-containing groups in organosolv
lignin (Shi et al. 2019).
The total content of OH groups in organosolv lignin and their nature (ali-
phatic, phenolic, or carboxyl) were determined using the NMR 31P method (Fig. 7)
(Table 4).
In contrast to conventional technical lignins, organosolv birch lignin is less con-
densed, contains no sulfur impurities, has a low ash content and a high content
of hydroxyl groups. This makes it possible to use organosolv birch lignin in the

Fig. 7  31P NMR spectrum of organosolv lignin isolated by catalytic peroxide fractionation of birch ligno-
cellulose

13
Wood Science and Technology (2023) 57:173–196 187

Table 4  Content of hydroxyl Hydroxyl groups δ 31P–NMR, ppm Content, mmol/g


groups in organosolv birch
lignin Aliphatic 150.0–145.5 3.66
Phenolic 136.6–144.7 2.01
Carboxyl 133.6–136.6 1.11

production of enterosorbents (Garyntseva et al. 2011) and binders (Kuznetsov et al.


2011), as well as for catalytic processing into liquid fuels (Kuznetsov et al. 2021b)
and functional polymers (Kuznetsov et al. 2020).

Preparation of nanocellulose materials from birch cellulose

Cellulosic nanomaterials are prepared, as a rule, by acid hydrolysis of cellulose fol-


lowed by the ultra-sound (US) treatment. Acid hydrolysis destroys the amorphous
areas in microfibrils, leaving the crystalline areas unchanged, and the US treatment
destroys interfibrillar hydrogen bonds with the formation of nanofibers (Kargarzaden
et al. 2012; Jasmani and Thielemans 2018). The size and degree of crystallinity of
cellulose nanofibrils depend on the time of hydrolysis and US treatment.
In a previous work (Kuznetsov et  al. 2021a), it was proposed to produce MFC
and NCC from MCC obtained by larch wood by one-step catalytic peroxide frac-
tionation. A similar approach was used in this work to produce nanocellulose mate-
rials from birch wood.
The nanocellulosic materials were obtained by controlled hydrolysis of birch
MCC with the H ­ 2SO4 solution (55 wt.%) for 2  h at a temperature of 25  °C under
continuous stirring. The suspension was centrifuged for 10  min and a precipitate
was separated from a centrifugate by decantation. After that, the precipitate was
added with distilled water and the US treatment was carried out for 15 min. Using
10 min centrifugation, the obtained suspension was separated into MFC (8000 rpm)
and NCC (12,000 rpm).
The yield, degree of polymerization (DP) and crystallitily index (CI) of MCC,
MFC, and NCC obtained are given in Table 5.
Sulfuric acid hydrolysis of MCC for 2  h with the subsequent US treatment
reduces the degree of its polymerization. The obtained MFC and NCC have DP val-
ues of 121 and 98, respectively. The yields of MFC (27.7 wt.%) and NCC (10.9
wt.%) are comparable with the literature data (Ditzel et  al. 2017; Nascimento and
Rezende 2018). Depending on the nature of a lignocellulosic material, conditions

Table 5  Yield, degree of Sample Yield*, wt.% Yield**, wt.% DP CI


polymerization, and crystallinity
index of the cellulose samples MCC – 38.8 ± 0.58 235 0.72
obtained from birch wood
MFC 42.0 ± 0.63 27.7 ± 0.42 121 0.78
NCC 16.8 ± 1.02 10.9 ± 0.16 98 0.69
*
 On abs. dry MCC, ** on abs. dry wood

13
188 Wood Science and Technology (2023) 57:173–196

for its pretreatment and acid hydrolysis, and methods used for the NCC isolation, the
NCC yield varies from 3.8 to 27.9% of the initial wood weight (Ditzel et al. 2017).
The MCC, MFC, and NCC samples obtained from birch wood were characterized
by FTIR, XRD, SEM, DLS, and elemental analysis.
The FTIR spectra of the MCC, MFC, and NCC samples (Fig. 8) are almost iden-
tical, which evidences the similarity of their structures typical of cellulose I (Fan
et  al. 2012; Nishiyama et  al. 2002). The spectra of the cellulose samples contain
no absorption bands characteristic of phenylpropane units of lignin (1605–1593,
1515–1495, and 1470–1460 ­cm− 1), which confirms its removal during the catalytic
peroxide delignification of birch wood. The absence of absorption bands in the spec-
tra of MFC and NCC, which are responsible for vibrations of ester bonds and acetyl
groups in hemicelluloses, points out their removal during hydrolysis and the US
treatment.
The XRD patterns of the MCC, MFC and NCC samples (Fig. 8) have maxima at
2θ angles of 15.2°, 16.2°, and 22.5° (Fig. 9) caused by the reflection from the 110,
101, and 002 planes of the crystal lattice of cellulose I (Park et al. 2010).
These peaks in the XRD patterns of all the samples indicate the preservation of
the structure of cellulose I during hydrolysis and US treatment of MCC. These treat-
ments of MCC facilitate the removal of amorphous areas in microfibrils, which is
accompanied by a decrease in the intensity of the maxima in the 2θ angle range
of 15.2°–16.2° on the XRD patterns of the MFC and NCC samples (Fig. 8). Acid
hydrolysis of cellulose leads to the rupture of polymer chains in the amorphous
areas of cellulose microfibrils, which is accompanied by an increase in the crys-
tallinity index from 0.70 (initial MCC) to 0.72 and 0.78 for MCC and MFC sam-
ples, respectively (Table  5). Similar effects were observed in Hindi (2017) during
preparation of nanocelluloses by acid hydrolysis. The US treatment reduces the CI
of the obtained NCC to 0.69 (Table 5). This can be explained by the fact that, during
the US treatment, the crystalline areas in nanocellulose are damaged (Correa et al.
2019) or decreased, which leads to a decrease in the calculated CI value.
It can be concluded from the analysis of the SEM image of the obtained cellu-
lose samples that acid hydrolysis and the US treatment of cellulose contribute to the
removal of residual hemicelluloses, destruction of amorphous areas in microfibrils,

Fig. 8  FTIR spectrum of samples MCC (1), MFC (2), and NCC (3) obtained from birch wood

13
Wood Science and Technology (2023) 57:173–196 189

Fig. 9  X-ray diffraction pattern of samples MCC (1), MFC (2), and NCC (3) obtained from birch wood

and cleavage of cellulose fibers into nanofibers by the rupture of interfibrillar hydro-
gen bonds (Fig. 10).
The MCC sample consists of loose fibers with numerous breaks and macropores
(Fig.  10a). The latter are probably formed as a result of the removal of hemicel-
luloses and lignin during peroxide delignification of wood. The MCC particle size
varies over a wide range (from 30 to 300 μm). The MFC consists of a set of fibers
of different sizes (Fig. 10b). The NCC consists of a network of smooth fibrils with
fairly uniform sizes (Fig. 10c).
To estimate the size distribution of nanofibers in the 0.01% aqueous suspensions
of MFC and NCC, the hydrodynamic particle diameter was measured by dynamic
light scattering (DLS) and the ζ potential of suspended particles using the electro-
phoretic mobility of the suspensions (Fig. 11). For the NCC sample, a narrow parti-
cle size distribution is observed and the average nanofiber size is 30.8 nm (curve 2
in Fig. 11A). The MFC sample is characterized by a wide particle size range: from
55 to 173 nm (curve 1 in Fig. 11A). The average MFC particle diameter is 91.3 nm.
The average value of the ζ potential is 24.4  mV for an aqueous suspension of
MFC (curve 1 in Fig. 11B) and − 30.3 mV for an aqueous suspension of NCC (curve

Fig. 10  SEM images of the MCC (a), MFC (b), and NCC (c) samples obtained from birch wood

13
190 Wood Science and Technology (2023) 57:173–196

Fig. 11  (A) Hydrodynamic particle diameter (B) and ζ potential of suspended particles of the MFC (1)
and NCC (2) samples

2 in Fig. 11B). The negative ζ value is related to the presence of negatively charged


sulfate groups on the surface nanoparticles. The MFC and NCC aqueous suspen-
sions have a ζ potential of − 24.4 and − 30.3 mV and are highly stable. It is well-
known that agglomeration of particles and separation of suspensions begin at the ζ
value lower than –15 mV (Jasmani and Adnan 2017).

Preparation of adsorbents from organosolv birch wood lignin

The adsorbents were produced by organosolv lignin treatments with 0.4% ­NaHCO3
solution at room temperature or with hot water at 95 °C with the subsequent neu-
tralization and freeze drying.
To evaluate the adsorption activity, substances of different molecular weights and
chemical natures were used as markers. Iodine and methylene blue model a class
of low-molecular-weight toxicants and gelatin, the protein-binding activity in the
sorption of protein pathological agents (microorganisms and their toxins, medium-
weight molecules, and bioactive endogenous intestinal polypeptides) (Nishiama
et al. 2002).
The results of the determination of some characteristics and the adsorption capac-
ity of sorbents obtained from birch wood organosolv lignin are given in Table 6. A
commercial CJSC Signtek Polyphepan enterosorbent (St. Petersburg) obtained from
hydrolytic lignin was chosen as a reference sample.
The data obtained are indicative of the high sorption activity of the adsorbents
obtained from birch wood organosolv lignin. The iodine adsorption capacity, which
characterizes the microporous structure of a sorbent, weakly depends on the lignin
processing technique used (it is 38.3 and 42.8%) and is comparable with the value
for the commercial Polyphepan enterosorbent (38.7%).
However, the method selected for obtaining the adsorbents affects their ability to
adsorb methylene blue and gelatin. The adsorbent obtained by the lignin process-
ing with a 0.4% ­NaHCO3 solution has higher adsorption capacity than the sorbent
obtained by the lignin processing with hot water. At the same time, both sorbents
exhibit higher (around 2 times) methylene blue and gelatin adsorption capacity than
the commercial Polyphepan enterosorbent.

13
Table 6  Sorption characteristics of the adsorbents obtained from birch organosolv lignin
Synthesis method Sorption of markers Characteristics of sorbents, wt.%
Wood Science and Technology (2023) 57:173–196

I2, % Methylene blue, mg/g Gelatin, mg/g Yield Content of water-solu- Ash content
ble substances

Treatment with 0.4% ­NaHCO3 38.3 ± 1.53 97.7 ± 2.93 236.7 ± 7.10 64.8 ± 1.94 4.3 ± 0.13 0.19 ± 0.008
Treatment with hot water 42.8 ± 1.28 72.6 ± 2.18 199.5 ± 5.98 89.4 ± 2.68 6.1 ± 0.18 0.15 ± 0.005
Polyphepan (Garyntseva et al. 2011) 38.7 ± 1.16 44.0 ± 1.32 115.0 ± 3.45 – – –
191

13
192 Wood Science and Technology (2023) 57:173–196

According to the pharmacological standards, the content of water-soluble sub-


stances in enterosorbents should be no more than 5% (Markelov et  al. 2008). The
data given in Table 6 show that this requirement is met by the adsorbent obtained by
lignin processing with an aqueous solution of sodium bicarbonate.

Fractionation of birch wood into valuable chemical products using catalytic


hydrolysis and delignification processes

Based on the results obtained, a new approach to the production of valuable chemi-
cal products (xylose, cellulose, nanocelluloses, lignin and enterosorbents) from
birch wood by integration of acid hydrolysis and peroxide delignification processes
catalyzed by solid catalysts (Fig. 12) is proposed.
The solid acid catalyst ­ZrO2/SO4−2 showed the highest activity among the studied
solid catalysts in hydrolysis of birch wood hemicelluloses at a temperature of 150 °C
(xylose yield 19.6% from wood weight).
By peroxide delignification of «hemicellulose-free» birch wood in the «formic
acid− water» medium at a temperature of 100 °C in the presence of T ­ iO2 solid cata-
lyst, microcrystalline cellulose (MCC) and organosolv lignin with respective yields
of 49.6 and 4.3% from wood weight were produced.

Birch Wood

Hydrolysis at 150°C over ZrO2/SO4-2 catalyst

Xylose Lignocellulose
(yield 19.6 wt %) (yield 75.2 wt %)

Peroxide delignification at 100°C


in HCOOH‒H2O medium
over TiO2 catalyst

Organosolv Lignin Microcrystalline Cellulose


(yield 4.3 wt %) (yield 49.6 wt %)

Hydrolysis with
Treatment with solution 55-% H2SO4,
of NaHCO3 or hot water
Microfibrillated cellulose
(yield 27.7 wt %)

Adsorbents Ultrasonic treatment


(yield 3.8 wt %)
Nanocrystalline
Cellulose
(yield 10.9 wt%)

Fig. 12  Biorefinery of birch wood based on the integration of heterogeneous catalytic hydrolysis and per-
oxide delignification processes (yields on abs. dry wood)

13
Wood Science and Technology (2023) 57:173–196 193

MFC and NCC were obtained from MCC with a yield of 27.7 and 10.9%, respec-
tively by controlled acid hydrolysis and subsequent US treatment.
Organosolv lignin treated with the NaHCO3 solution or hot water can be used as
an adsorbent, which is more efficient than the commercial enterosorbent Polyphepan
produced from hydrolyzed lignin.

Conclusion

In this study, it is proposed to perform the biorefinery of birch wood into xylose,
nanocelluloses, and adsorbents by integrating the environmentally friendly pro-
cesses of acid hydrolysis over Z­ rO2/SO4− 2 solid acid catalyst and peroxide delignifi-
cation in the «formic acid− water» medium over solid T ­ iO2 catalyst.
Under the optimal conditions of heterogeneous-catalytic hydrolysis of birch wood
(temperature of 150 °C, time of 1 h, LWR 24), xylose was obtained with a yield of
up to 19.6% of the wood weight.
The optimal conditions (temperature of 100  °C, 7.2 wt.% of H ­ 2O2, 40 wt.% of
HCOOH, LWR 15, time of 4 h) of peroxide delignification of «hemicellulose-free»
birch over T­ iO2 catalyst were experimentally established, which ensure a high yield
(41.2% from wood weight) of microcrystalline cellulose.
By hydrolysis of the microcrystalline cellulose with 55% ­H2SO4 for 2 h and sub-
sequent ultrasonic treatment for 45  min, microfibrillated cellulose with a yield of
31.8% of the microcrystalline cellulose weight and nanocrystalline cellulose with a
yield of 10.9% from MCC weight were produced. The samples of microfibrillated
and nanocrystalline celluloses have degrees of polymerization of 121 and 98 and
crystallinity indices of 0.78 and 0.69, respectively. The structure of microfibril-
lated and nanocrystalline celluloses consists of fibrils with average sizes of 91.3 and
30.8  nm, respectively. Their aqueous suspensions have ζ potentials of − 24.4 and
− 30.3 mV, respectively, and are highly stable.
The organosolv lignin formed in the process of birch wood peroxide delignifica-
tion with a yield of 4.3% from wood weight has low-molecular-weight (Mn = 806 g/
mol) and contains a significant amount of oxygen-containing functional groups. The
adsorbents obtained from organosolv lignin have a high sorption capacity for meth-
ylene blue (97.7 mg/g) and gelatin (236.7 mg/g), which exceeds the capacity of the
commercial Polyphepan enterosorbent (44 and 115 mg/g).
The products of birch wood biorefinery produced by the developed methods can
be used in food and chemical industries, medicine, veterinary, and in the synthesis
of new biopolymers and composites.
Acknowledgements The FTIR, NMR, GPC, SEM, and XRD investigations were carried out on the
equipment of the Krasnoyarsk Territorial Center for Collective Use, Krasnoyarsk Scientific Center, Sibe-
rian Branch of the Russian Academy of Sciences.

CRediT authorship contribution statement Conceptualization: BNK; methodology: BNK, IGS, and


NVG; formal analysis and investigations: IGS, NVG, AVP, AMS and EVG; writing an original draft:
NVG; writing a review and editing: BNK; funding acquisition: BNK.

13
194 Wood Science and Technology (2023) 57:173–196

Funding  This study was supported by the Russian Science Foundation, project no. 21–13-00250. https://​
rscf.​ru/​proje​ct/​21-​13-​00250.

Declarations 
Conflict of interest  The authors declare that they have no conflicts of interest.

References
ASTM D1795 (2013) Standard test method for intrinsic viscosity of cellulose, ASTM International, West
Conshohocken, PA
Cao Y, Chen SS, Zhang S, Sik Ok Y, Matsagar BM, Wu KCW, Tsang DCW (2019) Advances in lignin
valorization towards bio-based chemicals and fuels: Lignin biorefinery. Biores Technol 291:121878.
https://​doi.​org/​10.​1016/j.​biort​ech.​2019.​121878
Charreau H, Foresti ML, Va ́zquez A, (2013) Nanocellulose patents trends: a comprehensive review on
patents on cellulose nanocrystals, microfibrillated and bacterial cellulose. Recent Pat Nanotechnol
7(1):56–80. https://​doi.​org/​10.​2174/​18722​10511​30701​0056
Chemin M, Wirotius AL, Ham-Pichavant F, Chollet G, Da Silva PD, Petit-Conil M, Cramail H, Grelier
S (2015) Well-defined oligosaccharides by mild acidic hydrolysis of hemicelluloses. Eur Polymer J
66:190–197. https://​doi.​org/​10.​1016/j.​eurpo​lymj.​2015
Cherubini F (2010) The biorefinery concept: using biomass instead of oil for producing energy and chem-
icals. Energy Convers Manage 51:1412–1421. https://​doi.​org/​10.​1016/j.​encon​man.​2010.​01.​015
Corrêa AC, Carmona VB, Simão JA, Galvani F, Marconcini JM, Mattoso LHC (2019) Cellulose
nanocrystals from fibers of macauba (acrocomia aculeata) and gravata (bromelia balansae) from
brazilian pantanal. Polymers 11(11):1785. https://​doi.​org/​10.​3390/​polym​11111​785
Degirmenci V, Uner D, Cinlar B, Shanks BH, Yilmaz A, van Santen RA, Hensen EJM (2011) Sulfated
zirconia modified SBA-15 catalysts for cellobiose hydrolysis. Catal Lett 141:33–42. https://​doi.​org/​
10.​1007/​s10562-​010-​0466-1
Ditzel FI, Prestes E, Carvalho BM, Demiate IM, Pinheiro LA (2017) Nanocrystalline cellulose extracted
from pine wood and corncob. Carbohydr Polym 157:1577–1585. https://​doi.​org/​10.​1016/j.​carbp​ol.​
2016.​11
Du H, Liu W, Zhang M, Si Ch, Zhang X, Li B (2019) Cellulose nanocrystals and cellulose nanofibrils
based hydrogels for biomedical applications. Carbohydr Polym 209:130–144. https://​doi.​org/​10.​
1016/j.​carbp​ol.​2019.​01.​020
Fan M, Dai D, Huang B (2012) Fourier transform infrared spectroscopy for natural fibres. In: Salih S (ed)
Fourier transform - materials analysis. InTech. https://​doi.​org/​10.​5772/​35482
Garyntseva NV, Sudakova IG, Kuznetsov BN (2011) Properties of enterosorbents obtained from acetic
acid lignins of abies, aspen and birch wood. J Sib Fed Univ Chem 2(4):121–126
Garyntseva NV, Sudakova IG, Chudina AI, Malyar YuN, Kuznetsov BN (2019) Optimization of the pro-
cess of abies wood peroxide delignification in the medium “formic acid-water” in the presence of
­TiO2 catalyst. J Sib Fed Univ Chem 12(4):522–535. https://​doi.​org/​10.​17516/​1998-​2836-​0148
Gromov NV, Medvedeva TB, Pestunov AV, Taran OP (2018) The production of formic acid from pol-
ysaccharides and biomass via one-pot hydrolysis-oxidation in the presence of Mo-V-P heteropoly
acid catalyst. J Sib Fed Univ Chem 11(1):56–71. https://​doi.​org/​10.​17516/​1998-​2836-​0058
Hindi SSZ (2017) Nanocrystalline cellulose: synthesis from pruning waste of Zizyphusspina-christi and
characterization. Nanosci Nanotechnol Res 4(3):106–114. https://​doi.​org/​10.​12691/​nnr-4-​3-4
Jasmani L, Adnan Sh (2017) Preparation and characterization of nanocrystalline cellulose from Acacia
mangium and its reinforcement potential. Carbohydr Polym 161:166–171. https://​doi.​org/​10.​1016/j.​
carbp​ol.​2016.​12.​061
Jasmani L, Thielemans W (2018) Preparation of nanocellulose and its potential application. Forest Res
7:222. https://​doi.​org/​10.​4172/​2168-​9776.​10002​22
Kargarzadeh H, Ahmad I, Abdullah I, Dufresne A, Zainudin SY, Sheltami RM (2012) Effects of hydrol-
ysis conditions on the morphology, crystallinity, and thermal stability of cellulose nanocrystals
extracted from kenafbast fibers. Cellulose 19:855–866. https://​doi.​org/​10.​1007/​s10570-​012-​9684-6

13
Wood Science and Technology (2023) 57:173–196 195

Kuznetsov BN, Sudakova IG, Celzard A, Garyntseva NV, Ivanchenko NM, Petrov AV (2011) Binding
properties of lignins obtained at oxidative delignifiation of wood and straw. J Sib Fed Univ Chem
1(4):3–10
Kuznetsov BN, Yatsenkova OV, Chudina AI, Skripnikov AM, Kozlova SA, Garyntseva NV, Ches-
nokov NV (2014) Influence of mechanical and chemical activation of microcrystalline cellulose
on its structure and reactivity in hydrolysis with solid acid catalyst SBA-15. J Sib Fed Univ Chem
7(1):122–133
Kuznetsov BN, Malyar YuN, Kuznetsova SA, Grishechko LI, Kazachenko AS, Levdansky AV, Pestunov
AV, Boyandin AN, Celzard A (2016) Isolation, study and application of organosolv lignins (review).
J Sib Fed Univ Chem 4(9):454–482. https://​doi.​org/​10.​17516/​1998-​2836-​2016-9-​4-​454-​482
Kuznetsov BN, Sudakova IG, Garyntseva NV, Djakovitch L, Pinel C (2017) Kinetic studies and optimi-
zation of abies wood fractionation by hydrogen peroxide under mild conditions with ­TiO2 catalyst.
Reac Kinet Mech Cat 120:81–94. https://​doi.​org/​10.​1007/​s11144-​016-​1100-z
Kuznetsov BN, Vasilyeva NY, Kazachenko AS, Levdansky VA, Kondrasenko AA, Malyar Y, Skvortsova
GP, Lutoshkin MA (2020) Optimization of the process of abies ethanol lignin sulfation by sulfamic
acid – urea mixture in 1,4-dioxane medium. Wood Sci Technol 54:365–381. https://​doi.​org/​10.​1007/​
s00226-​020-​01157-6
Kuznetsov BN, Sudakova IG, Garyntseva NV, Tarabanko VE, Yatsenkova OV, Djakovitch L, Rataboul
F (2021a) Processes of catalytic oxidation for the production of chemicals from softwood biomass.
Catal Today 375:132–144. https://​doi.​org/​10.​1016/j.​cattod.​2020.​05.​044
Kuznetsov BN, Sharypov VI, Baryshnikov SV, Miroshnikova AV, Taran OP, Yakovlev VA, Lavrenov
AV, Djakovitch L (2021b) Catalytic hydrogenolysis of native and organosolv lignins of aspen wood
to liquid products in supercritical ethanol medium. Catal Today 379:114–123. https://​doi.​org/​10.​
1016/j.​cattod.​2020.​05.​048
Lavoine N, Desloges I, Dufresne A, Bras J (2012) Microfibrillated cellulose –Its barrier properties and
applications in cellulosic materials: a review. Carbohydr Polym 90:735–764. https://​doi.​org/​10.​
1016/j.​carbp​ol.​2012.​05.​026
Lepetit A, Drolet R, Tolnai B, Montplaisir D, Lucas R, Zerrouki R (2017) Microfibrillated cellulose with
sizing for rein-forcing composites with LDPE. Cellulose 24(10):4303–4312. https://​doi.​org/​10.​
1007/​s10570-​017-​1429-0
Markelov DA, Nitsak OV, Gerashchenko II (2008) Comparative study of the adsorption activity of medi-
cal sorbents. Pharm Chem J 42(7):30–33. https://​doi.​org/​10.​30906/​0023-​1134-​2008-​42-7-​30-​33
Melero JA, Grieken R, Morales G (2006) Advances in the synthesis and catalytic applications of orga-
nosulfonic-functionalized mesostructured materials. Chem Rev 106:3790–3812. https://​doi.​org/​10.​
1021/​cr050​994h
Moon RJ, Martini A, Nairn J, Simonsen J, Youngblood J (2011) Cellulose nanomaterials review: struc-
ture, properties and nanocomposites. Chem Soc Rev 40:3941–3994. https://​doi.​org/​10.​1039/​C0CS0​
0108B
Nascimento SA, Rezende CA (2018) Combined approaches to obtain cellulose nanocrystals, nanofibrils
and fermentable sugars from elephant grass. Carbohydr Polym 180:38–45. https://​doi.​org/​10.​1016/j.​
carbp​ol.​2017.​09
Nishiyama Y, Langan P, Chanzy H (2002) Crystal structure and hydrogen-bonding system in cellulose Iβ
from synchrotron x-ray and neutron fiber diffraction. J Am Chem Soc 124:9074–9082. https://​doi.​
org/​10.​1021/​ja025​7319
Park S, Baker JO, Himmel ME, Parilla PA, Jonson DK (2010) Cellulose crystallinity index: measurement
techniques and their impact on integrating cellulose performance. Biotechnol Biofuels 3:10. https://​
doi.​org/​10.​1186/​1754-​6834-3-​10
Shen X, Sun R (2021) Recent advances in lignocellulose prior-fractionation for biomaterials, biochemi-
cals, and bioenergy. Carbohyd Polym 261:117884. https://​doi.​org/​10.​1016/j.​carbp​ol.​2021.​11
Shi Z, Xu G, Deng J, Dong M, Murugadoss V, Liu Ch, Shao Q, Wu S, Guo Z (2019) Structural charac-
terization of lignin from D sinicus by FTIR and NMR techniques. Green Chem Lett Rev 12(3):235–
243. https://​doi.​org/​10.​1080/​17518​253.​2019.​16274​28
Sjöström E, Alern R (1999) Analytical methods of wood chemistry pulping and papermaking. Springer-
Verlag, Berlin
Takagaki A, Nishimura M, Nishimura S, Ebitani K (2011) Hydrolysis of sugars using magnetic silica
nanoparticles with sulfonic acid groups. Chem Lett 40:1195–1197. https://​doi.​org/​10.​1246/​cl.​2011.​
1195

13
196 Wood Science and Technology (2023) 57:173–196

Takkellapati S, Li T, Gonzalez MA (2018) An overview of biorefinery-derived platform chemicals from a


cellulose and hemicellulose biorefinery. Clean Techn Environ Policy 20:1615–1630. https://​doi.​org/​
10.​1007/​s10098-​018-​1568-5
Vilcocq L, Castilho PC, Carvalheiro F, Duarte LC (2014) Hydrolysis of oligosaccharides over solid acid
catalysts: a review. Chem Sus Chem 7:1010–1019. https://​doi.​org/​10.​1002/​cssc.​20130​0720
Wu C, Bing L, Li S, Yu D, Wang D (2016) Effect of coagulating agents on lignin and oligosaccharide
contents in pre-hydrolysis liquor obtained in the production of dissolving pulp from poplar residual
slabs. Bio Res 11(1):87–94. https://​doi.​org/​10.​15376/​biores.​11.1.​87-​94
Yatsenkova O, Chudina AI, Skripnikov AM, Chesnokov NV, Kuznetsov BN (2015) The influence of sul-
furic acid catalyst concentration on hydrolysis of birch wood hemicelluloses. J Sib Fed Univ Chem
8(2):211–221. https://​doi.​org/​10.​17516/​1998-​2836-​2015-8-​2-​211-​221
Zhou S, Liu P, Wang M, Zhao H, Yang J, Xu F (2016) Sustainable, reusable, and superhydrophobic
aerogels from micro-fibrillated cellulose for highly effective oil/water separation. ACS Sustainable
Chem Eng 4(12):6409–6416. https://​doi.​org/​10.​1021/​acssu​schem​eng.​6b010​75

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under
a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted
manuscript version of this article is solely governed by the terms of such publishing agreement and
applicable law.

13

You might also like