You are on page 1of 8

Article

pubs.acs.org/Biomac

Acidic Deep Eutectic Solvents As Hydrolytic Media for Cellulose


Nanocrystal Production
Juho Antti Sirviö,* Miikka Visanko, and Henrikki Liimatainen
Fibre and Particle Engineering Research Unit, University of Oulu, P.O. Box 4300, Oulu FI-90014, Finland
*
S Supporting Information

ABSTRACT: In this study, a new method to fabricate cellulose nanocrystals (CNCs) based on DES pretreatment of wood
cellulose fibers with choline chloride and organic acids are reported. Oxalic acid (anhydrous and dihydrate), p-toluenesulfonic
acid monohydrate, and levulinic acid were studied as acid components of DESs. DESs were formed at elevated temperatures
(60−100 °C) by combining choline chloride with organic acids and were then used to hydrolyze less ordered amorphous regions
of cellulose. All the DES treatments resulted in degradation of wood fibers into microsized fibers and after mechanically
disintegrating, CNCs were successfully obtained from choline chloride/oxalic acid dihydrate-treated fibers, whereas no liberation
of CNCs was observed with other DESs. The DES-produced CNCs had a width and length of 9−17 and 310−410 nm,
respectively. The crystallinity indexes (CrIs) and carboxylic acid content of the CNCs were 66−71% and 0.20−0.28 mmol/g,
respectively. CNCs exhibited good thermal stabilities (the onset thermal degradation temperatures ranged from 275−293 °C).
The demonstrated acidic DES method exhibits certain advantages over previously reported CNC productions, namely, milder
processing conditions and easily obtainable and relatively inexpensive biodegradable solvents with low toxicity (compared, e.g., to
ILs).

■ INTRODUCTION
Ionic liquids (ILs) can be utilized as an efficient reaction and
DESs typically consist of a hydrogen bond donor and
acceptor pair. The formation of strong hydrogen bonding and
treatment media for various biomasses.1,2 ILs contain large complexation between the components is assumed to prevent
organic cations and inorganic or organic anions having melting crystallization, consequently causing the drop in the melting
point lower than 100 °C.3 Compared with traditional molecular point and resulting in an eutectic mixture.6 A significant
solvents, they have several advantages, including high solvent amount of DESs can be attained directly from low toxic natural
capacity toward both organic and inorganic materials and components, such as choline chloride, urea, and glycerol.9
negligible vapor pressure, which reduces the volatile organic However, the toxicity of DES can be higher compared to its
compound (VOC) emissions. Despite their good properties, individual components,10 but the toxicity greatly depends on
ILs have several drawbacks, including synthesis mainly from oil- the studied organism.11,12 There is even evidence that the
based chemicals, toxicity, low biodegradability, low moisture formation of DESs can lower the toxic effect of individual
tolerance, and high cost.4,5 components (e.g., organic acids). DES systems consisting of
Deep eutectic solvents (DESs) are chemicals that are almost choline chloride and organic acids have been shown to exhibit
equivalent to ILs (DESs are even regarded as a subcategory of certain antimicrobial properties, yet they are well biocompat-
ILs according to some sources6) and exhibit similar properties, ible, and their toxicity is lower compared to some ILs based on
such as good solvent capacity and low vapor pressure.7 imidazolium and pyridinium.12 Choline chloride-based DESs
However, compared with ILs, DESs can be obtained by less are also readily biodegradable.13 Consequently, these favorable
demanding methods by simply heating and mixing two or more
components at an elevated temperature (mainly around 100 Received: June 20, 2016
°C). On a larger scale, DES can be efficiently synthesized using Revised: July 25, 2016
the twin screw extrusion process.8

© XXXX American Chemical Society A DOI: 10.1021/acs.biomac.6b00910


Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

Table 1. Composition of Different DES Systems, Reaction Conditions, Yields, Limiting Viscosity, and Carboxylic Acid Content
before and after Cellulose Treatments
molar reaction time reaction temp yieldd limiting viscosity carboxylic acid content
sample acid ratioa (h) (°C) (%) (dm3/kg) (mmol/g)
b
dissolving 503
pulp
DES1 anhydrous oxalic acid 1:2 2 100 72 104 0.25 ± 0.02
DES2 oxalic acid dihydrate 1:1 2 100 68 113 0.20 ± 0.04
DES3 oxalic acid dihydrate 1:1 4 100 76 104 0.27 ± 0.13
DES4 oxalic acid dihydrate 1:1 6 100 78 104 0.23 ± 0.07
DES5 oxalic acid dihydrate 1:1 2 120 73 100 0.23 ± 0.03
c
DES6 p-toluenesulfonic acid 1:1 2 60 70 108
monohydrate
c
DES7 p-toluenesulfonic acid 1:1 4 60 66 104
monohydrate
c
DES8 levulinic acid 1:2 2 100 88 267
a
Choline chloride/acid. No carboxylic acid groups detected. Carboxylic acid content not measured. Calculated from the mass of fibers after the
b c d

DES treatment vs the mass of the original fibers.

properties have encouraged many researchers to switch their are discussed elsewhere.38 Choline chloride and p-toluenesulfonic acid
interest from traditional molecular solvents and ILs to DESs. monohydrate were obtained from TCI (Germany), and oxalic acid
Cellulose nanocrystals (CNCs) are stiff, rod-like nanosized (anhydrous and dihydrate) and levulinic acid were purchased from
(length generally below a micrometer and width from a few Sigma-Aldrich (Germany). All the chemicals were used as received,
without further purification.
nanometers to tens of nanometers) cellulosic materials isolated Hydrolysis of Cellulose Pulp Using Deep Eutectic Solvent.
from natural biomasses, such as plant fibers.14 CNCs have Molar ratios between choline chloride and acids were 1:2 for
several unique properties such as lightweight, biodegradability, anhydrous oxalic acid39 and levulinic acid40 and 1:1 for oxalic acid
stiffness, a large surface area, and are produced from cellulose, dihydrate41 and p-toluenesulfonic acid monohydrate.42 All the DESs
which is renewable and the most abundant biopolymer.15 Due were produced by heating the mixtures at 100 °C, except in the case of
to the high reinforcement capability, CNCs are widely used in p-toluenesulfonic acid monohydrate, where a temperature of 60 °C
polymer nanocomposites to improve their properties.16 CNCs was used. DES pretreatments of cellulose were performed by adding
can also be used as a replacement for toxic oil-based chemicals 1.2 g of dried pulp into 120 g of DES during mixing for the desired
in mineral flotation, 17 as a stabilizer of oil-in-water time (2−4 h) at a predetermined temperature (60−120 °C). The
reaction conditions used are presented in Table 1. After mixing, the
emulsions18,19 and in biomedicine.20 Acidic hydrolysis using
reaction mixture was removed from the heating source (oil bath) and
aqueous mineral acids, such as sulfuric,21 hydrochloric,22 100 mL of deionized water was added. The pretreated pulp
hydrobromic,23 and phosphoric acids,24 or polyoxometalates25 suspensions were filtered and washed with 400 mL of deionized water.
has been the most common method to dissolve the amorphous Fabrication of Cellulose Nanocrystals from Deep Eutectic
regions of cellulose to yield CNCs. Several other methods, Solvent-Pretreated Cellulose Fibers. DES-pretreated aqueous
mostly based on different oxidation reactions, have also been cellulose fiber dispersions (0.5%) were first mixed at 11000 rpm
applied in CNC production.26−30 Recently, ILs were shown to with an Ultra-Turrax mixer (IKA T25, Germany) at a pH of 7 (the pH
act as an efficient swelling and hydrolysis medium to fabricate was adjusted using dilute NaOH solution) for 1 min. The fibers were
CNCs.31−34 DESs have so far been reported to have been used then disintegrated using a microfluidizer (Microfluidics M-110EH-30,
U.S.A.) to individualize CNCs. Three passes through 400 and 200 μm
in pretreatment media for cellulose saccharification35,36 and in chambers at a pressure of 1300 bar and then three passes through 200
the fabrication of cellulose nanofibers (CNFs;37 by retaining and 87 μm chambers at a pressure of 2000 bar were used.
amorphous cellulose regions intact, allowing the individualiza- Scanning Electron Microscopy. Scanning electron microscopy
tion of flexible and long nanosized filaments). (SEM, Zeiss Zigma HD VP, Germany) images of the freeze-dried
In this study, we report the results of DESs obtained from (liquid nitrogen and vacuum drying) samples filtered on a
choline chloride and organic acids as potential acidic hydrolytic polycarbonate membrane with a pore size of 0.2 μm were obtained.
solvents to fabricate CNCs from dissolving pulp. For this The accelerating voltage during imaging was 0.5 kV.
purpose, oxalic (anhydrous and dihydrate), p-toluenesulfonic Transmission Electron Microscopy. The morphological features
monohydrate, and levulinic acids were used as the acidic of the fabricated CNCs were analyzed with a Tecnai G2 Spirit
transmission electron microscope (FEI Europe, Eindhoven, The
components of DESs under different hydrolytic conditions
Netherlands). Samples were prepared by diluting each sample with
(temperature and time). The CNCs obtained after DES ultrapure water. A small droplet of the dilution was placed on top of a
pretreatment and mechanical microfluidization were charac- carbon-coated copper grid. The grid was first coated with polylysine by
terized by transmission electron microscopy (TEM), wide- applying a small droplet of 0.1% solution of polylysine on the top of
angle X-ray diffraction (WAXD), diffuse reflectance infrared the grid and allowing it to stand for 3 min. The excess polylysine was
Fourier transform spectrum (DRIFT), thermogravimetric removed from the grid by touching the droplet with a corner of a filter
analysis (TGA), and conductive titration. paper. The small droplet of the sample was then placed on the top of


the grid. The excess amount of the sample was removed from the grid
by touching the droplet with a corner of a filter paper. Negative
EXPERIMENTAL SECTION staining of the samples was performed by placing a droplet of uranyl
Materials. Dissolving cellulose pulp (softwood) was obtained as acetate (2% w/v) on top of each specimen. The excess uranyl acetate
dry sheets and used as a cellulose material after disintegration in was removed with filter paper, as described above. The grids were
deionized water. The disintegrated pulp was filtered, washed with dried at room temperature and analyzed at 100 kV under standard
ethanol, and dried at 60 °C for 24 h. The properties of cellulose pulp conditions. Images were captured using a Quemesa CCD camera, and

B DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

Figure 1. Schematic flow diagram of CNC production using DES. First step: hydrolysis of wood cellulose fibers in DES at elevated temperature;
Second step: washing and filtrating the DES-treated fibers using water and vacuum filtration; Third step: mechanical disintegration of DES-treated
and washed microfibers in water with microfluidizer.

Figure 2. (a) Dry cellulose pulp before and after DES treatments and washing and (b) examples of SEM images of cellulose pulp and fibers after
DES treatments.

iTEM image analysis software (Olympus Soft Imaging Solutions cellulose was performed using DRIFT. Spectra were collected with a
GMBH, Munster, Germany) was used to measure the width of the Bruker Vertex 80v spectrometer (U.S.A.) from freeze-dried samples.
individual nanocrystals. In total, 60 CNCs of each sample were Spectra were obtained in the 600−4000 cm−1 range, and 40 scans were
measured. The final results were averaged and standard errors were taken at a resolution of 2 cm−1 from each sample.
calculated. Limiting Viscosity. The effect of the DES pretreatment on the
X-ray Diffraction. The crystalline structure of the original pulp degree of polymerization was estimated using the limiting viscosity,
and CNCs was investigated using wide-angle X-ray diffraction measured in cupriethylenediamine solution according to the ISO 5351
(WAXD). Measurements were conducted on a Rigaku SmartLab 9 standard. Samples were freeze-dried prior to the measurements.
kW rotating anode diffractometer (Japan) using a Co Kα radiation (40 Thermogravimetric Analysis. TGA measurements were carried
kV, 135 mA; λ = 1.79030 nm). Samples were prepared by pressing out with a thermal analyzer Netzsch STA 409 PC (Germany)
tablets of freeze-dried celluloses to a thickness of 1 mm. Scans were apparatus in two different atmospheres: under nitrogen flow and under
taken over a 2θ (Bragg angle) range from 5−50° at a scanning speed air flow (dynamic air), both with a constant rate of 60 mL min−1. Each
of 10°/s, using a step of 0.5°. The degree of crystallinity in terms of the measurement was made using approximately 5 mg of the freeze-dried
CrI was calculated from the peak intensity of the main crystalline plane sample, which was heated from 25−600 °C at a scanning rate of 2 °C/
(200) diffraction (I200) at 26.2° and from the peak intensity at 22.0° min−1. The temperature of polymer degradation, Td, was taken as the
associated with the amorphous fraction of cellulose (Iam), according to temperature at the onset point of the weight loss in the TGA curve
eq 1:43 obtained.
⎛ I − Iam ⎞
CrI = ⎜ 200
⎝ I200 ⎠
⎟ · 100%
(1)
■ RESULTS AND DISCUSSION
The fabrication of CNCs is commonly based on the hydrolysis
and dissolution of amorphous regions of cellulose using strong
It should be noted that, due to the Co Kα radiation source, the
cellulose peaks have different diffraction angles compared to results acids, ILs, or oxidative treatments. Here, four different DES
obtained using the Cu Kα radiation source. systems containing choline chloride as a hydrogen bond donor
Diffuse Reflectance Infrared Fourier Transform Spectrosco- and organic acids as a hydrogen bond acceptor were studied as
py. Chemical characterization of raw cellulose and DES-pretreated potential new green media for CNC individualization.
C DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

Schematic illustration of the CNC production process is


presented in Figure 1. Table 1 presents the reaction conditions
used and the yields after DES pretreatments. All DES
pretreatments notably decreased the cellulose mass (yield of
66−88%), and the lowest yield was obtained after 2 and 4 h
reaction time with oxalic acid dihydrate and p-toluenesulfonic
acid-based DESs, respectively. However, the yield was seen to
increase slightly with an oxalic acid dihydrate-based system
when reaction time increased. This may be due to the
esterification of the hydroxyl groups of cellulose with oxalic
acid, either by the formation of monoester or cross-linked
diester. In particular, the formation of diester can increase the
yield because it can prevent the dissolution of the hydrolysis
product of cellulose. However, when p-toluenesulfonic acid was
used, the yield decreased when reaction time was increased
from 2−4 h, suggesting that DES systems do not possess Figure 3. DRIFT spectra of cellulose before and after DES
significant reactivity with cellulose. pretreatments. The superimposed peak of the carbonyl vibrations of
The fiber structure of the original dissolving pulp was ester and carboxylic acid of oxalic acid containing DES-pretreated
fibers is represented by a dashed line.
partially degraded during DES pretreatment (Figure 2a), as
noticed upon visual observation, and powder-like material was
obtained (except in the case of DES8, which remained a fluffy- titration. Titration was performed only for the oxalic acid-
like material similar to raw cellulose) after pretreatment and treated fibers, as based on the DRIFT spectra no significant
drying. Some of the products appeared as slightly greyish amount of chemical modification occurred during the DES6−8
(DES5), whereas the product obtained by DES1 was brown, treatments. All the oxalic acid (anhydrous and dihydrate) DES-
indicating the formation of chromophoric cellulose degradation pretreated fibers exhibited similar carboxylic acid content
products that could not be removed during the washing step. (0.20−0.27 mmol/g; Table 1; no detectable amount of the
The degradation of the fiber structure of cellulose was acid groups was observed in the original cellulose pulp).
confirmed by SEM (Figure 2b). Oxalic acid and p- Therefore, the results suggest the carboxylic acid content of
toluenesulfonic acid-based DES pretreatments resulted in the DES-pretreated fibers cannot be significantly adjusted by
formation of fragmented cellulose fiber particles, whereas DES8 increasing the temperature or the reaction time. This might
did not significantly alter the fiber structure of cellulose. be due to the overesterification of diesters.
Limiting viscosity was used to indicate the effect of the DES All the DES-pretreated fibers were mechanically disintegrated
pretreatment on the degree of polymerization of cellulose using a microfluidizer to liberate CNCs. All the samples
fibers. The original limiting viscosity of the dissolving pulp was underwent similar mechanical treatment expect for DES1,
decreased from 503 dm3/kg to approximately 100 dm3/kg which blocked the smallest interaction chamber (IXC; 87 μm),
with DES1−7 (Table 1). These results indicate strong cellulose and DES8, which blocked even the largest IXC (200 μm).
hydrolysis during the DES pretreatment, which is typical for the Further processing of both specimens was therefore discon-
mineral acid hydrolysis of cellulose during CNC production.44 tinued. Other samples passed through the microfluidizer
No significant differences between DES pretreatments existed, without any blocking. However, it was noted that instead of
except in the case of DES8, where limiting viscosity was almost forming a stable suspension, p-toluenesulfonic acid DES-
three times higher compared to other DES-pretreated fibers. pretreated fibers (DES6−7) formed a milky white suspension
However, the limiting viscosity of DES8 was still almost half in which particle sedimentation occurred almost immediately.
that of the original pulp. The differences in the limiting All oxalic acid dihydrate DES-pretreated fibers formed slightly
viscosities between DES1−7 and DES8 are most likely caused gel-like suspensions (Figure 4; DES2−5). These suspensions
by the strong acid characteristic of oxalic acid (pKa of 1.25 and remained stable at room temperature for several weeks without
4.14 in water45) and p-toluenesulfonic acid (pKa of −2.8 in any notable sedimentation, excluding DES5, for which minor
water46) compared to levulinic acid (pKa of 4.59 in water47). sedimentation was observed after two months of free settling at
The esterification of cellulose can take place in the presence room temperature.
of acids, such as carboxylic acids. The attachment of oxalic acid Due to the sedimentation of DES6 and 7 after disintegration,
to cellulose after DES pretreatment can be seen in the DRIFT only the stable specimens (DES2−5) were studied further.
spectra in Figure 3. Compared with the cellulose pulp Based on the TEM images (Figure 5, for higher resolution
spectrum, all the oxalic acid dihydrate DES-pretreated fibers images, see Supporting Information), all the oxalic acid
(DES2−5) exhibited a new band around 1740 cm−1, which is a dihydrate DES-pretreated samples (DES2−5) exhibited indi-
superimposed peak of the carbonyl vibrations of ester and vidually occurring short CNCs. Compared with CNCs
carboxylic acid. The overlapping of the ester and carboxylic acid obtained at 100 °C, DES5 had a thin, needle-like structure
carbonyl vibrations is typical of the half-esters of dicarboxylic (average width around 10 nm), whereas DES2−3 had slightly
acids of cellulose.48 The appearance of carbonyl stretching can thicker crystals (average width around 15 nm; Table 2). There
also be seen in the spectrum of anhydrous oxalic acid DES- was no significant difference within the specimens in the length
pretreated fibers (DES1), where no difference between original of CNCs; therefore, DES5 had the highest aspect ratio (36)
fibers and p-toluenesulfonic (DES6−7) and levulinic acids compared with the others (the aspect ratios of DES2−4 ranged
(DES8) was observed. from 21−28). The width of CNCs obtained after DES
The quantity of carboxylic acid groups on the DES- pretreatments were around 2-fold higher than those of the
pretreated fibers was determined using conductometric CNCs typically obtained from sulfuric acid hydrolysis of wood
D DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

Figure 4. Aqueous solutions (0.4%) of DES treated celluloses directly after mechanical disintegration (top row) and after free settling for two
months at room temperature (bottom row).

Figure 5. TEM images of CNCs obtained from DES-pretreated cellulose fibers: (a) DES2, (b) DES3, (c) DES4, and (d) DES5 (ground circles are
due to the sample preparation, most likely due to the air bubbles trapped in the cellulose solution).

Table 2. Average Dimensions and CrIs for CNCs Obtained


after DES Pretreatment and Mechanical Disintegration
(Error Represents Standard Error)
sample width length aspect ratio CrI
DES2 13.6 ± 1.1 390 ± 25 28 69
DES3 15.7 ± 1.3 337 ± 23 21 69
DES4 13.8 ± 0.7 364 ± 22 26 66
DES5 9.9 ± 0.7 353 ± 16 36 71

pulp, but the lengths of CNCs obtained here were significantly


longer, resulting in similar or even higher aspect ratio values
compared with the traditional acid hydrolysis method.49,50 It
should be noted that no sample fractionation steps were
utilized after mechanical disintegration and therefore a minor
Figure 6. XRD diffraction patterns of original cellulose pulp and
amount of longer cellulose nanofibrils was also observed, CNCs obtained from DES-pretreated cellulose fibers.
especially with samples (DES2−3) treated in the mildest
conditions. In addition, few thicker (around 200 nm) rod-like
particles were present in each sample. Therefore, it would be nature. The CrIs of DES2−4 were 66−71% (Table 2), which
beneficial to perform some final purification (e.g., using are slightly higher compared with original cellulose pulp (66%).
centrifuge) to obtain CNC solutions with higher uniformity. A small increase in the CrI is similar to the hydrochloride acid
The X-ray diffraction patterns of the CNCs and original hydrolytic treatments reported previously.51 However, the
cellulose pulp are presented in Figure 6. The diffraction obtained CrIs are lower compared with CNCs obtained by
patterns show the characteristic peaks for cellulose I, indicating acid hydrolysis but are in line with or higher than those
no rearrangement to other cellulose allomorphs occurred. The produced by oxidative treatments28,52−54 (it should be noted
presence of the diffraction patterns of cellulose I suggest that no that the Segal method used for CrI calculations is not suitable
dissolution of cellulose fibers took place during the DES for the comparison of different types of samples but rather
pretreatment and that hydrolysis was purely heterogeneous in quantifies changes within a single sample set55). It is plausible
E DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

that crystallinities of DES-pretreated fibers are decreased during The CNC solutions (0.1% in water) exhibited transmittances
the microfluidizer treatment, as it is well-known that ranging from 45−75% at a wavelength of 800 nm (Figure 8).
mechanical treatment can decrease the crystallinity of cellulose
due to strong shear forces during homogenization, which may
loosen the crystalline structure and result in peeling of the
cellulose chains on the crystallites.
The thermal stability of CNCs was studied using TGA
measurements at air and nitrogen atmosphere. It can be seen
from Figure 7 that the onset temperatures of CNCs ranged

Figure 8. UV−vis spectra of DES-pretreated and mechanically


disintegrated CNCs (0.1%).

The transmittance increased when a longer reaction time or a


higher temperature was used in the pretreatment stage,
indicating the higher uniformity of CNCs (presumably due to
the presence of fewer fiber aggregates after more intensive DES
pretreatments). The transmittances in the UV−vis results are
similar to those of the amino-modified CNCs previously
reported by sequential oxidation and reduction treatment28 but
are lower than those of CNCs obtained by periodate oxidation
followed by chlorite oxidation56 or reductive amination with
bisphosphonate containing amine.53 Relatively low trans-
mittance, especially at the lower wavelength, indicates that
some aggregated cellulose particles (particles that might not be
fully disintegrated during the mechanical treatment or
reaggregated CNCs) remained in the solution.
The CNCs are highly promising green nanomaterials for the
emerging bioeconomy to be used in numerous applications.
CNCs themselves have low toxicity at dilute solutions57,58 and
are biodegradable.59 Therefore, in respect of the whole material
production chain, it is important that CNCs can be produced
via environmentally sustainable methods and with few working
hazards. Oxalic acid itself is a moderately toxic chemical and
Figure 7. TGA curves of original pulp and CNCs prepared after DES even though DESs containing oxalic acid has some cytotoxicity
pretreatment (a) in air and (b) in nitrogen.
properties,13 low toxicity still indicates that choline chloride and
oxalic acid-based DES is biocompatible and readily biodegrad-
able.12 Therefore, the methodology introduced here can be
from 275−293 °C. Even though this is significantly lower seen as one of the most feasible ways to obtain CNCs.
compared to original cellulose pulp (the high thermal stability However, there is still room for improvement, and other even
of dissolving pulp is most likely due to the high molecular less toxic and cheaper DESs are waiting to be discovered, which
increases the feasibility of producing CNCs.


weight and the absence of a significant amount of hemi-
celluloses), it is still in the range of thermally stable CNCs
obtained by phosphoric acid hydrolysis24 and significantly CONCLUSIONS
higher than those obtained after sequential periodate oxidation DES based on choline chloride and oxalic acid dihydrate was
and reductive amination.53 The highest thermal stability was found to be an effective hydrolytic solvent for the production of
found for DES2, whereas there were no significant differences CNCs. DESs were used only as a pretreatment media, solvent
between the other CNCs. In a nitrogen atmosphere, the onset was removed by simple filtration and washing steps, and CNCs
of the main degradation temperature was slightly higher were produced mechanically from DES-free fibers. The use of
(around 300 °C for all CNCs) compared to the measurement DES as a pretreatment media thus eliminates cumbersome
in air. However, minor degradation (around 10% w) was separation of CNCs from the reaction mixture. It was found
observed after 200 °C. The residual mass at 600 °C for CNCs that even 2 h pretreatment time at 100 °C was suitable for the
was over two times higher compared with original cellulose fabrication of CNCs with a high aspect ratio, and the width
pulp. The formation of residual char is most likely due to the could be adjusted by increasing the reaction temperature. Due
dehydration of cellulose by acidic oxalic acid groups. to the low toxicity, biocompatibility, and biodegradability, DES
F DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

pretreatment can be seen as an environmentally friendly (19) Ojala, J.; Sirviö, J. A.; Liimatainen, H. Chem. Eng. J. 2016, 288,
method for CNC production. 312−320.


(20) Lin, N.; Dufresne, A. Eur. Polym. J. 2014, 59, 302−325.
ASSOCIATED CONTENT (21) Bondeson, D.; Mathew, A.; Oksman, K. Cellulose 2006, 13,
171−180.
*
S Supporting Information (22) Yu, H.; Qin, Z.; Liang, B.; Liu, N.; Zhou, Z.; Chen, L. J. Mater.
The Supporting Information is available free of charge on the Chem. A 2013, 1, 3938−3944.
ACS Publications website at DOI: 10.1021/acs.bio- (23) Sadeghifar, H.; Filpponen, I.; Clarke, S. P.; Brougham, D. F.;
mac.6b00910. Argyropoulos, D. S. J. Mater. Sci. 2011, 46, 7344−7355.
(24) Camarero Espinosa, S.; Kuhnt, T.; Foster, E. J.; Weder, C.
TEM images of CNCs with different magnitutes (PDF). Biomacromolecules 2013, 14, 1223−1230.


(25) Liu, Y.; Wang, H.; Yu, G.; Yu, Q.; Li, B.; Mu, X. Carbohydr.
AUTHOR INFORMATION Polym. 2014, 110, 415−422.
(26) Qin, Z.-Y.; Tong, G.-L.; Frank Chin, Y. C.; Zhou, J.-C.
Corresponding Author BioResources 2011, 6 (2), 1136−1146.
*Tel.: +358294482424. Fax: +358 855 323 27. E-mail: juho. (27) Leung, A. C. W.; Hrapovic, S.; Lam, E.; Liu, Y.; Male, K. B.;
sirvio@oulu.fi. Mahmoud, K. A.; Luong, J. H. T. Small 2011, 7, 302−305.
Notes (28) Visanko, M.; Liimatainen, H.; Sirviö, J. A.; Heiskanen, J. P.;
The authors declare no competing financial interest. Niinimäki, J.; Hormi, O. Biomacromolecules 2014, 15, 2769−2775.


(29) Tejado, A.; Alam, M. N.; Antal, M.; Yang, H.; van de Ven, T. G.
M. Cellulose 2012, 19, 831−842.
ACKNOWLEDGMENTS (30) Sirviö, J. A.; Visanko, M. T.; Heiskanen, J. P.; Liimatainen, H. J.
The following people are gratefully acknowledged for their Mater. Chem. A 2016, 4, 6368−6375.
contribution to this work: Ms. Vilma Haarala for her (31) Mao, J.; Heck, B.; Reiter, G.; Laborie, M.-P. Carbohydr. Polym.
contribution in the experimental part of the study, Dr. Ilkka 2015, 117, 443−451.
Miinalainen for his help with the TEM and SEM images, Mr. (32) Mao, J.; Osorio-Madrazo, A.; Laborie, M.-P. Cellulose 2013, 20,
1829−1840.
Tommi Kokkonen for TG analysis, and Mr. Sami Saukko for
(33) Man, Z.; Muhammad, N.; Sarwono, A.; Bustam, M. A.; Vignesh
his guidance with WAXD analysis. The facilities at the Center Kumar, M.; Rafiq, S. J. Polym. Environ. 2011, 19, 726−731.
of Microscopy and Nanotechnology at the University of Oulu (34) Tan, X. Y.; Abd Hamid, S. B.; Lai, C. W. Biomass Bioenergy
were utilized in this research.


2015, 81, 584−591.
(35) Xia, S.; Baker, G. A.; Li, H.; Ravula, S.; Zhao, H. RSC Adv. 2014,
REFERENCES 4, 10586−10596.
(1) Brandt, A.; Gräsvik, J.; Hallett, J. P.; Welton, T. Green Chem. (36) Kumar, A. K.; Parikh, B. S.; Pravakar, M. Environ. Sci. Pollut. Res.
2013, 15, 550−583. 2016, 23, 1−11.
(2) de Oliveira, H. F. N.; Farès, C.; Rinaldi, R. Chem. Sci. 2015, 6, (37) Sirviö, J. A.; Visanko, M.; Liimatainen, H. Green Chem. 2015, 17,
5215−5224. 3401−3406.
(3) Petkovic, M.; Seddon, K. R.; Rebelo, L. P. N.; Pereira, C. S. Chem. (38) Sirvio, J.; Hyvakko, U.; Liimatainen, H.; Niinimaki, J.; Hormi, O.
Soc. Rev. 2011, 40, 1383−1403. Carbohydr. Polym. 2011, 83, 1293−1297.
(4) Egorova, K. S.; Ananikov, V. P. ChemSusChem 2014, 7, 336−360. (39) Habibi, E.; Ghanemi, K.; Fallah-Mehrjardi, M.; Dadolahi-
(5) Liwarska-Bizukojc, E.; Maton, C.; Stevens, C. V. Biodegradation Sohrab, A. Anal. Chim. Acta 2013, 762, 61−67.
2015, 26, 453−463. (40) Maugeri, Z.; de María, P. D. RSC Adv. 2012, 2, 421−425.
(6) Hammond, O. S.; Bowron, D. T.; Edler, K. Green Chem. 2016, (41) Jablonský, M.; Škulcová, A.; Kamenská, L.; Vrška, M.; Šíma, J.
18, 2736−2744. BioResources 2015, 10, 8039−8047.
(7) Smith, E. L.; Abbott, A. P.; Ryder, K. S. Chem. Rev. 2014, 114, (42) Assanosi, A. A.; Farah, M. M.; Wood, J.; Al-Duri, B. RSC Adv.
11060−11082. 2014, 4, 39359−39364.
(8) Crawford, D. E.; Wright, L. A.; James, S. L.; Abbott, A. P. Chem. (43) Segal, L.; Creely, J. J.; Martin, A. E.; Conrad, C. M. Text. Res. J.
Commun. 2016, 52, 4215−4218. 1959, 29, 786−794.
(9) Dai, Y.; van Spronsen, J.; Witkamp, G.-J.; Verpoorte, R.; Choi, Y. (44) Sun, B.; Zhang, M.; Hou, Q.; Liu, R.; Wu, T.; Si, C. Cellulose
H. Anal. Chim. Acta 2013, 766, 61−68. 2016, 23, 439−450.
(10) de Morais, P.; Gonçalves, F.; Coutinho, J. A. P.; Ventura, S. P. (45) Knowles, C. F.; Hodgkinson, A. Analyst 1972, 97, 474−481.
M. ACS Sustainable Chem. Eng. 2015, 3, 3398−3404. (46) Guthrie, J. P. Can. J. Chem. 1978, 56, 2342−2354.
(11) Juneidi, I.; Hayyan, M.; Hashim, M. A. RSC Adv. 2015, 5, (47) Kopetzki, D.; Antonietti, M. Green Chem. 2010, 12, 656−660.
83636−83647. (48) Li, W. Y.; Jin, A. X.; Liu, C. F.; Sun, R. C.; Zhang, A. P.;
(12) Zhao, B.-Y.; Xu, P.; Yang, F.-X.; Wu, H.; Zong, M.-H.; Lou, W.- Kennedy, J. F. Carbohydr. Polym. 2009, 78, 389−395.
Y. ACS Sustainable Chem. Eng. 2015, 3, 2746−2755. (49) Sacui, I. A.; Nieuwendaal, R. C.; Burnett, D. J.; Stranick, S. J.;
(13) Radošević, K.; Cvjetko Bubalo, M.; Gaurina Srček, V.; Grgas, D.; Jorfi, M.; Weder, C.; Foster, E. J.; Olsson, R. T.; Gilman, J. W. ACS
Landeka Dragičević, T.; Radojčić Redovniković, I. Ecotoxicol. Environ. Appl. Mater. Interfaces 2014, 6, 6127−6138.
Saf. 2015, 112, 46−53. (50) Beck-Candanedo, S.; Roman, M.; Gray, D. G. Biomacromolecules
(14) Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, 2005, 6, 1048−1054.
M.; Gray, D.; Dorris, A. Angew. Chem., Int. Ed. 2011, 50, 5438−5466. (51) Palme, A.; Theliander, H.; Brelid, H. Carbohydr. Polym. 2016,
(15) Habibi, Y. Chem. Soc. Rev. 2014, 43, 1519−1542. 136, 1281−1287.
(16) Mariano, M.; El Kissi, N.; Dufresne, A. J. Polym. Sci., Part B: (52) Yang, H.; van de Ven, T. G. M. Cellulose 2016, 23, 1−11.
Polym. Phys. 2014, 52, 791−806. (53) Sirviö, J. A.; Hasa, T.; Ahola, J.; Liimatainen, H.; Niinimäki, J.;
(17) Laitinen, O.; Kemppainen, K.; Ä mmälä, A.; Antti Sirviö, J.; Hormi, O. Carbohydr. Polym. 2015, 133, 524−532.
Liimatainen, H.; Niinimäki, J. Ind. Eng. Chem. Res. 2014, 53, 20092− (54) Sirviö, J. A.; Visanko, M.; Laitinen, O.; Ä mmälä, A.; Liimatainen,
20098. H. Carbohydr. Polym. 2016, 136, 581−587.
(18) Zoppe, J. O.; Venditti, R. A.; Rojas, O. J. J. Colloid Interface Sci. (55) Ahvenainen, P.; Kontro, I.; Svedström, K. Cellulose 2016, 23,
2012, 369, 202−209. 1073−1086.

G DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX
Biomacromolecules Article

(56) Sirviö, J. A.; Honkaniemi, S.; Visanko, M.; Liimatainen, H. ACS


Appl. Mater. Interfaces 2015, 7, 19691−19699.
(57) Hanif, Z.; Ahmed, F. R.; Shin, S. W.; Kim, Y.-K.; Um, S. H.
Colloids Surf., B 2014, 119, 162−165.
(58) Du, L.; Arnholt, K.; Ripp, S.; Sayler, G.; Wang, S.; Liang, C.;
Wang, J.; Zhuang, J. Ecotoxicology 2015, 24, 2049−2053.
(59) Kümmerer, K.; Menz, J.; Schubert, T.; Thielemans, W.
Chemosphere 2011, 82, 1387−1392.

H DOI: 10.1021/acs.biomac.6b00910
Biomacromolecules XXXX, XXX, XXX−XXX

You might also like