You are on page 1of 9

Carbohydrate Polymers 213 (2019) 208–216

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Preparation of cellulose nanomaterials via cellulose oxalates T


⁎ ⁎
Jonatan Henschen , Dongfang Li, Monica Ek
Department of Fiber and Polymer Technology, KTH Royal Institute of Technology, Teknikringen 56-58, SE-100 44, Stockholm, Sweden

A R T I C LE I N FO A B S T R A C T

Keywords: Nanocellulose prepared from cellulose oxalate has been discussed as an alternative to other methods to prepare
Nanocellulose cellulose nanofibrils or crystals. The current work describes the use of a bulk reaction between pulp and oxalic
Cellulose oxalate acid dihydrate to prepare cellulose oxalate followed by homogenization to produce nanocellulose. The prepared
Oxalic acid dihydrate nanocellulose is on average 350 nm long and 3–4 nm wide, with particles of size and shape similar to both
Cellulose
cellulose nanofibrils and cellulose nanocrystals. Films prepared from this nanocellulose have a maximum tensile
Cellulose nanofibrils
Cellulose nanocrystals
stress of 140–200 MPa, strain at break between 3% and 5%, and oxygen permeability in the range of 0.3–0.5 cm3
μm m−2 day−1 kPa−1 at 50% relative humidity. The presented results illustrate that cellulose oxalates may be a
low-cost method to prepare nanocellulose with properties reminiscent of those of both cellulose nanofibrils and
cellulose nanocrystals, which may open up new application areas for cellulose nanomaterials.

1. Introduction fibrils and reduces the energy required to liberate them. One frequently
suggested application of CNFs and CNCs are as thin films. The differ-
As awareness of our impact on the planet is increasing, the demand ence in size between CNFs and CNCs makes the fabrication processes
for new sustainable materials has grown rapidly in recent years, as different. CNFs can be prepared either by solvent casting or by filtration
evidenced by the UN agenda 2030 (Assembly U.G., 2015). One raw over a fine membrane, while CNCs often is prepared by solvent casting.
material that has the potential to be used in many new materials is Films prepared from these materials have high transparency (Aulin,
wood and cellulosic pulp fibre derived from wood. By isolation of the Salazar-Alvarez, & Lindström, 2012; Saito et al., 2009), high specific
nanostructures that make up wood fibres, it is possible to obtain ma- strength (Saito et al., 2009; Syverud & Stenius, 2008) and are good
terials with highly interesting properties, such as high specific strength oxygen barriers. The properties of these films are however strongly
and stiffness (Tanpichai et al., 2012), high surface area (Moon, Martini, affected by humidity which for some cases means that they require
Nairn, Simonsen, & Youngblood, 2011), biodegradability (Jung et al., modification or mixing of the nanocellulose with other components to
2015) and low toxicity (Alexandrescu, Syverud, Gatti, & Chinga- reduce this effect (Aulin et al., 2012; Zhang, Zhang, Lu, & Deng, 2012).
Carrasco, 2013). Some suggested uses of these films are in food (Arora & Padua, 2010)
The nanomaterials isolated from pulp fibres are generally categor- and electronic (Jung et al., 2015) applications.
ized as either cellulose nanocrystals (CNCs) or cellulose nanofibrils In the pursuit of lowering production costs and changing the
(CNFs). CNCs from wood are rod-like particles with a length of properties of nanocellulose, many alternative pre-treatments and
50–500 nm and a width between 3 and 5 nm (Moon et al., 2011). These methods to produce these materials have been published. CNFs have
particles are obtained through acid hydrolysis (Rånby, 1951) of cellu- been produced with varying pre-treatments that either reduce the cel-
lose fibres to remove all amorphous and less ordered regions in the lulose chain length, e.g., enzymatic pre-treatment (Henriksson,
fibre, leaving only the crystalline regions of cellulose. This fabrication Henriksson, Berglund, & Lindstrom, 2007), or introduce electrostatic
process produces materials with high crystallinity ranging between 54 charges to cellulose, e.g., TEMPO-mediated oxidation (Saito, Kimura,
and 88% (Moon et al., 2011). CNFs, on the other hand, contain both the Nishiyama, & Isogai, 2007), phosphorylation and carboxymethylation
amorphous and crystalline regions of the nanostructures in wood. This (Wågberg et al., 2008). CNC has been produced with varying types and
makes the CNFs long and flexible with a length of 500–2000 nm and a concentrations of acid, including both organic and mineral acids, e.g.,
width between 4–20 nm (Moon et al., 2011). CNFs are produced oxalic acid (Chen, Zhu, Baez, Kitin, & Elder, 2016; Li, Henschen, & Ek,
through mechanical refining of pulp fibres, which is usually preceded 2017), citric acid (Spinella et al., 2016) and hydrochloric acid (Araki,
by a chemical pre-treatment that reduces the adhesive forces between Wada, Kuga, & Okano, 1998). At low concentrations, these acids


Corresponding authors.
E-mail addresses: hens@kth.se (J. Henschen), dongfan@kth.se (D. Li), monicaek@kth.se (M. Ek).

https://doi.org/10.1016/j.carbpol.2019.02.056
Received 15 November 2018; Received in revised form 20 January 2019; Accepted 16 February 2019
Available online 18 February 2019
0144-8617/ © 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

catalyse the hydrolysis of the amorphous regions of cellulose fibril ag- Table 1
gregates. At higher concentrations, they can also react with cellulose to The prepared samples, their names and reaction parameters. All reactions were
introduce new functionality. Oxalic acid has been used both at low and performed by mixing 20 g of pulp with 77 g of oxalic acid dihydrate at 110 °C.
high concentrations to produce nanocellulose in different systems. The Name Raw material Reaction time (min) Washing
first publication describing the production of nanocellulose using
highly concentrated oxalic acid used molten oxalic acid dihydrate (Ek SWD35E Dissolving pulp 35 Ethanol
SWD35A Dissolving pulp 35 Acetone
et al., 2015); later publications using oxalic acid as a 70% water solu-
SWD60E Dissolving pulp 60 Ethanol
tion (Chen et al., 2016) and the dissolution of oxalic acid in eutectic SWD60A Dissolving pulp 60 Acetone
solvents (Sirvio, Visanko, & Liimatainen, 2016) have appeared. Oxalic SWK35E Kraft pulp 35 Ethanol
acid dihydrate contains 29% water of crystallization and has a melting SWK35A Kraft pulp 35 Acetone
SWK60E Kraft pulp 60 Ethanol
point of 101 °C, making reactions with pulp fibres to produce cellulose
SWK60A Kraft pulp 60 Acetone
oxalates without additional solvent possible. The authors have pre-
viously published a short communication where molten oxalate dihy-
drate was used to prepare cellulose oxalate. This material was shortly (2014). Initially, all samples were passed through two large chambers
shown to be a suitable derivative to prepare CNCs at high yield using of 400 μm and 200 μm connected in series. Following this, they were
ultrasonication (Li et al., 2017). The current paper further investigates passed multiples times through a 200 μm and a 100 μm interaction
the possibility to prepare high-quality cellulosic nanomaterials using chamber connected in series. The samples were passed 1, 3 or 5 times
oxalic acid dihydrate from different types of pulp, reaction conditions through the small chambers, and are denoted as 1p, 3p or 5p, respec-
and more sustainable washing procedures. Using microfluidization, tively. The pressure was set to 925 bar when passing the samples
nanocellulose was prepared with particles resembling both CNFs and through the large chambers and 1,600 bar when passing through the
CNCs. These materials characterized and found suitable to prepare films large chambers.
using vacuum filtration with good mechanical and oxygen barrier
properties.
2.2.2. Characterization
2. Experimental Fourier transform infrared spectroscopy (FTIR) spectra of the cel-
lulose oxalates were measured with a PerkinElmer Spectrum 2000 FT-
2.1. Materials IR (PerkinElmer, USA) equipped with a heat-controlled single reflection
attenuated total reflection (ATR) accessory (Golden Gate heat-con-
Softwood sulphite dissolving pulp sheets with a cellulose content of trolled); 32 scans were recorded for each spectrum.
≥96 wt % were supplied by Domsjö fabriker (Aditya Birla, Domsjö, The nanocellulose content, represented by the amount of colloidal
Sweden). Softwood kraft pulp sheets (Imperial Anchor) with a cellulose particles, was calculated for SWD35A_5p and SWK35A_5p. This was
content of 84 wt % were supplied by Holmen AB (Iggesund, Sweden). done by diluting the nanocellulose to 2 g L−1 and dispersing them using
Oxalic acid dihydrate (≥99%) was purchased from Sigma Aldrich a T18 UltraTurrax (IKA, Germany) at 16,000 RPM for 10 min before the
(Merck KGaA, Darmstadt, Germany). Ethanol (96%) and acetone suspensions were centrifuged at 2,500 RCF for 1 h. The solid content of
(≥99.5%) were supplied by VWR International AB (Stockholm, the supernatant was determined and compared to the solid content
Sweden). Commercially available food-grade olive oil was acquired at a prior to centrifugation. The total nanocellulose yield was calculated by
local supermarket (ICA, Stockholm, Sweden). multiplying the nanocellulose content with the gravimetric yield.
The free carboxyl content (FCC) of each cellulose oxalate was de-
2.2. Methods termined by conductometric titration as previously described (Habibi,
Chanzy, & Vignon, 2006; Li et al., 2017). A suspension was prepared by
2.2.1. Preparation of cellulose oxalate mixing 100 mg of each cellulose oxalate with 100 ml of water and 10 ml
The dry pulp sheets were torn by hand into pieces of approximately of a 0.01 M NaCl solution. The suspension was then stirred for 1 h and
1 × 1 cm2. An evaporation flask was filled with 20 g of the torn pulp titrated with 0.01 M NaOH. All titrations were carried out under con-
and 77 g of oxalic acid dihydrate. To improve mixing, 30 glass balls (Ø: stant nitrogen bubbling. The calculation of the content of free carboxyl
15–16 mm) were added to the flask. The flask was attached to a rotary groups is based on the equation below:
evaporator and lowered into a heating bath set to 110 °C while rotated CNaOH × VNaOH
for either 35 or 60 min. After the desired time was reached, the mixture Free carboxyl content =
m
was lifted from the oil bath and allowed to cool at room temperature
while still rotating. After cooling down and solidifying, either ethanol where CNaOH is the exact concentration (mol/l) of the NaOH solution,
or acetone was added to dissolve excess oxalic acid, and the solution VNaOH is the exact volume (l) of the NaOH solution used for titration
was allowed to mix overnight. The added solvent was removed by va- before the conductivity increased from the plateau of the titration curve
cuum filtration, and the remaining solids were washed further by fil- and m is the dry weight (g) of the oxalate sample.
tering either acetone or ethanol through the material until the filtrate The morphologies of SWK35E_3p and SWD60E_5p were studied by
reached a conductivity below 10 μS cm−1. At that point, the remaining adsorbing the samples on silica wafers and imaging them using scan-
solids were collected as cellulose oxalate and allowed to dry at 40 °C in ning electron microscopy (SEM) and atomic force microscopy (AFM).
an oven. The product was then obtained as a dry powder. The different The samples were adsorbed onto the silicon wafers using a binding
samples are denoted as seen in Table 1. The gravimetric yield of cel- layer of polyethylenimine. After washing and plasma treating the wa-
lulose oxalate was calculated based on the dry weight of the pulp fibres fers they were dipped in polyethylenimine solution (1 g L−1), rinsed
and cellulose oxalate. with deionized water, dried in a flow of nitrogen, quickly dipped in
To prepare nanocellulose out of cellulose oxalates, two of the dilute suspensions of nanocellulose, rinsed with deionized waster and
samples (SWK35E and SWD60E) were dispersed in deionized water at a dried in a flow of nitrogen. Just before imaging, the samples using an S-
concentration of approximately 2 wt %. To fully dissociate the car- 4800 field emission SEM (Hitachi, Tokyo, Japan), they were coated
boxylic acids and aid the fibrillation, the pH of the dispersions was with a 5 nm platinum/palladium coating in a 208HR high-resolution
adjusted to pH 9–10 using 0.1 M sodium hydroxide before they were sputter coater (Cressington, Watford, UK) to suppress specimen char-
mechanically disintegrated using a microfluidizer (M-110EH, ging. AFM images were acquired using tapping mode on a Multimode
Microfluidics Corp, United States)similarly as described by Khan et al. IIIa instrument (Bruker, Santa Barbara, CA, United States). The images

209
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

were recorded under ambient conditions (23 °C and 50% relative hu-
midity) in air using Scanassyst cantilevers (Bruker, Santa Barbara, CA,
United States) with a nominal resonance frequency of 70 kHz and a
spring constant of 0.4 N m−1. The images captured by AFM were used
to determine the dimensions of the nanocellulose. The width was Fig. 1. Potential esterification reaction of the cellulose with oxalic acid dihy-
measured using Nanoscope Analysis software (Bruker, Santa Barbara, drate; during the reaction, the cellulose also undergoes acid-catalysed hydro-
CA, United States) by measuring the height of 223 (SWK35E_3p) and lysis (not shown).
307 (SWD60E_5p) particles. The length was measured using ImageJ
(Schneider, Rasband, & Eliceiri, 2012) by tracking 89 (SWK35E_3p) and more sterically available for chemical reactions, which makes them
59 (SWD60E_5p) particles. more labile than those in cellulose in acid-catalysed hydrolysis and
Thermogravimetric analysis (TGA) was performed on a Mettler dehydration. By washing the acid/pulp mixture with filtration it is
Toledo TGA/SDTA 851e to study the thermal properties of the cellulose possible to eliminate the need for dialysis, this also facilitates recycling
oxalates. Approximately 4 mg of each sample was heated from 30 to the acid as it will be diluted less during washing. Both high demands for
600 °C at a rate of 10 °C min−1 under N2 flow (flow rate 50 ml min−1). chemical recovery and long washing procedures are two mayor ob-
The data were analysed by Mettler-STARe software. stacles when preparing conventional sulphuric acid based CNCs (Nelson
X-ray diffraction (XRD) was performed on a PANalytical X’Pert PRO et al., 2016). Preparation of CNCs using hydrochloric acid can also be
MRD X-ray diffractometer (The Netherlands) equipped with a done without the need for dialysis, unlike cellulose oxalate, the re-
PANalytical X’Celerator detector. Diffractograms were collected in a sulting material however has low surface charge and subsequently poor
range of 2θ = 10 to 50°. Cu Kα radiation was monochromatized with a colloidal stability (Yu et al., 2013).
nickel filter. The crystallinity index was estimated as described by Oxalic acid has the potential to participate in the simultaneous
Segal, Creely, Martin, and Conrad (1959). hydrolysis and esterification of cellulose (Fig. 1) (Li et al., 2017), due to
the fibre structure it is difficult for the acid to fully penetrate the fibre
2.2.3. Films and react with all cellulose chains. Esterification is most likely to occur
Films were prepared from nanocellulose suspensions (0.2 wt %) on C6-OH due to its greater steric accessibility than C2-OH and C3-OH
dispersed using a T18 UltraTurrax (IKA, Germany) at 14,000 RPM for (Fox, Li, Xu, & Edgar, 2011). Esterification with a difunctional car-
20 min before they were degassed by sonication in an ULTRAsonik 28X boxylic acid can result in both the introduction of charged acid groups
ultrasonic bath (NeyDental, Inc., Bloomfield, CT, USA) for 20 min. and crosslinking between cellulose chains. Nevertheless, our previous
Films of ˜50 g m−2 were produced using vacuum filtration over a work has indicated that most reacted oxalic acid leaves a charged group
0.45 μm PVDF Durapore membrane filter (Merck Millipore) in a Rapid- on cellulose (Li et al., 2017).
Köthen sheet former (Paper Testing Instruments, Austria). Water was The oxalate functionality contains two carbonyl groups, which can
drained from the suspensions until a stable gel-like material was left. At be observed using FTIR. In all samples, a broad adsorption at
this point, a second membrane was placed on top of the gel, and both 1739 cm−1 corresponding to the C]O stretching of the carbonyl group
membranes were placed between two sheet-former carrier boards. The was observed. This indicates successful esterification of the cellulose
whole assembly was dried for 20 min at 93 °C and 95 kPa reduced oxalates. In Fig. 2, the spectra of the raw materials and SWK35E,
pressure. SWK35 A, SWD60A and SWK60A are shown; spectra for all samples are
The oxygen permeability of the above CNF films was measured found in the supplementary information (ESI).
using a Mocon Oxtran 2/20 (Modern Controls Inc., Minneapolis, MN) The signal of carboxyl acids (−COOH) and ester bonds (−COO−)
instrument equipped with a coulometric sensor. The samples were overlaps and is difficult to differentiate in the FTIR spectra. Moreover,
masked using aluminium foil, leaving a round area of 5 cm2 exposed for no clear trends were observed in the intensities of the carbonyl signals
the measurement. The measurements were conducted at 23 °C and at of the prepared samples. To determine the effect of the reaction time,
50% RH. Each sample was conditioned prior to the measurement. Two raw material, and washing liquor on the degree of esterification, con-
replicates for every sample were measured. The tensile properties of the ductometric titrations were performed to quantify the free carboxyl
films were evaluated on an Instron Universal testing machine 5944
fitted with a 500 N load. The initial distance between the test grips was
20 mm, and the width of each specimen was 5 mm. At least 5 specimens
from each sample were tested. The separation rate was 2 mm min−1.
The samples were conditioned and tested at 23 °C and 50% RH. The
stress-strain curve was recorded for each test, and the data were aver-
aged over all specimens. Young’s modulus was automatically calculated
by Bluehill version 3.72 (Illinois Tool Works Inc., Illinois, USA) soft-
ware.

3. Results and discussion

Cellulose oxalates were prepared by mixing pulp with molten oxalic


acid dihydrate. As the reaction progressed, the appearance of the pulp
changed from a white to a greyish colour, followed by the torn pulp
sheet pieces losing their original shape as the mixture turned into a dark
yellow or brown paste. As the reaction cooled, the oxalic acid solidified,
and ethanol or acetone was added to remove excess oxalic acid. For the
experiments, both dissolving pulp and kraft pulp were used as raw
materials (Table 1). The observed colour change was more prominent
for samples prepared from kraft pulp, presumably due to its higher Fig. 2. FTIR spectra of a selection of the cellulose oxalates and raw materials.
amount of hemicellulose. Owing to the less ordered structural con- The vertical line in the inset is at wavenumber 1739 cm−1 corresponding to the
formation of hemicellulose compared to that of cellulose (Cai, Zhang, C]O stretching of the carbonyl group. Spectra for all the remaining samples are
Charles, & Wyman, 2014), the hydroxyl groups of hemicellulose are available in the supplementary information (ESI).

210
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

Table 2 was considerably lower at 0.62 mmol g−1, corresponding to an esti-


The yield, free carboxyl content, degree of substitution and thermal degradation mated degree of substitution (DS) of 0.13-0.16. No clear trends among
onset temperature for the cellulose oxalates. the FCCs of the cellulose oxalates washed with the same type of solvent
Yield [%] Free carboxyl Estimated degree Thermal could be observed despite the differences in reaction time and type of
content [mmol of substitution degradation onset raw materials. As has been reported for the esterification of cellulose by
g−1] temperature [°C] oxalic acid dihydrate, the DS on cellulose does not change when the
reaction time is prolonged from 30 to 60 min (Li et al., 2017). In our
SWD35E 93.8 0.86 0.15 175
SWD35A 97.1 1.05 0.18 176 previous work, we discussed that when no procedures for cellulose
SWD60E 85.4 0.62 0.11 175 activation (e.g., swelling) are carried out, substitution only occurs on
SWD60A 96.4 1.08 0.19 175 the accessible surfaces of cellulose fibrils. The highest theoretical DS of
SWK35E 83.6 0.92 0.16 177
cellulose in this case is 0.05 (Larsson, Hult, Wickholm, Pettersson, &
SWK35A 96.0 1.10 0.19 173
SWK60E 85.6 0.97 0.17 176
Iversen, 1999; Pu, Ziemer, & Ragauskas, 2006). In the current study, all
SWK60A 99.3 1.04 0.18 N/A DS values exceeded this value, which indicates the occurrence of es-
terification on both accessible and inaccessible surfaces of the cellulose
fibrils. Nevertheless, as the reaction time increased from 35 min to
60 min, esterification could not proceed further. This could be attrib-
uted to the strong intermolecular interactions among the cellulose
chains, which limited the migration of molten oxalic acid further to-
wards the crystalline regions of cellulose fibrils. Other methods to
prepare nanocellulose by reacting oxalic acid with pulp fibres produced
materials with FCC between 0.11 – 0.39 mmol g-1 (Chen et al., 2016;
Sirvio et al., 2016), this is considerably lower than the values observed
using the current procedure. It is believed that the higher FCC is ob-
tained due to not adding water to quench the reaction. At elevated
temperatures, introducing additional water in this reaction is favorable
for the reverse reaction (hydrolysis on the formed ester bonds) but not
the forward reaction (esterification), due to the change in chemical
equilibrium. An increased FCC will facilitate the fibrillation when
preparing the nanocellulose.
In addition, the FCCs of samples washed with acetone were slightly
Fig. 3. TG (left axis) and DTG (right axis) curves for SWD60A at 10 °C min−1
under N2 flow. The curves for the other samples are similar and are available in higher than those of samples washed with ethanol. During washing, it is
the supplementary information (ESI). hypothesized that smaller fractions of cellulose oxalates with high DS
can interact rather well with polar solvents and thus can produce
smaller particles. The particles may be fine enough to either be en-
content of each sample.
trapped in the filter paper or possibly pass through it. Consequently,
As shown in Table 2, the FCC of all samples ranged between
these fractions were removed from the samples after washing. As
0.86–1.1 mmol g−1, except for SWD60E, which for unknown reasons
ethanol is more polar than acetone, this is more likely to occur when

Fig. 4. Nanocellulose suspension at approximately 1.5 wt % prepared from the cellulose oxalates passed through the small chambers in the microfluidizer for either 1
passes (top) or 3 passes (bottom). SWK35E_3p was never collected.

211
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

Fig. 5. Representative SEM (top) and AFM (middle) micrographs of nanocellulose adsorbed onto silica surfaces from suspensions of SWK35E_3p (left) and
SWD60E_5p (right). Distribution (bottom) width (left) and length (right) of SWD60E_5p and SWK35E_3p measured using AFM.

washing cellulose oxalate with ethanol. with ethanol resulted in lower yields in all samples, which could be
As reported earlier (Li et al., 2017), prolonging the reaction time attributed to the removal of small fractions of hydrolysed and esterified
from 35 min to 60 min when using dissolving pulp slightly decreases the cellulose and hemicellulose during washing, as discussed above.
yield (Table 2). This is thought to be due to increased hydrolysis with Thermal degradation of cellulose oxalates was detected by TGA to
possible dissolution of some of the degradation products. This trend is determine the onset temperature and the maximum degradation tem-
not observed when using kraft pulp as a raw material; for kraft pulp, the peratures of the main mass-loss regions. Three different mass-loss re-
yield increases slightly when increasing the reaction time from 30 min gions could be observed (Fig. 3). The first mass-loss region (30–130 °C)
to 60 min. There is no clear difference in the gravimetric yield between could be attributed to the loss of physically adsorbed water in cellulose
the samples prepared from kraft or dissolving pulp, despite having oxalates (Moriana, Vilaplana, Karlsson, & Ribes, 2014). The thermal
different cellulose contents. The dissolving pulp has a cellulose con- degradation onset temperatures of all the studied cellulose oxalates
tent > 96%, while kraft pulp contains 84% cellulose. This difference were 173–177 °C, which showed the beginning of the second main
indicates that the samples prepared from kraft pulp not only contain mass-loss region (175–280 °C). This could be related to the decom-
cellulose oxalate but also hemicellulose, which has been hydrolysed position of chemically attached oxalate groups as well as the dehy-
and esterified by oxalic acid. dration of cellulose. The third main mass-loss region (280–375 °C) in-
The yield of cellulose oxalates was affected by whether either dicated the depolymerization of cellulose (Peng et al., 2013).
ethanol or acetone was used to remove the excess oxalic acid. Washing Considering the onset temperatures, the cellulose oxalates and

212
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

aggregates present in the gel. This was determined for SWD35A_5p and
SWK35A_5p. After dilution and centrifugation, the content was de-
termined to be 93% and 98% for SWD35A_5p and SWK35A_5p, re-
spectively. The lower content for SWD35A_5p is likely connected to the
presence of larger aggregates, which reduce the viscosity and scatter
light, as observed when visually comparing the gels after homo-
genization. The total nanocellulose yield (from raw material to dis-
persed nanocellulose) was determined to 90% and 94% for SWD35A_5p
and SWK35A_5p, respectively.
Preparing nanocellulose from dry cellulose oxalate reduces the need
for biocides as it is possible to fibrillate the nanocellulose on demand. It
may also be open up new processes to fibrillate nanocellulose, for ex-
ample in combination with other processing steps when preparing
formulations. Other studies have shown great potential to increase the
yield when producing nanocellulose by first producing CNCs followed
by preparing CNFs from the solid residue remaining after the reaction
(Chen et al., 2016; Wang, Zhu, & Considine, 2013; Wang, Chen, Zhu, &
Yang, 2017). These studies have used centrifugation followed by dia-
lysis for washing the material, and in some cases required a substantial
Fig. 6. Films prepared from a) SWK35A_5p and b) SWD35A_5p produced by number of passes in a microfluidizer to prepare the CNFs. The current
vacuum filtration over a membrane.
paper is able to simplify the washing step by using only filtration, and
to prepare a nanocellulose gel with only one pass through the 200 μm
presumably the resulting nanocellulose can be used as reinforcements and 100 μm interaction chamber of the microfluidizer. Simplifying the
in thermoplastics, such as polystyrene (PS, melting temperature process and reducing the required energy to fibrillate is crucial in order
74–105 °C), low-density polyethylene (LDPE, melting temperature to decrease the production cost in an industrial process.
103–110 °C) and high-density polyethylene (HDPE melting temperature Samples of SWK35A_5p and SWD35A_5p were used to produce
125–132 °C). films; both of these samples formed even and transparent films with a
When homogenizing the cellulose oxalates, they easily pass through thickness of 30 μm when they were vacuum filtered and pressed
the 100 μm without clogging. At approximately 1.5 wt %, the samples (Fig. 6). SEM images of the films (Fig. 7) show a closely packed fibril
prepared from kraft pulp resulted in thick gels, while the samples structure. The SWD35A_5p and SWK35A_5p films had crystallinity in-
prepared from dissolving pulp were slightly less viscous and had a dexes of 73.6% and 78.5%, respectively (Table 3). The crystallinity
higher opacity (Fig. 4). Gels could be made from all samples with high indexes of the raw materials were measured to be 70.9% and 58.1% for
viscosity after one pass through the microfluidizer. After homogeniza- the dissolving pulp and kraft pulp, respectively. The crystallinity of
tion of the cellulose oxalates, the difference in colour between the both samples increased after the reaction as amorphous regions in the
different raw materials, as described above, is clearly apparent. Mi- cellulose were hydrolysed, which was especially evident in the sample
crographs of the homogenized cellulose oxalates (Fig. 5) show that the prepared from kraft pulp.
material consists of short fibrils with average lengths of 0.34 μm Films produced from sulphuric or hydrochloric acid CNCs are often
(SWK35E) and 0.37 μm (SWD60E) and average widths of 3.2 nm produced using solvent casting because the particles are small enough
(SWK35E) and 4.3 nm (SWD60E). The width of the particles is com- to pass through most filter membranes. Solvent casting is a very slow
parable to that commonly found for CNCs and in the lower range for process that, together with poor film properties, limits the use of films
CNFs. The length, however, has a large range that spans lengths typical produced from CNCs. The nanocellulose prepared in the current work
for both CNFs and CNCs. The analysed samples contain many particles was prepared using vacuum filtration over a 0.43 μm membrane
that are similar to the shape and length (50–500 nm) (Moon et al., without any substantial material loss. Films produced from SWD35A_5p
2011) of CNCs and a considerable number of particles that are longer had a lower tensile strength and lower elongation at break compared to
(up to 1.1 μm) than common CNCs and shaped similar to flexible CNFs. those produced from SWK35A_5p (Table 3). In the present work, the
It should be noted that the number of long particles is likely under- suspension of SWK35A_5p showed a higher transparency than the
estimated because it is easier to adsorb, identify and measure small suspension of SWD35A_5p at the same consistency, which indicates a
particles than it is long and entangled particles when analysing the more efficient separation of nanosized cellulose fibrils in the former
images. than the latter (Moser, Lindström, & Henriksson, 2015). The more ef-
The nanocellulose content is an indication of the amount of larger ficient separation of nanofibrils in the suspension enabled the exposure

Fig. 7. SEM micrographs of films prepared from SWD35A_5p (left) and SWK35A_5p (right). Note the difference in magnification between the two images.

213
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

Crystallinity index, mechanical properties and oxygen barrier properties of the films prepared from SWD35A_5p and SWK35A_5p. Data are presented as the mean values. The error corresponds to the confidence interval,
of more surface area for bonding between the nanofibrils during film

Oxygen permeability [cm3 μm m−2 day−1 kPa−1] 50 %


formation. SWK35 A also has a higher crystallinity index, which may

0.14–5.03(Aulin, Gällstedt, & Lindström, 2010, 2012;


contribute to its higher mechanical strength. Consequently, the inter-
facial interactions between the nanofibrils were stronger in the film

Naderi et al., 2016; Syverud & Stenius, 2008)


made from SWK35 A than the one made from SWD35 A, as represented
by the superior tensile strength of the former relative to the latter. The
mechanical strength of the prepared films is comparable with many
other films prepared from CNFs, as is the strain at break and elastic
modulus.
In general, the films based on the nanocellulose in the current work
have comparable tensile properties to those reported for neat nano-
cellulose films (tensile strength: 30–240 MPa; elastic modulus:
1–17.5 GPa) (Alain, Jean-Yves, & Vignon, 1997; Fukuzumi, Saito,
Iwata, Kumamoto, & Isogai, 2009; Henriksson, Berglund, Isaksson,
Lindström, & Nishino, 2008; Iwamoto, Abe, & Yano, 2008; Iwamoto,
0.54
0.31

Nakagaito, & Yano, 2007; Iwamoto, Nakagaito, Yano, & Nogi, 2005;
RH

Leitner, Hinterstoisser, Wastyn, Keckes, & Gindl, 2007; Henriksson &


Berglund, 2007; Nakagaito, Iwamoto, & Yano, 2005; Nakagaito & Yano,
2008; Nogi, Iwamoto, Nakagaito, & Yano, 2009; Rampinelli, Di Landro,
6–14.9(Moon et al., 2011;

& Fujii, 2010; Saito et al., 2009; Stelte & Sanadi, 2009; Svagan, Samir, &
6-15(Moon et al., 2011)

Berglund, 2007; Svagan, Hedenqvist, & Berglund, 2009; Syverud &


Reising et al., 2012)

Stenius, 2008; Yano & Nakahara, 2004), as well as higher tensile


Modulus [GPa]

strength and elastic modulus than bio-based films based on non-cellu-


10.6 (+-0.4)
10.2 (+-0.4)

losic polysaccharides (e.g., starch, hemicelluloses, and pectin) (tensile


strength: 9.8–74 MPa; elastic modulus: 0.8–2.4 GPa) (Cao, Chen,
Chang, Stumborg, & Huneault, 2008; Edlund, Ryberg, & Albertsson,
2010; Le Normand, Moriana, & Ek, 2014; Mikkonen et al., 2010) and
commercial petroleum-based polymers (e.g., polyethylene, poly-
Tensile strain at maximum

2–10(Moon et al., 2011)

propylene, polyvinyl chloride, and polyamide) (tensile strength:


0.6(Reising et al., 2012)

8–165 MPa; elastic modulus: 0.2–4.1 GPa) (Mangaraj, Goswami, &


tensile stress [%]

Mahajan, 2009).
The films produced from SWD35A_5p and SWK35A_5p both showed
3.0 (+-0.4)
5.0 (+-0.6)

low oxygen permeability (Table 3), and the kraft-based film had slightly
lower oxygen permeability than the dissolving-based film. This can be
attributed to the higher degree of fibrillation and higher crystallinity
index of the samples produced from kraft pulp. Fewer large aggregates
Maximum tensile stress [MPa]

increases the path required for gas molecules to travel through the film.
95–240(Moon et al., 2011)

The reported permeability is comparable to that of many others at 50%


RH and is often regarded as sufficient for many applications.
70(Reising, Moon, &
Youngblood, 2012)

4. Conclusions
142 (+-5)
197 (+-7)

The high-yield preparation of cellulose oxalate through the reaction


between oxalic acid and cellulose was studied. During the reaction,
simultaneous esterification and acid hydrolysis occurs, resulting in a
highly charged (0.6–1.1 mmol/g) cellulose derivative with a crystal-
66.2–82.1(Peng et al., 2013; Tejado, Alam, Antal,

linity index of approximately 75%. The cellulose oxalate was prepared


Yang, & van de Ven, 2012; Zhao et al., 2013)

as a dry powder, after homogenization all samples produces nano-


cellulose gels after one pass through the 100 μm interaction chamber of
the microfluidizer. The particle length of the resulting nanocellulose
varies greatly and contains particles which resemble the size and shape
of both CNF and CNC. It was shown that it is possible to prepare films
54–88(Moon et al., 2011)

with mechanical and barrier properties similar to many other CNFs


through filtration. Using the described method it is possible to prepare
Crystallinity index

nanocellulose without the need for extensive dialysis or centrifugation


or extensive homogenization, and with very high yield. The nano-
cellulose has great potential for uses where the size of traditional CNFs
may present problems, such as when the particle length results in a too
high viscosity, and where CNCs are too small, such as when preparing
74
79

films though filtration.


alpha = 0.05.

Typical CNC

Typical CNF

Conflicts of interest
films

films
SWD35A
SWK35A
Table 3

The authors are shareholders in FineCell Sweden AB, a company


working in commercializing nanocellulose.

214
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

Acknowledgments Jung, Y. H., Chang, T., Zhang, H., Yao, C., Zheng, Q., Yang, V. W., ... Ma, Z. (2015). High-
performance green flexible electronics based on biodegradable cellulose nanofibril
paper. Nature Communications, 6, 7170. https://doi.org/10.1038/ncomms8170.
This research was made possible through funding by Södra Research Khan, A., Vu, K. D., Chauve, G., Bouchard, J., Riedl, B., & Lacroix, M. (2014).
Foundation, Sweden, Gunnar and Birgitta Nordin's foundation, Sweden Optimization of microfluidization for the homogeneous distribution of cellulose na-
(grant no. GFS2016-0274) managed by The Royal Swedish Academy of nocrystals (CNCs) in biopolymeric matrix. Cellulose, 21(5), 3457–3468. https://doi.
org/10.1007/s10570-014-0361-9.
Agriculture and Forestry (KSLA) and Ångpanneföreningen's Foundation
Larsson, P. T., Hult, E.-L., Wickholm, K., Pettersson, E., & Iversen, T. (1999). CP/MAS
for Research and Development (ÅFORSK), Sweden(grant no. 17-521). 13C-NMR spectroscopy applied to structure and interaction studies on cellulose I.
Solid State Nuclear Magnetic Resonance, 15(1), 31–40. https://doi.org/10.1016/
Appendix A. Supplementary data S0926-2040(99)00044-2.
Le Normand, M., Moriana, R., & Ek, M. (2014). The bark biorefinery: A side-stream of the
forest industry converted into nanocomposites with high oxygen-barrier properties.
Supplementary material related to this article can be found, in the Cellulose, 21(6), 4583–4594. https://doi.org/10.1007/s10570-014-0423-z.
online version, at doi:https://doi.org/10.1016/j.carbpol.2019.02.056. Leitner, J., Hinterstoisser, B., Wastyn, M., Keckes, J., & Gindl, W. (2007). Sugar beet
cellulose nanofibril-reinforced composites. Cellulose, 14(5), 419–425. https://doi.
org/10.1007/s10570-007-9131-2.
References Li, D., Henschen, J., & Ek, M. (2017). Esterification and hydrolysis of cellulose using
oxalic acid dihydrate in a solvent-free reaction suitable for preparation of surface-
functionalised cellulose nanocrystals with high yield. Green Chemistry, 19(23),
Alain, D., Jean-Yves, C., & Vignon, M. R. (1997). Mechanical behavior of sheets prepared
5564–5567. https://doi.org/10.1039/C7GC02489D.
from sugar beet cellulose microfibrils. Journal of Applied Polymer Science, 64(6),
Mangaraj, S., Goswami, T. K., & Mahajan, P. V. (2009). Applications of plastic films for
1185–1194. https://doi.org/10.1002/(SICI)1097-4628(19970509)64:6<1185::AID-
modified atmosphere packaging of fruits and vegetables: A review. Food Engineering
APP19>3.0.CO;2-V.
Reviews, 1(2), 133. https://doi.org/10.1007/s12393-009-9007-3.
Alexandrescu, L., Syverud, K., Gatti, A., & Chinga-Carrasco, G. (2013). Cytotoxicity tests
Mikkonen, K. S., Mathew, A. P., Pirkkalainen, K., Serimaa, R., Xu, C., Willför, S., ...
of cellulose nanofibril-based structures. Cellulose, 20(4), 1765–1775. https://doi.org/
Tenkanen, M. (2010). Glucomannan composite films with cellulose nanowhiskers.
10.1007/s10570-013-9948-9.
Cellulose, 17(1), 69–81. https://doi.org/10.1007/s10570-009-9380-3.
Araki, J., Wada, M., Kuga, S., & Okano, T. (1998). Flow properties of microcrystalline
Moon, R. J., Martini, A., Nairn, J., Simonsen, J., & Youngblood, J. (2011). Cellulose
cellulose suspension prepared by acid treatment of native cellulose. Colloids and
nanomaterials review: Structure, properties and nanocomposites. Chemical Society
Surfaces A: Physicochemical and Engineering Aspects, 142(1), 75–82. https://doi.org/
Reviews, 40(7), 3941–3994. https://doi.org/10.1039/C0CS00108B.
10.1016/S0927-7757(98)00404-X.
Moriana, R., Vilaplana, F., Karlsson, S., & Ribes, A. (2014). Correlation of chemical,
Arora, A., & Padua, G. W. (2010). Review: Nanocomposites in food packaging. Journal of
structural and thermal properties of natural fibres for their sustainable exploitation.
Food Science, 75(1), R43–R49. https://doi.org/10.1111/j.1750-3841.2009.01456.x.
Carbohydrate Polymers, 112, 422–431. https://doi.org/10.1016/j.carbpol.2014.06.
Assembly U.G (2015). Transforming our world : the 2030 Agenda for Sustainable
009.
Development. Retrieved fromhttps://sustainabledevelopment.un.org/post2015/
Moser, C., Lindström, M. E., & Henriksson, G. (2015). Toward industrially feasible
transformingourworld.
methods for following the process of manufacturing cellulose nanofibers.
Aulin, C., Gällstedt, M., & Lindström, T. (2010). Oxygen and oil barrier properties of
BioResources, 10.
microfibrillated cellulose films and coatings. Cellulose, 17(3), 559–574. https://doi.
Naderi, A., Lindström, T., Flodberg, G., Sundström, J., Junel, K., Runebjörk, A., ...
org/10.1007/s10570-009-9393-y.
Erlandsson, J. (2016). Phosphorylated nanofibrillated cellulose: production and
Aulin, C., Salazar-Alvarez, G., & Lindström, T. (2012). High strength, flexible and
propertyes. Nordic Pulp and Paper Research Journal, 31(1), 20. https://doi.org/10.
transparent nanofibrillated cellulose–nanoclay biohybrid films with tunable oxygen
3183/npprj-2016-31-01-p020-029.
and water vapor permeability. Nanoscale, 4(20), 6622–6628. https://doi.org/10.
Nakagaito, A. N., & Yano, H. (2008). Toughness enhancement of cellulose nanocompo-
1039/C2NR31726E.
sites by alkali treatment of the reinforcing cellulose nanofibers. Cellulose, 15(2),
Cai, C. M., Zhang, T., Charles, R. K., & Wyman, E. (2014). Integrated furfural production
323–331. https://doi.org/10.1007/s10570-007-9168-2.
as a renewable fuel and chemical platform from lignocellulosic biomass. Journal of
Nakagaito, A. N., Iwamoto, S., & Yano, H. (2005). Bacterial cellulose: The ultimate nano-
Chemical Technology & Biotechnology, 89(1), 2–10. https://doi.org/10.1002/jctb.
scalar cellulose morphology for the production of high-strength composites. Applied
4168.
Physics A, 80(1), 93–97. https://doi.org/10.1007/s00339-004-2932-3.
Cao, X., Chen, Y., Chang, P. R., Stumborg, M., & Huneault, M. A. (2008). Green com-
Nelson, K., Retsina, T., Iakovlev, M., van Heiningen, A., Deng, Y., Shatkin, J. A., &
posites reinforced with hemp nanocrystals in plasticized starch. Journal of Applied
Mulyadi, A. (2016). American process: Production of Low cost nanocellulose for re-
Polymer Science, 109(6), 3804–3810. https://doi.org/10.1002/app.28418.
newable, advanced materials applications. In L. D. Madsen, & E. B. Svedberg (Eds.).
Chen, L., Zhu, J. Y., Baez, C., Kitin, P., & Elder, T. (2016). Highly thermal-stable and
materials research for manufacturing: An industrial perspective of turning materials into
functional cellulose nanocrystals and nanofibrils produced using fully recyclable
New products (pp. 267–302). Cham: Springer International Publishing.
organic acids. Green Chemistry, 18(13), 3835–3843. https://doi.org/10.1039/
Nogi, M., Iwamoto, S., Nakagaito, A. N., & Yano, H. (2009). Optically transparent na-
C6GC00687F.
nofiber paper. Advanced Materials, 21(16), 1595–1598. https://doi.org/10.1002/
Edlund, U., Ryberg, Y. Z., & Albertsson, A.-C. (2010). Barrier films from renewable for-
adma.200803174.
estry waste. Biomacromolecules, 11(9), 2532–2538. https://doi.org/10.1021/
Peng, Y., Gardner, D. J., Han, Y., Kiziltas, A., Cai, Z., & Tshabalala, M. A. (2013).
bm100767g.
Influence of drying method on the material properties of nanocellulose I:
Ek, M., Henschen, J., & Li, D. (2015). SE Patent No. SE 539317.
Thermostability and crystallinity. Cellulose, 20(5), 2379–2392. https://doi.org/10.
Fox, S. C., Li, B., Xu, D., & Edgar, K. J. (2011). Regioselective esterification and ether-
1007/s10570-013-0019-z.
ification of cellulose: A review. Biomacromolecules, 12(6), 1956–1972. https://doi.
Pu, Y., Ziemer, C., & Ragauskas, A. J. (2006). CP/MAS 13C NMR analysis of cellulase
org/10.1021/bm200260d.
treated bleached softwood kraft pulp. Carbohydrate Research, 341(5), 591–597.
Fukuzumi, H., Saito, T., Iwata, T., Kumamoto, Y., & Isogai, A. (2009). Transparent and
https://doi.org/10.1016/j.carres.2005.12.012.
high gas barrier films of cellulose nanofibers prepared by TEMPO-Mediated oxida-
Rampinelli, G., Di Landro, L., & Fujii, T. (2010). Characterization of biomaterials based on
tion. Biomacromolecules, 10(1), 162–165. https://doi.org/10.1021/bm801065u.
microfibrillated cellulose with different modifications. Journal of Reinforced Plastics
Habibi, Y., Chanzy, H., & Vignon, M. R. (2006). TEMPO-mediated surface oxidation of
and Composites, 29(12), 1793–1803. https://doi.org/10.1177/0731684409335453.
cellulose whiskers. Cellulose, 13(6), 679–687. https://doi.org/10.1007/s10570-006-
Rånby, B. G. (1951). Fibrous macromolecular systems. Cellulose and muscle. The col-
9075-y.
loidal properties of cellulose micelles. Discussions of the Faraday Society, 11, 158–164.
Henriksson, M., & Berglund, L. A. (2007). Structure and properties of cellulose nano-
Reising, A. B., Moon, R. J., & Youngblood, J. P. (2012). Effect of particle alignment on
composite films containing melamine formaldehyde. Journal of Applied Polymer
mechanical properties of neat cellulose nanocrystal films. J-for-Journal of Science &
Science, 106(4), 2817–2824. https://doi.org/10.1002/app.26946.
Technology for Forest Products and Processes, 2(6), 32–41.
Henriksson, M., Berglund, L. A., Isaksson, P., Lindström, T., & Nishino, T. (2008).
Saito, T., Hirota, M., Tamura, N., Kimura, S., Fukuzumi, H., Heux, L., ... Isogai, A. (2009).
Cellulose nanopaper structures of high toughness. Biomacromolecules, 9(6),
Individualization of nano-sized plant cellulose fibrils by direct surface carboxylation
1579–1585. https://doi.org/10.1021/bm800038n.
using TEMPO catalyst under neutral conditions. Biomacromolecules, 10(7),
Henriksson, M., Henriksson, G., Berglund, L. A., & Lindstrom, T. (2007). An en-
1992–1996. https://doi.org/10.1021/bm900414t.
vironmentally friendly method for enzyme-assisted preparation of microfibrillated
Saito, T., Kimura, S., Nishiyama, Y., & Isogai, A. (2007). Cellulose nanofibers prepared by
cellulose (MFC) nanofibers. European Polymer Journal, 43(8), 3434–3441. https://doi.
TEMPO-mediated oxidation of native cellulose. Biomacromolecules, 8(8), 2485–2491.
org/10.1016/j.eurpolymj.2007.05.038.
https://doi.org/10.1021/bm0703970.
Iwamoto, S., Abe, K., & Yano, H. (2008). The effect of hemicelluloses on wood pulp na-
Schneider, C. A., Rasband, W. S., & Eliceiri, K. W. (2012). NIH Image to ImageJ: 25 years
nofibrillation and nanofiber network characteristics. Biomacromolecules, 9(3),
of image analysis. Nature Methods, 9(7), 671–675.
1022–1026. https://doi.org/10.1021/bm701157n.
Segal, L., Creely, J. J., Martin, A. E., Jr., & Conrad, C. M. (1959). An empirical method for
Iwamoto, S., Nakagaito, A. N., & Yano, H. (2007). Nano-fibrillation of pulp fibers for the
estimating the degree of crystallinity of native cellulose using the X-ray dif-
processing of transparent nanocomposites. Applied Physics A, 89(2), 461–466. https://
fractometer. Textile Research Journal, 29(10), 786–794. https://doi.org/10.1177/
doi.org/10.1007/s00339-007-4175-6.
004051755902901003.
Iwamoto, S., Nakagaito, A. N., Yano, H., & Nogi, M. (2005). Optically transparent com-
Sirvio, J. A., Visanko, M., & Liimatainen, H. (2016). Acidic deep eutectic solvents as
posites reinforced with plant fiber-based nanofibers. Applied Physics A, 81(6),
hydrolytic media for cellulose nanocrystal production. Biomacromolecules, 17(9),
1109–1112. https://doi.org/10.1007/s00339-005-3316-z.
3025–3032. https://doi.org/10.1021/acs.biomac.6b00910.

215
J. Henschen, et al. Carbohydrate Polymers 213 (2019) 208–216

Spinella, S., Maiorana, A., Qian, Q., Dawson, N. J., Hepworth, V., McCallum, S. A., ... Wågberg, L., Decher, G., Norgren, M., Lindström, T., Ankerfors, M., & Axnäs, K. (2008).
Gross, R. A. (2016). Concurrent cellulose hydrolysis and esterification to prepare a The build-up of polyelectrolyte multilayers of microfibrillated cellulose and cationic
surface-modified cellulose nanocrystal decorated with carboxylic acid moieties. ACS polyelectrolytes. Langmuir, 24(3), 784–795. https://doi.org/10.1021/la702481v.
Sustainable Chemistry & Engineering, 4(3), 1538–1550. https://doi.org/10.1021/ Wang, Q. Q., Zhu, J. Y., & Considine, J. M. (2013). Strong and optically transparent films
acssuschemeng.5b01489. prepared using cellulosic solid residue recovered from cellulose nanocrystals pro-
Stelte, W., & Sanadi, A. R. (2009). Preparation and characterization of cellulose nanofi- duction waste stream. ACS Applied Materials & Interfaces, 5(7), 2527–2534.
bers from two commercial hardwood and softwood pulps. Industrial & Engineering Wang, R. B., Chen, L. H., Zhu, J. Y., & Yang, R. D. (2017). Tailored and integrated pro-
Chemistry Research, 48(24), 11211–11219. https://doi.org/10.1021/ie9011672. duction of carboxylated cellulose nanocrystals (CNC) with nanofibrils (CNF) through
Svagan, A. J., Samir, M. A. S. A., & Berglund, L. A. (2007). Biomimetic polysaccharide maleic acid hydrolysis. Chemnanomat, 3(5), 328–335.
nanocomposites of high cellulose content and high toughness. Biomacromolecules, Yano, H., & Nakahara, S. (2004). Bio-composites produced from plant microfiber bundles
8(8), 2556–2563. https://doi.org/10.1021/bm0703160. with a nanometer unit web-like network. Journal of Materials Science, 39(5),
Svagan, A. J., Hedenqvist, M. S., & Berglund, L. (2009). Reduced water vapour sorption in 1635–1638. https://doi.org/10.1023/B:JMSC.0000016162.43897.0a.
cellulose nanocomposites with starch matrix. Composites Science and Technology, Yu, H. Y., Qin, Z. Y., Liang, B. L., Liu, N., Zhou, Z., & Chen, L. (2013). Facile extraction of
69(3), 500–506. https://doi.org/10.1016/j.compscitech.2008.11.016. thermally stable cellulose nanocrystals with a high yield of 93% through hydrochloric
Syverud, K., & Stenius, P. (2008). Strength and barrier properties of MFC films. Cellulose, acid hydrolysis under hydrothermal conditions. Journal of Materials Chemistry A,
16(1), 75. https://doi.org/10.1007/s10570-008-9244-2. 1(12), 3938–3944. https://doi.org/10.1039/c3ta01150j.
Tanpichai, S., Quero, F., Nogi, M., Yano, H., Young, R. J., Lindström, T., & Eichhorn, S. J. Zhang, W., Zhang, Y., Lu, C., & Deng, Y. (2012). Aerogels from crosslinked cellulose
(2012). Effective young’s modulus of bacterial and microfibrillated cellulose fibrils in nano/micro-fibrils and their fast shape recovery property in water. Journal of
fibrous networks. Biomacromolecules, 13(5), 1340–1349. https://doi.org/10.1021/ Materials Chemistry, 22(23), 11642–11650. https://doi.org/10.1039/C2JM30688C.
bm300042t. Zhao, J., Zhang, W., Zhang, X., Zhang, X., Lu, C., & Deng, Y. (2013). Extraction of cel-
Tejado, A., Alam, M. N., Antal, M., Yang, H., & van de Ven, T. G. M. (2012). Energy lulose nanofibrils from dry softwood pulp using high shear homogenization.
requirements for the disintegration of cellulose fibers into cellulose nanofibers. Carbohydrate Polymers, 97(2), 695–702. https://doi.org/10.1016/j.carbpol.2013.05.
Cellulose, 19(3), 831–842. https://doi.org/10.1007/s10570-012-9694-4. 050.

216

You might also like