ACS Publications. Most Trusted. Most Cited. Most Read
Lytic Polysaccharide Monooxygenase-Assisted Preparation of Oxidized-Cellulose Nanocrystals with a High Carboxyl Content from the Tunic of Marine Invertebrate Ciona intestinalis
My Activity

Figure 1Loading Img
  • Open Access
Research Article

Lytic Polysaccharide Monooxygenase-Assisted Preparation of Oxidized-Cellulose Nanocrystals with a High Carboxyl Content from the Tunic of Marine Invertebrate Ciona intestinalis
Click to copy article linkArticle link copied!

  • Anthi Karnaouri
    Anthi Karnaouri
    Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
  • Blanca Jalvo
    Blanca Jalvo
    Department of Materials and Environmental Chemistry, Stockholm University, Stockholm 10691, Sweden
    More by Blanca Jalvo
  • Philipp Moritz
    Philipp Moritz
    Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
  • Leonidas Matsakas
    Leonidas Matsakas
    Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
  • Ulrika Rova
    Ulrika Rova
    Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
    More by Ulrika Rova
  • Oliver Höfft
    Oliver Höfft
    Institute of Electrochemistry, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
  • Georgia Sourkouni
    Georgia Sourkouni
    Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
  • Wolfgang Maus-Friedrichs
    Wolfgang Maus-Friedrichs
    Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
  • Aji P. Mathew*
    Aji P. Mathew
    Department of Materials and Environmental Chemistry, Stockholm University, Stockholm 10691, Sweden
    *Email: [email protected]. Tel.:+46 (0) 8161256.
  • Paul Christakopoulos*
    Paul Christakopoulos
    Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
    *Email: [email protected]. Tel.: +46 (0) 920 492510.
Open PDFSupporting Information (1)

ACS Sustainable Chemistry & Engineering

Cite this: ACS Sustainable Chem. Eng. 2020, 8, 50, 18400–18412
Click to copy citationCitation copied!
https://doi.org/10.1021/acssuschemeng.0c05036
Published December 8, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Abstract

Click to copy section linkSection link copied!

The tunicate species Ciona intestinalis is a fast-growing marine invertebrate animal that contains cellulose in its outer part—the tunic. The high crystallinity and microfibril aspect ratio of tunicate cellulose make it an excellent starting material for the isolation of nanocellulose. In the present work, tunic from C. intestinalis was subjected to organosolv pretreatment followed by bleaching and acid-hydrolysis steps for the isolation of nanocrystals. Applying an intermediate enzymatic treatment step with a lytic polysaccharide monooxygenase (LPMO) from the thermophilic fungus Thermothelomyces thermophila was proved to facilitate the isolation of nanocellulose and to improve the overall process yield, even when the bleaching step was omitted. LPMOs are able to oxidatively cleave the glycosidic bonds of a polysaccharide substrate, either at the C1 and/or C4 position, with the former leading to introduction of carboxylate moieties. X-ray photoelectron spectroscopy analysis showed a significant increase in the atomic percentage of the C═O/O–C–O and O–C═O bonds upon the addition of LPMO, while the obtained nanocrystals exhibited higher thermal stability compared to the untreated ones. Moreover, an enzymatic post-treatment with LPMOs was performed to additionally functionalize the cellulose nanocrystals. Our results demonstrate that LPMOs are promising candidates for the enzymatic modification of cellulose fibers, including the preparation of oxidized-nanocellulose, and offer great perspectives for the production of novel biobased nanomaterials.

Copyright © 2020 American Chemical Society

Synopsis

Organosolv fractionation combined with enzymatic treatment with LPMOs allows for the isolation of nanocrystals with a high carboxyl content from Ciona intestinalis tunic.

Introduction

Click to copy section linkSection link copied!

The decrease of oil-based resources prompts the use of biomass as a feedstock to produce renewable materials, which are expected to be environmentally friendly and offer improved properties, such as biodegradability and enhanced biocompatibility. Cellulose is a natural renewable, biodegradable, and biocompatible polymer with abundant availability, which makes it a very promising substrate for numerous applications. Being a hierarchical material, cellulose can be used for the preparation of two types of biobased nanomaterials viz. cellulose nanofibrils and cellulose nanocrystals (CNCs), each one with distinct properties regarding their shape, size, and morphology. (1) Both these nanoparticles have found application in various fields, including polymer reinforcement, production of antimicrobial and medical materials, biosensors, hydrogels, drilling fluids, and drug delivery systems. (2−4) Cellulose nanofibrils are processed by mechanical treatment, including homogenization or grinding with or without other pretreatment steps that provide lateral cleavage of cellulose fibers into nanofibrils with diameters less than 10 nm and lengths in the micron scale. (5) The isolation of cellulose into highly crystalline nanoscale materials, namely, CNCs, requires chemical treatments such as acid hydrolysis to remove the amorphous regions in the cellulose chains and yield fibrils that typically have diameter < 10 nm and lengths in the nanometer range, for example, 150–300 nm for wood-based nanocrystals. (6)
The isolation of cellulose into CNCs has become a major research focus because of their attractive combinations of physicochemical characteristics, such as high crystallinity and good processability, as well as high specific strength and modulus, biocompatibility, and biodegradability, among others. (5,7) The dimensions and the crystallinity of these nanocrystals depend on the origin of the cellulose fibers as well as the procedure employed to obtain them because variations in the CNC extraction process lead mainly to different CNC properties. (8,9) Wood and annual plants are the commonly used natural resources for cellulose and nanoscale cellulose. Although cellulose can be produced from the cell wall of certain algae and bacteria, tunicates, which live in the oceans, are the only known animal species that produces cellulose in the outer tissues. Tunic, the mantle of these animals, contains cellulose of high crystallinity (80–90%) and a high aspect ratio, called tunicin, which offers excellent polymer reinforcement properties at very low concentrations, thus indicating its excellent chemical and material applications. (10) Tunicin nanocrystals (TNCs), which are usually isolated through sulphuric acid hydrolysis and are functionalized with -HSO3 groups that provide anionic surface charges, (11) are well known to be much longer than those of plant origin, exist predominantly in the Iβ polymorph, and have a very large aspect ratio, high crystallinity, and a high specific surface area. (12) Nanocrystals from tunicates belonging to species Styela clava and Halocynthia roretzi and nanofibrils from Ciona intestinalis have been successfully isolated and characterized, (12−14) while Mathew et al. have demonstrated the efficient reinforcement of TNCs in biopolymer matrices as plasticized starch. (11,15)
Apart from mechanical and chemical processes, enzyme-mediated pretreatment approaches in aqueous environments have received increasing attention for the production of cellulose nanofibrils and nanocrystals partially because of the high substrate specificity of enzymes. (16) Cellulases, mainly endoglucanases and exoglucanases, as well as xylanases, have been shown to play a key role in improving cellulose nanofiber production by removing the amorphous regions, thus facilitating refining and fibrillation. (17,18) Accessory enzymes such as lytic polysaccharide monooxygenases (LPMOs) can increase the hydrolytic performance of cellulase cocktails by facilitating the accessibility of the enzymes to the cellulosic component. (19−21) LPMOs represent a powerful class of enzymes that are able to oxidatively cleave cellulose and other recalcitrant polysaccharides, and they possess a type II copper active center coordinated by a histidine brace. (22−24) Apart from boosting the production of fermentable sugars, LPMOs have been recently reported to be able to introduce carboxyl groups to the substrate, (20,25) thus increasing the electrostatic repulsion of cellulose fibers and, in this way, promoting the fibrillation process toward the formation of cellulose nanofibrils through decreasing the energy requirements of mechanical treatment. (26−28) In a similar way, LPMOs could partially or entirely substitute the need for harsh chemical treatments that lead to fiber modification and swelling, such as bleaching.
The ability of LPMOs to introduce carboxylate moieties on the surface of cellulose makes them promising candidates as biocatalysts for the biofunctionalization of nanocellulose. Oxidized nanocelluloses have gained much attention not only in the biomedical field because of their antimicrobial properties (29) but also as scaffolds for the grafting of various molecules to produce modified materials, (30) as there is increasing interest in replacing petrol-based or synthetic polymers by sustainable, biobased materials. Until now, a chemoenzymatic modification process using laccase as the biocatalyst and TEMPO as the enhancer (laccase–TEMPO system) is the only one that has been applied to introduce carbonyl and carboxyl groups at the C6 position. (31) A different mechanism of oxidation is employed by LPMOs because these catalysts are able to catalyze cleavage of the polysaccharide substrates by oxidation at the C1 and/or C4 position, yielding products with either an additional carboxyl group (lactonic acids) or gemdiols (hydrated form of corresponding 4-ketolaldoses), respectively. (32) It has been recently reported that an AA11 LPMO from Fusarium fujikuroi is able to catalyze the introduction of carboxyl groups on α-chitin, without affecting significantly the crystallinity of the fibers. (33) In a similar way, LPMOs can be utilized for the modification of cellulose surface properties toward the production of oxidized materials with defined properties. (18,28,34)
In the current study, tunic from the marine invertebrate C. intestinalis was utilized, for the first time, as a feedstock for the preparation of CNCs with a high carboxyl content. The available amount of C. intestinalis cellulose is approximately 560,000 tons (dry weight) at full production in Sweden. Tunicates have an extremely high growth capacity, and they are predicted to become much more common in the coming years as a result of climate change and marine eutrophication; therefore, exploitation of their potential for novel applications sounds promising. To achieve the goal of nanocrystal production, we employed a strategy based on organosolv (OS) pretreatment for the initial fractionation of the material and the recovery of a cellulose-rich fraction, combined with the use of LPMOs as an additional enzymatic treatment step, together with a modification of the widely used acid hydrolysis extraction, to introduce carboxyl groups on the polysaccharide substrates and prepare oxidized TNCs. A post-treatment with LPMOs was also assessed for the modification of the isolated TNCs. Our results demonstrate that pretreatment with LPMOs improved the isolation of TNCs, even when the bleaching step was omitted, thus simplifying and reducing the costs of the extraction process, while post-treatment with LPMOs could further increase the carboxyl content and the negative surface charge of TNCs.

Experimental Section

Click to copy section linkSection link copied!

Preparation of the Tunic Material

OS Pretreatment of Tunic

The whole tunic from the marine invertebrate C. intestinalis, harvested in South Sweden, was provided by Marine Feed Sweden AB. It was thoroughly washed with fresh water, air-dried in an oven (80 °C), and used as the raw material. OS pretreatment took place in a high-pressure hydrothermal reactor, as previously described. (35) A suspension of biomass and aqueous EtOH solution (60% v/v), with a liquid to solid ratio of 10, was treated at 175 °C for 60 min and subsequently cooled down below 40 °C. After pretreatment, the solid fraction was separated from the pretreatment liquor by vacuum filtration, washed with fresh aqueous EtOH solution (60% v/v), dried in an oven at 50 °C, and stored at room temperature. Structural carbohydrate analysis of untreated and pretreated solids was conducted following the NREL protocol. (36) The monomeric sugars produced were quantified by ion-exchange chromatography using an Aminex HPX-87 N column with a microguard column at 85 °C (Bio-Rad Laboratories, Hercules, CA, USA), specified for the analysis of samples with high salinity, using ultrapure H2O (Type I) at a flow rate of 0.6 mL/min as the mobile phase. Inorganic ash and protein contents were determined according to the procedures previously described. (35)

Enzymatic Treatment of Tunic and Analysis of the Product Profile

For the enzymatic pretreatment of the tunic, MtLPMO9H (XP_003661787 and MYCTH_46583) from Thermothelomyces thermophila (previously known as Myceliophthora thermophila) was used. The enzyme was heterologously produced in Pichia pastoris and purified to its homogeneity, as previously described. (37) This LPMO exhibits a double C1/C4-oxidizing activity, although it is the C1-oxidizing effect that is the most important in the context of this study. The enzymatic treatment of tunic took place at 50 mM sodium acetate, pH 6.0, and 50 °C, for 18 h. Ascorbic acid was used as the electron donor at a final concentration of 5 mM. The initial solid loading was 4.5% w/v, and the enzyme loading was 30 mg/g of substrate. At the end of the enzymatic reaction, the tunic was recovered after centrifugation and was washed extensively with distilled H2O and Triton-X (0.1% w/v) to remove the residual enzyme that has been bound to the biomass. The supernatant was analyzed for the oxidized products of the MtLPMO activity on tunic. The soluble oxidized products of MtLPMO were first detected by high-performance anion exchange chromatography (HPAEC), using an ICS 5000SP system (Dionex, Thermo Fisher Scientific Inc.), as previously described. (37) Protein concentration was identified by the Lowry method, using the Folin phenol reagent. (38)
To further identify the reaction products, liquid chromatography-electrospray ionization tandem mass spectrometry (LC-ESI MS/MS) analysis was used. Prior to the analysis, 20 μL of each sample were diluted with 20 μL 0.1% v/v trifluoroacetic acid (TFA) and applied to a C-18 Ziptip packed with 15 μL graphitized carbon, which was pre-equilibrated with 3 × 25 μL 90% v/v acetonitrile (AcCN) containing 0.5% v/v TFA and 3 × 25 μL 0.5% v/v TFA. The Ziptip was washed with an additional 3 × 25 μL 0.5% v/v TFA after applying samples. The oxidized samples were eluted with 3 × 25 μL 40% v/v AcCN containing 0.5% v/v TFA. The eluent was dried in SpeedVac and reconstituted with 10 μL H2O. To reduce the samples, 20 μL of each sample were incubated with 20 μL of 1 M NaBH4 and 20 mM NaOH at 50 °C overnight. Salts were removed from the samples as described previously. (39) 3 μL of each sample was applied to analysis. The sugars were separated on a column (10 cm × 250 μm) packed in-house with 5 μm porous graphite particles (Hypercarb, Thermo-Hypersil, Runcorn, UK) and eluted with an acetonitrile gradient (Buffer A, 10 mM ammonium bicarbonate; Buffer B, 10 mM ammonium bicarbonate in 80% v/v AcCN). The gradient (0–45% Buffer B) was eluted for 46 min followed by a washing step with 100% Buffer B and equilibrated with Buffer A for the next 24 min. The samples were analyzed in both negative and positive-ion modes on an LTQ linear ion trap mass spectrometer (Thermo Electron, San José, CA), with an IonMax standard electrospray ionization source equipped with a stainless-steel needle kept at −3.5 kV. Compressed air was used as the nebulizer gas. The heated capillary was kept at 270 °C, and the capillary voltage was −50 kV. Full scan (m/z 380–2000, two microscans, maximum 100 ms, and a target value of 30,000) was performed followed by data-dependent MS2 scans (two microscans, maximum 100 ms, and a target value of 10,000) with a normalized collision energy of 35%, an isolation window of 2.5 units, activation q = 0.25, and an activation time of 30 ms. The threshold for MS2 was set to 300 counts. Data acquisition and processing were conducted with Xcalibur software (Version 2.0.7).

Isolation of TNCs

The TNC suspension was prepared by sulfuric acid hydrolysis following a modification of the process, as previously reported. (15) The aqueous tunicate suspension was mixed with H2SO4 to reach a final acid/H2O concentration of 55 wt % at 70 °C under vigorous stirring. The hydrolysis was quenched by adding excess of distilled water after 30 min and allowed to cool down to room temperature. The suspension was successively centrifuged at 5000 rpm for 15 min at 4 °C to concentrate the CNCs and to remove the excess of aqueous acid. The isolated nanocrystals were rinsed and dialyzed in distilled H2O until pH 4.5 was reached. The suspension was sonicated using a Qsonica Q500, 500-watt Sonicator at 75% output for 5 min.
In the case of those samples that were subjected to the additional bleaching step prior to sulphuric acid hydrolysis, a 5 wt % KOH solution was added to the samples, and the mixture was boiled under stirring for 6 h to swell the fibers, facilitate the infiltration of the bleaching solution, and lead to the removal of the residual cellulose-bound protein fraction and other tunic compounds. The final solution was extensively washed with distilled H2O and was subsequently bleached upon addition of 0.19 M sodium chlorite into 0.3 M sodium acetate buffer, at 80 °C for 6 h, under agitation. The samples were washed with distilled H2O, and then, a 5 wt % KOH solution was added, and the mixture was incubated at room temperature for 24 h. The samples were washed with distilled H2O, and the final product was centrifuged for 15 min at 10000 rpm (15,344 xg) at 4 °C and resuspended in distilled H2O. This process was repeated twice, and the TNC recovery yield was determined as previously described by Bondeson et al. (40) Briefly, recovery yield was calculated as the % of the initial weight of TNCs after hydrolysis. Approximately 1/4 of the total volume was freeze-dried and compared to the initial weight. Isolated TNCs were characterized as described below.

LPMO Post-treatment of TNCs

To examine whether enzymatic post-treatment could further increase the carboxyl content of isolated TNCs, an additional incubation with MtLPMO for 8 h took place. The initial concentration of TNCs was 0.5 wt %, and the enzyme loading was 10 mg/g of substrate. All other reaction conditions (buffer system, time, and agitation), as well as the TNC recovery and characterization process, were the same as those followed for TNCs prior to LPMO post-treatment.

Characterization of TNCs

Morphology and Size

Topographical surface images were captured using an atomic force microscope (Veeco Multimode V, USA) operating in tapping mode. For the analysis, prior to examination, a droplet of diluted suspension of the sample was deposited on a carbon microgrid and was allowed to dry. The diameter measurements were performed by employing the Nanoscope V software and analyzing the height images. The structure of the nanocrystals obtained was observed using a scanning electron microscopy (SEM) JEOL JSM-7000F instrument with an accelerating voltage of 3 kV. The samples were sputter-coated with a thin layer of gold for 80 s before SEM visualization. The particle length of the nanocrystals and the surface charge, namely, the zeta potential (ζ-potential) value of the TNC suspensions, were determined using dynamic light scattering (DLS) and electrophoretic light scattering in a Zetasizer Nano ZS (Malvern Instruments, Malvern, UK) apparatus. HCl and NaOH were used to adjust the pH to the required value (pH 7.0).

Thermal Analysis

The thermal behavior of the nanocrystals was assessed by thermogravimetric analysis (TGA), using TA Instruments Discovery thermal analyzer, to measure the mass transformation as a function of temperature, for an interval of 30–600 °C, at a heating rate of 5 °C/min. The samples were exposed to air gas at a flow rate of 20 mL/min. The reported values for each sample represent the average of three different measurements.

X-ray Photoelectron Spectroscopy (XPS) Analysis

For XPS measurements, 50 μL of diluted TNC suspension were deposited on a gold substrate and allowed to dry. The freshly prepared samples were analyzed by XPS in an ultrahigh vacuum chamber, with a base pressure below 5 · 10–10 mbar. The detailed apparatus setup was described previously. (41,42) All samples were examined once, whereby XPS has a general detection limit of 0.1 at % and a measurement accuracy of 10%. (43) For MgKα X-ray generation, a commercial X-ray source (RS40B1, Prevac, Rogów, Poland) was used, and the X-rays hit the surface under an angle of 80° to the surface normal. The emitted photoelectrons were detected using a hemispheric analyzer (type Leybold EA10/100) under an angle of 10° to the surface normal. The pass energy for the analyzer was kept constant at 80 eV for survey spectra and at 40 eV for detailed spectra. All XP spectra are displayed as a function of binding energy with respect to the Fermi level. Because of a constant shift in all components assigned to the cellulose, XP spectra have been charge-corrected using the aliphatic C–C binding at 285.0 eV as a reference. For quantitative XPS analysis, a linear background subtraction for the photoelectron peaks was employed. The peak components were deconvoluted by Voigt-type profiles with CasaXPS software (version 2.3.16 Pre-rel 1.4, CasaXPS Ltd., Teignmouth, United Kingdom).

X-ray Diffraction (XRD) Analysis

The crystallinity index (CrI) of the tunicate CNCs was determined by XRD using a PANalytical Empyrean X-ray diffractometer, with a smallest increment of 0.0001° and angular reproducibility <0.0002°, equipped with a PixCel3D detector and a graphite monochromator. CuKÜ1 radiation with λ = 1.540598 at 45 kV and 40 mA was used in the 2θ range of 5–90° at a scanning speed of 0.026°/s. CrI was assessed by the peak height method suggested by Segal. (44)

Evaluation of the TNC Effect on Bacterial Growth

The ability of isolated TNCs to inhibit the growth of Bacillus subtilis subsp. spizizenii (DSM 347) and Escherichia coli (DSM 1103) was assayed by the growth curve method. Exponential-phase cultures were grown overnight in Casein Soya and Luria-Bertani liquid broth for B. subtilis and E. coli, respectively. Cultures were subsequently diluted to an optical density at 600 nm (OD600) of 0.0138 (108 cells/mL) and placed in 15 mL sterile tubes containing the appropriate nutrient broth supplemented with the different TNC samples at a final concentration of 0.25% w/v. The bacterial growth in nutrient broth in the absence of TNCs was used as a positive control, while the negative control consisted of bacteria incubated in pure TNC solution in distilled H2O. Nutrient broth was used as a blank. Cultures were incubated at 37 °C, for E. coli, and 30 °C, for B. subtilis for 24 h, under agitation (160 rpm). Samples were taken at different time intervals (2, 4, 6, 8, 18, 20, and 24 h), and the growth was monitored spectrophotometrically by measuring the OD600. Triplicate experiments were carried out for each bacterial culture.

Results and Discussion

Click to copy section linkSection link copied!

Pretreatment Steps for the Preparation of TNCs

In the present study, the cellulose-rich tunic of C. intestinalis was employed as a starting material for the isolation of nanocrystals. TNCs have been obtained after subjecting biomass to a combined process using two novel and independent pretreatment methods, either followed or not by bleaching, prior to the acid hydrolysis step, as summarized in Figure 1. These methods included: (i) an initial OS pretreatment method, by employing a low-toxic aqueous ethanol solution, which has been previously reported to be very efficient for the fractionation of tunic and the removal of proteins and lipids (33) and (ii) an additional enzymatic treatment with LPMO that has been demonstrated to promote fibrillation and facilitate the isolation of cellulose nanofibrils from other feedstocks, such as cotton linters, (34) birchwood Kraft pulp, (27) and spruce biomass. (28) The tunicin without any pretreatment was also processed in parallel and used as the control.

Figure 1

Figure 1. Flow diagram of the experimental process.

According to the results from our previous study, (35) after the OS pretreatment of tunic, the cellulose, protein, and ash contents of the material were 48.5, 10.1, and 32.3 wt %, respectively. The results show selective recovery of cellulose in the solids (the cellulose matrix remains intact) and solubilization of protein and ash, which underline the suitability of this process as an initial step for the isolation of TNCs. OS was selected in our case not only for its ability to achieve efficient fractionation of different biomass feedstocks (35,45,46) but also because it has been reported to promote swelling of cellulose fibrils (47) without decreasing the crystallinity. (48)
Prior to the isolation of TNCs, incubation of the samples with a fungal MtLPMO was employed as a second pretreatment step. The soluble products of the LPMO reaction on the OS-pretreated material were analyzed to verify the activity of the enzyme on the substrate. Control reactions in the absence of LPMOs with 5 mM ascorbic acid and copper equimolar to the LPMO added were also performed. HPAEC data are depicted in Figure 2 and show the presence of nonoxidized and oxidized sugars. MtLPMO belongs to the auxiliary activity 9 (AA9) family of the CAZy database (http://www.cazy.org/AA9.html) (49) and exhibits a double regioselectivity mode of action; (37) the enzyme releases a mix of C1, C4, and C1/C4 double oxidized oligosaccharides from OS-pretreated tunic. LC-ESI MS/MS analysis (Figure S1, Supporting Information) occurred based on reports in the literature, (50) and it reveals the formation of C1 and C4-oxidized products in hydrated form (aldonic acids and gemdiols) with a DP = 2–5, in nonhydrated form (lactones and 4-ketoaldoses) with a DP = 3–6, as well as nonoxidized cello-oligosaccharides with varying size (DP = 3–8), as shown in Table 1. Traces of oxidized pentose oligosaccharides were also detected, which probably originate from the activity of MtLPMO on sulfated xylans that have been observed in tunic of C. intestinalis. (51) After incubation with MtLPMO, the solubilization of the material was <3 wt %, verifying that the soluble products represented only a small fraction of the initial substrate; this is in accordance with the already known mode of action of LPMOs, being reported to create nicking points on the surface of fibers and introduce carboxyl groups on the crystalline areas of cellulose. (52) The C1-oxidative activity of MtLPMO on the OS-pretreated tunic sample makes it a promising candidate biocatalyst to facilitate the isolation of TNCs. The tunicin of invertebrates has been shown to form a very dense structure of bundled fibrils embedded in the tunic matrix. (10) The increase of surface charge upon addition of LPMOs might enable the disruption of this rigid structure.

Figure 2

Figure 2. HPAEC chromatograph of soluble products detected after activity of MtLPMO on OS-pretreated tunic. The profile of the products reveals a mixed C1/C4-oxidizing activity. The products of blank reactions in the absence of enzyme (blank reaction 1, without enzyme, only ascorbic acid, and copper equivalent to LPMOs added) and electron donor (blank reaction 2 and only MtLPMO) are also depicted.

Table 1. Structure of Oxidized and Nonoxidized Products Detected in the Reaction Medium after the Action of MtLPMO on OS-Pretreated Tunic Biomass as the Substrate
[M + Na]structure of oligosaccharides (DP)
C1 and C4-oxidized products in hydrated form (aldonic acids and gemdiols)
381-Glcβ1-4Glc- (DP2)
543-Glcβ1-4Glcβ1-4Glc- (DP3)
705-Glcβ1-4Glcβ1-4Glcβ1-4Glc- (DP4)
867-Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc- (DP5)
C1 and C4-oxidized products in nonhydrated form (lactones and 4-ketoaldoses)
525-Glcβ1-4Glcβ1-4Glc- (DP3)
687-Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP4)
849-Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP5)
1011-Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP6)
nonoxidized products
527Glcβ1-4Glcβ1-4Glc (DP3)
689Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP4)
851Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP5)
1013Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP6)
1175Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP7)
1337Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glcβ1-4Glc (DP8)
437Xylβ1-4Xylβ1-4Xyl (DP3)
569Xyl1-4Xylβ1-4Xylβ1-4Xyl (DP4)
701Xyl1-4Xylβ1-4Xylβ1-4Xylβ1-4Xyl (DP5)
833Xyl1-4Xyl1-4Xylβ1-4Xylβ1-4Xylβ1-4Xyl (DP6)

Isolation, Recovery, and Properties of Isolated TNCs after Pretreatment

Untreated (Sample A), OS-pretreated (Sample B), and OS/LPMO-pretreated (Sample C) materials were all subjected to acid hydrolysis, in either the presence (Samples A, B, and C) or the absence (Samples Bw/o and Cw/o) of an additional intermediate bleaching step. Upon addition of sulphuric acid, disruption and hydrolysis of amorphous regions surrounding the embedded parts within the cellulose microfibrils occurred, leaving the microcrystalline segments intact. The recovery yield, final concentration of the TNC suspension, and ζ-potential of the TNCs isolated are summarized in Table 2, while the particle length is described in Table S1, Supporting information.
Table 2. Recovery Yield, % Wt. Concentration, and ζ-Potential Values of the TNCs Isolated from Tunic Biomass by Different Pretreatment Methodsa.
 isolated TNCs after pretreatment
samplerecovery yield (wt %)concentration (wt %)ζ-potential, pH 7 (mV)
untreated material (A)63.30.48–26.0 ± 2.3
OS-pretreated (B)72.90.55–22.9 ± 1.2
OS-pretreated w/o bleaching (Bw/o)26.80.24–16 ± 0.7
OS/LPMO-pretreated (C)91.20.54–22.1 ± 1.2
OS/LPMO-pretreated w/o bleaching (Cw/o)85.10.52–27.8 ± 0.8
a

The letters in the left column refer to the corresponding pictures of Figure 1

TNCs were successfully isolated with high recovery yields (>60 wt %) from almost all biomass samples, as observed at Table 2; the only exception was observed when OS constituted the only treatment that had been applied prior to acid hydrolysis (Sample Bw/o), where the recovery yield was very low (26.8 wt %). Bleaching step enabled the removal of inorganic ash content and further removal of the residual protein left after OS pretreatment. Because the amount of protein was low (10.1 wt %) as mentioned above, it is possible that the enhancing effect of bleaching in non-LPMO treated pulps (Samples A, B) occurred because of the removal of inorganic ash. The low nanocrystal yield indicates that the isolation of nanocrystals in the presence of inorganic structures is hampered. This can be possibly attributed to the presence of calcium in these structures, which was further confirmed by scanning electron microscopy-energy-dispersive spectrometry (SEM-EDS) analysis (Figure S2C, Supporting information); these results are also in agreement with previously reported data from ash analysis of the OS-pretreated material, which shows the presence of calcium at a concentration of 33.1 mg/g. (35)
OS pretreatment enabled fractionation of tunic biomass; however, although in small amounts, the presence of protein, as well as the inorganic ash, negatively affected the TNC isolation process in the absence of the bleaching step. In the case of LPMO pretreatment though, the enzyme facilitated the disruption of cellulose fibrils, therefore promoting the isolation of TNCs. Even when omitting the bleaching step, the OS/LPMO-pretreated material (Sample Cw/o) led to 85.1 wt % recovery yield, whereas the corresponding bleached material showed even higher recovery (91.2 wt %). Our results regarding the LPMO efficiency in facilitating fibril separation are in accordance with those previously reported in the literature; introducing a pretreatment step with a fungal PaLPMO9E from Podospora anserina enabled further processing of birchwood fibers for the production of cellulose nanofibrils, (26) while softwood kraft cellulose was more prone to mechanical fibrillation after activity of PaLPMO9H from the same fungus. (27) The great potential of LPMOs as environmentally friendly catalysts for the preparation of cellulose nanofibers from spruce is also highlighted in the pioneered work by Koskela et al. (28) Apart from reducing the use of harsh chemicals, the ability to isolate TNCs in the absence of the bleaching step may offer another additional advantage; because bleaching leads to removal of inorganic compounds, such as calcium and silica, omitting this step leads to production of TNCs decorated with inorganic residues, which offer great potential for further applications, such as in the preparation of hydrogels and hybrid biomaterials through cross-linking. (53,54)

DLS Measurements and Determination of ζ-Potential

ζ-Potential is used to estimate the electrostatic repulsion between particles and thus to determine the colloidal stability of the TNCs. The ζ-potential is a function of the surface charge density of the material and the particle size. As shown in Table 2, with the exception of the OS-pretreated material w/o bleaching (Sample Bw/o), all TNCs exhibit a ζ-potential absolute value over 20 mV, which can be translated into stable colloidal suspensions. (55) In the absence of LPMO treatment (Samples A, B, and Bw/o), the results indicate that TNCs have electrically negative surface charge, which can be attributed to the presence of sulphate half-esters introduced during sulphuric acid hydrolysis. (56) In the case of the LPMO-pretreated Sample Cw/o, the ζ-potential value is slightly higher probably because of the addition of carboxylate moieties at the C1 position. As mentioned above and has been previously reported, (37)MtLPMO is able to perform both C1 and C4 oxidation, resulting in the formation of a carboxylic acid or a gemdiol, respectively. The presence of LPMO-introduced C1-carboxyl groups on the surface of the material generates electrostatic repulsion that hampers the aggregation of TNCs and increases the colloidal stability (57) by improving solubility and dispersion of the fibers. Similar results after LPMO treatment of the fibers are reported in other studies, (27,28,34) while Hu et al. managed to efficiently produce from softwood a self-stabilized suspension of nanofibers with a ζ-potential value of −60 mV upon the addition of AA9 together with cellulases. (18) Other types of oxidative treatment have been shown to increase the electrostatic repulsion of the subsequently isolated nanofibers; apart from the widely used TEMPO-oxidation, (58) and the combination of OS treatment in conjunction with a mild oxidative process has also been reported in the literature for the isolation of CNCs from Eucalyptus wood, where oxidative irradiation produced pulps with a higher absolute ζ-potential value. (59) Surprisingly, the absolute value of the ζ-potential upon LPMO treatment in the absence of bleaching (Sample Cw/o) was higher than that upon LPMO treatment with bleaching (Sample Cw/o). These results could probably reflect the effect of other compounds that remain in the structure in the absence of bleaching and affect the surface charge density of the material. To elucidate the effect of different pretreatment steps on the final composition and overall structure of the samples, further research is required.

Atomic Force Microscopy (AFM) Imaging and Measurement of the TNC Diameter

The structure and size of the TNCs were both analyzed by AFM (Figure 3A). AFM images obtained from diluted TNC suspensions show the presence of well-isolated nanoscale TNCs, except for the nanocrystals originating from the unbleached OS-pretreated material (Sample Bw/o), where residual structures are present, as verified by SEM-EDS analysis (Figure S2C, Supporting information) and discussed above. These structures are expected to originate from the inorganic ash in the biomass, which was not completely removed during the processing route followed. All the samples displayed a rodlike shape, as expected and reported before for TNCs, (10,11) with length ranging from 600 nm up to almost 1 μm. The images show that the TNCs from the unbleached OS/LPMO-pretreated material (Sample Cw/o) (Figure 3A-a) are more individualized compared to the other TNCs. This could be attributed to the presence of carboxyl groups introduced by LPMOs that, together with the sulphate groups on the surface of the nanocrystals because of acid hydrolysis, promote repulsive forces that keep the nanocrystals apart from each other. Because broadening effects may occur during scanning because of the tip geometry, the nanocrystal diameter was determined from the AFM height images by employing the Nanoscope V software. As a mean value of triplicates, the average diameter of TNCs is presented in Figure 3B. The diameter varies from 4.45 nm ± 1.07 to 26.08 nm ± 6.63, which corresponds to the OS/LPMO-pretreated material w/o bleaching (Figure 3A-e) and OS-pretreated material w/o bleaching (Figure 3A-c), respectively. Results from DLS analysis (Table S1, Supporting information) showed that LPMO-pretreatment did not affect the length of the TNCs produced (Samples C and Cw/o). On the contrary, the width was shown to be reduced, either in the absence (Sample Cw/o) or presence (Sample C) of the bleaching step, yielding more fine and thin nanocrystals. Our results are in accordance with those previously reported, (27,34) verifying the decrease of the nanocellulose diameter upon LPMO treatment as a result of fiber disruption induced by the enzyme oxidative activity.

Figure 3

Figure 3. (A) AFM images of TNCs isolated from the untreated material (A), OS-pretreated material (B), OS-pretreated material w/o bleaching (Bw/o), OS/LPMO-pretreated material (C), and OS/LPMO-pretreated material w/o bleaching (Cw/o). (B) TNC diameter measured with Nanoscope V software.

AFM characterization and recovery yield of isolated TNCs (Table 2 and Figure 3) showed that when the material is pretreated with LPMO, the bleaching step is not of pivotal importance to obtain nanocrystals (Sample Cw/o). LPMOs promote the liberation and disintegration of cellulose fibers and enhance the subsequent isolation of nanocrystals from tunic. TNCs after LPMO pretreatment are thinner, longer, and with higher electrically negative charge. As mentioned above, the samples obtained after MtLPMO treatment (Samples C and Cw/o) had an absolute ζ-potential value above 20 mV (Table 2), indicating that there is no agglomeration because of the sufficient mutual repulsion resulting in colloidal stability. Although the effect of LPMOs has been evaluated mainly toward the production of cellulose nanofibers, (26−28) both in the cases of preparing nanofibrils and nanocrystals, C1-oxidizing LPMOs seem to result in fiber delamination into thinner and shorter structures while introducing C1-carboxylate groups. The unique property of LPMO-catalyzed oxidation compared to other oxidation methods such as TEMPO is that, apart from the fiber repulsion because of these carboxylate groups that facilitate the colloidal stability of the final product, these enzymes also cause chain cleavage and disruption in the crystalline areas, which might also enhance the subsequent acid hydrolysis toward the isolation of CNCs.

Thermal Stability of TNCs

The data regarding the TGA of the TNCs are presented in Figure 4. The onset degradation and the peak degradation temperature as well as the wt % solid residue for each sample are summarized in Table 3.

Figure 4

Figure 4. TGA graphs of the TNCs isolated from tunic biomass after different treatments.

Table 3. Thermal Stability Data and CrI of the Isolated TNCsa
 thermal stability parametersCrI (%)
sampleonset temperature (T0) (°C)peak degradation temperature (Tmax) (°C)residue (%)
untreated material (A)167. ± 10.8280.7 ± 11.41.4 ± 0.683
OS-pretreated (B)220.6 ± 8.0276.1 ± 24.036.1 ± 5.279
OS-pretreated w/o bleaching (Bw/o)33.5 ± 5.6290.6 ± 5.960.7 ± 3.0N.D.
OS/LPMO-pretreated (C)291.6 ± 15.9326.3 ± 11.216.3 ± 10.381
OS/LPMO-pretreated w/o bleaching (Cw/o)207.6 ± 12.4265.3 ± 15.643.2 ± 2.676
a

Ν.D. not determined.

The TNCs isolated from the untreated material (Sample A) showed the typical TGA curve for CNCs obtained through sulphuric acid hydrolysis. (60−63) The average onset temperature of degradation of this sample was 167.72 ± 10.80 °C (Table 3), displaying one of the lowest thermal stabilities among all samples, following the TNCs isolated from the unbleached OS-pretreated material (Sample Bw/o). The introduction of the sulphate groups during the acid hydrolysis with sulphuric acid affects the thermal stability of the TNCs because of the dehydration reaction, thus rendering them more susceptible to degradation when the temperature increases. (62,63) Comparatively, TNCs isolated from the OS/LPMO-pretreated material (Sample C) presented the highest thermal stability (Table 3). Zhao et al. obtained similar results when they compared the effect of enzymatic hydrolysis to that of acid hydrolysis for the isolation of nanocrystals. (10) Our results suggest that the pretreated material, with the exception of the unbleached OS-pretreated one (Sample Bw/o), has undergone an effective “cleaning process” where the amount of sulphate groups on the nanocrystal surface has been decreased, thus leading to a higher thermal stability. The high percentage of residues observed in OS- and OS/LPMO-pretreated samples w/o bleaching (Samples Bw/o and Cw/o) is attributed to the inorganic ash, being the highest for the TNCs isolated from unbleached OS-pretreated tunicate biomass (Table 3). For these samples, the starting decomposition temperature has a lower value than for the corresponding bleached samples because of decomposition of proteins and other compounds that remain after the isolation process (proteins and lipids). The high thermal stability of the OS/LPMO-pretreated material (Sample C) is also related to the high crystallinity of the sample, as will be discussed below.

Determination of the CrI

Determination of the CrI of the samples is shown in Table 3. For the untreated material, the CrI (0.83) was similar to that previously mentioned in the literature for tunicate nanocrystals isolated after sulphuric acid hydrolysis (0.8) (12) and higher than that for nanocrystals isolated from wood biomass (0.60) (12,59) or bacterial cellulose (0.72). (12) Comparison of the results of the OS-pretreated material (Sample B) and OS/LPMO-pretreated material (Sample C) shows that addition of LPMOs slightly increases the crystallinity of the sample; this increase has been suggested to be related more to the C4-oxidation activity of LPMOs. (19) In another study, treatment with C1-oxidizing LPMOs did not significantly affect the CrI. (34) The OS/LPMO-pretreated material w/o bleaching (Sample Cw/o) exhibits lower crystallinity (0.76) than the corresponding sample after bleaching (0.81) possibly because of the residual compounds that remained after the isolation process.

XPS Analysis and Determination of the Carboxyl Content of TNCs

The XPS survey spectra for all samples (Figure S3, Supporting Information) present the two main characteristic peaks for cellulose, which can be assigned to carbon (285 eV) and oxygen (533 eV). (64) A more detailed analysis of the spectrum of the TNCs from the untreated material (Sample A) also reveals the presence of low amounts (<1 at. %) of sulfur with the S2p peak at 170 eV (Figure S4, Supporting information). This binding energy is characteristic of sulfate groups. (65) Pretreated samples, either after OS or OS/LPMO treatment (Samples B and C), contain only low amounts of SO42– groups compared to the untreated one. This further confirms the results of TGA, where it was demonstrated that the pretreatment apparently “prevents” the introduction of sulphate groups on the substrate, thus increasing the thermal stability. In accordance with the ash analysis data that are reported above, some of the spectra also show very small proportions (<1 at. %) of Na, Si, and Ca, which, however, are not taken into account for an evaluation of the detailed spectra. The molybdenum peaks in the spectra are due to the metallic sample holder used for the analysis.
The detailed spectra of the C1s signal in Figure 5A show the oxidizing effect of the pretreatments. According to the XPS studies on nanocellulose by Mathew et al., (66) the following molecular species were used for the fitting of the C1s detailed peak (full width at half-maximum value, FWHM, is also provided):
1.

–C–C– / −C–H– (285.0 eV, FWHM 1.8 eV)

2.

–C–O– (287.0 eV, FWHM 1.6 eV)

3.

–C═O/O–C–O– (288.3 eV, FWHM 1.6 eV) and

4.

–O–C═O– (289.6 eV, FWHM 1.8 eV)

Figure 5

Figure 5. (A) XPS detailed spectra of the C 1 s region before (solid line) and after (dashed line) the LPMO post-treatment of the untreated material (A), OS-pretreated material (B), and OS/LPMO-pretreated material (C). (B) Stoichiometric amount of the molecular species.

A representative example for a fitted C1s detail peak can be found in Figure S5, Supporting Information. These species were adapted to all C1s detailed spectra, and Figure 5B shows the corresponding relative amount of each molecular species (functional group) present in the surface region. The peak shift to higher binding energies in the curves of the OS-pretreated material (Sample B), OS/LPMO-pretreated material (Sample C), and OS/LPMO-pretreated material w/o bleaching (Sample Cw/o) (Figure S6, Supporting information) indicates an increased proportion of oxidized −C═O/O–C═O bonds compared to the untreated material (Sample A), as a result of OS and LPMO treatment. Because the OS/LPMO-pretreated material w/o bleaching (Sample Cw/o) also exhibits this oxidation, it can be assumed that the LPMO-treatment could efficiently substitute the bleaching process toward the successful isolation of nanocellulose. However, it should be mentioned that a lower degree of oxidation could be observed for other unbleached samples prepared in the same manner. Further investigations have to be carried out here to find the reason for these differences. The presence of carboxyl groups on the surface results in strong electrostatic repulsion between TNCs in water, promoting the formation of CNCs, and offer great potential for further functionalization.

Evaluation of the TNC Effect on Bacterial Growth

Although oxidized cellulose can be used as a scaffold for grafting applications toward the development of materials with efficient antimicrobial properties, (67,68) the oxidized nanocellulose itself has shown an antibacterial effect, which is mainly bacteriostatic rather than bactericidal, and it is attributed to the presence of carboxylate groups. (69,70) Our aim was to study the effect of the oxidized nanocellulose prepared in this work on the growth of Gram-positive and Gram-negative bacteria, using B. subtilis and E. coli, respectively, as model organisms. To assess the bacteriostatic properties of nanocrystals from C. intestinalis tunic on the growth of the bacterial cultures, the optical density at 600 nm was measured at different time intervals after the incubation of the cultures with 0.25% w/v TNCs from the untreated (Sample A) and pretreated materials (Samples B, C, and Cw/o). As shown in Figure 6, growth rates of B. subtilis (Figure 6A) and E. coli (Figure 6B) were reduced upon addition of TNCs, compared to the corresponding cultures in the absence of TNCs. In general, the effect of the TNCs on B. subtilis resulted in a higher growth reduction than in E. coli, as expected, because Gram-positive bacteria are more sensitive to potential toxic agents than Gram-negative bacteria, because of the lack of the protective outer membrane in the former ones. (71,72) All bleached materials, including the unpretreated one, resulted in a similar reduction of the growth rate, whereas in the case of the OS/LPMO-pretreated material without bleaching (Sample Cw/o) such reduction was less pronounced. Τhis can be possibly attributed to the presence of inorganic compounds, such as calcium (Figure S2C, Supporting information) in the TNCs from unbleached biomass. Ca2+ has been identified as an intracellular signaling molecule in prokaryotes, related to optimal growth and morphology. (73,74) Our results demonstrate the bacteriostatic effect of TNCs on the two different bacterial strains tested, which could be partially hampered by the presence of Ca2+ when the bleaching step is omitted. Further studies need to be performed to elucidate and understand the decrease of the bacterial growth rate in the presence of TNCs and the possible effect of calcium ions.

Figure 6

Figure 6. Growth curve of B. subtilis and E. coli in the presence of 0.25% w/v TNCs: untreated material (A), OS-pretreated material (B), OS/LPMO-pretreated material (C), and OS/LPMO-pretreated material w/o bleaching (Cw/o).

LPMO Post-treatment of TNCs to Increase the Carboxyl Content

The isolated TNCs were subjected to a post-treatment with MtLPMO to evaluate whether it would be possible to further increase the carboxyl content. The effects of oxidative post-treatment of isolated nanocrystals with LPMOs have not been described earlier. TNCs from the OS-pretreated material w/o bleaching were not further processed, as this sample had a very low recovery yield and showed agglomeration. As depicted in Table 4, ζ-potential values turned more electrically negative after the LPMO post-treatment, confirming the extra charge introduction by the LPMO enzyme. This originates from the additional carboxyl groups added in the C1 position after the activity of LPMOs. (52) To exclude any possible surface modification caused by the addition of ascorbic acid and free copper, the isolated nanocrystals from the untreated sample (A) were treated with 5 mM ascorbic acid and 0.18 μM CuSO4 which is equivalent to the amount of MtLPMO in the reaction. The samples were analyzed by XPS, and the results were compared to those upon addition of MtLPMO (Sample Apost-treated). The results depicted in Figure S7, Supporting information showed that the amount of carboxylate groups does not change after the treatment with ascorbic acid and free copper, corroborating the idea that the introduction of carboxylate groups, which is related to facilitating the isolation of nanocrystals, is attributed exclusively to the activity of LPMOs.
Table 4. Surface Charge and Crystallinity of TNCs after LPMO Post-treatment
sampleζ-potential, pH 7 (mV)CrI %
untreated material (Apost-treated)–34.2 ± 1.763
OS-pretreated (Bpost-treated)–36.2 ± 276
OS/LPMO-pretreated (Cpost-treated)N.D.a75
OS/LPMO-pretreated w/o bleaching (Cw/o - post-treated)–35.9 ± 2.872
a

Ν.D. not determined.

All TNCs obtained after LPMO post-treatment had an absolute ζ-potential value above 30 mV, higher than those prior to incubation with LPMO. These results indicate that there is no agglomeration, which means that sufficient mutual repulsion led to colloidal stability. OS-pretreated TNCs (Sample Bpost-treated) displayed the highest charge density, with a ζ-potential of −36.2 mV. Post-treatment of TNCs with LPMO led to a further increase of carbonyl and carboxyl groups, resulting in negative charge, while it does not affect the length of the TNCs.
Crystallinity of TNCs was decreased after the LPMO post-treatment probably because of the oxidative cleavage of the nanocrystals after the activity of LPMOs (Table 4). The CrI reduction was more profound in the case of the untreated material and the bleached OS/LPMO-pretreated material. This is probably attributed to the higher crystallinity of the samples prior to LPMO post-treatment because highly crystalline cellulose seems to be preferred by certain LPMOs. (75) The structure and size of the TNCs after the post-treatment with LPMOs were analyzed by AFM (Figure 7A). AFM images obtained from diluted post-treated TNC suspensions show, in general, lower aggregation between the nanocrystals and that the TNCs conserved their rodlike shape and length compared to the ones described before. As it was observed for the pretreated samples, the TNCs from the OS/LPMO post-treated material w/o bleaching (Sample (Cw/o - post-treated) are more individualized and thinner compared to the other post-treated TNCs. As it was commented for the pretreated nanocrystals, this could be attributed to the introduction of carboxyl groups introduced by the LPMO enzymatic post-treatment on the surface of the nanocrystals, together with the ones already being added during the pretreatment, thus promoting higher repulsive forces. The promotion of lateral cleavage of nanocrystals to single crystals also can be expected during this post-treatment step. The average diameter of LPMO post-treated TNCs is shown in Figure 7B, and it varies from 5.53 ± 1.07 nm, for the OS/LPMO-pretreated material w/o bleaching (Sample Cw/o - post-treated), to 10.22 ± 0.25 for the OS/LPMO-pretreated material (Sample Cpost-treated).

Figure 7

Figure 7. (A) AFM images of the untreated material (Apost-treated), OS-pretreated material (Bpost-treated), OS/LPMO-pretreated material (Cpost-treated), and OS/LPMO-pretreated material w/o bleaching (Cw/o - post-treated) and (B) TNC diameter measurements after LPMO post-treatment, as identified by the Nanoscope V software.

To confirm the presence of additional functional carboxyl and carbonyl groups, TNCs after post-treatment with LPMOs were again investigated by XPS. The C1s detailed spectra were deconvolved, as described above, and are shown in Figure 5. The bars depicted in Figure 5B (shaded) present the relative amounts of functional groups after the LPMO post-treatment. The results show that the additional enzymatic treatment further increased the amount of −C═O/O–C═O bonds in all cases, confirming the results from the AFM and ζ-potential measurements. The oxidizing effect of LPMOs on TNCs that have already been enzymatically treated is lower. The XPS analysis of the OS/LPMO-pretreated material w/o bleaching showed very different results in repeated measurements; therefore, it is not listed in Figure 5. Further investigations must be carried out. Our results show the potential of LPMOs to achieve introduction of carboxyl groups on the surface of nanocrystals, which can theoretically occur in a controlled manner by applying different reaction times and/or enzyme concentrations. The above offers great perspectives for the design of one-pot enzymatic reactions with great selectivity for the production of functionalized nanomaterials.

Conclusions

Click to copy section linkSection link copied!

In the present work, a novel process that combines OS fractionation together with LPMO pretreatment is presented for the efficient isolation of CNCs from the C. intestinalis tunic. Our results show that TNCs are obtained in high yields, and they have high crystallinity, good thermal properties, and a high carboxyl content. MtLPMO is able to introduce both carboxylic and ketone groups to the substrate, as expected by the mixed C1/C4 oxidative activity of the enzyme. The presence of carboxylate ions on the surface results in strong electrostatic repulsion between TNCs in water, facilitating the nanocrystal formation even in the absence of the bleaching step. Moreover, for the first time, a post-enzymatic treatment with LPMOs is tested to additionally functionalize the CNCs, individualize the single crystals, and enhance the colloidal stability. This developed route demonstrates an energy and resource-efficient one-pot synthesis process that can be extended to other biomass substrates for isolating highly functional and negatively charged CNCs for various applications.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssuschemeng.0c05036.

  • Mass spectrometry data of degradation products generated by MtLPMO, SEM figures, XPS spectra, relative amount of the molecular species, and particle length data for TNCs isolated following different treatments (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Anthi Karnaouri - Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
    • Blanca Jalvo - Department of Materials and Environmental Chemistry, Stockholm University, Stockholm 10691, Sweden
    • Philipp Moritz - Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, GermanyOrcidhttp://orcid.org/0000-0002-7582-101X
    • Leonidas Matsakas - Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, SwedenOrcidhttp://orcid.org/0000-0002-3687-6173
    • Ulrika Rova - Biochemical Process Engineering, Chemical Engineering, Department of Civil, Environmental and Natural Resources Engineering, Luleå University of Technology, Luleå 97187, Sweden
    • Oliver Höfft - Institute of Electrochemistry, Clausthal University of Technology, Clausthal-Zellerfeld 38678, GermanyOrcidhttp://orcid.org/0000-0002-1313-3166
    • Georgia Sourkouni - Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
    • Wolfgang Maus-Friedrichs - Clausthal Centre of Material Technology, Clausthal University of Technology, Clausthal-Zellerfeld 38678, Germany
  • Author Contributions

    A.K. and B.J. equally contributed to this work.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

Part of this work was conducted within the frame of IKY-DAAD (2018-2019) ″Enzymatic superficial modification of natural and synthetic polymers and their spectroscopic analysis″ project, funded by a bilateral agreement between German Academic Exchange Service and Greek State Scholarship Foundation. Fredrik Norén from Marine Feed Sweden AB is greatly acknowledged for providing the tunic material.

Abbreviations

Click to copy section linkSection link copied!

TNC

tunicate nanocrystal

LPMO

lytic polysaccharide monooxygenase

OS

organosolv

DLS

dynamic light scattering

AFM

atomic force microscopy

XPS

X-ray photoelectron spectroscopy

XRD

X-ray diffraction

SEM

scanning electron microscopy

TGA

thermogravimetric analysis

References

Click to copy section linkSection link copied!

This article references 75 other publications.

  1. 1
    Xu, X.; Liu, F.; Jiang, L.; Zhu, J. Y.; Haagenson, D.; Wiesenborn, D. P. Cellulose nanocrystals vs. cellulose nanofibrils: a comparative study on their microstructures and effects as polymer reinforcing agents. ACS Appl. Mater. Interfaces 2013, 5, 29993009,  DOI: 10.1021/am302624t
  2. 2
    Lin, N.; Dufresne, A. Nanocellulose in biomedicine: Current status and future prospect. Eur. Polym. J. 2014, 59, 302325,  DOI: 10.1016/j.eurpolymj.2014.07.025
  3. 3
    Salas, C.; Nypelö, T.; Rodriguez-Abreu, C.; Carrillo, C.; Rojas, O. J. Nanocellulose properties and applications in colloids and interfaces. 2014. Curr. Opin. Colloid Interface Sci. 2014, 19, 383396,  DOI: 10.1016/j.cocis.2014.10.003
  4. 4
    Lee, K.-Y.; Aitomäki, Y.; Berglund, L. A.; Oksman, K.; Bismarck, A. On the use of nanocellulose as reinforcement in polymer matrix composites. Compos. Sci. Technol. 2014, 105, 1527,  DOI: 10.1016/j.compscitech.2014.08.032
  5. 5
    Klemm, D.; Kramer, F.; Moritz, S.; Lindstrom, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: a new family of nature-based materials. Angew. Chem., Int. Ed. 2011, 50, 54385466,  DOI: 10.1002/anie.201001273
  6. 6
    Habibi, Y.; Lucia, L. A.; Rojas, O. J. Cellulose nanocrystals: chemistry, self-assembly, and applications. Chem. Rev. 2010, 110, 34793500,  DOI: 10.1021/cr900339w
  7. 7
    Mariano, M.; El Kissi, N.; Dufresne, A. Cellulose nanocrystals and related nanocomposites: review of some properties and challenges. J. Polym. Sci., Part B: Polym. Phys. 2014, 52, 791806,  DOI: 10.1002/polb.23490
  8. 8
    Jonoobi, M.; Oladi, R.; Davoudpour, Y.; Oksman, K.; Dufrense, A.; Hamzeh, Y.; Davoodi, R. Different preparation methods and properties of nanostructured cellulose from various natural resources and residues: a review. Cellulose 2015, 22, 935969,  DOI: 10.1007/s10570-015-0551-0
  9. 9
    Börjesson, M.; Westman, G. Crystalline Nanocellulose — Preparation, Modification, and Properties, Cellulose-Fundamental Aspects and Current Trends, Matheus Poletto and Heitor Luiz Ornaghi Junior, IntechOpen, December 9th 2015.  DOI: 10.5772/61899 . Available from: https://www.intechopen.com/books/cellulose-fundamental-aspects-and-current-trends/crystalline-nanocellulose-preparation-modification-and-properties
  10. 10
    Zhao, Y.; Li, J. Excellent chemical and material cellulose from tunicates: diversity in cellulose production yield and chemical and morphological structures from different tunicate species. Cellulose 2014, 21, 34273441,  DOI: 10.1007/s10570-014-0348-6
  11. 11
    Mathew, A. P.; Dufresne, A. Morphological investigation of nanocomposites from sorbitol plasticized starch and tunicin whiskers. Biomacromolecules 2002, 3, 609617,  DOI: 10.1021/bm0101769
  12. 12
    Zhao, Y.; Moser, C.; Lindström, M. E.; Henriksson, G.; Li, J. Cellulose nanofibers from softwood, hardwood, and tunicate: preparation–structure–film performance interrelation. Appl. Mater. Interfaces 2017, 9, 1350813519,  DOI: 10.1021/acsami.7b01738
  13. 13
    Sacui, I. A.; Nieuwendaal, R. C.; Burnett, D. J.; Stranick, S. J.; Jorfi, M.; Weder, C.; Foster, E. J.; Olsson, R. T.; Gilman, J. W. Comparison of the properties of cellulose nanocrystals and cellulose nanofibrils isolated from bacteria, tunicate, and wood processed using acid, enzymatic, mechanical, and oxidative methods. ACS Appl. Mater. Interfaces 2014, 6, 61276138,  DOI: 10.1021/am500359f
  14. 14
    Elazzouzi-Hafraoui, S.; Nishiyama, Y.; Putaux, J.-L.; Heux, L.; Dubreuil, F.; Rochas, C. The shape and size distribution of crystalline nanoparticles prepared by acid hydrolysis of native cellulose. Biomacromolecules 2008, 9, 5765,  DOI: 10.1021/bm700769p
  15. 15
    Mathew, A. P.; Thielemans, W.; Dufresne, A. Mechanical properties of nanocomposites from sorbitol plasticized starch and tunicin whiskers. J. Appl. Polym. Sci. 2008, 109, 40654074,  DOI: 10.1002/app.28623
  16. 16
    Squinca, P.; Bilatto, S.; Badino, A. C.; Farinas, C. S. Nanocellulose production in future biorefineries: an integrated approach using tailor-made enzymes. ACS Sustainable Chem. Eng. 2020, 8, 22772286,  DOI: 10.1021/acssuschemeng.9b06790
  17. 17
    Henriksson, M.; Henriksson, G.; Berglund, L. A.; Lindström, T. An environmentally friendly method for enzyme-assisted preparation of microfibrillated cellulose (MFC) nanofibers. Eur. Polym. J. 2007, 43, 34343441,  DOI: 10.1016/j.eurpolymj.2007.05.038
  18. 18
    Hu, J.; Tian, D.; Renneckar, S.; Saddler, J. N. Enzyme-mediated nanofibrillation of cellulose by the synergistic actions of an endoglucanase, lytic polysaccharide monooxygenase (LPMO) and xylanase. Sci. Rep. 2018, 8, 3195,  DOI: 10.1038/s41598-018-21016-6
  19. 19
    Hu, J.; Arantes, V.; Pribowo, A.; Saddler, J. N. The synergistic action of accessory enzymes enhances the hydrolytic potential of a “cellulase mixture” but is highly substrate specific. Biotechnol. Biofuels 2013, 6, 112,  DOI: 10.1186/1754-6834-6-112
  20. 20
    Westereng, B.; Ishida, T.; Vaaje-Kolstad, G.; Wu, M.; Eijsink, V. G.; Igarashi, K.; Samejima, M.; Ståhlberg, J.; Horn, S. J.; Sandgren, M. The putative endoglucanase PcGH61D from Phanerochaete chrysosporium is a metal-dependent oxidative enzyme that cleaves cellulose. PLoS One 2011, 6, e27807  DOI: 10.1371/journal.pone.0027807
  21. 21
    Harris, P. V.; Welner, D.; McFarland, K. C.; Re, E.; Navarro Poulsen, J. C.; Brown, K.; Salbo, R.; Ding, H.; Vlasenko, E.; Merino, S.; Xu, F.; Cherry, J.; Larsen, S.; Lo Leggio, L. Stimulation of lignocellulosic biomass hydrolysis by proteins of Glycoside Hydrolase Family 61: Structure and function of a large, enigmatic family. Biochemistry 2010, 49, 33053316,  DOI: 10.1021/bi100009p
  22. 22
    Vaaje-Kolstad, G.; Westereng, B.; Horn, S. J.; Liu, Z.; Zhai, H.; So̷rlie, M.; Eijsink, V. G. An oxidative enzyme boosting the enzymatic conversion of recalcitrant polysaccharides. Science 2010, 330, 219222,  DOI: 10.1126/science.1192231
  23. 23
    Forsberg, Z.; Vaaje-Kolstad, G.; Westereng, B.; Bunæs, A. C.; Stenstro̷m, Y.; MacKenzie, A.; So̷rlie, M.; Horn, S. J.; Eijsink, V. G. Cleavage of cellulose by a CBM33 protein. Protein Sci. 2011, 20, 14791483,  DOI: 10.1002/pro.689
  24. 24
    Quinlan, R. J.; Sweeney, M. D.; Lo Leggio, L.; Otten, H.; Poulsen, J. C.; Johansen, K. S.; Krogh, K. B.; Jo̷rgensen, C. I.; Tovborg, M.; Anthonsen, A.; Tryfona, T.; Walter, C. P.; Dupree, P.; Xu, F.; Davies, G. J.; Walton, P. H. Insights into the oxidative degradation of cellulose by a copper metalloenzyme that exploits biomass components. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 1507915084,  DOI: 10.1073/pnas.1105776108
  25. 25
    Beeson, W. T.; Phillips, C. M.; Cate, J. H. D.; Marletta, M. A. Oxidative cleavage of cellulose by fungal copper-dependent polysaccharide monooxygenases. J. Am. Chem. Soc. 2011, 134, 890892,  DOI: 10.1021/ja210657t
  26. 26
    Moreau, C.; Tapin-Lingua, S.; Grisel, S.; Gimbert, I.; Le Gall, S.; Petit-Conil, M.; Berrin, J. G.; Cathala, B.; Villares, A. Lytic polysaccharide monooxygenases (LPMOs) facilitate cellulose nanofibrils production. Biotechnol. Biofuels. 2019, 12, 156,  DOI: 10.1186/s13068-019-1501-0
  27. 27
    Villares, A.; Moreau, C.; Bennati-Granier, C.; Garajova, S.; Foucat, L.; Falourd, X.; Saake, B.; Berrin, J. G.; Cathala, B. Lytic polysaccharide monooxygenases disrupt the cellulose fibers structure. Sci. Rep. 2017, 7, 40262,  DOI: 10.1038/srep40262
  28. 28
    Koskela, S.; Wang, S.; Xu, D.; Yang, X.; Li, K.; Berglund, L.; McKee, L. S.; Bulone, V.; Zhou, Q. Lytic polysaccharide monooxygenase (LPMO) mediated production of ultra-fine cellulose nanofibres from delignified softwood fibres. Green Chem. 2019, 21, 59245933,  DOI: 10.1039/C9GC02808K
  29. 29
    Spangler, D.; Rothenburger, S.; Nguyen, K.; Jampani, H.; Weiss, S.; Bhende, S. In vitro antimicrobial activity of oxidized regenerated cellulose against antibiotic-resistant microorganisms. Surg. Infect. (Larchmt) 2003, 4, 255262,  DOI: 10.1089/109629603322419599
  30. 30
    Wohlhauser, S.; Delepierre, G.; Labet, M.; Morandi, G.; Thielemans, W.; Weder, C.; Zoppe, J. O. Grafting polymers from cellulose nanocrystals: synthesis, properties and applications. Macromolecules 2018, 51, 61576189,  DOI: 10.1021/acs.macromol.8b00733
  31. 31
    Quintana, E.; Roncero, M. B.; Vidal, T.; Valls, C. Cellulose oxidation by Laccase-TEMPO treatments. Carbohydr. Polym. 2017, 157, 14881495,  DOI: 10.1016/j.carbpol.2016.11.033
  32. 32
    Vaaje-Kolstad, G.; Forsberg, Z.; Loose, J. S.; Bissaro, B.; Eijsink, V. G. Structural diversity of lytic polysaccharide monooxygenases. Curr. Opin. Struct. Biol. 2017, 44, 6776,  DOI: 10.1016/j.sbi.2016.12.012
  33. 33
    Wang, D.; Li, J.; Salazar-Alvarez, G.; McKee, L. S.; Srivastava, V.; Sellberg, J. A.; Bulome, V.; Hsieh, Y. S. Y. Production of functionalised chitins assisted by fungal lytic polysaccharide monooxygenase. Green Chem. 2018, 20, 20912100,  DOI: 10.1039/C8GC00422F
  34. 34
    Valls, C.; Pastor, F. I. J.; Roncero, M. B.; Vidal, T.; Diaz, P.; Martinez, J.; Valenzuela, S. V. Assessing the enzymatic effects of cellulases and LPMO in improving mechanical fibrillation of cotton linters. Biotechnol. Biofuels 2019, 12, 161,  DOI: 10.1186/s13068-019-1502-z
  35. 35
    Hrůzová, K.; Matsakas, L.; Karnaouri, A.; Norén, F.; Rova, U.; Christakopoulos, P. Second-generation biofuel production from the marine filter feeder Ciona intestinalis. ACS Sustainable Chem. Eng. 2020, 8, 83738380,  DOI: 10.1021/acssuschemeng.0c02417
  36. 36
    Sluiter, J. B.; Ruiz, R. O.; Scarlata, C. J.; Sluiter, A. D.; Templeton, D. W. Compositional analysis of lignocellulosic feedstocks. 1. Review and description of methods. J. Agric. Food Chem. 2010, 58, 90439053,  DOI: 10.1021/jf1008023
  37. 37
    Karnaouri, A.; Muraleedharan, M. N.; Dimarogona, M.; Topakas, E.; Rova, U.; Sandgren, M.; Christakopoulos, P. Recombinant expression of thermostable processive MtEG5 endoglucanase and its synergism with MtLPMO from Myceliophthora thermophila during the hydrolysis of lignocellulosic substrates. Biotechnol. Biofuels 2017, 10, 126,  DOI: 10.1186/s13068-017-0813-1
  38. 38
    Lowry, O. H.; Rosebrough, N. J.; Lewis Farr, A.; Randall, R. J. Protein Measurement with the Folin phenol reagent. J. Biol. Chem. 1951, 193, 265275
  39. 39
    Schulz, B. L.; Packer, N. H.; Karlsson, N. G. Small-scale analysis of O-linked oligosaccharides from glycoproteins and mucins separated by gel electrophoresis. Anal. Chem. 2002, 74, 60886097,  DOI: 10.1021/ac025890a
  40. 40
    Bondeson, D.; Mathew, A.; Oksman, K. Optimization of the isolation of nanocrystals from microcrystalline cellulose by acid hydrolysis. Cellulose 2006, 13, 171180,  DOI: 10.1007/s10570-006-9061-4
  41. 41
    Klarhöfer, L.; Roos, B.; Viöl, W.; Höfft, O.; Dieckhoff, S.; Kempter, V.; Maus-Friedrichs, W. Valence band spectroscopy on lignin. Holzforschung 2008, 62.  DOI: 10.1515/HF.2008.116
  42. 42
    Krischok, S.; Höfft, O.; Günster, J.; Stultz, J.; Goodman, D. W.; Kempter, V. H2O interaction with bare and Li-precovered TiO2: Studies with electron spectroscopies (MIES and UPS (HeI and II)). Surf. Sci. 2001, 495, 818,  DOI: 10.1016/S0039-6028(01)01570-9
  43. 43
    Hercules, D. M.; Cox, L. E.; Onisick, S.; Nichols, G. D.; Carver, J. C. Electron spectroscopy (ESCA). Use for trace analysis. Anal. Chem. 1973, 45, 19731975,  DOI: 10.1021/ac60333a037
  44. 44
    Nam, S.; French, A. D.; Condon, B. D.; Concha, M. Segal crystallinity index revisited by the simulation of X-ray diffraction patterns of cotton cellulose Iβ and cellulose II. Carbohydr. Polym. 2016, 135, 19,  DOI: 10.1016/j.carbpol.2015.08.035
  45. 45
    Thoresen, P. P.; Matsakas, L.; Rova, U.; Christakopoulos, P. Recent advances in organosolv fractionation: Towards biomass fractionation technology of the future. Bioresour. Technol. 2020, 306, 123189,  DOI: 10.1016/j.biortech.2020.123189
  46. 46
    Kalogiannis, K. G.; Karnaouri, A.; Michailof, C.; Tzika, A. M.; Asimakopoulou, G.; Topakas, E.; Lappas, A. A. OxiOrganosolv: A novel acid free oxidative organosolv fractionation for lignocellulose fine sugar streams. Bioresour. Technol. 2020, 313, 123599,  DOI: 10.1016/j.biortech.2020.123599
  47. 47
    Mantanis, G. I.; Young, R. A.; Rowell, R. M. Swelling of compressed cellulose fiber webs in organic liquids. Cellulose 1995, 2, 122,  DOI: 10.1007/BF00812768
  48. 48
    Sannigrahia, P.; Miller, S. J.; Ragauskas, A. J. Effects of organosolv pretreatment and enzymatic hydrolysis on cellulose structure and crystallinity in Loblolly pine. Carbohydr. Res. 2010, 345, 965970,  DOI: 10.1016/j.carres.2010.02.010
  49. 49
    Levasseur, A.; Drula, E.; Lombard, V.; Coutinho, P. M.; Henrissat, B. Expansion of the enzymatic repertoire of the CAZy database to integrate auxiliary redox enzymes. Biotechnol. Biofuels 2013, 6, 41,  DOI: 10.1186/1754-6834-6-41
  50. 50
    Silva, C. d. O. G.; Teixeira, T. S.; Rodrigues, K. B.; Souza, A. A.; Monclaro, A. V.; Mendes, T. D. Combination of MALDI-TOF MS and UHPLC-ESI-MS for the characterization of lytic polysaccharide monooxygenase activity. Anal. Methods 2020, 12, 149161,  DOI: 10.1039/c9ay01774g
  51. 51
    Zhao, Y.; Li, J. Ascidian bioresources: common and variant chemical compositions and exploitation strategy – examples of Halocynthia roretzi, Styela plicata, Ascidia sp. and Ciona intestinalis. Z. Naturforschung C 2016, 71, 165180,  DOI: 10.1515/znc-2016-0012
  52. 52
    Eibinger, M.; Ganner, T.; Bubner, P.; Rošker, S.; Kasciddvgdgracher, D.; Haltrich, D.; Ludwig, R.; Plank, H.; Nidetzky, B. Cellulose surface degradation by a lytic polysaccharide monooxygenase and its effect on cellulase hydrolytic efficiency. J. Biol. Chem. 2014, 289, 3592935938,  DOI: 10.1074/jbc.M114.602227
  53. 53
    Siqueira, P.; Siqueira, É.; De Lima, A. E.; Siqueira, G.; Pinzón-Garcia, A. D.; Lopes, A. P.; Segura, M. E. C.; Isaac, A.; Pereira, F. V.; Botaro, V. R. Three-dimensional stable alginate-nanocellulose gels for biomedical applications: towards tunable mechanical properties and cell growing. Nanomaterials 2019, 9, 78,  DOI: 10.3390/nano9010078
  54. 54
    Xie, K.; Yu, Y.; Shi, Y. Synthesis and characterization of cellulose/silica hybrid materials with chemical crosslinking. Carbohydr. Polym. 2009, 78, 799805,  DOI: 10.1016/j.carbpol.2009.06.019
  55. 55
    Bhattacharjee, S. DLS and zeta potential - What they are and what they are not?. J. Controlled Release 2016, 235, 337351,  DOI: 10.1016/j.jconrel.2016.06.017
  56. 56
    Johnston, L. J.; Jakubek, Z. J.; Beck, S.; Araki, J.; Cranston, E. D.; Danumah, C.; Fox, D.; Li, H.; Wang, J.; Mester, Z.; Moores, A.; Rabb, S. A.; Murphy, K. E.; Stephan, C. Determination of sulfur and sulfate half-ester content in cellulose nanocrystals: an interlaboratory comparison. Metrologia 2018, 55, 872,  DOI: 10.1088/1681-7575/aaeb60
  57. 57
    Bennati-Granier, C.; Garajova, S.; Champion, C.; Grisel, S.; Haon, M.; Zhou, S.; Fanuel, M.; Ropartz, D.; Rogniaux, H.; Gimbert, I.; Record, E.; Berrin, J. G. Substrate specificity and regioselectivity of fungal AA9 lytic polysaccharide monooxygenases secreted by Podospora anserina. Biotechnol. Biofuels 2015, 8, 90,  DOI: 10.1186/s13068-015-0274-3
  58. 58
    Isogai, A.; Saito, T.; Fukuzumi, H. TEMPO-oxidized cellulose nanofibers. Nanoscale 2011, 3, 7185,  DOI: 10.1039/C0NR00583E
  59. 59
    Zhang, R.; Liu, Y. High energy oxidation and organosolv solubilization for high yield isolation of cellulose nanocrystals (CNC) from Eucalyptus hardwood. Sci. Rep. 2018, 8, 16505,  DOI: 10.1038/s41598-018-34667-2
  60. 60
    Oksman, K.; Etang, J. A.; Mathew, A. P.; Jonobi, M. Cellulose nanowhiskers separated from a bio-residue from wood bioethanol production. Biomass Bioenergy 2011, 35, 146152,  DOI: 10.1016/j.biombioe.2010.08.021
  61. 61
    Herrera, M. A.; Mathew, A. P.; Oksman, K. Characterization of cellulose nanowhiskers: A comparison of two industrial bioresidues. Iop Conf. Ser.: Mater. Sci. Eng. 2012, 31, 012006  DOI: 10.1088/1757-899X/31/1/012006
  62. 62
    Roman, M.; Winter, W. T. Effect of sulfate groups from sulfuric acid hydrolysis on the thermal degradation behavior of bacterial cellulose. Biomacromolecules 2004, 5, 16711677,  DOI: 10.1021/bm034519+
  63. 63
    Rosa, M. F.; Medeiros, E. S.; Malmonge, J. A.; Gregorski, K. S.; Wood, D. F.; Mattoso, L. H. C.; Glenn, G.; Orts, W. J.; Imam, S. H. Cellulose nanowhiskers from coconut husk fibers: Effect of preparation conditions on their thermal and morphological behavior. Carbohydr. Polym. 2010, 81, 8392,  DOI: 10.1016/j.carbpol.2010.01.059
  64. 64
    Stevens, J. S.; Schroeder, S. L. M. Quantitative analysis of saccharides by X-ray photoelectron spectroscopy. Surf. Interface Anal. 2009, 41, 453462,  DOI: 10.1002/sia.3047
  65. 65
    Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy. Physical Electronics 1995, ISBN: 9780964812413
  66. 66
    Mathew, A. P.; Oksman, K.; Karim, Z.; Liu, P.; Khan, S. A.; Naseri, N. Process scale up and characterization of wood cellulose nanocrystals hydrolysed using bioethanol pilot plant. Ind. Crops Prod. 2014, 58, 212219,  DOI: 10.1016/j.indcrop.2014.04.035
  67. 67
    Missoum, K.; Sadocco, P.; Causio, J.; Belgacem, M. N.; Bras, J. Antibacterial activity and biodegradability assessment of chemically grafted nanofibrillated cellulose. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 45, 477483,  DOI: 10.1016/j.msec.2014.09.037
  68. 68
    Tavakolian, M.; Okshevsky, M.; van de Ven, T. G. M.; Tufenkji, N. Developing antibacterial nanocrystalline cellulose using natural antibacterial agents. ACS Appl. Mater. Interfaces 2018, 10, 3382733838,  DOI: 10.1021/acsami.8b08770
  69. 69
    Vytrasova, J.; Tylsova, A.; Brozkova, I.; Cervenka, L.; Pejchalova, M.; Havelka, P. Antimicrobial effect of oxidized cellulose salts. J. Ind. Microbiol. Biotechnol. 2008, 35, 12471252,  DOI: 10.1007/s10295-008-0421-y
  70. 70
    Abaev, I. K.; Kaputskiĭ, V. E.; Adarchenko, A. A.; Sobeshchukh, O. P. Mechanism of the antibacterial action of monocarboxycellulose and other ion-exchange derivatives of cellulose. Antibiot. Med. Biotekhnol. 1986, 31, 624628
  71. 71
    Saini, S.; Falco, Ç. Y.; Belgacem, M. N.; Bras, J. Surface cationized cellulose nanofibrils for the production of contact active antimicrobial surfaces. Carbohydr. Polym. 2016, 135, 239247,  DOI: 10.1016/j.carbpol.2015.09.002
  72. 72
    Breijyeh, Z.; Jubeh, B.; Karaman, R. Resistance of Gram-negative bacteria to current antibacterial agents and approaches to resolve it. Molecules 2020, 25, 1340,  DOI: 10.3390/molecules25061340
  73. 73
    Herbaud, M. L.; Guiseppi, A.; Denizot, F.; Haiech, J.; Kilhoffer, M. C. Calcium signalling in Bacillus subtilis. Biochim. Biophys. Acta 1998, 1448, 212226,  DOI: 10.1016/s0167-4889(98)00145-1
  74. 74
    Onoda, T.; Enokizono, J.; Kaya, H.; Oshima, A.; Freestone, P.; Norris, V. Effects of calcium and calcium chelators on growth and morphology of Escherichia coli L-form NC-7. J. Bacteriol. 2000, 182, 14191422,  DOI: 10.1128/jb.182.5.1419-1422.2000
  75. 75
    Valenzuela, S. V.; Valls, C.; Schink, V.; Sánchez, D.; Blanca Roncero, M.; Diaz, P.; Martinez, J.; Javier Pastor, F. I. Differential activity of lytic polysaccharide monooxygenases on celluloses of different crystallinity. Effectiveness in the sustainable production of cellulose nanofibrils. Carbohydr. Polym. 2019, 207, 5967,  DOI: 10.1016/j.carbpol.2018.11.076

Cited By

Click to copy section linkSection link copied!

This article is cited by 16 publications.

  1. Koar Chorozian, Anthi Karnaouri, Antonis Karantonis, Maria Souli, Evangelos Topakas. Characterization of a Dual Cellulolytic/Xylanolytic AA9 Lytic Polysaccharide Monooxygenase from Thermothelomyces thermophilus and its Utilization Toward Nanocellulose Production in a Multi-Step Bioprocess. ACS Sustainable Chemistry & Engineering 2022, 10 (27) , 8919-8929. https://doi.org/10.1021/acssuschemeng.2c02255
  2. Salla Koskela, Shennan Wang, Peter Matthew Paul Fowler, Fangchang Tan, Qi Zhou. Structure and Self-Assembly of Lytic Polysaccharide Monooxygenase-Oxidized Cellulose Nanocrystals. ACS Sustainable Chemistry & Engineering 2021, 9 (34) , 11331-11341. https://doi.org/10.1021/acssuschemeng.1c02407
  3. Margaux Grellier, Céline Moreau, Johnny Beaugrand, Sacha Grisel, Jean-Guy Berrin, Bernard Cathala, Ana Villares. Action of AA9 lytic polysaccharide monooxygenase enzymes on different cellulose allomorphs. International Journal of Biological Macromolecules 2024, 275 , 133429. https://doi.org/10.1016/j.ijbiomac.2024.133429
  4. Muhammad Syukri Bin Mohamad Misenan, Mohamad Saman Ahmad Farabi, Zulkipli Nor Akhlisah, Nor Ain Fatihah Azlisham. Preparation, characterization and application of nanocellulose from tunicate for electronic applications. 2024, 295-319. https://doi.org/10.1016/B978-0-443-22414-0.00016-8
  5. Xiaotong Wang, Jing Guo, Hongyi Ren, Jing Jin, Hui He, Peng Jin, Zhengguo Wu, Yonghua Zheng. Research progress of nanocellulose-based food packaging. Trends in Food Science & Technology 2024, 143 , 104289. https://doi.org/10.1016/j.tifs.2023.104289
  6. Olanrewaju Raji, Vincent G. H. Eijsink, Emma Master, Zarah Forsberg. Modularity impacts cellulose surface oxidation by a lytic polysaccharide monooxygenase from Streptomyces coelicolor. Cellulose 2023, 30 (17) , 10783-10794. https://doi.org/10.1007/s10570-023-05551-8
  7. Saül Llàcer Navarro, Monika Tõlgo, Lisbeth Olsson, Tiina Nypelö. Carboxylation of sulfated cellulose nanocrystals by family AA9 lytic polysaccharide monooxygenases. Cellulose 2023, 30 (15) , 9331-9347. https://doi.org/10.1007/s10570-023-05440-0
  8. Muhammad Mujtaba, Leonardo Fernandes Fraceto, Mahyar Fazeli, Sritama Mukherjee, Susilaine Maira Savassa, Gerson Araujo de Medeiros, Anderson do Espírito Santo Pereira, Sandro Donnini Mancini, Juha Lipponen, Francisco Vilaplana. Lignocellulosic biomass from agricultural waste to the circular economy: a review with focus on biofuels, biocomposites and bioplastics. Journal of Cleaner Production 2023, 402 , 136815. https://doi.org/10.1016/j.jclepro.2023.136815
  9. Hong He, Hui Teng, Fengping An, Yiwei Wang, Renhui Qiu, Lei Chen, Hongbo Song. Nanocelluloses review: Preparation, biological properties, safety, and applications in the food field. Food Frontiers 2023, 4 (1) , 85-99. https://doi.org/10.1002/fft2.208
  10. Zhipeng Qiu, Xixi Zeng, Jinchuan Xu, Bo Zheng, Ling Chen. Regioselective C6-OH oxidation of starch by laccase-TEMPO-system: A multi-scale structure evolution and water absorption properties study. Industrial Crops and Products 2023, 193 , 116148. https://doi.org/10.1016/j.indcrop.2022.116148
  11. Xiaoyu Lv, Jianing Han, Min Liu, Hui Yu, Kaihua Liu, Yifan Yang, Yi Sun, Panpan Pan, Zhenlin Liang, Lirong Chang, Jingdi Chen. Overview of preparation, modification, and application of tunicate-derived nanocellulose. Chemical Engineering Journal 2023, 452 , 139439. https://doi.org/10.1016/j.cej.2022.139439
  12. Lingfeng Long, Yun Hu, Fubao Sun, Wa Gao, Zhikui Hao, Heng Yin. Advances in lytic polysaccharide monooxygenases with the cellulose-degrading auxiliary activity family 9 to facilitate cellulose degradation for biorefinery. International Journal of Biological Macromolecules 2022, 219 , 68-83. https://doi.org/10.1016/j.ijbiomac.2022.07.240
  13. Dileswar Pradhan, Amit K. Jaiswal, Swarna Jaiswal. Emerging technologies for the production of nanocellulose from lignocellulosic biomass. Carbohydrate Polymers 2022, 285 , 119258. https://doi.org/10.1016/j.carbpol.2022.119258
  14. Laleh Solhi, James Li, Joyce Li, Nathan M. I. Heyns, Harry Brumer. Oxidative enzyme activation of cellulose substrates for surface modification. Green Chemistry 2022, 24 (10) , 4026-4040. https://doi.org/10.1039/D2GC00393G
  15. Anthi Karnaouri, Koar Chorozian, Dimitrios Zouraris, Antonis Karantonis, Evangelos Topakas, Ulrika Rova, Paul Christakopoulos. Lytic polysaccharide monooxygenases as powerful tools in enzymatically assisted preparation of nano-scaled cellulose from lignocellulose: A review. Bioresource Technology 2022, 345 , 126491. https://doi.org/10.1016/j.biortech.2021.126491
  16. Madhu Nair Muraleedharan, Anthi Karnaouri, Maria Piatkova, Maria-Ximena Ruiz-Caldas, Leonidas Matsakas, Bing Liu, Ulrika Rova, Paul Christakopoulos, Aji P. Mathew. Isolation and modification of nano-scale cellulose from organosolv-treated birch through the synergistic activity of LPMO and endoglucanases. International Journal of Biological Macromolecules 2021, 183 , 101-109. https://doi.org/10.1016/j.ijbiomac.2021.04.136

ACS Sustainable Chemistry & Engineering

Cite this: ACS Sustainable Chem. Eng. 2020, 8, 50, 18400–18412
Click to copy citationCitation copied!
https://doi.org/10.1021/acssuschemeng.0c05036
Published December 8, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Article Views

2257

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Flow diagram of the experimental process.

    Figure 2

    Figure 2. HPAEC chromatograph of soluble products detected after activity of MtLPMO on OS-pretreated tunic. The profile of the products reveals a mixed C1/C4-oxidizing activity. The products of blank reactions in the absence of enzyme (blank reaction 1, without enzyme, only ascorbic acid, and copper equivalent to LPMOs added) and electron donor (blank reaction 2 and only MtLPMO) are also depicted.

    Figure 3

    Figure 3. (A) AFM images of TNCs isolated from the untreated material (A), OS-pretreated material (B), OS-pretreated material w/o bleaching (Bw/o), OS/LPMO-pretreated material (C), and OS/LPMO-pretreated material w/o bleaching (Cw/o). (B) TNC diameter measured with Nanoscope V software.

    Figure 4

    Figure 4. TGA graphs of the TNCs isolated from tunic biomass after different treatments.

    Figure 5

    Figure 5. (A) XPS detailed spectra of the C 1 s region before (solid line) and after (dashed line) the LPMO post-treatment of the untreated material (A), OS-pretreated material (B), and OS/LPMO-pretreated material (C). (B) Stoichiometric amount of the molecular species.

    Figure 6

    Figure 6. Growth curve of B. subtilis and E. coli in the presence of 0.25% w/v TNCs: untreated material (A), OS-pretreated material (B), OS/LPMO-pretreated material (C), and OS/LPMO-pretreated material w/o bleaching (Cw/o).

    Figure 7

    Figure 7. (A) AFM images of the untreated material (Apost-treated), OS-pretreated material (Bpost-treated), OS/LPMO-pretreated material (Cpost-treated), and OS/LPMO-pretreated material w/o bleaching (Cw/o - post-treated) and (B) TNC diameter measurements after LPMO post-treatment, as identified by the Nanoscope V software.

  • References


    This article references 75 other publications.

    1. 1
      Xu, X.; Liu, F.; Jiang, L.; Zhu, J. Y.; Haagenson, D.; Wiesenborn, D. P. Cellulose nanocrystals vs. cellulose nanofibrils: a comparative study on their microstructures and effects as polymer reinforcing agents. ACS Appl. Mater. Interfaces 2013, 5, 29993009,  DOI: 10.1021/am302624t
    2. 2
      Lin, N.; Dufresne, A. Nanocellulose in biomedicine: Current status and future prospect. Eur. Polym. J. 2014, 59, 302325,  DOI: 10.1016/j.eurpolymj.2014.07.025
    3. 3
      Salas, C.; Nypelö, T.; Rodriguez-Abreu, C.; Carrillo, C.; Rojas, O. J. Nanocellulose properties and applications in colloids and interfaces. 2014. Curr. Opin. Colloid Interface Sci. 2014, 19, 383396,  DOI: 10.1016/j.cocis.2014.10.003
    4. 4
      Lee, K.-Y.; Aitomäki, Y.; Berglund, L. A.; Oksman, K.; Bismarck, A. On the use of nanocellulose as reinforcement in polymer matrix composites. Compos. Sci. Technol. 2014, 105, 1527,  DOI: 10.1016/j.compscitech.2014.08.032
    5. 5
      Klemm, D.; Kramer, F.; Moritz, S.; Lindstrom, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: a new family of nature-based materials. Angew. Chem., Int. Ed. 2011, 50, 54385466,  DOI: 10.1002/anie.201001273
    6. 6
      Habibi, Y.; Lucia, L. A.; Rojas, O. J. Cellulose nanocrystals: chemistry, self-assembly, and applications. Chem. Rev. 2010, 110, 34793500,  DOI: 10.1021/cr900339w
    7. 7
      Mariano, M.; El Kissi, N.; Dufresne, A. Cellulose nanocrystals and related nanocomposites: review of some properties and challenges. J. Polym. Sci., Part B: Polym. Phys. 2014, 52, 791806,  DOI: 10.1002/polb.23490
    8. 8
      Jonoobi, M.; Oladi, R.; Davoudpour, Y.; Oksman, K.; Dufrense, A.; Hamzeh, Y.; Davoodi, R. Different preparation methods and properties of nanostructured cellulose from various natural resources and residues: a review. Cellulose 2015, 22, 935969,  DOI: 10.1007/s10570-015-0551-0
    9. 9
      Börjesson, M.; Westman, G. Crystalline Nanocellulose — Preparation, Modification, and Properties, Cellulose-Fundamental Aspects and Current Trends, Matheus Poletto and Heitor Luiz Ornaghi Junior, IntechOpen, December 9th 2015.  DOI: 10.5772/61899 . Available from: https://www.intechopen.com/books/cellulose-fundamental-aspects-and-current-trends/crystalline-nanocellulose-preparation-modification-and-properties
    10. 10
      Zhao, Y.; Li, J. Excellent chemical and material cellulose from tunicates: diversity in cellulose production yield and chemical and morphological structures from different tunicate species. Cellulose 2014, 21, 34273441,  DOI: 10.1007/s10570-014-0348-6
    11. 11
      Mathew, A. P.; Dufresne, A. Morphological investigation of nanocomposites from sorbitol plasticized starch and tunicin whiskers. Biomacromolecules 2002, 3, 609617,  DOI: 10.1021/bm0101769
    12. 12
      Zhao, Y.; Moser, C.; Lindström, M. E.; Henriksson, G.; Li, J. Cellulose nanofibers from softwood, hardwood, and tunicate: preparation–structure–film performance interrelation. Appl. Mater. Interfaces 2017, 9, 1350813519,  DOI: 10.1021/acsami.7b01738
    13. 13
      Sacui, I. A.; Nieuwendaal, R. C.; Burnett, D. J.; Stranick, S. J.; Jorfi, M.; Weder, C.; Foster, E. J.; Olsson, R. T.; Gilman, J. W. Comparison of the properties of cellulose nanocrystals and cellulose nanofibrils isolated from bacteria, tunicate, and wood processed using acid, enzymatic, mechanical, and oxidative methods. ACS Appl. Mater. Interfaces 2014, 6, 61276138,  DOI: 10.1021/am500359f
    14. 14
      Elazzouzi-Hafraoui, S.; Nishiyama, Y.; Putaux, J.-L.; Heux, L.; Dubreuil, F.; Rochas, C. The shape and size distribution of crystalline nanoparticles prepared by acid hydrolysis of native cellulose. Biomacromolecules 2008, 9, 5765,  DOI: 10.1021/bm700769p
    15. 15
      Mathew, A. P.; Thielemans, W.; Dufresne, A. Mechanical properties of nanocomposites from sorbitol plasticized starch and tunicin whiskers. J. Appl. Polym. Sci. 2008, 109, 40654074,  DOI: 10.1002/app.28623
    16. 16
      Squinca, P.; Bilatto, S.; Badino, A. C.; Farinas, C. S. Nanocellulose production in future biorefineries: an integrated approach using tailor-made enzymes. ACS Sustainable Chem. Eng. 2020, 8, 22772286,  DOI: 10.1021/acssuschemeng.9b06790
    17. 17
      Henriksson, M.; Henriksson, G.; Berglund, L. A.; Lindström, T. An environmentally friendly method for enzyme-assisted preparation of microfibrillated cellulose (MFC) nanofibers. Eur. Polym. J. 2007, 43, 34343441,  DOI: 10.1016/j.eurpolymj.2007.05.038
    18. 18
      Hu, J.; Tian, D.; Renneckar, S.; Saddler, J. N. Enzyme-mediated nanofibrillation of cellulose by the synergistic actions of an endoglucanase, lytic polysaccharide monooxygenase (LPMO) and xylanase. Sci. Rep. 2018, 8, 3195,  DOI: 10.1038/s41598-018-21016-6
    19. 19
      Hu, J.; Arantes, V.; Pribowo, A.; Saddler, J. N. The synergistic action of accessory enzymes enhances the hydrolytic potential of a “cellulase mixture” but is highly substrate specific. Biotechnol. Biofuels 2013, 6, 112,  DOI: 10.1186/1754-6834-6-112
    20. 20
      Westereng, B.; Ishida, T.; Vaaje-Kolstad, G.; Wu, M.; Eijsink, V. G.; Igarashi, K.; Samejima, M.; Ståhlberg, J.; Horn, S. J.; Sandgren, M. The putative endoglucanase PcGH61D from Phanerochaete chrysosporium is a metal-dependent oxidative enzyme that cleaves cellulose. PLoS One 2011, 6, e27807  DOI: 10.1371/journal.pone.0027807
    21. 21
      Harris, P. V.; Welner, D.; McFarland, K. C.; Re, E.; Navarro Poulsen, J. C.; Brown, K.; Salbo, R.; Ding, H.; Vlasenko, E.; Merino, S.; Xu, F.; Cherry, J.; Larsen, S.; Lo Leggio, L. Stimulation of lignocellulosic biomass hydrolysis by proteins of Glycoside Hydrolase Family 61: Structure and function of a large, enigmatic family. Biochemistry 2010, 49, 33053316,  DOI: 10.1021/bi100009p
    22. 22
      Vaaje-Kolstad, G.; Westereng, B.; Horn, S. J.; Liu, Z.; Zhai, H.; So̷rlie, M.; Eijsink, V. G. An oxidative enzyme boosting the enzymatic conversion of recalcitrant polysaccharides. Science 2010, 330, 219222,  DOI: 10.1126/science.1192231
    23. 23
      Forsberg, Z.; Vaaje-Kolstad, G.; Westereng, B.; Bunæs, A. C.; Stenstro̷m, Y.; MacKenzie, A.; So̷rlie, M.; Horn, S. J.; Eijsink, V. G. Cleavage of cellulose by a CBM33 protein. Protein Sci. 2011, 20, 14791483,  DOI: 10.1002/pro.689
    24. 24
      Quinlan, R. J.; Sweeney, M. D.; Lo Leggio, L.; Otten, H.; Poulsen, J. C.; Johansen, K. S.; Krogh, K. B.; Jo̷rgensen, C. I.; Tovborg, M.; Anthonsen, A.; Tryfona, T.; Walter, C. P.; Dupree, P.; Xu, F.; Davies, G. J.; Walton, P. H. Insights into the oxidative degradation of cellulose by a copper metalloenzyme that exploits biomass components. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 1507915084,  DOI: 10.1073/pnas.1105776108
    25. 25
      Beeson, W. T.; Phillips, C. M.; Cate, J. H. D.; Marletta, M. A. Oxidative cleavage of cellulose by fungal copper-dependent polysaccharide monooxygenases. J. Am. Chem. Soc. 2011, 134, 890892,  DOI: 10.1021/ja210657t
    26. 26
      Moreau, C.; Tapin-Lingua, S.; Grisel, S.; Gimbert, I.; Le Gall, S.; Petit-Conil, M.; Berrin, J. G.; Cathala, B.; Villares, A. Lytic polysaccharide monooxygenases (LPMOs) facilitate cellulose nanofibrils production. Biotechnol. Biofuels. 2019, 12, 156,  DOI: 10.1186/s13068-019-1501-0
    27. 27
      Villares, A.; Moreau, C.; Bennati-Granier, C.; Garajova, S.; Foucat, L.; Falourd, X.; Saake, B.; Berrin, J. G.; Cathala, B. Lytic polysaccharide monooxygenases disrupt the cellulose fibers structure. Sci. Rep. 2017, 7, 40262,  DOI: 10.1038/srep40262
    28. 28
      Koskela, S.; Wang, S.; Xu, D.; Yang, X.; Li, K.; Berglund, L.; McKee, L. S.; Bulone, V.; Zhou, Q. Lytic polysaccharide monooxygenase (LPMO) mediated production of ultra-fine cellulose nanofibres from delignified softwood fibres. Green Chem. 2019, 21, 59245933,  DOI: 10.1039/C9GC02808K
    29. 29
      Spangler, D.; Rothenburger, S.; Nguyen, K.; Jampani, H.; Weiss, S.; Bhende, S. In vitro antimicrobial activity of oxidized regenerated cellulose against antibiotic-resistant microorganisms. Surg. Infect. (Larchmt) 2003, 4, 255262,  DOI: 10.1089/109629603322419599
    30. 30
      Wohlhauser, S.; Delepierre, G.; Labet, M.; Morandi, G.; Thielemans, W.; Weder, C.; Zoppe, J. O. Grafting polymers from cellulose nanocrystals: synthesis, properties and applications. Macromolecules 2018, 51, 61576189,  DOI: 10.1021/acs.macromol.8b00733
    31. 31
      Quintana, E.; Roncero, M. B.; Vidal, T.; Valls, C. Cellulose oxidation by Laccase-TEMPO treatments. Carbohydr. Polym. 2017, 157, 14881495,  DOI: 10.1016/j.carbpol.2016.11.033
    32. 32
      Vaaje-Kolstad, G.; Forsberg, Z.; Loose, J. S.; Bissaro, B.; Eijsink, V. G. Structural diversity of lytic polysaccharide monooxygenases. Curr. Opin. Struct. Biol. 2017, 44, 6776,  DOI: 10.1016/j.sbi.2016.12.012
    33. 33
      Wang, D.; Li, J.; Salazar-Alvarez, G.; McKee, L. S.; Srivastava, V.; Sellberg, J. A.; Bulome, V.; Hsieh, Y. S. Y. Production of functionalised chitins assisted by fungal lytic polysaccharide monooxygenase. Green Chem. 2018, 20, 20912100,  DOI: 10.1039/C8GC00422F
    34. 34
      Valls, C.; Pastor, F. I. J.; Roncero, M. B.; Vidal, T.; Diaz, P.; Martinez, J.; Valenzuela, S. V. Assessing the enzymatic effects of cellulases and LPMO in improving mechanical fibrillation of cotton linters. Biotechnol. Biofuels 2019, 12, 161,  DOI: 10.1186/s13068-019-1502-z
    35. 35
      Hrůzová, K.; Matsakas, L.; Karnaouri, A.; Norén, F.; Rova, U.; Christakopoulos, P. Second-generation biofuel production from the marine filter feeder Ciona intestinalis. ACS Sustainable Chem. Eng. 2020, 8, 83738380,  DOI: 10.1021/acssuschemeng.0c02417
    36. 36
      Sluiter, J. B.; Ruiz, R. O.; Scarlata, C. J.; Sluiter, A. D.; Templeton, D. W. Compositional analysis of lignocellulosic feedstocks. 1. Review and description of methods. J. Agric. Food Chem. 2010, 58, 90439053,  DOI: 10.1021/jf1008023
    37. 37
      Karnaouri, A.; Muraleedharan, M. N.; Dimarogona, M.; Topakas, E.; Rova, U.; Sandgren, M.; Christakopoulos, P. Recombinant expression of thermostable processive MtEG5 endoglucanase and its synergism with MtLPMO from Myceliophthora thermophila during the hydrolysis of lignocellulosic substrates. Biotechnol. Biofuels 2017, 10, 126,  DOI: 10.1186/s13068-017-0813-1
    38. 38
      Lowry, O. H.; Rosebrough, N. J.; Lewis Farr, A.; Randall, R. J. Protein Measurement with the Folin phenol reagent. J. Biol. Chem. 1951, 193, 265275
    39. 39
      Schulz, B. L.; Packer, N. H.; Karlsson, N. G. Small-scale analysis of O-linked oligosaccharides from glycoproteins and mucins separated by gel electrophoresis. Anal. Chem. 2002, 74, 60886097,  DOI: 10.1021/ac025890a
    40. 40
      Bondeson, D.; Mathew, A.; Oksman, K. Optimization of the isolation of nanocrystals from microcrystalline cellulose by acid hydrolysis. Cellulose 2006, 13, 171180,  DOI: 10.1007/s10570-006-9061-4
    41. 41
      Klarhöfer, L.; Roos, B.; Viöl, W.; Höfft, O.; Dieckhoff, S.; Kempter, V.; Maus-Friedrichs, W. Valence band spectroscopy on lignin. Holzforschung 2008, 62.  DOI: 10.1515/HF.2008.116
    42. 42
      Krischok, S.; Höfft, O.; Günster, J.; Stultz, J.; Goodman, D. W.; Kempter, V. H2O interaction with bare and Li-precovered TiO2: Studies with electron spectroscopies (MIES and UPS (HeI and II)). Surf. Sci. 2001, 495, 818,  DOI: 10.1016/S0039-6028(01)01570-9
    43. 43
      Hercules, D. M.; Cox, L. E.; Onisick, S.; Nichols, G. D.; Carver, J. C. Electron spectroscopy (ESCA). Use for trace analysis. Anal. Chem. 1973, 45, 19731975,  DOI: 10.1021/ac60333a037
    44. 44
      Nam, S.; French, A. D.; Condon, B. D.; Concha, M. Segal crystallinity index revisited by the simulation of X-ray diffraction patterns of cotton cellulose Iβ and cellulose II. Carbohydr. Polym. 2016, 135, 19,  DOI: 10.1016/j.carbpol.2015.08.035
    45. 45
      Thoresen, P. P.; Matsakas, L.; Rova, U.; Christakopoulos, P. Recent advances in organosolv fractionation: Towards biomass fractionation technology of the future. Bioresour. Technol. 2020, 306, 123189,  DOI: 10.1016/j.biortech.2020.123189
    46. 46
      Kalogiannis, K. G.; Karnaouri, A.; Michailof, C.; Tzika, A. M.; Asimakopoulou, G.; Topakas, E.; Lappas, A. A. OxiOrganosolv: A novel acid free oxidative organosolv fractionation for lignocellulose fine sugar streams. Bioresour. Technol. 2020, 313, 123599,  DOI: 10.1016/j.biortech.2020.123599
    47. 47
      Mantanis, G. I.; Young, R. A.; Rowell, R. M. Swelling of compressed cellulose fiber webs in organic liquids. Cellulose 1995, 2, 122,  DOI: 10.1007/BF00812768
    48. 48
      Sannigrahia, P.; Miller, S. J.; Ragauskas, A. J. Effects of organosolv pretreatment and enzymatic hydrolysis on cellulose structure and crystallinity in Loblolly pine. Carbohydr. Res. 2010, 345, 965970,  DOI: 10.1016/j.carres.2010.02.010
    49. 49
      Levasseur, A.; Drula, E.; Lombard, V.; Coutinho, P. M.; Henrissat, B. Expansion of the enzymatic repertoire of the CAZy database to integrate auxiliary redox enzymes. Biotechnol. Biofuels 2013, 6, 41,  DOI: 10.1186/1754-6834-6-41
    50. 50
      Silva, C. d. O. G.; Teixeira, T. S.; Rodrigues, K. B.; Souza, A. A.; Monclaro, A. V.; Mendes, T. D. Combination of MALDI-TOF MS and UHPLC-ESI-MS for the characterization of lytic polysaccharide monooxygenase activity. Anal. Methods 2020, 12, 149161,  DOI: 10.1039/c9ay01774g
    51. 51
      Zhao, Y.; Li, J. Ascidian bioresources: common and variant chemical compositions and exploitation strategy – examples of Halocynthia roretzi, Styela plicata, Ascidia sp. and Ciona intestinalis. Z. Naturforschung C 2016, 71, 165180,  DOI: 10.1515/znc-2016-0012
    52. 52
      Eibinger, M.; Ganner, T.; Bubner, P.; Rošker, S.; Kasciddvgdgracher, D.; Haltrich, D.; Ludwig, R.; Plank, H.; Nidetzky, B. Cellulose surface degradation by a lytic polysaccharide monooxygenase and its effect on cellulase hydrolytic efficiency. J. Biol. Chem. 2014, 289, 3592935938,  DOI: 10.1074/jbc.M114.602227
    53. 53
      Siqueira, P.; Siqueira, É.; De Lima, A. E.; Siqueira, G.; Pinzón-Garcia, A. D.; Lopes, A. P.; Segura, M. E. C.; Isaac, A.; Pereira, F. V.; Botaro, V. R. Three-dimensional stable alginate-nanocellulose gels for biomedical applications: towards tunable mechanical properties and cell growing. Nanomaterials 2019, 9, 78,  DOI: 10.3390/nano9010078
    54. 54
      Xie, K.; Yu, Y.; Shi, Y. Synthesis and characterization of cellulose/silica hybrid materials with chemical crosslinking. Carbohydr. Polym. 2009, 78, 799805,  DOI: 10.1016/j.carbpol.2009.06.019
    55. 55
      Bhattacharjee, S. DLS and zeta potential - What they are and what they are not?. J. Controlled Release 2016, 235, 337351,  DOI: 10.1016/j.jconrel.2016.06.017
    56. 56
      Johnston, L. J.; Jakubek, Z. J.; Beck, S.; Araki, J.; Cranston, E. D.; Danumah, C.; Fox, D.; Li, H.; Wang, J.; Mester, Z.; Moores, A.; Rabb, S. A.; Murphy, K. E.; Stephan, C. Determination of sulfur and sulfate half-ester content in cellulose nanocrystals: an interlaboratory comparison. Metrologia 2018, 55, 872,  DOI: 10.1088/1681-7575/aaeb60
    57. 57
      Bennati-Granier, C.; Garajova, S.; Champion, C.; Grisel, S.; Haon, M.; Zhou, S.; Fanuel, M.; Ropartz, D.; Rogniaux, H.; Gimbert, I.; Record, E.; Berrin, J. G. Substrate specificity and regioselectivity of fungal AA9 lytic polysaccharide monooxygenases secreted by Podospora anserina. Biotechnol. Biofuels 2015, 8, 90,  DOI: 10.1186/s13068-015-0274-3
    58. 58
      Isogai, A.; Saito, T.; Fukuzumi, H. TEMPO-oxidized cellulose nanofibers. Nanoscale 2011, 3, 7185,  DOI: 10.1039/C0NR00583E
    59. 59
      Zhang, R.; Liu, Y. High energy oxidation and organosolv solubilization for high yield isolation of cellulose nanocrystals (CNC) from Eucalyptus hardwood. Sci. Rep. 2018, 8, 16505,  DOI: 10.1038/s41598-018-34667-2
    60. 60
      Oksman, K.; Etang, J. A.; Mathew, A. P.; Jonobi, M. Cellulose nanowhiskers separated from a bio-residue from wood bioethanol production. Biomass Bioenergy 2011, 35, 146152,  DOI: 10.1016/j.biombioe.2010.08.021
    61. 61
      Herrera, M. A.; Mathew, A. P.; Oksman, K. Characterization of cellulose nanowhiskers: A comparison of two industrial bioresidues. Iop Conf. Ser.: Mater. Sci. Eng. 2012, 31, 012006  DOI: 10.1088/1757-899X/31/1/012006
    62. 62
      Roman, M.; Winter, W. T. Effect of sulfate groups from sulfuric acid hydrolysis on the thermal degradation behavior of bacterial cellulose. Biomacromolecules 2004, 5, 16711677,  DOI: 10.1021/bm034519+
    63. 63
      Rosa, M. F.; Medeiros, E. S.; Malmonge, J. A.; Gregorski, K. S.; Wood, D. F.; Mattoso, L. H. C.; Glenn, G.; Orts, W. J.; Imam, S. H. Cellulose nanowhiskers from coconut husk fibers: Effect of preparation conditions on their thermal and morphological behavior. Carbohydr. Polym. 2010, 81, 8392,  DOI: 10.1016/j.carbpol.2010.01.059
    64. 64
      Stevens, J. S.; Schroeder, S. L. M. Quantitative analysis of saccharides by X-ray photoelectron spectroscopy. Surf. Interface Anal. 2009, 41, 453462,  DOI: 10.1002/sia.3047
    65. 65
      Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy. Physical Electronics 1995, ISBN: 9780964812413
    66. 66
      Mathew, A. P.; Oksman, K.; Karim, Z.; Liu, P.; Khan, S. A.; Naseri, N. Process scale up and characterization of wood cellulose nanocrystals hydrolysed using bioethanol pilot plant. Ind. Crops Prod. 2014, 58, 212219,  DOI: 10.1016/j.indcrop.2014.04.035
    67. 67
      Missoum, K.; Sadocco, P.; Causio, J.; Belgacem, M. N.; Bras, J. Antibacterial activity and biodegradability assessment of chemically grafted nanofibrillated cellulose. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 45, 477483,  DOI: 10.1016/j.msec.2014.09.037
    68. 68
      Tavakolian, M.; Okshevsky, M.; van de Ven, T. G. M.; Tufenkji, N. Developing antibacterial nanocrystalline cellulose using natural antibacterial agents. ACS Appl. Mater. Interfaces 2018, 10, 3382733838,  DOI: 10.1021/acsami.8b08770
    69. 69
      Vytrasova, J.; Tylsova, A.; Brozkova, I.; Cervenka, L.; Pejchalova, M.; Havelka, P. Antimicrobial effect of oxidized cellulose salts. J. Ind. Microbiol. Biotechnol. 2008, 35, 12471252,  DOI: 10.1007/s10295-008-0421-y
    70. 70
      Abaev, I. K.; Kaputskiĭ, V. E.; Adarchenko, A. A.; Sobeshchukh, O. P. Mechanism of the antibacterial action of monocarboxycellulose and other ion-exchange derivatives of cellulose. Antibiot. Med. Biotekhnol. 1986, 31, 624628
    71. 71
      Saini, S.; Falco, Ç. Y.; Belgacem, M. N.; Bras, J. Surface cationized cellulose nanofibrils for the production of contact active antimicrobial surfaces. Carbohydr. Polym. 2016, 135, 239247,  DOI: 10.1016/j.carbpol.2015.09.002
    72. 72
      Breijyeh, Z.; Jubeh, B.; Karaman, R. Resistance of Gram-negative bacteria to current antibacterial agents and approaches to resolve it. Molecules 2020, 25, 1340,  DOI: 10.3390/molecules25061340
    73. 73
      Herbaud, M. L.; Guiseppi, A.; Denizot, F.; Haiech, J.; Kilhoffer, M. C. Calcium signalling in Bacillus subtilis. Biochim. Biophys. Acta 1998, 1448, 212226,  DOI: 10.1016/s0167-4889(98)00145-1
    74. 74
      Onoda, T.; Enokizono, J.; Kaya, H.; Oshima, A.; Freestone, P.; Norris, V. Effects of calcium and calcium chelators on growth and morphology of Escherichia coli L-form NC-7. J. Bacteriol. 2000, 182, 14191422,  DOI: 10.1128/jb.182.5.1419-1422.2000
    75. 75
      Valenzuela, S. V.; Valls, C.; Schink, V.; Sánchez, D.; Blanca Roncero, M.; Diaz, P.; Martinez, J.; Javier Pastor, F. I. Differential activity of lytic polysaccharide monooxygenases on celluloses of different crystallinity. Effectiveness in the sustainable production of cellulose nanofibrils. Carbohydr. Polym. 2019, 207, 5967,  DOI: 10.1016/j.carbpol.2018.11.076
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssuschemeng.0c05036.

    • Mass spectrometry data of degradation products generated by MtLPMO, SEM figures, XPS spectra, relative amount of the molecular species, and particle length data for TNCs isolated following different treatments (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.