Next Article in Journal
A Comparative Thermoacoustic Insulation Study of Silica Aerogels Reinforced with Reclaimed Textile Fibres: Cotton, Polyester and Wool
Next Article in Special Issue
Highly Conductive, Anti-Freezing Hemicellulose-Based Hydrogels Prepared via Deep Eutectic Solvents and Their Applications
Previous Article in Journal
Hydrogel-Based Biosensors for Effective Therapeutics
Previous Article in Special Issue
Preparation and Characterization of Porous Cellulose Acetate Nanofiber Hydrogels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Cellulose-Based Hydrogels Prepared by Ionic Liquid-Based Processes

Department of Materials Science and Bioengineering, School of Engineering, Nagaoka University of Technology, Nagaoka 940-2188, Niigata, Japan
Gels 2023, 9(7), 546; https://doi.org/10.3390/gels9070546
Submission received: 1 June 2023 / Revised: 15 June 2023 / Accepted: 22 June 2023 / Published: 5 July 2023
(This article belongs to the Special Issue Advances in Cellulose-Based Hydrogels (2nd Edition))

Abstract

:
This review summarizes the recent advances in preparing cellulose hydrogels via ionic liquid-based processes and the applications of regenerated cellulose hydrogels/iongels in electrochemical materials, separation membranes, and 3D printing bioinks. Cellulose is the most abundant natural polymer, which has attracted great attention due to the demand for eco-friendly and sustainable materials. The sustainability of cellulose products also depends on the selection of the dissolution solvent. The current state of knowledge in cellulose preparation, performed by directly dissolving in ionic liquids and then regenerating in antisolvents, as described in this review, provides innovative ideas from the new findings presented in recent research papers and with the perspective of the current challenges.

Graphical Abstract

1. Introduction

The development of products with durability, repairability, and reusability is a sustainable way to mitigate environmental impacts and move toward a circular economy. The impact on the environment has manifested from the continuous use of petroleum-based resources that must be substituted with alternative renewable and eco-friendly resources [1,2]. Biomass-based raw materials are a promising replacement of non-renewable, petroleum-based materials for advancing the efforts toward achieving a sustainable society. In order to achieve the desired levels of sustainable growth and to mitigate environmental pollution, biomass has been applied to reduce wastes and develop value-added products [3]. Expanding the use of such biomass-based materials does not only address the societal need for low-cost materials, but also reduces carbon footprints [4,5]. As a consequence of the abundance of renewable biomass materials in nature, its rational use is of great significance for the sustainable development of the chemical industry [6].
Cellulose is a biomass source with the potential to replace fossil fuels through its use in the production of value-added chemicals such as 5-hydroxymethylfurfural and levulinic acid [7]. It is a natural polymer available from diverse sources such as plants and from the extracellular production of several microbial genera, particularly Gluconacetobacter xylinus [8,9]. Cellulose can be obtained from wastes such as paper [10], textile [11,12,13], oil palm frond [14], and crop straw [15,16]. Due to its availability, biodegradability, non-toxicity, hydrophilic nature, inexpensive cost, and multifunctionality, cellulose is attractive and shows promise for utilization in sustainable material engineering. The processibility of cellulose allows it to be formed into various types of materials, for example, into hydrogels [17] and/or applied in its dry form as aerogels [18]. Cellulose hydrogels (jelly-like solids) are obtained by the regeneration or crosslinking of cellulose in solutions into a 3D shape, which are different from the non-cross-linked gel-like suspension of nano- or micro-fibrillated cellulose [19,20,21]. Due to the presence of hydroxyl groups in cellulose, the cross-linking in the hydrogel is formed through hydrogen bonding, which is reconstructed during the cellulose regeneration and solvent removal process [22]. Numerous hydroxyl groups in cellulose and other hydrophilic groups, e.g., carboxyl and amino groups, present in the cellulose derivatives, enable the entrapment of large volumes of water [23,24]. Depending on their applications in biotechnology, cosmetology, and ecology, a variety of sizes, from macro to nanogels, such as membranes, films, fibers, microspheres, or nanoparticles, can be fabricated by various interactions [25,26]. Those interactions are hydrogen bonding, electrostatic interactions, van der Waals forces, and physical entanglements of the cellulose molecules and/or their blends [25,26]. One of the main limitations of processing native cellulose is that it is difficult to dissolve in conventional organic or inorganic solvents due to the abundance of hydrogen bonds. The cellulose derivatives, e.g., carboxymethylcellulose, are much easier to dissolve, but multiple production steps are required to produce such derivatives [27]. The native cellulose is a semi-crystalline polymer comprising a high order of crystalline regions and a lower order of amorphous regions [28]. The rich hydrogen bonds between and within cellulose polymer chains are beneficial for the mechanical properties of cellulose crystals which greatly enhances mechanical strength of highly crystalline cellulose materials [29]. However, the resultant strong and complex 3D inter- and intramolecular hydrogen bonding network contributes to the chain rigidity and insolubility of cellulose, which must be disrupted for dissolution [30,31]. The bonding disruption can be achieved by the derivatization of hydroxyl groups of cellulose or comparative formation of a stronger hydrogen bond with other strong hydrogen-bonding solvents [30,31]. The solvents that have been used in the sol–gel process of cellulose are: N-dimethylacetamide (DMAc)/LiCl [32]; tetrabutylammonium fluoride/dimethyl sulfoxide (TBAF/DMSO) [33]; alkali/urea (NaOH/urea/H2O, LiOH/thiourea/H2O, etc.) [34,35]; N-methyl morpholine oxide (NMMO) [36,37]; NMMO/DMSO [38]; ZnCl2 aqueous solutions [39,40,41,42]; ZnCl2 molten salt hydrates [43]; molten hydrates of LiClO4 NaSCN/KSCN/LiSCN and LiCl/ZnCl2 [44]; molten salt hydrates of ZnBr2 and FeCl3 [45]; LiBr trihydrate molten salt [46]; and ethylenediamine solutions of NaSCN, KSCN, or NaI [44,47]. However, harsh dissolution conditions, such as high temperature, strong corrosiveness from the alkali metal hydroxide, and long processing time, have been reported [48]. Direct non-derivatizing dissolution of cellulose by the NMMO-based Lyocell process is commercialized. Although it is less-time consuming, the thermal stability and side reaction of NMMO, which decreases the performance of hydrogel fibers, remains a concern [49]. The ZnCl2 aqueous solutions and inorganic molten salt hydrates have been of interest as cellulose dissolution solvents in recent years due to their ease of preparation and low cost [41,42,50]. However, the solubility of cellulose in such inorganic salts is low and requires a high temperature or long dissolution time, causing energy consumption and hydrolysis [50,51].
Ionic liquids (ILs) are considered to be desirable eco-friendly solvents and have been used to replace the organic solvents for cellulose dissolution under milder conditions. A regular crystalline structure is difficult to form because of the asymmetry of ILs and the reduced electrostatic forces between the cations and anions [52]. Therefore, they can be liquid at low or room temperature. The ILs can form comparatively stronger hydrogen bonding with the native molecular hydrogen bond network within cellulose structures, causing hydrogen bonding disruption and dissolution. Due to their main advantages besides their high cellulose dissolution ability, ILs have low vapor pressure and are efficiently recycled in order to avoid their propagation in ecosystems. Cellulosic solutions in ILs can also be converted into ionogels or hydrogels by simple operations such as regeneration in cellulose antisolvents [53]. Based on the simplicity and eco-friendliness of IL-based methods, this review summarized the findings of the research conducted in the past decade on fabricating cellulose hydrogels by using processes under pure IL, IL/cosolvent, and other new IL systems. Regarding the importance of continuous application for the green solvent, IL, in the direct dissolution process, the cellulose hydrogels and related products are truly sustainable according to the circular economy because their processing is also sustainable.

2. Ionic Liquids Used in Dissolution of Cellulose

2.1. Pure Ionic Liquid Systems

Anions and cations of ILs strongly affect cellulose solubility. The ILs containing cations with an unsaturated heterocyclic ring have the ability to dissolve cellulose because of their π electron delocalization of the ring. Such cations are active to interact with cellulose by providing more space for anions to form hydrogen bonds with cellulose. In contrast, the larger volume of ILs containing cations with the saturated heterocyclic ring slows down cations and anions transfer, which is less favorable for cellulose dissolution [54]. The options for cellulose dissolution have been ILs with aromatic imidazolium or pyridinium cations [55,56]. The imidazolium salts, such as 1-allyl-3-methyl imidazolium chloride (AmimCl), 1-buthyl-3-methyl imidazolium chloride (BmimCl), and 1-ethyl-3-methylimidazolium acetate (EmimAc), have received attention [57,58]. A series of substituted imidazolium cations, paired with lactate or glycolate anions, was developed by Meenatchi et al. [59]. For 5% w/w of cellulose solution, 1-ethyl imidazolium lactate showed the highest dissolution power within 20 min at 80 °C because of its higher number of acidic protons (C-2, C-4, C-5) which, in turn, favor the formation of more hydrogen bonds with oxygen atoms of cellulose. The replacement of the acidic C-2 proton by a methyl group (2-methylimidazolium) reduced the dissolution power, which suggested the substitution decreased the hydrogen bond interactions [59].
The anions also play a vital role in the breakage of the native cellulose hydrogen bond network because of the strong electronegativity, allowing anions to form strong hydrogen bonds with the hydroxide groups of cellulose. After being combined with various cations possessing alkyl functional groups, Cl, acetate (Ac), dimethyl phosphate (DMP), and diethyl phosphate (DEP) anions (acidic protons) strongly exhibit the ability to dissolve cellulose [60]. In contrast, weak hydrogen-bonding anions, e.g., BF4, PF6, and Tf2N, are not suitable for cellulose dissolution [61,62]. The Cl-based ILs are widely adopted as a small-sized hydrophilic hydrogen bond acceptor, for instance BmimCl is frequently used due to its thermoplasticity and thermal processability [63,64,65,66]. The EmimAc provides better solubility for cellulose due to its lower viscosity (~140 mPas at 25 °C and 10 mPa s at 80 °C) and to the fact that it forms easily processable and stable cellulose solutions; it is also less toxic (LD50 > 2000 mg/kg) and less corrosive than BmimCl [67,68,69,70]. Therefore, it has been a choice for creating regenerated cellulose hydrogels, especially in the form of fibers, because it requires less energy during the shaping process [71]. The AmimCl is also able to dissolve cellulose at a higher concentration than BmimCl [55,72], owing to its high Cl concentration, electrochemical stability, and compatibility of hydrogen bond acceptors providing more interaction sites for hydroxyl groups of cellulose [73,74]. AmimCl is an effective IL for cellulose dissolution and hydrogel fabrication. The regenerated cellulose produced from AmimCl has the highest crystallinity, tensile strength, and transparency compared with those from BmimCl and EmimAc [72]. Even though AmimCl shows greater cellulose dissolution capability than EmimDEP [75], the content of depolymerized cellulose increases more with time and temperature in AmimCl than in EmimDEP [75]. Xu et al. [76] studied Amim(MeO)PHO2 IL for cellulose dissolution and decomposition. To increase the ionicity of this IL, Na2PHO3 inorganic salt, having an anion structure similar to that of Amim(MeO)PHO2, was added to form a composite solvent system (Na2PHO3/Amim(MeO)PHO2) and facilitate cellulose dissolution. The Na2PHO3 addition reduced H+ concentration of the solvent system, which inhibited the decomposition of cellulose. The lower viscosity of Na2PHO3/Amim(MeO)PHO2 solvent caused 22% w/w cellulose to dissolve at 80 °C over 48 h. With the increase in Na2PHO3 content, the degree of polymerization and thermal stability of the regenerated cellulose increased, but the crystallinity decreased relative to that of the original microcrystalline cellulose (MCC), because the long molecular chains exhibited difficulty in forming new hydrogen bonds during regeneration [76].
One of the advantages of imidazolium-based ILs is their capability to dissolve cellulose even at room temperature. However, some drawbacks are the high viscosity and difficulty in cellulose dispersion due to the strong association between the anions and cations [77]. When ILs are used in the presence of lignin in the pulp or impurities, side reactions such as acetylation may occur at high temperatures above 100 °C [78]. Therefore, the researchers have studied new ILs for efficient cellulose dissolution. Li et al. [79] developed a series of superbase-derived ILs such as: 1,8-diazabicyclo [5.4.0] undec-7-ene (DBU) including 1,8-diazabicyclo [5.4.0] undec-7-enium (DBUH) carboxylate, e.g., OAc−; and 1,5-diazabicyclo [4.3.0] non-5-ene (DBN) including 1,5-diazabicyclo [4.3.0] non-5-enium (DBNH) carboxylate with lower viscosity and high cellulose solubility. The DBUH ethoxyacetate has the higher cellulose dissolution capacity (14.8% w/w wood pulp at 80 °C) [80], but lower corrosion to steels than EmimAc and AmimCl [81]. On the other hand, some organic N-oxide-based solutions can dissolve cellulose efficiently, but the drawbacks are the instability of the N-oxides and the high dissolution temperature leading to the partial degradation of cellulose. Lui et al. [82] prepared DBN coupled with two N-oxides, i.e., pyridine N-oxide (PyO) and 2-picoline-N-oxide (PiO). The solvents dissolved cellulose efficiently at temperatures lower than 80 °C. The solvent DBN/PyO, at molar ratio of DBN and PyO of four, showed the cellulose solubility of 10.9 and 14.1% w/w at 50 and 70 °C, respectively. Galamba et al. [83] studied benzethonium- and didecyldimethylammonium- based ILs combined with short alkyl carboxylate anions to dissolve MCC at a concentration of 4% w/w. The polymeric cellulose hydrogel structures varied depending on the type of ILs and the ratios between cellulose and IL. However, after regeneration, some ILs remained in the hydrogel structures, which was advantageous in terms of the antibacterial/antimicrobial response of the hydrogels. Some new IL systems for cellulose dissolution, as reported in recent studies, are shown in Figure 1.

2.2. Ionic Liquid–Cosolvent Systems

One of the problems with using ILs to process cellulose is the high viscosity of the cellulose/IL solutions. The use of IL/cosolvent mixtures is an option that can modify the properties of the IL solvent by improving the rheological properties of the cellulose mixture, to reduce the viscosity of the solvent system and to accelerate the mass transfer process. The solvation of cation and anion by the cosolvents enhances the production of more free anions from anion–cation pairs, which readily interact with cellulose [84,85]. The frequently used cosolvents are polar aprotic organic solvents like DMSO, dimethylformamide (DMF), dimethylacetamide (DMAc), and dimethylimidazolinone (DMI). Zhou et al. [86] studied the cellulose solution state and formation mechanism by using these four cosolvents with EmimAc, BmimAc, AmimCl, and BmimCl. Acetate-based ILs and a high cosolvent content facilitated the formation of a molecularly dispersed state. In contrast, a state of coexistence of single molecular chains and undissolved cellulose microdomains gradually converted into a molecularly dispersed state when using chloride based ILs or a low content of cosolvents. The cosolvents DMF and DMAc promoted the molecular dispersion of cellulose better than DMSO and DMI. In IL/cosolvent systems with a high hydrogen-bonding basicity (>0.92) and a large amount of small ion cluster structures, cellulose favors the formation of a molecularly dispersed state. Cellulose dissolution capacity can be increased at lower processing temperature. Importantly, this IL/cosolvent system reduce the quantity of IL used in cellulose dissolution [87]. Other recent studies also indicate the advantages of the mixed IL/cosolvent system for cellulose dissolution that time, temperature, and viscosity of cellulose mixture in IL solutions are reduced by the inclusion of a cosolvent [88,89]. Hawkins et al. [88] studied the dissolution time of flax fibers in EmimAc as a function of temperature and DMSO concentration. The dissolution rate was maximized when using an equal amount of EmimAc to DMSO, whereby the dissolution was about three times faster than that in the pure IL. The independence of an activation energy (Ea), required for the dissolution of flax fibers in EmimAc, on DMSO concentration suggests that flax fibers dissolution is not primarily a viscosity-driven process. The result is related to the work of Chen et al. [90]; the dissolution is not only governed by viscosity, but also by solvent power.
Due to its low cost, low toxicity, and low viscosity, DMSO is a suitable candidate for a cosolvent. A series of IL/DMSO systems have been studied as shown in Table 1. The efficiency of IL/DMSO mixtures in cellulose dissolution and derivatization depends on the structures of ILs [91] and temperature [92]. With increasing temperature, the specific interaction between cellulose and the solvent increases [92]. Ferreira et al. [91] studied the effect of the presence of an ether linkage in the sidechain of imidazolium acetate (ImAc) using RMenImAc, R = 1-butyl or 2-methoxyethyl, n = 1 or 2 and the mixtures with DMSO. The mixture of BmimAc–DMSO was more efficient for the dissolution of MCC and acylation of cellulose because the mixture was less viscous, more basic, and formed stronger hydrogen bonds with cellobiose. Using the corresponding ILs with C2–CH3 instead of C2–H (1-butyl-2,3-dimethylimidazolium acetate and 1-(2-methoxyethyl)-2,3-dimethylimidazolium acetate) increased the concentration of dissolved cellulose [91]. Increasing the amount of DMSO does not only improve cellulose solubilization, but also leads to the formation of a more pronounced macroporous structure of the hydrogel [93].
In designing a new IL solvent system for cellulose dissolution, reversible/switchable reactions with CO2, a well-recognized greenhouse gas, have been extended [100,101,102]. Xie et al. [100] reported that the CO2-based reversible IL/DMSO solvent could dissolve MCC up to 10% w/w under mild conditions at 60 °C and a CO2 pressure of 2.0 MPa for 2 h, when 1,1,3,3-tetramethyl guanidine (TMG) was used as a base, in conjunction with methanol, ethanol, propanol, butanol, and ethylene glycol (Figure 2A). Yang et al. [103] also reported that DBU/DMSO could promote the dissolution of cellulose pulp at the relative low temperature of 50 °C in the presence of CO2. This is because the addition of CO2 causes the reversibility or switch-ability of this solvent system by changing the polarity from nonpolar to polar and reversing to its initial nonpolar state upon the release of CO2 [104,105]. Cellulose dissolves in this solvent system by nonderivative and derivative approaches (Figure 2B). In the nonderivative approach, simple alcohols such as methanol, hexanol, or ethylene glycol, in the presence of strong organic bases such as DBU, react with CO2 leading to the formation a DMSO–carbonate species solvent that can solubilize cellulose [100]. In the derivative approach, the superbase activates cellulose first and, thereafter, reacts with CO2 to form a DMSO-soluble cellulose carbonate intermediate [106].
Since the reversible reaction between cellulose, organic base, and CO2 is exothermic, the resulting high temperature is not suitable for cellulose dissolution. Therefore, the temperature used in most studies is in the range of 30–60 °C. At 30 °C, increasing the pressure of CO2 above 20 bar (2 MPa) increases in the carbonate formation and cellulose solubilization kinetics in 10 min, whereas at 40 bar (4 MPa), the carbonate formation is reached within 5 min. On the other hand, for CO2 pressures below 2 MPa, the solubilization of 3% w/w MCC is slower but occurs within 15 min [107]. At 30 °C and CO2 pressures of 0.2 MPa, the solubility of MCC in DMSO/DBU at weight ratio of DBU of 0.1 can reach 9% w/w. However, increasing the weight ratio of DBU to 0.5 slightly inhibits the MCC solubility. This indicates that the DBU amount influences MCC dissolution behavior [108]. At the higher temperature range of 40–60 °C, the dissolution of cellulose derived from a cassava pulp waste in DBU and ethylene glycol, which produces a switchable polarity solvent containing the cation of DBU (DBUH+) and the anion of ethylenedicarbonate (O2COCH2CH2OCO22−) in the mixture with CO2, can be achieved in less than 30 min under CO2 pressure of more than 5.0 MPa [109]. Under 1 atm (0.1 MPa) in an open system, the obtained cellulose is stable and suitable for the preparation of cellulose hydrogel fiber by wet spinning [110]. Besides TMG and DBU, clear cellulose solutions were obtained in 1,5,7-Triazabicyclo [4.4.0] dec-5-ene (TBD) under 1 MPa CO2 at 50 °C for 3 h [111], and 2-tert-butyl-1,1,3,3-tetramethylguanidine (BTMG) under 1 atm (0.1 MPa) CO2 at room temperature for 5 min [112], but cellulose was not dissolved in triethyl amine (TEA) [111]. Instead of DMSO, other organic solvents with low Henry’s constants, such as propylene carbonate and sulfolane, are used as CO2 absorbents to increase CO2 absorption capacity in the solvent, so that the cellulose dissolution is enhanced. With the addition of propylene carbonate, the paper cellulose dissolution can be increased from 1.0 to 5.0% w/w, while those for microcrystalline cellulose and corncob cellulose are increased from 5.0 to 10.0 and 7.0% w/w, respectively, at 60 °C and a CO2 pressure of 0.5 MPa. The chemical and crystalline structures of the regenerated cellulose is not affected by the addition of these CO2 absorbents [113].

3. Regeneration of Cellulose in Antisolvents

The reorganization of the hydrogen bond network during the regeneration is a crucial step to obtaining the desired hydrogel properties. Cellulose regeneration occurs through a physical sol–gel process by using polar protic antisolvents, including ethanol, methanol, and water, that donate and accept hydrogen bonds to disrupt the favorable cellulose–IL interactions. Concomitantly, strong cellulose–cellulose interactions are promoted. The anion of the ILs affects the coagulation kinetics in the cellulose regeneration process. For example, in the case of ILs with Ac− anion, the addition of water as an antisolvent triggers the hydrogen bond destruction between cellulose and Ac− anion, and simultaneously induces the formation of hydrogen bonds between cellulose and cellulose, while Ac− forms a hydrogen bond with water. After the cellulose chains aggregate, the gelation is then induced, resulting in the precipitation of cellulose from IL/protic solvent mixtures [114,115]. This gelation occurs without a chemical crosslinker, because of the high entanglement density of cellulose [116]. The gelation of cellulose to hydrogels can be performed in different shapes (Figure 3).
While the mechanism of cellulose dissolution has been systematically investigated [117], cellulose regeneration phenomenon is inherently difficult to study. Zhou et al. [118] investigated the effect of different antisolvents on the regeneration from MCC/BmimCl solution using ATR–FTIR. The H–bond interaction between the chlorine ion and antisolvents is the driving force for cellulose regeneration. The crystallinity of the regenerated cellulose related to the H–bond acidity and basicity of the antisolvents [119,120]. For the phase separation of IL and cellulose in regeneration using water as the antisolvent from a MCC/EmimAc solution, a two-step model, involving the destruction and reconstruction of H–bonds based on the time-dependent FTIR spectroscopy, was proposed by Liu et al. [121]. In the first step, skeletons of the cellulose gel are built after the double diffusion of free IL and water, inducing structural change in cellulose regeneration. In the second step, the interaction between cellulose and IL is destroyed as water diffuses. The H–bond between the released anion of EmimAc and water is constructed, and the H–bond network among the cellulose hydroxyl groups is then reconstructed [121]. Relying on the double diffusion of antisolvents, a surfactant-induced and diffusion-driven method can be used to fabricate hierarchical porous cellulose hydrogel with permeability [122]. The mechanism of cellulose regeneration is also investigated by theoretical calculation and molecular dynamics (MD) simulation [123,124,125]. Fu et al. [126] studied the mechanism of cellulose regeneration from theoretical calculation by analyzing the structures and hydrogen bonds of BmimAc-nH2O and cellobiose-ILs-nH2O (n = 0–6) clusters. When numbers of water molecules (n) were ≥5 in the BmimAc-nH2O clusters, the solvent-separated ion pairs played a dominant role in the system. The cation and anion interaction could be separated by water in reducing cellulose dissolution by ILs, when numbers of water molecules were increased. Furthermore, a more stable structure with high hydration in an aqueous solution of BmimAc-nH2O and cellobiose-ILs (n = 0–6) clusters was found. In the presence of water molecules in the system, the hydrogen bonds among H2O, the hydroxyl group of cellulose, and the oxygen of acetate were formed. Therefore, cellulose regeneration was promoted by the decrease of the interactions between cellulose and ILs [126]. After the dissolution and regeneration process, the transition from cellulose I to II occurs [127,128]. In the CO2/DBU/DMSO solvation system, cellulose that is regenerated through integrated thermally induced CO2 release is cellulose IVI crystalline structure [102].
Figure 3. Ionic liquid-based preparation processes of cellulose hydrogels in the form of bulk (A) (Reprinted with permission from ref. [95] Copyright © 2023, Elsevier, B.V.), fiber (B) (Reprinted with permission from ref. [129] Copyright © 2017 Wiley Periodicals, Inc.), membrane (C) (Reprinted with permission from ref. [130] Copyright © 2017 Wiley Periodicals, Inc.), and nonwoven mesh (D) (Reprinted with permission from ref. [131] Copyright © 2019 Wiley Periodicals, Inc.).
Figure 3. Ionic liquid-based preparation processes of cellulose hydrogels in the form of bulk (A) (Reprinted with permission from ref. [95] Copyright © 2023, Elsevier, B.V.), fiber (B) (Reprinted with permission from ref. [129] Copyright © 2017 Wiley Periodicals, Inc.), membrane (C) (Reprinted with permission from ref. [130] Copyright © 2017 Wiley Periodicals, Inc.), and nonwoven mesh (D) (Reprinted with permission from ref. [131] Copyright © 2019 Wiley Periodicals, Inc.).
Gels 09 00546 g003
The characteristics of the final hydrogel are related to the regeneration technique and composition of the regeneration bath [132]. The density of the regenerated cellulose hydrogel depends on the type of antisolvent that determines the mixing rate between the IL and the antisolvent during the regeneration process [104] (Figure 4A). The reduced regeneration rate of cellulose chains by the low mixing rate results in an ordered structure and a transparent cellulose hydrogel. Hence, mechanical characteristics are also affected [133,134]. Therefore, finding a suitable antisolvent is a key parameter. Water is mostly used due to the most environmentally friendly and inexpensive antisolvent. The regenerated cellulose using water had a higher crystallinity and better thermal stability than that using ethanol [135]. Due to the hydrotropic nature of the more bulky cations of IL, cellulose regeneration may require more hydrophobic mixtures (e.g., water–alcohols) for a more efficient regeneration of pure cellulose [136]. The reconstitution of cellulose molecules with various antisolvents affecting the multiscale structures and properties of cellulose has been studied. Gao et al. [137] employed different coagulation baths of ethanol, methanol, NaOH, and H2SO4 aqueous solutions to prepared cellulose hydrogel from TMGH OOCOCH2CH2OCOO/DMSO (the mole fraction of IL in the mixed solvent is 0.2). The hydrogel regenerated from alcohol had a more amorphous crystal structure, a smoother surface morphology, and higher transparency properties than those regenerated from 5% w/w NaOH and 5% w/w H2SO4 aqueous solutions (Figure 4B). The hydrogel from ethanol exhibited good thermostability and mechanical properties (tensile strength and strain of 56.2 MPa and 20.4%, respectively, and a Young’s modulus of 901 MPa). Water vapor permeability of 8.7 × 10−3 g μm/m2 day kPa and oxygen permeability of about 4.0 cm3 μm/m2 day atm were obtained. The hydrogels regenerated from aqueous solutions of alkali and the acid-coagulating baths shielded oxygen in order to permeate [137]. In the case of regenerating cellulose from an EmimDEP solution, the regenerated cellulose fiber hydrogel shows differences in its internal structure and mechanical properties due to the different diffusion rates between the IL solution and coagulation bath containing 10% w/w of IL and the antisolvent (Figure 4C). In comparison with the pure water, the absolute ethanol slightly reduces the crystallinity and orientation, whereas the elongation slightly increases [138]. The coagulating bath containing IL has consequences for an up-scaled continuous process, such as the possibility to regenerate cellulose in an IL-concentrated bath. This minimizes solvent recycling costs using, for example, the thermal separation of IL and water. A closed-loop process operation, performing single filament spinning in IL/H2O mixtures, was simulated and studied by Guizani et al. [139]. Spinnability and preserved mechanical properties of the hydrogel fibers of cellulose dissolved in DBNHAc were achieved in a coagulation bath containing up to 30% w/w IL, although the fibrillar structure was less pronounced in comparison to a pure water bath. At 45 % of IL in the bath, the spinnability was possible but resulted in lower quality fibers because the mechanical properties of the hydrogel fibers started to deteriorate [139].
A switchable coagulant is also used to regenerate cellulose from the IL solution, and it is easier to remove from the IL [140,141]. This approach can minimize the direct evaporation process in the recovery of hydrophilic ILs from an aqueous solution, which is energy consuming. The CO2 is used as a coagulation solvent for the separation of cellulose and IL [140]. Since CO2 is miscible in the IL, such as BmimAc, but immiscible with cellulose, after depressurization and removal of CO2, the cellulose precipitated as a transparent gel [142]. The yield of the regenerated cellulose (60%), prepared from 10% w/w MCC in BmimAc, increases after 3 h, under 6.6 MPa compressed CO2, and at 25 °C [143]. However, more cellulose can be regenerated from the IL/cosolvent systems than this neat IL system [108,143]. The degree of polymerization (DP) value of the regenerated cellulose under the neat system decreases from 282 of MCC to 202, depending on the reaction time of the compressed CO2, as well as the pressure and temperature. The DP is higher than those regenerated by a cosolvent system (DP = 157–175) and ethanol antisolvent (DP = 163) [143], but slightly lower than those regenerated using water (DP = 210) [144]. Upon the reaction of amine with CO2, the non-ionic amine is changed to ionic carbamate, which can be employed in the formation of switchable solvents [145]. These reactions of amines with CO2 in the field of CO2 capture and storage have been studied recently [146,147]. The CO2 capture occurs by the formation of an insoluble monoethanolamine–carbamate, 2-hydroxyethylammonium (2-hydroxyethyl) carbamate. By heating and/or reducing the pressure, the captured CO2 is released from the carbamate [141]. This approach to coagulate cellulose in an IL solution, by the formation of a carbamate salt having strong hydrogen bond donors and acceptors, can disrupt the hydrogen bonding between cellulose and the IL anion, resulting in coagulation. The reversible coagulant approach also facilitates the recovery of the IL to avoid using the commonly used liquid antisolvent, thus saving on the cost and energy necessary for liquid evaporation [141].

4. Applications

4.1. Electrochemical and Electronic Materials

Active research on electrochemical energy storage has been stimulated due to the increasing need for energy storage devices. Depending on the balance between energy and power density, electrochemical energy storage is categorized into batteries and supercapacitors. These devices possess a long cycling life and instantaneously store and release energy, allowing the potential for high efficiency energy storage devices [148,149]. The more environmentally friendly production of energy storage device technologies is a matter of interest in the use of biopolymer-based hydrogel electrolytes [150]. For instance, cellulose-based supercapacitors have gained attention due to their mechanical strength, environmental friendliness, and light weight [151,152,153]. Low-cost cellulose also decreases the total electrode cost, which is beneficial for producing a competitive electrode material for electricity storage on a large scale [154].
The addition of graphene allows the cellulose hydrogel to improve the thermal and electrical conductivity suitable for energy storage applications, whereas the cellulose hydrogel provides graphene at a low-cost template for large-scale applications. Graphene, a one-atom-thick two-dimensional sp2-bonded carbon material, has high thermal conductivity (~2000–4000 W/m/K) and mechanical properties. It is a promising filler to improve the thermally conductive and mechanical properties of polymer matrices, which has been frequently reported in recent years [155,156,157,158,159]. Xu et al. [160] reported the use of wood pulp cellulose as the raw material to prepare cellulose/graphene composite hydrogels by dissolving in BmimCl as the medium to dissolve cellulose and disperse graphene, and then coagulating the hydrogel in water. Enhanced mechanical strength and thermal stability was observed in comparison with pure cellulose hydrogels. Their appearances and mechanical properties are shown in Figure 5A–D. With 0.5% w/w of graphite oxide in cellulose, about four times higher compressive Young’s modulus was obtained due to an organized alignment of cellulose chains on graphene sheets and an interfacial adhesion between graphene and cellulose chains. However, at the higher doping, more graphene led to enhanced mechanical properties of the hydrogel because the uniformly dispersed graphene/cellulose had a typical 3D network structure. This made the cellulose/graphene composite hydrogels become more compact. It is also reported that the mechanical strength of the composite hydrogel via π–π stacking interactions is much higher than that of the pure cellulose hydrogel via hydrogen bonds [160]. The graphene is functionalized with IL by cation–π or π–π and electrostatic interactions [161]. The technique containing graphene and IL can improve the performance of the corresponding composite hydrogel in the wider applications of supercapacitors [148], electrocatalysts [162], and sensors [163].
The aqueous phase in hydrogels, and the liquid-like transport properties of the hydrogels, enable their fast diffusion for conductivity [166]. When the hydrogels are treated with graphene, or other conductive materials such as polypyrene, they can transfer electrons [164]. Liang et al. [164] dissolved MCC in BmimCl solution and crosslinked by N,N′-methylene bisacrylamide to obtain cellulose hydrogel. The hydrogel was then immersed in a mixed aqueous solution of ferric trichloride and sodium p-toluenesulfonate, and in the pyrrole solution, to form a conductive hydrogel. The mechanical strength was also improved in comparison to the typical cellulose hydrogel (Figure 5E,F). Such ionic conductive hydrogels, prepared by immersion/swelling in electrolyte solutions, have also attracted an increasing interest in terms of batteries [165,167] and solar cells [168]. The cellulose hydrogel has properties superior to some commercial products, such as transparency, thermal stability, and ionic conductivity (Figure 5G–J).
In modern electronics, for instance, in human wearable sensors and sensors in soft robotics, in the development of soft electrolytes with high stretchability, compressibility, and ion transport are in demand [169,170]. As triboelectric nanogenerators, mechanical energies in human activities can be turned into electricity because of the functions of stretchability, transparency, ionic conductivity, and even self-healing behaviors [171,172]. This also make them suitable for wearable electronics [173]. Various polymer systems, such as PAA/PDMAPS/IL hydrogel as triboelectric nanogenerator [174], HPC/PVC/NaCl hydrogel as an artificial nerve allowing the passing of stable electrical signals [175], PVA/PAM/NaCl hydrogel as a high-sensitivity strain sensor [176], cellulose nanofibers/PVA/AlCl3 hydrogel as artificial electronic skins [177], and cellulose nanofibers/PAM–AA/LiCl hydrogel as a muscle-like sensor for human motion monitoring [178], have been reported. For these applications, the use of hydrogels has two drawbacks, which are water retention ability and narrow temperature range [179,180]. As water evaporates from a hydrogel, the ionic conductivity of the hydrogel decreases due to the evaporation of water at high temperatures [181]. At subzero temperatures, the hydrogels become rigid and restrict the ionic transport [182]. Hence, hydrogels are not suitable for long-term applications in dry, subzero or hot environments. Ionogels containing IL can be used as electrodes in electronic devices at high temperatures or in dry environments due to their nonflammability and negligible vapor pressure [174]. The ionogels can be prepared from various types of cellulose, e.g., micro and non-crystalline, bacterial cellulose, and cellulose derivatives [183]. Cellulose ionogels are a promising solid electrolyte due to their high conductivity (0.01–8 S/m), wide electrochemical potential window, and thermal stability up to 200 °C [167,184,185]. Since the properties of the ionogel are related to the properties of the IL, imidazolium cations such as Amim+ and Bmim+, having high electrical conductivity and the ability to dissolve cellulose and to form gel for electrochemical applications, are the focus of most studies [185,186,187,188]. Zhao et al. [188] developed a topology-tunable dynamic hydrogel containing cellulose, BmimCl, and water. The prepared gel exhibited tunable properties of mechanical strength, ionic conductivity, viscoelasticity, and self-healing. With 32% w/w of water, the microstructure switched to a dense and compact network, giving the gel good stretchability, robust toughness, a high ionic conductivity, transparency, and biocompatibility. This demonstrated that this hydrogel had a potential for use in electronic skins and intelligent devices [188].

4.2. Membranes for Water Treatments

In membrane technology for water and wastewater treatments, cellulose-based membranes are well known [189]. Cellulose composite nanofiltration membranes show high rejection performance for the separation of different dyes. Anokhina et al. [190] studied the filtration of two anionic dyes, Orange II (350 g/mol) and Remazol Brilliant Blue R (626 g/mol) by applying a cellulose membrane prepared using IL as a membrane casting solution. The high concentration of cellulose in IL led to the formation of a denser cellulose-based selective layer in the membrane. In addition, the separation selectivity indicated by rejection factors of the two dyes (R350 and R626, respectively) varied during filtration through the membranes prepared from an 8% w/w cellulose solution in varying EmimAc and DMSO ratios. At 90% w/w EmimAc in DMSO, R350 and R626 were 40% and 65%, respectively. The rejection factors of the two dyes increased with an increase in cellulose concentration in the casting solution due to the formation of a denser porous structure of the cellulose layer. At 16% w/w cellulose in 50% w/w EmimAc in DMSO, R350 and R626 were 61 and 82%, respectively [190], whereas, in the system of 14% w/w cellulose in pure EmimAc, R350 and R626 were 24 and 31%, respectively [96]. In the diluted IL system (14% w/w cellulose in 50% w/w EmimAc in DMSO), replacement of water with alcohols as antisolvents in regeneration step can alter the membrane permeability and rejection of Orange II and Remazol Brilliant Blue R [96]. Ilyin et al. [96] reported that methanol led to a fivefold increase in membrane permeability of DMF as compared to that of the membrane regenerated in water (PDMF = 0.23 kg/m2/h/atm), but decreased its rejection to 35–47% due to an increase in the porosity of the membrane. Ethanol yielded a membrane with lower permeability, but better rejection (56–71%). As compared to methanol, ethanol had a lower diffusion rate, leading to the contraction of the forming film, reducing its pore size. Isopropanol provided an even lower diffusion rate and delayed structural changes of membrane during phase separation and eliminated the effects of network contraction. However, the film had better porosity and permeability compared to that regenerated by ethanol [96]. The dense layer of the membrane and the size exclusion of the dye are the two main factors of dye rejection [191]. Solute diffusion is the transport mechanism through the cellulose-based membranes. In addition, the surface charge and hydrophilicity are the dominant factors of the antifouling capability of the membranes [191]. Esfahani et al. [191] studied 10% w/w bamboo cellulose-based membranes and prepared them using BmimCl and coagulated them in water, showing the high rejection (>80%) of crystal violet and 60–70% for methylene blue and methylene orange at a dye concentration of 10 ppm in water. This is because of the dense layer and hydrophilicity of the membrane (Figure 6A–C). The membrane prepared from 3% bamboo cellulose showed low rejection (35%) of the three dyes, but high water flux of 600 L/m2/h, respectively. The dye separation efficiency of the cellulose membranes also depends on pH, temperature, ionic charge and concentration, and reuse cycle (Figure 6D). Comparing ~2–3% w/w lemon peel cellulose [192] and pineapple peel cellulose [193] prepared in BmimCl and coagulated in water, the adsorption capacities of methylene blue are 25 and 57 mg/g from the initial dye concentration of 50 and 100 ppm, respectively.

4.3. Bioink for 3D Printing

Three-dimensional (3D) printing, known as rapid prototyping technology (RP) [194], additive manufacturing technology (AM) [195], or solid freeform fabrication (SFF) [196], is the technique for constructing solid objects by depositing several layers of material in sequence. Due to the capability of fabricating cost-effective and customizable objects, 3D printing has been recognized as a promising technology [195]. With development in science and technology, it is sufficiently mature to be applied to lower cost materials by open-source software [197]. To date, this technology has been applied in the medical [198], pharmaceutical [199], education [200], clothing [201], and food [202] industries. These 3D-printed objects have proven to have effective applications in cell culture and drug delivery [203,204,205]. One of the primary reasons for using biopolymers in 3D printing is the dispersion stability and flow behavior of inks, as well as structural and thermal properties of the printed objects [206,207]. The desired mechanical properties of 3D-printed constructions are conferred by inducing self-assembled structures and strengthening the interlayer bonding [206]. There has been an increase in the number of studies using cellulose or cellulose combined with other biopolymers in 3D printing [208,209,210,211]. The alignment of CNC and other anisotropic particles during 3D printing mimics the multiscale architectures of biological materials such as wood, plant stems, bone, and mollusk shells [212]. Such particle alignment by shear stresses (shear-inducing flow) is intrinsically present in extrusion-based 3D printing processes [210,213]. However, the concentration of CNC that can be added to the inks remains very limited [214], which is appropriate for the printing of cellulose micropatterns on flat substrates [215] or the printing of liquid resins into CNC-laden complex geometries [216,217]. It is not currently feasible to print 3D objects at high cellulose content without nozzle clogging [218]. The content of cellulose nanofiber that can be used in the ink is even lower than that of CNC, due to the high amount of surface hydroxyl groups and entanglements via hydrogen bonding between the fibers leading to an increased yield stress and the high pressure needed for the extrusion of the ink [219,220]. The ILs allow higher concentrations of cellulose in inks compared to water-based systems [221].
Markstedt et al. [221] showed the possibility of printing membrane or spatially tailored 3D hydrogels by using cellulose dissolved in EmimAc, followed by coagulation in water. The hollow cylindrical structures were created by complex patterns of 2D structures and multilayered printing constructs. High cellulose concentrations and molecular weight were shear-thinning and favorable for printing. A solution of 4% w/w cellulose, dispensed at the pressure for the extrusion of 6 bars with a 12.7 mm long needle, having an inner diameter of 0.41 mm at a flow rate of 10 μL/min, enabled the prints to obtain an acceptable resolution. The rates of coagulation had to be considered carefully in the material construction and selection of printing parameters. If the coagulation occurred too quickly it would result in poor adhesion between printed layers and collapse, whilst coagulation that was too slow led to height limitations [221]. By using 4% w/w bleached kraft pulp cellulose dissolved in EmimDMP, obtaining a viscosity of 740 Pa·s and a yield stress of 2.5 kPa, Hopson et al. [222] reported that the printed ionogels and the hydrogels had a good print resolution and fidelity due to the low volatility of IL. The results were comparable to the hydrogel prepared by the commercial ink (Cellink XploreTM) [222]. Because of their increased viscosity and yield stress, caused by the high cellulose content and viscose IL in inks, cosolvents, such as 1-butanol and DMSO, can modulate the printability [223]. For instance, DMSO in the range of 41–47% w/w with EmimAc or BmimAc yields ink within the printable range for inkjet printing of 1.0–4.8% w/w cellulose (printability parameter <10 at 55 °C). By using water as an antisolvent, the regenerated ink hydrogel on a PET substrate exhibits structural distortion, whereas there is no such issues on glass [223]. Using 30–40% w/w water as a cosolvent with bacterial cellulose dissolved in choline acrylate (a polymerizable IL) was studied by Fedotovam et al. [224,225]. The addition of water exhibited the printability and stability of inks. The solvent viscosity and yield stress were decreased from 2230 Pa (anhydrous system) to 97–160 Pa. The decreased viscosity of the dispersion caused by the water–acrylate complexes led to the destruction of the network of ionic bonds between IL cations and anions. However, cellulose in this binary solvent was nearly a suspension of the cellulose nanofibers, not a solution. The MD simulations provided information that water changed the structural organization of the hydration of acrylic ions at 30–40% of water. The IL/water ratio for 3D printability related to the ion–dipole interactions between IL components and between cellulose–IL in the ink, which decreased with the addition of water. This IL/water ratio was enough for the stabilization of the dispersion, while the morphology and crystalline structure of the cellulose nanofibers were preserved. When further increasing the water content (50% w/w), the yield stress increased due to the amorphization, swelling, and agglomeration of cellulose nanofibers [224,225].

5. Conclusions and Outlook

This review summarizes the recent studies that examined the preparation of cellulose hydrogels and related products by dissolving cellulose in ILs and then regenerating it in an antisolvent such as water or alcohol. Overall, these studies indicate that the properties of the hydrogels can be tuned through varying the concentrations of cellulose in the solution, the type of ILs/cosolvents, and the compositions of antisolvents in the regeneration process. Imidazolium ILs, paired with acetate and chlorine, have been widely used in cellulose dissolution. However, the development of new ILs for higher cellulose dissolution are preferable, where dissolving at room temperature, the lower viscosity of cellulose solution, the low decomposition/depolymerization of cellulose, and low cost, have been challenges. The cosolvents, such as polar aprotic organic solvents, have been mixed with the recently developed imidazolium ILs to achieve this goal. The CO2 switchable solvent system for cellulose dissolution has been explored as a new platform for cellulose processing. The dissolution of cellulose is efficient, even though it is performed at room temperature. Importantly, low IL content has been shown to be effective. This can address the remaining problem of using IL in terms of cost, since ILs are more expensive than other solvents. However, cellulose solubilization conditions in this CO2 solvent need to be optimized since the cellulose dissolution depends on cellulose source, type/amount of the superbases, pressure of CO2, temperature, and time. In addition, the preparation of cellulose hydrogels in their various forms, such as membranes and fibers regenerated from cellulose dissolved in these CO2 solvents, has not systematically and widely studied. Therefore, it is of great interest to study the conditions of the regeneration process, such as composition, temperature of the coagulation bath, and time to reach the gel state.
The recycling and purification of ILs are essential for an economical and sustainable cellulose hydrogel manufacturing process. Water or other antisolvents added to regenerate cellulose need to be efficiently separated from the ILs. Several methods have been reported, from simple evaporation to more complex phase separation processes, but most studies in this field are still preliminary. In-depth knowledge is essential for improving the recyclability and long-term chemical stability of the ILs. The separation process should be efficient, easy to operate under mild conditions, and energy-saving, which are beneficial for the scaling-up of the manufacturing process. Due to the capability of recycling and purification, the cellulose-based hydrogels, prepared by using the IL solvent, will truly follow the 17 Sustainable Development Goals (SDGs) outlined in the resolution of the United Nations that highlight the need for green and sustainable solvents.

Funding

This review received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Acknowledgments

A full fee waiver of the article processing charge was granted for this article to be published in Gels. This work was supported by Nagaoka University of Technology, Japan.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Xia, Z.; Li, J.; Zhang, J.; Zhang, X.; Zheng, X.; Zhang, J. Processing and valorization of cellulose, lignin and lignocellulose using ionic liquids. J. Bioresour. Bioprod. 2020, 5, 79–95. [Google Scholar] [CrossRef]
  2. Yeboah, W.O.; Kwofie, E.M.; Wang, D. Circular bioeconomy potential of rice husk as a bioplastic resource: Techno-environmental assessment. Bioresour. Technol. Rep. 2022, 20, 101248. [Google Scholar] [CrossRef]
  3. Ahrens, T.; Drescher-Hartung, S.; Anne, O. Sustainability of future bioenergy production. Waste Manag. 2017, 67, 1–2. [Google Scholar] [CrossRef] [PubMed]
  4. Gan, W.; Wang, Y.; Xiao, S.; Gao, R.; Shang, Y.; Xie, Y.; Liu, J.; Li, J. Magnetically Driven 3D Cellulose Film for Improved Energy Efficiency in Solar Evaporation. ACS Appl. Mater. Interfaces 2021, 13, 7756–7765. [Google Scholar] [CrossRef] [PubMed]
  5. Nie, S.; Hao, N.; Zhang, K.; Xing, C.; Wang, S. Cellulose nanofibrils-based thermally conductive composites for flexible electronics: A mini review. Cellulose 2020, 27, 4173–4187. [Google Scholar] [CrossRef]
  6. Klemm, D.; Cranston, E.D.; Fischer, D.; Gama, M.; Kedzior, S.A.; Kralisch, D.; Kramer, F.; Kondo, T.; Lindström, T.; Nietzsche, S.; et al. Nanocellulose as a natural source for groundbreaking applications in materials science: Today’s state. Mater. Today 2018, 21, 720–748. [Google Scholar] [CrossRef] [Green Version]
  7. Li, X.; Xu, R.; Yang, J.; Nie, S.; Liu, D.; Liu, Y.; Si, C. Production of 5-hydroxymethylfurfural and levulinic acid from lignocellulosic biomass and catalytic upgradation. Ind. Crops Prod. 2019, 130, 184–197. [Google Scholar] [CrossRef]
  8. Taokaew, S.; Nakson, N.; Zhang, X.; Kongklieng, P.; Kobayashi, T. Biotransformation of okara extracted protein to nanocellulose and chitin by Gluconacetobacter xylinus and Bacillus pumilus. Bioresour. Technol. Rep. 2022, 17, 100904. [Google Scholar] [CrossRef]
  9. Taokaew, S.; Nakson, N.; Thienchaimongkol, J.; Kobayashi, T. Enhanced production of fibrous bacterial cellulose in Gluconacetobacter xylinus culture medium containing modified protein of okara waste. J. Biosci. Bioeng. 2023, 135, 71–78. [Google Scholar] [CrossRef]
  10. Long, L.-Y.; Weng, Y.-X.; Wang, Y.-Z. Cellulose Aerogels: Synthesis, Applications, and Prospects. Polymers 2018, 10, 623. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, L.; Huang, S.; Wang, Y. Recycling of Waste Cotton Textile Containing Elastane Fibers through Dissolution and Regeneration. Membranes 2022, 12, 355. [Google Scholar] [CrossRef]
  12. Zhong, X.; Li, R.; Wang, Z.; Wang, Y.; Wang, W.; Yu, D. Highly flexible, transparent film prepared by upcycle of wasted jute fabrics with functional properties. Process Saf. Environ. Prot. 2021, 146, 718–725. [Google Scholar] [CrossRef]
  13. Ma, Y.; Zeng, B.; Wang, X.; Byrne, N. Circular Textiles: Closed Loop Fiber to Fiber Wet Spun Process for Recycling Cotton from Denim. ACS Sustain. Chem. Eng. 2019, 7, 11937–11943. [Google Scholar] [CrossRef]
  14. Selvakumaran, N.; Lazim, A. Superabsorbent hydrogel from extracted oil palm frond waste cellulose using microwave irradiation for cadmium ion removal from aqueous solution. Chem. Chem. Technol. 2019, 13, 518–525. [Google Scholar] [CrossRef]
  15. Miao, X.; Lin, J.; Bian, F. Utilization of discarded crop straw to produce cellulose nanofibrils and their assemblies. J. Bioresour. Bioprod. 2020, 5, 26–36. [Google Scholar] [CrossRef]
  16. Su, R.; Wang, F.; Ding, J.; Li, Q.; Zhou, W.; Liu, Y.; Gao, B.; Yue, Q. Magnetic hydrogel derived from wheat straw cellulose/feather protein in ionic liquids as copper nanoparticles carrier for catalytic reduction. Carbohydr. Polym. 2019, 220, 202–210. [Google Scholar] [CrossRef] [PubMed]
  17. Pang, J.; Wu, M.; Zhang, Q.; Tan, X.; Xu, F.; Zhang, X.; Sun, R. Comparison of physical properties of regenerated cellulose films fabricated with different cellulose feedstocks in ionic liquid. Carbohydr. Polym. 2015, 121, 71–78. [Google Scholar] [CrossRef]
  18. Wang, H.; Qian, H.; Luo, Z.; Zhang, K.; Shen, X.; Zhang, Y.; Zhang, M.; Liebner, F. ZCIS/ZnS QDs fluorescent aerogels with tunable emission prepared from porous 3D nanofibrillar bacterial cellulose. Carbohydr. Polym. 2019, 224, 115173. [Google Scholar] [CrossRef]
  19. Afsahi, G.; Dimic-Misic, K.; Gane, P.; Budtova, T.; Maloney, T.; Vuorinen, T. The investigation of rheological and strength properties of NFC hydrogels and aerogels from hardwood pulp by short catalytic bleaching (Hcat). Cellulose 2018, 25, 1637–1655. [Google Scholar] [CrossRef]
  20. Nechyporchuk, O.; Belgacem, M.N.; Bras, J. Production of cellulose nanofibrils: A review of recent advances. Ind. Crops Prod. 2016, 93, 2–25. [Google Scholar] [CrossRef]
  21. Siró, I.; Plackett, D. Microfibrillated cellulose and new nanocomposite materials: A review. Cellulose 2010, 17, 459–494. [Google Scholar] [CrossRef]
  22. Dutta, S.D.; Patel, D.K.; Lim, K.-T. Functional cellulose-based hydrogels as extracellular matrices for tissue engineering. J. Biol. Eng. 2019, 13, 55. [Google Scholar] [CrossRef]
  23. Wu, C.; McClements, D.J.; Li, L.; He, M.; Li, Y.; Teng, F. Fabrication of composite hydrogels by assembly of okara cellulose nanofibers and gum Arabic in ionic liquids: Structure and properties. J. Mol. Liq. 2022, 349, 118132. [Google Scholar] [CrossRef]
  24. Zainal, S.H.; Mohd, N.H.; Suhaili, N.; Anuar, F.H.; Lazim, A.M.; Othaman, R. Preparation of cellulose-based hydrogel: A review. J. Mater. Res. Technol. 2021, 10, 935–952. [Google Scholar] [CrossRef]
  25. Tian, W.; Gao, X.; Zhang, J.; Yu, J.; Zhang, J. Cellulose nanosphere: Preparation and applications of the novel nanocellulose. Carbohydr. Polym. 2022, 277, 118863. [Google Scholar] [CrossRef] [PubMed]
  26. Song, M.-H.; Pham, T.P.T.; Yun, Y.-S. Ionic liquid-assisted cellulose coating of chitosan hydrogel beads and their application as drug carriers. Sci. Rep. 2020, 10, 13905. [Google Scholar] [CrossRef]
  27. Liu, Y.; Ahmed, S.; Sameen, D.E.; Wang, Y.; Lu, R.; Dai, J.; Li, S.; Qin, W. A review of cellulose and its derivatives in biopolymer-based for food packaging application. Trends Food Sci. Technol. 2021, 112, 532–546. [Google Scholar] [CrossRef]
  28. Kim, S.H.; Lee, C.M.; Kafle, K. Characterization of crystalline cellulose in biomass: Basic principles, applications, and limitations of XRD, NMR, IR, Raman, and SFG. Korean J. Chem. Eng. 2013, 30, 2127–2141. [Google Scholar] [CrossRef]
  29. Wohlert, M.; Benselfelt, T.; Wågberg, L.; Furó, I.; Berglund, L.A.; Wohlert, J. Cellulose and the role of hydrogen bonds: Not in charge of everything. Cellulose 2022, 29, 1–23. [Google Scholar] [CrossRef]
  30. Cao, Y.; Zhang, R.; Cheng, T.; Guo, J.; Xian, M.; Liu, H. Imidazolium-based ionic liquids for cellulose pretreatment: Recent progresses and future perspectives. Appl. Microbiol. Biotechnol. 2017, 101, 521–532. [Google Scholar] [CrossRef]
  31. Xu, H.; Zhang, D.; Wu, F.; Cao, R. Deep oxidative desulfurization of fuels based on [C4mimCl]CoCl2 ionic liquid oxone solutions at room temperature. Fuel 2017, 208, 508–513. [Google Scholar] [CrossRef]
  32. Han, Q.; Gao, X.; Zhang, H.; Chen, K.; Peng, L.; Jia, Q. Preparation and comparative assessment of regenerated cellulose films from corn (Zea mays) stalk pulp fines in DMAc/LiCl solution. Carbohydr. Polym. 2019, 218, 315–323. [Google Scholar] [CrossRef]
  33. Choe, D.; Kim, Y.M.; Nam, J.E.; Nam, K.; Shin, C.S.; Roh, Y.H. Synthesis of high-strength microcrystalline cellulose hydrogel by viscosity adjustment. Carbohydr. Polym. 2018, 180, 231–237. [Google Scholar] [CrossRef]
  34. From, M.; Larsson, P.T.; Andreasson, B.; Medronho, B.; Svanedal, I.; Edlund, H.; Norgren, M. Tuning the properties of regenerated cellulose: Effects of polarity and water solubility of the coagulation medium. Carbohydr. Polym. 2020, 236, 116068. [Google Scholar] [CrossRef]
  35. Qiu, C.; Zhu, K.; Zhou, X.; Luo, L.; Zeng, J.; Huang, R.; Lu, A.; Liu, X.; Chen, F.; Zhang, L.; et al. Influences of Coagulation Conditions on the Structure and Properties of Regenerated Cellulose Filaments via Wet-Spinning in LiOH/Urea Solvent. ACS Sustain. Chem. Eng. 2018, 6, 4056–4067. [Google Scholar] [CrossRef]
  36. Peng, J.; Li, Y.; Liu, X.; Ke, G.; Song, D.; Wu, S.; Xu, W.; Zhu, K. Cellulose film with air barrier and moisture-conducting character fabricated by NMMO. J. Mater. Sci. 2021, 56, 18313–18326. [Google Scholar] [CrossRef]
  37. Zhang, J.; Tominaga, K.; Yamagishi, N.; Gotoh, Y. Comparison of Regenerated Cellulose Fibers Spun from Ionic Liquid Solutions with Lyocell Fiber. J. Fiber Sci. Technol. 2020, 76, 257–266. [Google Scholar] [CrossRef]
  38. Gorbacheva, S.N.; Yadykova, A.Y.; Ilyin, S.O. A novel method for producing cellulose nanoparticles and their possible application as thickeners for biodegradable low-temperature greases. Cellulose 2021, 28, 10203–10219. [Google Scholar] [CrossRef]
  39. Zhao, J.; Deng, M.; Li, S.; Guan, Z.; Xia, Y.; Yang, J.; Lin, X. Room temperature preparation of cellulose nanocrystals with high yield via a new ZnCl2 solvent system. Carbohydr. Polym. 2022, 278, 118946. [Google Scholar] [CrossRef] [PubMed]
  40. Tong, Z.; Meng, J.; Liu, S.; Liu, Y.; Zeng, S.; Wang, L.; Xia, Q.; Yu, H. Room temperature dissolving cellulose with a metal salt hydrate-based deep eutectic solvent. Carbohydr. Polym. 2021, 272, 118473. [Google Scholar] [CrossRef]
  41. Shu, L.; Zhang, X.-F.; Wang, Z.; Yao, J. Structure reorganization of cellulose hydrogel by green solvent exchange for potential plastic replacement. Carbohydr. Polym. 2022, 275, 118695. [Google Scholar] [CrossRef]
  42. Zhang, X.-F.; Ma, X.; Hou, T.; Guo, K.; Yin, J.; Wang, Z.; Shu, L.; He, M.; Yao, J. Inorganic Salts Induce Thermally Reversible and Anti-Freezing Cellulose Hydrogels. Angew. Chem. Int. Ed. 2019, 58, 7366–7370. [Google Scholar] [CrossRef]
  43. Shen, C.; Hu, C.; Zhang, W.; Lin, X.; Qi, W.; Zhang, Z.; Gu, J. Acidified ZnCl2 molten salt hydrate systems as hydrolytic media for cellulose I and II nanocrystal production: From rods to spheres. Cellulose 2022, 29, 7629–7647. [Google Scholar] [CrossRef]
  44. Fischer, S.; Leipner, H.; Thümmler, K.; Brendler, E.; Peters, J. Inorganic Molten Salts as Solvents for Cellulose. Cellulose 2003, 10, 227–236. [Google Scholar] [CrossRef]
  45. Sun, L.; Han, J.; Wu, J.; Huang, W.; Li, Y.; Mao, Y.; Wang, L.; Wang, Y. Cellulose pretreatment with inorganic salt hydrate: Dissolution, regeneration, structure and morphology. Ind. Crops Prod. 2022, 180, 114722. [Google Scholar] [CrossRef]
  46. Huang, Z.; Liu, C.; Feng, X.; Wu, M.; Tang, Y.; Li, B. Effect of regeneration solvent on the characteristics of regenerated cellulose from lithium bromide trihydrate molten salt. Cellulose 2020, 27, 9243–9256. [Google Scholar] [CrossRef]
  47. Hattori, K.; Arai, A. Preparation and Hydrolysis of Water-Stable Amorphous Cellulose. ACS Sustain. Chem. Eng. 2016, 4, 1180–1186. [Google Scholar] [CrossRef]
  48. Zhou, Y.; Zhang, X.; Zhang, J.; Cheng, Y.; Wu, J.; Yu, J.; Zhang, J. Molecular weight characterization of cellulose using ionic liquids. Polym. Test. 2021, 93, 106985. [Google Scholar] [CrossRef]
  49. Wang, S.; Lu, A.; Zhang, L. Recent advances in regenerated cellulose materials. Prog. Polym. Sci. 2016, 53, 169–206. [Google Scholar] [CrossRef]
  50. Sen, S.; Martin, J.D.; Argyropoulos, D.S. Review of Cellulose Non-Derivatizing Solvent Interactions with Emphasis on Activity in Inorganic Molten Salt Hydrates. ACS Sustain. Chem. Eng. 2013, 1, 858–870. [Google Scholar] [CrossRef]
  51. Przypis, M.; Wawoczny, A.; Gillner, D. Biomass and Cellulose Dissolution—The Important Issue in Renewable Materials Treatment. Appl. Sci. 2023, 13, 1055. [Google Scholar] [CrossRef]
  52. Wang, Y.; He, H.; Wang, C.; Lu, Y.; Dong, K.; Huo, F.; Zhang, S. Insights into Ionic Liquids: From Z-Bonds to Quasi-Liquids. JACS Au 2022, 2, 543–561. [Google Scholar] [CrossRef] [PubMed]
  53. Hopson, C.; Rigual, V.; Domínguez, J.C.; Alonso, M.V.; Oliet, M.; Rodríguez, F. A new approach for the use of cellulose-rich solids from biorefinery in the formulation of gel-like materials. Ind. Crops Prod. 2022, 186, 115230. [Google Scholar] [CrossRef]
  54. Li, Y.; Liu, X.; Zhang, Y.; Jiang, K.; Wang, J.; Zhang, S. Why Only Ionic Liquids with Unsaturated Heterocyclic Cations Can Dissolve Cellulose: A Simulation Study. ACS Sustain. Chem. Eng. 2017, 5, 3417–3428. [Google Scholar] [CrossRef]
  55. Zhang, J.; Xu, L.; Yu, J.; Wu, J.; Zhang, X.; He, J.; Zhang, J. Understanding cellulose dissolution: Effect of the cation and anion structure of ionic liquids on the solubility of cellulose. Sci. China Chem. 2016, 59, 1421–1429. [Google Scholar] [CrossRef]
  56. Paiva, T.; Echeverria, C.; Godinho, M.H.; Almeida, P.L.; Corvo, M.C. On the influence of imidazolium ionic liquids on cellulose derived polymers. Eur. Polym. J. 2019, 114, 353–360. [Google Scholar] [CrossRef]
  57. Jadhav, S.; Ganvir, V.; Singh, M.K.; Shanmuganathan, K. Synthesis of N-oxyethylene substituted imidazolium-based zwitterions as a recyclable solvent for cellulose dissolution. Cellulose 2023, 30, 87–109. [Google Scholar] [CrossRef]
  58. Morita, T.; Okada, H.; Yamada, T.; Hidaka, R.; Ueki, T.; Niitsuma, K.; Kitazawa, Y.; Watanabe, M.; Nishikawa, K.; Higashi, K. A study combining magic-angle spinning NMR and small-angle X-ray scattering on the interaction in the mixture of poly(benzyl methacrylate) and ionic liquid 1-ethyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide. Phys. Chem. Chem. Phys. 2022, 24, 26575–26582. [Google Scholar] [CrossRef]
  59. Meenatchi, B.; Renuga, V.; Manikandan, A. Cellulose dissolution and regeneration using various imidazolium based protic ionic liquids. J. Mol. Liq. 2017, 238, 582–588. [Google Scholar] [CrossRef]
  60. Liu, Y.-R.; Thomsen, K.; Nie, Y.; Zhang, S.-J.; Meyer, A.S. Predictive screening of ionic liquids for dissolving cellulose and experimental verification. Green Chem. 2016, 18, 6246–6254. [Google Scholar] [CrossRef]
  61. Lefroy, K.S.; Murray, B.S.; Ries, M.E. Effect of Oil on Cellulose Dissolution in the Ionic Liquid 1-Butyl-3-methyl Imidazolium Acetate. ACS Omega 2022, 7, 37532–37545. [Google Scholar] [CrossRef]
  62. Wahlström, R.M.; Suurnäkki, A. Enzymatic hydrolysis of lignocellulosic polysaccharides in the presence of ionic liquids. Green Chem. 2015, 17, 694–714. [Google Scholar] [CrossRef] [Green Version]
  63. Haq, M.A.; Habu, Y.; Yamamoto, K.; Takada, A.; Kadokawa, J.-I. Ionic liquid induces flexibility and thermoplasticity in cellulose film. Carbohydr. Polym. 2019, 223, 115058. [Google Scholar] [CrossRef] [PubMed]
  64. Idenoue, S.; Oga, Y.; Hashimoto, D.; Yamamoto, K.; Kadokawa, J.I. Preparation of Reswellable Amorphous Porous Celluloses through Hydrogelation from Ionic Liquid Solutions. Materials 2019, 12, 3249. [Google Scholar] [CrossRef] [Green Version]
  65. Wang, Y.; Xie, T.; Zhang, J.; Dang, B.; Li, Y. Green fabrication of an ionic liquid-activated lignocellulose flame-retardant composite. Ind. Crops Prod. 2022, 178, 114602. [Google Scholar] [CrossRef]
  66. Hu, Y.; Guo, Q.; Liu, P.; Zhu, R.; Lu, F.; Ramaswamy, S.; Wu, Y.; Xu, F.; Zhang, X. Fabrication of Novel Cellulose-Based Antibacterial Film Loaded with Poacic Acid against Staphylococcus Aureus. J. Polym. Environ. 2021, 29, 745–754. [Google Scholar] [CrossRef]
  67. Spörl, J.M.; Batti, F.; Vocht, M.-P.; Raab, R.; Müller, A.; Hermanutz, F.; Buchmeiser, M.R. Ionic Liquid Approach Toward Manufacture and Full Recycling of All-Cellulose Composites. Macromol. Mater. Eng. 2018, 303, 1700335. [Google Scholar] [CrossRef] [Green Version]
  68. Chen, F.; Sawada, D.; Hummel, M.; Sixta, H.; Budtova, T. Unidirectional All-Cellulose Composites from Flax via Controlled Impregnation with Ionic Liquid. Polymers 2020, 12, 1010. [Google Scholar] [CrossRef]
  69. Duchemin, B.J.C.; Mathew, A.P.; Oksman, K. All-cellulose composites by partial dissolution in the ionic liquid 1-butyl-3-methylimidazolium chloride. Compos. Part A Appl. Sci. Manuf. 2009, 40, 2031–2037. [Google Scholar] [CrossRef]
  70. Khakalo, A.; Tanaka, A.; Korpela, A.; Hauru, L.K.J.; Orelma, H. All-Wood Composite Material by Partial Fiber Surface Dissolution with an Ionic Liquid. ACS Sustain. Chem. Eng. 2019, 7, 3195–3202. [Google Scholar] [CrossRef] [Green Version]
  71. Vinogradova, Y.S.; Chen, J.Y. Micron- and nano-cellulose fiber regenerated from ionic liquids. J. Text. Inst. 2016, 107, 472–476. [Google Scholar] [CrossRef]
  72. Zheng, X.; Huang, F.; Chen, L.; Huang, L.; Cao, S.; Ma, X. Preparation of transparent film via cellulose regeneration: Correlations between ionic liquid and film properties. Carbohydr. Polym. 2019, 203, 214–218. [Google Scholar] [CrossRef] [PubMed]
  73. Xue, Y.; Li, W.; Yang, G.; Lin, Z.; Qi, L.; Zhu, P.; Yu, J.; Chen, J. Strength Enhancement of Regenerated Cellulose Fibers by Adjustment of Hydrogen Bond Distribution in Ionic Liquid. Polymers 2022, 14, 2030. [Google Scholar] [CrossRef]
  74. Jiang, J.; Xiao, Y.; Huang, W.; Gong, P.; Peng, S.; He, J.; Fan, M.; Wang, K. An insight into the influence of hydrogen bond acceptors on cellulose/1-allyl-3-methyl imidazolium chloride solution. Carbohydr. Polym. 2017, 178, 295–301. [Google Scholar] [CrossRef] [PubMed]
  75. Zhou, L.; Pan, F.; Liu, Y.; Kang, Z.; Zeng, S.; Nie, Y. Study on the regularity of cellulose degradation in ionic liquids. J. Mol. Liq. 2020, 308, 113153. [Google Scholar] [CrossRef]
  76. Xu, K.; Xiao, Y.; Cao, Y.; Peng, S.; Fan, M.; Wang, K. Dissolution of cellulose in 1-allyl-3-methylimidazolium methyl phosphonate ionic liquid and its composite system with Na2PHO3. Carbohydr. Polym. 2019, 209, 382–388. [Google Scholar] [CrossRef]
  77. Fumino, K.; Peppel, T.; Geppert-Rybczyńska, M.; Zaitsau, D.H.; Lehmann, J.K.; Verevkin, S.P.; Köckerling, M.; Ludwig, R. The influence of hydrogen bonding on the physical properties of ionic liquids. Phys. Chem. Chem. Phys. 2011, 13, 14064–14075. [Google Scholar] [CrossRef] [PubMed]
  78. Zweckmair, T.; Hettegger, H.; Abushammala, H.; Bacher, M.; Potthast, A.; Laborie, M.-P.; Rosenau, T. On the mechanism of the unwanted acetylation of polysaccharides by 1,3-dialkylimidazolium acetate ionic liquids: Part 1—Analysis, acetylating agent, influence of water, and mechanistic considerations. Cellulose 2015, 22, 3583–3596. [Google Scholar] [CrossRef]
  79. Li, X.; Li, H.; Ling, Z.; Xu, D.; You, T.; Wu, Y.-Y.; Xu, F. Room-Temperature Superbase-Derived Ionic Liquids with Facile Synthesis and Low Viscosity: Powerful Solvents for Cellulose Dissolution by Destroying the Cellulose Aggregate Structure. Macromolecules 2020, 53, 3284–3295. [Google Scholar] [CrossRef]
  80. Wang, X.; Zheng, W.; Guo, Z.; Nawaz, H.; You, T.; Li, X.; Xu, F. Rheological characteristics of novel cellulose/superbase-derived ionic liquid solutions and the coagulation process towards regenerated cellulose films. Green Chem. 2023, 25, 1597–1610. [Google Scholar] [CrossRef]
  81. Li, X.; Wang, X.; Nawaz, H.; Zhang, J.; Chen, X.; Cheng, P.; You, T.; Ramaswamy, S.; Xu, F. The choice of ionic liquid ions to mitigate corrosion impacts: The influence of superbase cations and electron-donating carboxylate anions. Green Chem. 2022, 24, 2114–2128. [Google Scholar] [CrossRef]
  82. Liu, Q.; Yu, H.; Mu, T.; Xue, Z.; Xu, F. Robust superbase-based emerging solvents for highly efficient dissolution of cellulose. Carbohydr. Polym. 2021, 272, 118454. [Google Scholar] [CrossRef] [PubMed]
  83. Galamba, J.; Alves, V.D.; Jordão, N.; Neves, L.A. Development of cellulose-based polymeric structures using dual functional ionic liquids. RSC Adv. 2021, 11, 39278–39286. [Google Scholar] [CrossRef] [PubMed]
  84. Zhao, Y.; Liu, X.; Wang, J.; Zhang, S. Insight into the Cosolvent Effect of Cellulose Dissolution in Imidazolium-Based Ionic Liquid Systems. J. Phys. Chem. B 2013, 117, 9042–9049. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, A.; Zhang, Y.; Zhao, Y.; Wang, J. Cellulose dissolution at ambient temperature: Role of preferential solvation of cations of ionic liquids by a cosolvent. Carbohydr. Polym. 2013, 92, 540–544. [Google Scholar] [CrossRef]
  86. Zhou, Y.; Zhang, X.; Yin, D.; Zhang, J.; Mi, Q.; Lu, H.; Liang, D.; Zhang, J. The solution state and dissolution process of cellulose in ionic-liquid-based solvents with different hydrogen-bonding basicity and microstructures. Green Chem. 2022, 24, 3824–3833. [Google Scholar] [CrossRef]
  87. Sánchez, P.B.; Tsubaki, S.; Pádua, A.A.H.; Wada, Y. Kinetic analysis of microwave-enhanced cellulose dissolution in ionic solvents. Phys. Chem. Chem. Phys. 2020, 22, 1003–1010. [Google Scholar] [CrossRef]
  88. Hawkins, J.E.; Liang, Y.; Ries, M.E.; Hine, P.J. Time temperature superposition of the dissolution of cellulose fibres by the ionic liquid 1-ethyl-3-methylimidazolium acetate with cosolvent dimethyl sulfoxide. Carbohydr. Polym. Technol. Appl. 2021, 2, 100021. [Google Scholar] [CrossRef]
  89. Shamsuri, A.A.; Abdan, K.; Jamil, S.N.A.M. Properties and applications of cellulose regenerated from cellulose/imidazolium-based ionic liquid/co-solvent solutions: A short review. e-Polymers 2021, 21, 869–880. [Google Scholar] [CrossRef]
  90. Chen, F.; Sawada, D.; Hummel, M.; Sixta, H.; Budtova, T. Swelling and dissolution kinetics of natural and man-made cellulose fibers in solvent power tuned ionic liquid. Cellulose 2020, 27, 7399–7415. [Google Scholar] [CrossRef]
  91. Ferreira, D.C.; Oliveira, M.L.; Bioni, T.A.; Nawaz, H.; King, A.W.T.; Kilpeläinen, I.; Hummel, M.; Sixta, H.; El Seoud, O.A. Binary mixtures of ionic liquids-DMSO as solvents for the dissolution and derivatization of cellulose: Effects of alkyl and alkoxy side chains. Carbohydr. Polym. 2019, 212, 206–214. [Google Scholar] [CrossRef] [Green Version]
  92. Wittmar, A.S.M.; Koch, D.; Prymak, O.; Ulbricht, M. Factors Affecting the Nonsolvent-Induced Phase Separation of Cellulose from Ionic Liquid-Based Solutions. ACS Omega 2020, 5, 27314–27322. [Google Scholar] [CrossRef]
  93. Satani, H.; Kuwata, M.; Shimizu, A. Simple and environmentally friendly preparation of cellulose hydrogels using an ionic liquid. Carbohydr. Res. 2020, 494, 108054. [Google Scholar] [CrossRef]
  94. Kikuchi, K.; Kaneko, K.; Seonju, J.; Fukaya, R.; Yamada, M.; Ishii, H.; Inoue, T.; Shimizu, A. Influence of gelation temperature on physicochemical properties of cellulose hydrogels prepared from ionic liquid/DMSO solution. J. Mol. Liq. 2023, 376, 121465. [Google Scholar] [CrossRef]
  95. Ilyin, S.O.; Kostyuk, A.V.; Anokhina, T.S.; Melekhina, V.Y.; Bakhtin, D.S.; Antonov, S.V.; Volkov, A.V. The Effect of Non-Solvent Nature on the Rheological Properties of Cellulose Solution in Diluted Ionic Liquid and Performance of Nanofiltration Membranes. Int. J. Mol. Sci. 2023, 24, 8057. [Google Scholar] [CrossRef]
  96. Layek, S.; Banerjee, P.; Sarkar, N. An insight into the dissolution of cellulose in 1-butyl-3-methylimidazolium chloride-DMSO binary Mixture: Exploring the dynamics of rhodamine 6G and fluorescein. J. Mol. Liq. 2021, 339, 116817. [Google Scholar] [CrossRef]
  97. Liu, Y.; Jing, S.; Carvalho, D.; Fu, J.; Martins, M.; Cavaco-Paulo, A. Cellulose Dissolved in Ionic Liquids for Modification of the Shape of Keratin Fibers. ACS Sustain. Chem. Eng. 2021, 9, 4102–4110. [Google Scholar] [CrossRef]
  98. Müller, K.; Zollfrank, C. Ionic liquid aided solution-precipitation method to prepare polymer blends from cellulose with polyesters or polyamide. Eur. Polym. J. 2020, 133, 109743. [Google Scholar] [CrossRef]
  99. Lei, L.; Lindbråthen, A.; Sandru, M.; Gutierrez, M.T.; Zhang, X.; Hillestad, M.; He, X. Spinning Cellulose Hollow Fibers Using 1-Ethyl-3-methylimidazolium Acetate–Dimethylsulfoxide Co-Solvent. Polymers 2018, 10, 972. [Google Scholar] [CrossRef] [Green Version]
  100. Xie, H.; Yu, X.; Yang, Y.; Zhao, Z.K. Capturing CO2 for cellulose dissolution. Green Chem. 2014, 16, 2422–2427. [Google Scholar] [CrossRef]
  101. Jin, L.; Gan, J.; Hu, G.; Cai, L.; Li, Z.; Zhang, L.; Zheng, Q.; Xie, H. Preparation of Cellulose Films from Sustainable CO2/DBU/DMSO System. Polymers 2019, 11, 994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zhang, L.; Shi, W.; Sheng, H.; Feng, S.; Yao, M.; Chen, P.; Zheng, Q.; Xie, H. Unique CO2-switched cellulose solution properties in the CO2/DBU/DMSO solvent system and the preparation of regenerated materials. Green Chem. 2021, 23, 5856–5865. [Google Scholar] [CrossRef]
  103. Yang, Y.; Song, L.; Peng, C.; Liu, E.; Xie, H. Activating cellulose via its reversible reaction with CO2 in the presence of 1,8-diazabicyclo[5.4.0]undec-7-ene for the efficient synthesis of cellulose acetate. Green Chem. 2015, 17, 2758–2763. [Google Scholar] [CrossRef]
  104. Onwukamike, K.N.; Lapuyade, L.; Maillé, L.; Grelier, S.; Grau, E.; Cramail, H.; Meier, M.A.R. Sustainable Approach for Cellulose Aerogel Preparation from the DBU–CO2 Switchable Solvent. ACS Sustain. Chem. Eng. 2019, 7, 3329–3338. [Google Scholar] [CrossRef] [Green Version]
  105. Zhang, L.; Shi, W.; Wang, J.; Jin, L.; Hu, G.; Zheng, Q.; Xie, H.; Chen, P. Unique gelation and rheological properties of the cellulose/CO2-based reversible ionic liquid/DMSO solutions. Carbohydr. Polym. 2019, 222, 115024. [Google Scholar] [CrossRef]
  106. Zhang, Q.; Oztekin, N.S.; Barrault, J.; De Oliveira Vigier, K.; Jérôme, F. Activation of Microcrystalline Cellulose in a CO2-Based Switchable System. ChemSusChem 2013, 6, 593–596. [Google Scholar] [CrossRef]
  107. Onwukamike, K.N.; Tassaing, T.; Grelier, S.; Grau, E.; Cramail, H.; Meier, M.A.R. Detailed Understanding of the DBU/CO2 Switchable Solvent System for Cellulose Solubilization and Derivatization. ACS Sustain. Chem. Eng. 2018, 6, 1496–1503. [Google Scholar] [CrossRef] [Green Version]
  108. Wang, J.; Xue, Z.; Yan, C.; Li, Z.; Mu, T. Fine regulation of cellulose dissolution and regeneration by low pressure CO2 in DMSO/organic base: Dissolution behavior and mechanism. Phys. Chem. Chem. Phys. 2016, 18, 32772–32779. [Google Scholar] [CrossRef]
  109. Nanta, P.; Skolpap, W.; Kasemwong, K.; Shimoyama, Y. Dissolution and modification of cellulose using high-pressure carbon dioxide switchable solution. J. Supercrit. Fluids 2017, 130, 84–90. [Google Scholar] [CrossRef]
  110. Li, J.; Lu, S.; Liu, F.; Qiao, Q.; Na, H.; Zhu, J. Structure and Properties of Regenerated Cellulose Fibers Based on Dissolution of Cellulose in a CO2 Switchable Solvent. ACS Sustain. Chem. Eng. 2021, 9, 4744–4754. [Google Scholar] [CrossRef]
  111. Xu, Q.; Song, L.; Zhang, L.; Hu, G.; Du, J.; Liu, E.; Zheng, Q.; Liu, Y.; Li, N.; Xie, H. Organocatalytic Cellulose Dissolution and In Situ Grafting of ϵ-Caprolactone via ROP in a Reversible DBU/DMSO/CO2 System. ChemistrySelect 2017, 2, 7128–7134. [Google Scholar] [CrossRef]
  112. Wang, C.-G.; Li, N.; Wu, G.; Lin, T.T.; Lee, A.M.X.; Yang, S.-W.; Li, Z.; Luo, D.H.-K. Carbon Dioxide Mediated Cellulose Dissolution and Derivatization to Cellulose Carbonates in a Low-pressure System. Carbohydr. Polym. Technol. Appl. 2022, 3, 100186. [Google Scholar] [CrossRef]
  113. Zhang, J.; Mi, S.; Liu, F.; Qiao, Q.; Na, H.; Zhu, J. Enhanced dissolution of cellulose in CO2 switchable system by solvent with low Henry’s constants as CO2 absorbent. Cellulose 2022, 29, 6745–6758. [Google Scholar] [CrossRef]
  114. Gupta, K.M.; Hu, Z.; Jiang, J. Molecular insight into cellulose regeneration from a cellulose/ionic liquid mixture: Effects of water concentration and temperature. RSC Adv. 2013, 3, 4425–4433. [Google Scholar] [CrossRef]
  115. Olsson, C.; Idström, A.; Nordstierna, L.; Westman, G. Influence of water on swelling and dissolution of cellulose in 1-ethyl-3-methylimidazolium acetate. Carbohydr. Polym. 2014, 99, 438–446. [Google Scholar] [CrossRef]
  116. Berton, P.; Shen, X.; Rogers, R.D.; Shamshina, J.L. 110th Anniversary: High-Molecular-Weight Chitin and Cellulose Hydrogels from Biomass in Ionic Liquids without Chemical Crosslinking. Ind. Eng. Chem. Res. 2019, 58, 19862–19876. [Google Scholar] [CrossRef]
  117. Remsing, R.C.; Swatloski, R.P.; Rogers, R.D.; Moyna, G. Mechanism of cellulose dissolution in the ionic liquid 1-n-butyl-3-methylimidazolium chloride: A 13C and 35/37Cl NMR relaxation study on model systems. Chem. Commun. 2006, 12, 1271–1273. [Google Scholar] [CrossRef]
  118. Fan, Z.; Chen, J.; Guo, W.; Ma, F.; Sun, S.; Zhou, Q. Anti-solvents tuning cellulose nanoparticles through two competitive regeneration routes. Cellulose 2018, 25, 4513–4523. [Google Scholar] [CrossRef]
  119. Fan, Z.; Chen, J.; Guo, W.; Ma, F.; Sun, S.; Zhou, Q. Crystallinity of regenerated cellulose from [Bmim]Cl dependent on the hydrogen bond acidity/basicity of anti-solvents. RSC Adv. 2017, 7, 41004–41010. [Google Scholar] [CrossRef] [Green Version]
  120. Berga, L.; Bruce, I.; Nicol, T.W.J.; Holding, A.J.; Isobe, N.; Shimizu, S.; Walker, A.J.; Reid, J.E.S.J. Cellulose dissolution and regeneration using a non-aqueous, non-stoichiometric protic ionic liquid system. Cellulose 2020, 27, 9593–9603. [Google Scholar] [CrossRef]
  121. Liu, B.; Li, W.; Xu, Y.; Zhang, H.; Cai, R.; Guo, Z.; Zhou, L.; Zhang, J.; Yuan, Y. Mechanism of cellulose regeneration from its ionic liquid solution as revealed by infrared spectroscopy. Polymer 2022, 257, 125280. [Google Scholar] [CrossRef]
  122. Zhang, S.; Wang, J.; Shen, F.; Wang, D.; Zhang, S. Efficient removal of metal ions from the ionic liquid aqueous solution by selective electrodialysis. Sep. Purif. Technol. 2022, 295, 121322. [Google Scholar] [CrossRef]
  123. Peng, H.; Wang, S.; Xu, H.; Dai, G. Preparations, properties, and formation mechanism of novel cellulose hydrogel membrane based on ionic liquid. J. Appl. Polym. Sci. 2018, 135, 45488. [Google Scholar] [CrossRef]
  124. Jedvert, K.; Idström, A.; Köhnke, T.; Alkhagen, M. Cellulosic nonwovens produced via efficient solution blowing technique. J. Appl. Polym. Sci. 2020, 137, 48339. [Google Scholar] [CrossRef] [Green Version]
  125. Fan, Z.; Chen, J.; Sun, S.; Zhou, Q. Surfactant-assisted fabrication of ultra-permeable cellulose gels with macro channels and insights on regeneration of cellulose from ionic liquids. J. Mol. Liq. 2019, 280, 64–70. [Google Scholar] [CrossRef]
  126. Heasman, P.; Mehandzhiyski, A.Y.; Ghosh, S.; Zozoulenko, I. A computational study of cellulose regeneration: All-atom molecular dynamics simulations. Carbohydr. Polym. 2023, 311, 120768. [Google Scholar] [CrossRef]
  127. Pang, J.; Mehandzhiyski, A.Y.; Zozoulenko, I. A computational study of cellulose regeneration: Coarse-grained molecular dynamics simulations. Carbohydr. Polym. 2023, 313, 120853. [Google Scholar] [CrossRef]
  128. Ding, Z.-D.; Chi, Z.; Gu, W.-X.; Gu, S.-M.; Liu, J.-H.; Wang, H.-J. Theoretical and experimental investigation on dissolution and regeneration of cellulose in ionic liquid. Carbohydr. Polym. 2012, 89, 7–16. [Google Scholar] [CrossRef]
  129. Fu, L.; Ju, Z.; Yu, M.; Luo, H.; Zhang, C.; Zhang, X.; Cheng, H.; Zheng, M.; Jin, L.; Ge, C. Cellulose Regeneration in Imidazolium-Based Ionic Liquids and Antisolvent Mixtures: A Density Functional Theory Study. ACS Omega 2022, 7, 42170–42180. [Google Scholar] [CrossRef]
  130. Sun, L.; Chen, J.Y.; Jiang, W.; Lynch, V. Crystalline characteristics of cellulose fiber and film regenerated from ionic liquid solution. Carbohydr. Polym. 2015, 118, 150–155. [Google Scholar] [CrossRef]
  131. Liu, Y.; Nie, Y.; Pan, F.; Zhou, L.; Ji, X.; Kang, Z.; Zhang, S. Study on ionic liquid/cellulose/coagulator phase diagram and its application in green spinning process. J. Mol. Liq. 2019, 289, 111127. [Google Scholar] [CrossRef]
  132. Sangtarashani, S.M.H.; Rahmaninia, M.; Behrooz, R.; Khosravani, A. Lignocellulosic hydrogel from recycled old corrugated container resources using ionic liquid as a green solvent. J. Environ. Manag. 2020, 270, 110853. [Google Scholar] [CrossRef] [PubMed]
  133. Ingildeev, D.; Effenberger, F.; Bredereck, K.; Hermanutz, F. Comparison of direct solvents for regenerated cellulosic fibers via the lyocell process and by means of ionic liquids. J. Appl. Polym. Sci. 2013, 128, 4141–4150. [Google Scholar] [CrossRef]
  134. Zhang, D.; Wang, Y.; Shi, Z. Influences of diffusion coefficient of 1-allyl-3-methylimidazolium chloride on structure and properties of regenerated cellulose fiber obtained via dry-jet-wet spinning. J. Appl. Polym. Sci. 2019, 136, 47609. [Google Scholar] [CrossRef]
  135. Tan, X.; Chen, L.; Li, X.; Xie, F. Effect of anti-solvents on the characteristics of regenerated cellulose from 1-ethyl-3-methylimidazolium acetate ionic liquid. Int. J. Biol. Macromol. 2019, 124, 314–320. [Google Scholar] [CrossRef] [Green Version]
  136. Hauru, L.K.J.; Hummel, M.; King, A.W.T.; Kilpeläinen, I.; Sixta, H. Role of Solvent Parameters in the Regeneration of Cellulose from Ionic Liquid Solutions. Biomacromolecules 2012, 13, 2896–2905. [Google Scholar] [CrossRef]
  137. Guo, Y.; Cai, L.; Guo, G.; Xie, H.; Zhang, L.; Jin, L.; Liang, S.; Hu, L.; Xu, Q.; Zheng, Q. Cellulose Membranes from Cellulose CO2-Based Reversible Ionic Liquid Solutions. ACS Sustain. Chem. Eng. 2021, 9, 11847–11854. [Google Scholar] [CrossRef]
  138. Wang, B.; Nie, Y.; Kang, Z.; Liu, X. Effects of coagulating conditions on the crystallinity, orientation and mechanical properties of regenerated cellulose fibers. Int. J. Biol. Macromol. 2023, 225, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  139. Guizani, C.; Larkiala, S.; Moriam, K.; Sawada, D.; Elsayed, S.; Rantasalo, S.; Hummel, M.; Sixta, H. Air gap spinning of a cellulose solution in [DBNH][OAc] ionic liquid with a novel vertically arranged spinning bath to simulate a closed loop operation in the Ioncell® process. J. Appl. Polym. Sci. 2021, 138, 49787. [Google Scholar] [CrossRef]
  140. To, T.Q.; Kenny, C.; Cheong, S.; Aldous, L. Carbon dioxide as a pH-switch anti-solvent for biomass fractionation and pre-treatment with aqueous hydroxide solutions. Green Chem. 2017, 19, 2129–2134. [Google Scholar] [CrossRef] [Green Version]
  141. Shamshina, J.L.; Qin, Y.; Belmore, K.; Daly, D.T.; Rogers, R.D. Switchable carbamate coagulants to improve recycling ionic liquid from biomass solutions. Green Chem. Eng. 2021, 2, 384–391. [Google Scholar] [CrossRef]
  142. Barber, P.S.; Griggs, C.S.; Gurau, G.; Liu, Z.; Li, S.; Li, Z.; Lu, X.; Zhang, S.; Rogers, R.D. Coagulation of chitin and cellulose from 1-ethyl-3-methylimidazolium acetate ionic-liquid solutions using carbon dioxide. Angew. Chem. Int. Ed. 2013, 52, 12350–12353. [Google Scholar] [CrossRef] [PubMed]
  143. Sun, X.; Chi, Y.; Mu, T. Studies on staged precipitation of cellulose from an ionic liquid by compressed carbon dioxide. Green Chem. 2014, 16, 2736–2744. [Google Scholar] [CrossRef]
  144. Xu, A.-R.; Guo, X.; Ma, J. Properties of Cellulose Regenerated from Powerful 1-Butyl-3-methylimidazolium Acetate/Dimethyl Sulfoxide Solvent. J. Macromol. Sci. Part B 2016, 55, 559–565. [Google Scholar] [CrossRef]
  145. Ebrahiminejadhasanabadi, M.; Nelson, W.M.; Naidoo, P.; Mohammadi, A.H.; Ramjugernath, D. Investigation of Mixed MEA-Based Solvents Featuring Ionic Liquids and NMP for CO2 Capture: Experimental Measurement of CO2 Solubility and Thermophysical Properties. J. Chem. Eng. Data 2021, 66, 899–914. [Google Scholar] [CrossRef]
  146. Gelles, T.; Lawson, S.; Rownaghi, A.A.; Rezaei, F. Recent advances in development of amine functionalized adsorbents for CO2 capture. Adsorption 2020, 26, 5–50. [Google Scholar] [CrossRef]
  147. Varghese, A.M.; Karanikolos, G.N. CO2 capture adsorbents functionalized by amine–bearing polymers: A review. Int. J. Greenh. Gas Control 2020, 96, 103005. [Google Scholar] [CrossRef]
  148. Li, J.; Yun, X.; Hu, Z.; Xi, L.; Li, N.; Tang, H.; Lu, P.; Zhu, Y. Three-dimensional nitrogen and phosphorus co-doped carbon quantum dots/reduced graphene oxide composite aerogels with a hierarchical porous structure as superior electrode materials for supercapacitors. J. Mater. Chem. A 2019, 7, 26311–26325. [Google Scholar] [CrossRef]
  149. Yun, X.; Wu, S.; Li, J.; Li, L.; Zhou, J.; Lu, P.; Tang, H.; Zhu, Y. Facile synthesis of crystalline RuSe2 nanoparticles as a novel pseudocapacitive electrode material for supercapacitors. Chem. Commun. 2019, 55, 12320–12323. [Google Scholar] [CrossRef]
  150. Kasprzak, D.; Galiński, M. Chitin and chitin-cellulose composite hydrogels prepared by ionic liquid-based process as the novel electrolytes for electrochemical capacitors. J. Solid State Electrochem. 2021, 25, 2549–2563. [Google Scholar] [CrossRef]
  151. Koga, H.; Nogi, M.; Isogai, A. Ionic Liquid Mediated Dispersion and Support of Functional Molecules on Cellulose Fibers for Stimuli-Responsive Chromic Paper Devices. ACS Appl. Mater. Interfaces 2017, 9, 40914–40920. [Google Scholar] [CrossRef] [PubMed]
  152. Morales, A.; Seelam, S.; Love, S.A.; O’Malley, S.M.; Hu, X.; Salas-de la Cruz, D. Reduced graphene oxide influences morphology and thermal properties of silk/cellulose biocomposites. Int. J. Biol. Macromol. 2023, 236, 123971. [Google Scholar] [CrossRef] [PubMed]
  153. Chandrasekaran, S.; Sotenko, M.; Cruz-Izquierdo, A.; Rymansaib, Z.; Iravani, P.; Kirwan, K.; Scott, J.L. Preparation of Printable and Biodegradable Cellulose-Laponite Composite for Electronic Device Application. J. Polym. Environ. 2021, 29, 17–27. [Google Scholar] [CrossRef]
  154. Huang, Q.; Yang, Y.; Chen, R.; Wang, X. High performance fully paper-based all-solid-state supercapacitor fabricated by a papermaking process with silver nanoparticles and reduced graphene oxide-modified pulp fibers. EcoMat 2021, 3, e12076. [Google Scholar] [CrossRef]
  155. Guo, Y.; Xu, G.; Yang, X.; Ruan, K.; Ma, T.; Zhang, Q.; Gu, J.; Wu, Y.; Liu, H.; Guo, Z. Significantly enhanced and precisely modeled thermal conductivity in polyimide nanocomposites with chemically modified graphene via in situ polymerization and electrospinning-hot press technology. J. Mater. Chem. C 2018, 6, 3004–3015. [Google Scholar] [CrossRef]
  156. Guo, Y.; Yang, X.; Ruan, K.; Kong, J.; Dong, M.; Zhang, J.; Gu, J.; Guo, Z. Reduced Graphene Oxide Heterostructured Silver Nanoparticles Significantly Enhanced Thermal Conductivities in Hot-Pressed Electrospun Polyimide Nanocomposites. ACS Appl. Mater. Interfaces 2019, 11, 25465–25473. [Google Scholar] [CrossRef]
  157. Chen, Y.; Hou, X.; Kang, R.; Liang, Y.; Guo, L.; Dai, W.; Nishimura, K.; Lin, C.-T.; Jiang, N.; Yu, J. Highly flexible biodegradable cellulose nanofiber/graphene heat-spreader films with improved mechanical properties and enhanced thermal conductivity. J. Mater. Chem. C 2018, 6, 12739–12745. [Google Scholar] [CrossRef]
  158. Ahmed, A.; Adak, B.; Bansala, T.; Mukhopadhyay, S. Green Solvent Processed Cellulose/Graphene Oxide Nanocomposite Films with Superior Mechanical, Thermal, and Ultraviolet Shielding Properties. ACS Appl. Mater. Interfaces 2020, 12, 1687–1697. [Google Scholar] [CrossRef]
  159. Zhang, K.; Ketterle, L.; Järvinen, T.; Lorite, G.S.; Hong, S.; Liimatainen, H. Self-assembly of graphene oxide and cellulose nanocrystals into continuous filament via interfacial nanoparticle complexation. Mater. Des. 2020, 193, 108791. [Google Scholar] [CrossRef]
  160. Xu, M.; Huang, Q.; Wang, X.; Sun, R. Highly tough cellulose/graphene composite hydrogels prepared from ionic liquids. Ind. Crops Prod. 2015, 70, 56–63. [Google Scholar] [CrossRef]
  161. Zhang, D.; Peng, L.; Shi, N.; Yu, Y.; Min, Y.; Epstein, A.J. Self-assembled high-performance graphene oxide fibers using ionic liquid as coagulating agent. J. Mater. Sci. 2017, 52, 7698–7708. [Google Scholar] [CrossRef]
  162. Murugesan, B.; Pandiyan, N.; Arumugam, M.; Veerasingam, M.; Sonamuthu, J.; Jeyaraman, A.R.; Samayanan, S.; Mahalingam, S. Two dimensional graphene oxides converted to three dimensional P, N, F and B, N, F tri-doped graphene by ionic liquid for efficient catalytic performance. Carbon 2019, 151, 53–67. [Google Scholar] [CrossRef]
  163. Zhou, K.; Chen, C.; Lei, M.; Gao, Q.; Nie, S.; Liu, X.; Wang, S. Reduced graphene oxide-based highly sensitive pressure sensor for wearable electronics via an ordered structure and enhanced interlayer interaction mechanism. RSC Adv. 2020, 10, 2150–2159. [Google Scholar] [CrossRef]
  164. Liang, X.; Qu, B.; Li, J.; Xiao, H.; He, B.; Qian, L. Preparation of cellulose-based conductive hydrogels with ionic liquid. React. Funct. Polym. 2015, 86, 1–6. [Google Scholar] [CrossRef]
  165. Wan, J.; Zhang, J.; Yu, J.; Zhang, J. Cellulose Aerogel Membranes with a Tunable Nanoporous Network as a Matrix of Gel Polymer Electrolytes for Safer Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 24591–24599. [Google Scholar] [CrossRef] [PubMed]
  166. Morelle, X.P.; Illeperuma, W.R.; Tian, K.; Bai, R.; Suo, Z.; Vlassak, J.J. Highly Stretchable and Tough Hydrogels below Water Freezing Temperature. Adv. Mater. 2018, 30, 1801541. [Google Scholar] [CrossRef]
  167. Chereddy, S.; Aguirre, J.; Dikin, D.; Wunder, S.L.; Chinnam, P.R. Gel Electrolyte Comprising Solvate Ionic Liquid and Methyl Cellulose. ACS Appl. Energy Mater. 2020, 3, 279–289. [Google Scholar] [CrossRef] [Green Version]
  168. Khanmirzaei, M.H.; Ramesh, S.; Ramesh, K. Hydroxypropyl Cellulose Based Non-Volatile Gel Polymer Electrolytes for Dye-Sensitized Solar Cell Applications using 1-methyl-3-propylimidazolium iodide ionic liquid. Sci. Rep. 2015, 5, 18056. [Google Scholar] [CrossRef] [Green Version]
  169. Dai, C.; Ye, C.; Ren, J.; Yang, S.; Cao, L.; Yu, H.; Liu, S.; Shao, Z.; Li, J.; Chen, W.; et al. Humanoid Ionotronic Skin for Smart Object Recognition and Sorting. ACS Mater. Lett. 2023, 5, 189–201. [Google Scholar] [CrossRef]
  170. Ge, D.; Yu, H.-Y.; Miao, Z.; He, X.; Abdalkarim, S.Y.H. Intrinsically Conductive Bifunctional Nanocellulose-Reinforced Robust and Self-Healable Electronic Skin: Deep Insights into Multiple Bonding Network, Property Reinforcement, and Sensing Mechanism. ACS Sustain. Chem. Eng. 2023, 11, 1157–1167. [Google Scholar] [CrossRef]
  171. Fan, F.-R.; Tian, Z.-Q.; Lin Wang, Z. Flexible triboelectric generator. Nano Energy 2012, 1, 328–334. [Google Scholar] [CrossRef]
  172. Hu, Y.; Zhang, M.; Qin, C.; Qian, X.; Zhang, L.; Zhou, J.; Lu, A. Transparent, conductive cellulose hydrogel for flexible sensor and triboelectric nanogenerator at subzero temperature. Carbohydr. Polym. 2021, 265, 118078. [Google Scholar] [CrossRef] [PubMed]
  173. Li, T.; Wang, Y.; Li, S.; Liu, X.; Sun, J. Mechanically Robust, Elastic, and Healable Ionogels for Highly Sensitive Ultra-Durable Ionic Skins. Adv. Mater. 2020, 32, 2002706. [Google Scholar] [CrossRef] [PubMed]
  174. Sun, L.; Chen, S.; Guo, Y.; Song, J.; Zhang, L.; Xiao, L.; Guan, Q.; You, Z. Ionogel-based, highly stretchable, transparent, durable triboelectric nanogenerators for energy harvesting and motion sensing over a wide temperature range. Nano Energy 2019, 63, 103847. [Google Scholar] [CrossRef]
  175. Zhou, Y.; Wan, C.; Yang, Y.; Yang, H.; Wang, S.; Dai, Z.; Ji, K.; Jiang, H.; Chen, X.; Long, Y. Highly Stretchable, Elastic, and Ionic Conductive Hydrogel for Artificial Soft Electronics. Adv. Funct. Mater. 2019, 29, 1806220. [Google Scholar] [CrossRef]
  176. Chen, G.; Huang, J.; Gu, J.; Peng, S.; Xiang, X.; Chen, K.; Yang, X.; Guan, L.; Jiang, X.; Hou, L. Highly tough supramolecular double network hydrogel electrolytes for an artificial flexible and low-temperature tolerant sensor. J. Mater. Chem. A 2020, 8, 6776–6784. [Google Scholar] [CrossRef]
  177. Hu, K.; He, P.; Zhao, Z.; Huang, L.; Liu, K.; Lin, S.; Zhang, M.; Wu, H.; Chen, L.; Ni, Y. Nature-inspired self-powered cellulose nanofibrils hydrogels with high sensitivity and mechanical adaptability. Carbohydr. Polym. 2021, 264, 117995. [Google Scholar] [CrossRef]
  178. Geng, L.; Hu, S.; Cui, M.; Wu, J.; Huang, A.; Shi, S.; Peng, X. Muscle-inspired double-network hydrogels with robust mechanical property, biocompatibility and ionic conductivity. Carbohydr. Polym. 2021, 262, 117936. [Google Scholar] [CrossRef]
  179. Park, S.; Parida, K.; Lee, P.S. Deformable and Transparent Ionic and Electronic Conductors for Soft Energy Devices. Adv. Energy Mater. 2017, 7, 1701369. [Google Scholar] [CrossRef] [Green Version]
  180. Rong, Q.; Lei, W.; Huang, J.; Liu, M. Low Temperature Tolerant Organohydrogel Electrolytes for Flexible Solid-State Supercapacitors. Adv. Energy Mater. 2018, 8, 1801967. [Google Scholar] [CrossRef]
  181. Liu, T.; Liu, M.; Dou, S.; Sun, J.; Cong, Z.; Jiang, C.; Du, C.; Pu, X.; Hu, W.; Wang, Z.L. Triboelectric-Nanogenerator-Based Soft Energy-Harvesting Skin Enabled by Toughly Bonded Elastomer/Hydrogel Hybrids. ACS Nano 2018, 12, 2818–2826. [Google Scholar] [CrossRef] [PubMed]
  182. Ye, Y.; Zhang, Y.; Chen, Y.; Han, X.; Jiang, F. Cellulose Nanofibrils Enhanced, Strong, Stretchable, Freezing-Tolerant Ionic Conductive Organohydrogel for Multi-Functional Sensors. Adv. Funct. Mater. 2020, 30, 2003430. [Google Scholar] [CrossRef]
  183. Hopson, C.; Villar-Chavero, M.M.; Domínguez, J.C.; Alonso, M.V.; Oliet, M.; Rodriguez, F. Cellulose ionogels, a perspective of the last decade: A review. Carbohydr. Polym. 2021, 274, 118663. [Google Scholar] [CrossRef]
  184. Zhang, J.; Wu, J.; Yu, J.; Zhang, X.; He, J.; Zhang, J. Application of ionic liquids for dissolving cellulose and fabricating cellulose-based materials: State of the art and future trends. Mater. Chem. Front. 2017, 1, 1273–1290. [Google Scholar] [CrossRef]
  185. Zhu, M.; Yu, L.; He, S.; Hong, H.; Liu, J.; Gan, L.; Long, M. Highly Efficient and Stable Cellulose-Based Ion Gel Polymer Electrolyte for Solid-State Supercapacitors. ACS Appl. Energy Mater. 2019, 2, 5992–6001. [Google Scholar] [CrossRef]
  186. Uto, T.; Yamamoto, K.; Kadokawa, J.-i. Cellulose Crystal Dissolution in Imidazolium-Based Ionic Liquids: A Theoretical Study. J. Phys. Chem. B 2018, 122, 258–266. [Google Scholar] [CrossRef]
  187. Kowalczuk, J.; Bielejewski, M.; Tritt-Goc, J. Ionic liquid dynamics and electrical conductivity under confinement within micro and nanocellulose ionogels. Cellulose 2023, 30, 3551–3567. [Google Scholar] [CrossRef]
  188. Zhao, D.; Zhu, Y.; Cheng, W.; Xu, G.; Wang, Q.; Liu, S.; Li, J.; Chen, C.; Yu, H.; Hu, L. A Dynamic Gel with Reversible and Tunable Topological Networks and Performances. Matter 2020, 2, 390–403. [Google Scholar] [CrossRef] [Green Version]
  189. Livazovic, S.; Li, Z.; Behzad, A.R.; Peinemann, K.V.; Nunes, S.P. Cellulose multilayer membranes manufacture with ionic liquid. J. Membr. Sci. 2015, 490, 282–293. [Google Scholar] [CrossRef] [Green Version]
  190. Anokhina, T.S.; Pleshivtseva, T.S.; Ignatenko, V.Y.; Antonov, S.V.; Volkov, A.V. Fabrication of composite nanofiltration membranes from cellulose solutions in an [Emim]OAc–DMSO mixture. Pet. Chem. 2017, 57, 477–482. [Google Scholar] [CrossRef]
  191. Esfahani, M.R.; Taylor, A.; Serwinowski, N.; Parkerson, Z.J.; Confer, M.P.; Kammakakam, I.; Bara, J.E.; Esfahani, A.R.; Mahmoodi, S.N.; Koutahzadeh, N.; et al. Sustainable Novel Bamboo-Based Membranes for Water Treatment Fabricated by Regeneration of Bamboo Waste Fibers. ACS Sustain. Chem. Eng. 2020, 8, 4225–4235. [Google Scholar] [CrossRef]
  192. Dai, H.; Chen, Y.; Ma, L.; Zhang, Y.; Cui, B. Direct regeneration of hydrogels based on lemon peel and its isolated microcrystalline cellulose: Characterization and application for methylene blue adsorption. Int. J. Biol. Macromol. 2021, 191, 129–138. [Google Scholar] [CrossRef]
  193. Dai, H.; Huang, Y.; Zhang, Y.; Zhang, H.; Huang, H. Green and facile fabrication of pineapple peel cellulose/magnetic diatomite hydrogels in ionic liquid for methylene blue adsorption. Cellulose 2019, 26, 3825–3844. [Google Scholar] [CrossRef]
  194. Zhou, Y.; Liao, S.; Chu, Y.; Yuan, B.; Tao, X.; Hu, X.; Wang, Y. An injectable bioink with rapid prototyping in the air and in-situ mild polymerization for 3D bioprinting. Biofabrication 2021, 13, 045026. [Google Scholar] [CrossRef] [PubMed]
  195. Valino, A.D.; Dizon, J.R.C.; Espera, A.H.; Chen, Q.; Messman, J.; Advincula, R.C. Advances in 3D printing of thermoplastic polymer composites and nanocomposites. Prog. Polym. Sci. 2019, 98, 101162. [Google Scholar] [CrossRef]
  196. Sears, N.A.; Wilems, T.S.; Gold, K.A.; Lan, Z.; Cereceres, S.N.; Dhavalikar, P.S.; Foudazi, R.; Cosgriff-Hernandez, E.M. Hydrocolloid Inks for 3D Printing of Porous Hydrogels. Adv. Mater. Technol. 2019, 4, 1800343. [Google Scholar] [CrossRef]
  197. He, X.; Lu, Q. Design and fabrication strategies of cellulose nanocrystal-based hydrogel and its highlighted application using 3D printing: A review. Carbohydr. Polym. 2023, 301, 120351. [Google Scholar] [CrossRef] [PubMed]
  198. Murphy, S.V.; Atala, A. 3D bioprinting of tissues and organs. Nat. Biotechnol. 2014, 32, 773–785. [Google Scholar] [CrossRef]
  199. Mohan, D.; Teong, Z.K.; Sajab, M.S.; Kamarudin, N.H.; Kaco, H. Intact Fibrillated 3D-Printed Cellulose Macrofibrils/CaCO3 for Controlled Drug Delivery. Polymers 2021, 13, 1912. [Google Scholar] [CrossRef]
  200. AbouHashem, Y.; Dayal, M.; Savanah, S.; Štrkalj, G. The application of 3D printing in anatomy education. Med. Educ. Online 2015, 20, 29847. [Google Scholar] [CrossRef] [PubMed]
  201. Vanderploeg, A.; Lee, S.-E.; Mamp, M. The application of 3D printing technology in the fashion industry. Int. J. Fash. Des. Technol. Educ. 2017, 10, 170–179. [Google Scholar] [CrossRef]
  202. He, C.; Zhang, M.; Fang, Z. 3D printing of food: Pretreatment and post-treatment of materials. Crit. Rev. Food Sci. Nutr. 2020, 60, 2379–2392. [Google Scholar] [CrossRef] [PubMed]
  203. Arafat, B.; Wojsz, M.; Isreb, A.; Forbes, R.T.; Isreb, M.; Ahmed, W.; Arafat, T.; Alhnan, M.A. Tablet fragmentation without a disintegrant: A novel design approach for accelerating disintegration and drug release from 3D printed cellulosic tablets. Eur. J. Pharm. Sci. 2018, 118, 191–199. [Google Scholar] [CrossRef]
  204. Xu, W.; Molino, B.Z.; Cheng, F.; Molino, P.J.; Yue, Z.; Su, D.; Wang, X.; Willför, S.; Xu, C.; Wallace, G.G. On Low-Concentration Inks Formulated by Nanocellulose Assisted with Gelatin Methacrylate (GelMA) for 3D Printing toward Wound Healing Application. ACS Appl. Mater. Interfaces 2019, 11, 8838–8848. [Google Scholar] [CrossRef] [Green Version]
  205. Li, Q.; Guan, X.; Cui, M.; Zhu, Z.; Chen, K.; Wen, H.; Jia, D.; Hou, J.; Xu, W.; Yang, X.; et al. Preparation and investigation of novel gastro-floating tablets with 3D extrusion-based printing. Int. J. Pharm. 2018, 535, 325–332. [Google Scholar] [CrossRef] [PubMed]
  206. Shahbazi, M.; Jäger, H. Current Status in the Utilization of Biobased Polymers for 3D Printing Process: A Systematic Review of the Materials, Processes, and Challenges. ACS Appl. Bio Mater. 2021, 4, 325–369. [Google Scholar] [CrossRef] [PubMed]
  207. Li, N.; Qiao, D.; Zhao, S.; Lin, Q.; Zhang, B.; Xie, F. 3D printing to innovate biopolymer materials for demanding applications: A review. Mater. Today Chem. 2021, 20, 100459. [Google Scholar] [CrossRef]
  208. Li, L.; Zhu, Y.; Yang, J. 3D bioprinting of cellulose with controlled porous structures from NMMO. Mater. Lett. 2018, 210, 136–138. [Google Scholar] [CrossRef]
  209. Hausmann, M.K.; Siqueira, G.; Libanori, R.; Kokkinis, D.; Neels, A.; Zimmermann, T.; Studart, A.R. Complex-Shaped Cellulose Composites Made by Wet Densification of 3D Printed Scaffolds. Adv. Funct. Mater. 2020, 30, 1904127. [Google Scholar] [CrossRef]
  210. Hausmann, M.K.; Rühs, P.A.; Siqueira, G.; Läuger, J.; Libanori, R.; Zimmermann, T.; Studart, A.R. Dynamics of Cellulose Nanocrystal Alignment during 3D Printing. ACS Nano 2018, 12, 6926–6937. [Google Scholar] [CrossRef]
  211. Mulakkal, M.C.; Trask, R.S.; Ting, V.P.; Seddon, A.M. Responsive cellulose-hydrogel composite ink for 4D printing. Mater. Des. 2018, 160, 108–118. [Google Scholar] [CrossRef]
  212. Studart, A.R. Additive manufacturing of biologically-inspired materials. Chem. Soc. Rev. 2016, 45, 359–376. [Google Scholar] [CrossRef]
  213. Siqueira, G.; Kokkinis, D.; Libanori, R.; Hausmann, M.K.; Gladman, A.S.; Neels, A.; Tingaut, P.; Zimmermann, T.; Lewis, J.A.; Studart, A.R. Cellulose Nanocrystal Inks for 3D Printing of Textured Cellular Architectures. Adv. Funct. Mater. 2017, 27, 1604619. [Google Scholar] [CrossRef]
  214. Gauss, C.; Pickering, K.L.; Muthe, L.P. The use of cellulose in bio-derived formulations for 3D/4D printing: A review. Compos. Part C 2021, 4, 100113. [Google Scholar] [CrossRef]
  215. Roman, M.; Navarro, F. Deposition of Cellulose Nanocrystals by Inkjet Printing. In Model Cellulosic Surfaces; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 2009; Volume 1019, pp. 157–171. [Google Scholar]
  216. Li, V.C.F.; Mulyadi, A.; Dunn, C.K.; Deng, Y.; Qi, H.J. Direct Ink Write 3D Printed Cellulose Nanofiber Aerogel Structures with Highly Deformable, Shape Recoverable, and Functionalizable Properties. ACS Sustain. Chem. Eng. 2018, 6, 2011–2022. [Google Scholar] [CrossRef]
  217. Li, V.C.-F.; Kuang, X.; Mulyadi, A.; Hamel, C.M.; Deng, Y.; Qi, H.J. 3D printed cellulose nanocrystal composites through digital light processing. Cellulose 2019, 26, 3973–3985. [Google Scholar] [CrossRef]
  218. Amini, M.; Kamkar, M.; Ahmadijokani, F.; Ghaderi, S.; Rojas, O.J.; Hosseini, H.; Arjmand, M. Mapping 3D Printability of Ionically Cross-Linked Cellulose Nanocrystal Inks: Architecting from Nano- to Macroscale Structures. Biomacromolecules 2023, 24, 775–788. [Google Scholar] [CrossRef]
  219. Shin, S.; Hyun, J. Rheological properties of cellulose nanofiber hydrogel for high-fidelity 3D printing. Carbohydr. Polym. 2021, 263, 117976. [Google Scholar] [CrossRef]
  220. Ma, T.; Lv, L.; Ouyang, C.; Hu, X.; Liao, X.; Song, Y.; Hu, X. Rheological behavior and particle alignment of cellulose nanocrystal and its composite hydrogels during 3D printing. Carbohydr. Polym. 2021, 253, 117217. [Google Scholar] [CrossRef]
  221. Markstedt, K.; Sundberg, J.; Gatenholm, P. 3D Bioprinting of Cellulose Structures from an Ionic Liquid. 3D Print. Addit. Manuf. 2014, 1, 115–121. [Google Scholar] [CrossRef]
  222. Hopson, C.; Rigual, V.; Alonso, M.V.; Oliet, M.; Rodriguez, F. Eucalyptus bleached kraft pulp-ionic liquid inks for 3D printing of ionogels and hydrogels. Carbohydr. Polym. 2023, 313, 120897. [Google Scholar] [CrossRef] [PubMed]
  223. Gunasekera, D.H.A.T.; Kuek, S.; Hasanaj, D.; He, Y.; Tuck, C.; Croft, A.K.; Wildman, R.D. Three dimensional ink-jet printing of biomaterials using ionic liquids and co-solvents. Faraday Discuss. 2016, 190, 509–523. [Google Scholar] [CrossRef] [Green Version]
  224. Fedotova, V.S.; Sokolova, M.P.; Vorobiov, V.K.; Sivtsov, E.V.; Lukasheva, N.V.; Smirnov, M.A. Water Influence on the Physico-Chemical Properties and 3D Printability of Choline Acrylate—Bacterial Cellulose Inks. Polymers 2023, 15, 2156. [Google Scholar] [CrossRef] [PubMed]
  225. Fedotova, V.S.; Sokolova, M.P.; Vorobiov, V.K.; Sivtsov, E.V.; Ribeiro, M.C.C.; Smirnov, M.A. Synthesis and Physicochemical Properties of Acrylate Anion Based Ionic Liquids. Polymers 2022, 14, 5148. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of 1-allyl-3-methylimidazolium methyl phosphonate (Amim (MeO)PHO2) (A) (Reprinted with permission from ref. [76] Copyright © 2019, Elsevier, B.V.), 1,5-diazabicyclo [4.3.0] non-5-ene (DBN), pyridine N-oxide (PyO) and 2-picoline N-oxide (PiO) with graphic dissolution mechanism of cellulose (B) (Reprinted with permission from ref. [82] Copyright © 2021, Elsevier, B.V.), and benzethonium (BE)- and didecyldimethylammonium (DDA)-based ILs combined with short alkyl carboxylate anions (C) (Reprinted with permission from ref. [83] Copyright © 2021, Royal Society of Chemistry).
Figure 1. Chemical structures of 1-allyl-3-methylimidazolium methyl phosphonate (Amim (MeO)PHO2) (A) (Reprinted with permission from ref. [76] Copyright © 2019, Elsevier, B.V.), 1,5-diazabicyclo [4.3.0] non-5-ene (DBN), pyridine N-oxide (PyO) and 2-picoline N-oxide (PiO) with graphic dissolution mechanism of cellulose (B) (Reprinted with permission from ref. [82] Copyright © 2021, Elsevier, B.V.), and benzethonium (BE)- and didecyldimethylammonium (DDA)-based ILs combined with short alkyl carboxylate anions (C) (Reprinted with permission from ref. [83] Copyright © 2021, Royal Society of Chemistry).
Gels 09 00546 g001
Figure 2. Reaction of 1,1,3,3-tetramethyl guanidine (TMG) with methanol, ethanol, propanol, butanol, and ethylene glycol in the presence of CO2 and dimethyl sulfoxide (DMSO) to form reversible ILs (TMGHO2COR) (A). Pictures illustrate 5% w/w of microcrystalline cellulose (MCC) solution in the reversible ionic compound (RIC) solvents at the different molar fractions in DMSO after CO2 capture (XRIC). (Reprinted with permission from ref. [100] Copyright © 2014, Royal Society of Chemistry). Nonderivative and derivative approaches of the CO2 switchable 1,8-diazabicyclo [5.4.0] undec-7-ene (DBU) super base solvent system (B). Picture illustrates 10% w/w of MCC solution in superbase (DBU)/CO2/DMSO solvent. (Reprinted with permission from ref. [104] Copyright © 2019, American Chemical Society).
Figure 2. Reaction of 1,1,3,3-tetramethyl guanidine (TMG) with methanol, ethanol, propanol, butanol, and ethylene glycol in the presence of CO2 and dimethyl sulfoxide (DMSO) to form reversible ILs (TMGHO2COR) (A). Pictures illustrate 5% w/w of microcrystalline cellulose (MCC) solution in the reversible ionic compound (RIC) solvents at the different molar fractions in DMSO after CO2 capture (XRIC). (Reprinted with permission from ref. [100] Copyright © 2014, Royal Society of Chemistry). Nonderivative and derivative approaches of the CO2 switchable 1,8-diazabicyclo [5.4.0] undec-7-ene (DBU) super base solvent system (B). Picture illustrates 10% w/w of MCC solution in superbase (DBU)/CO2/DMSO solvent. (Reprinted with permission from ref. [104] Copyright © 2019, American Chemical Society).
Gels 09 00546 g002
Figure 4. Morphologies of cellulose hydrogels coagulated in different compositions of the regeneration bath. SEM images of cellulose hydrogels prepared from 5% w/w microcrystalline cellulose (MCC) in 1,5-diazabicyclo [4.3.0] non-5-ene (DBN)–CO2 solvent system and coagulated using water, methanol, ethanol, isopropanol, and no coagulating solvent (A1A5) (Reprinted with permission from ref. [104] Copyright © 2019, American Chemical Society). AFM images (B1B4) and photographs (C1C4) showing transparency of cellulose hydrogels prepared from 5% w/w MCC in 1,1,3,3-tetramethyl guanidine (TMG)/dimethyl sulfoxide (DMSO) solvent system and coagulated using ethanol (B1,C1), methanol (B2,C2), 5% w/w NaOH (B3,C3), and 5% w/w H2SO4 (B4,C4) aqueous solutions (Reprinted with permission from ref. [137] Copyright © 2021, American Chemical Society). SEM images of cellulose hydrogels prepared from 12% w/w in 1-ethyl-3-methylimidazolium diethyl phosphate (EmimDEP) and coagulated in water (D1), ethanol (D2), and mixed coagulating solvents containing ethanol:water:IL of 7:2:1 (D3) and 1:8:1 (D4) (Reprinted with permission from ref. [138] Copyright © 2022, Elsevier, B.V.).
Figure 4. Morphologies of cellulose hydrogels coagulated in different compositions of the regeneration bath. SEM images of cellulose hydrogels prepared from 5% w/w microcrystalline cellulose (MCC) in 1,5-diazabicyclo [4.3.0] non-5-ene (DBN)–CO2 solvent system and coagulated using water, methanol, ethanol, isopropanol, and no coagulating solvent (A1A5) (Reprinted with permission from ref. [104] Copyright © 2019, American Chemical Society). AFM images (B1B4) and photographs (C1C4) showing transparency of cellulose hydrogels prepared from 5% w/w MCC in 1,1,3,3-tetramethyl guanidine (TMG)/dimethyl sulfoxide (DMSO) solvent system and coagulated using ethanol (B1,C1), methanol (B2,C2), 5% w/w NaOH (B3,C3), and 5% w/w H2SO4 (B4,C4) aqueous solutions (Reprinted with permission from ref. [137] Copyright © 2021, American Chemical Society). SEM images of cellulose hydrogels prepared from 12% w/w in 1-ethyl-3-methylimidazolium diethyl phosphate (EmimDEP) and coagulated in water (D1), ethanol (D2), and mixed coagulating solvents containing ethanol:water:IL of 7:2:1 (D3) and 1:8:1 (D4) (Reprinted with permission from ref. [138] Copyright © 2022, Elsevier, B.V.).
Gels 09 00546 g004
Figure 5. Appearance and properties of cellulose hydrogels prepared by using ILs, e.g., 1-buthyl-3-methyl imidazolium chloride (BmimCl) and 1-allyl-3-methyl imidazolium chloride (AmimCl), and an addition of graphene oxide (AD), polypyrrole (PPy) (E,F), and liquid electrolyte (GJ). Images of composite cellulose hydrogels containing 5% w/w of cellulose and 0, 0.1, 1, and 10% w/w of graphene oxide before (A) and after (B) freeze-drying. The freeze-dried composite hydrogel supporting a counterpoise (C). Compressive stress–strain curves of the cellulose hydrogels containing 0 and 0.5% w/w of graphene oxide (D). (Reprinted with permission from ref. [160] Copyright © 2015, Elsevier, B.V.). Images (E) and compressive stress–strain curves (F) of the cellulose hydrogels containing 2.5% w/w of MCC and PPy. (Reprinted with permission from ref. [164] Copyright © 2014, Elsevier, B.V.). Image (G) and thermal properties (H,I) of cellulose hydrogel containing 4% w/w of cellulose (CAM-4). Electrolyte uptake and ionic conductivity of cellulose hydrogels containing 2–10% w/w of cellulose compared with those of Celgard 2400 commercial separator (J). (Reprinted with permission from ref. [165] Copyright © 2017, American Chemical Society).
Figure 5. Appearance and properties of cellulose hydrogels prepared by using ILs, e.g., 1-buthyl-3-methyl imidazolium chloride (BmimCl) and 1-allyl-3-methyl imidazolium chloride (AmimCl), and an addition of graphene oxide (AD), polypyrrole (PPy) (E,F), and liquid electrolyte (GJ). Images of composite cellulose hydrogels containing 5% w/w of cellulose and 0, 0.1, 1, and 10% w/w of graphene oxide before (A) and after (B) freeze-drying. The freeze-dried composite hydrogel supporting a counterpoise (C). Compressive stress–strain curves of the cellulose hydrogels containing 0 and 0.5% w/w of graphene oxide (D). (Reprinted with permission from ref. [160] Copyright © 2015, Elsevier, B.V.). Images (E) and compressive stress–strain curves (F) of the cellulose hydrogels containing 2.5% w/w of MCC and PPy. (Reprinted with permission from ref. [164] Copyright © 2014, Elsevier, B.V.). Image (G) and thermal properties (H,I) of cellulose hydrogel containing 4% w/w of cellulose (CAM-4). Electrolyte uptake and ionic conductivity of cellulose hydrogels containing 2–10% w/w of cellulose compared with those of Celgard 2400 commercial separator (J). (Reprinted with permission from ref. [165] Copyright © 2017, American Chemical Society).
Gels 09 00546 g005
Figure 6. Properties and efficiencies cellulose membranes prepared by dissolving bamboo cellulose (AC) and lemon peel cellulose mixed with raw lemon peel (D) in 1-buthyl-3-methyl imidazolium chloride (BmimCl) as the casting solution. Dynamic contact angle (A), zeta potential (B), and effectiveness (C) of membranes prepared from bamboo cellulose at the concentration of 3, 5, and 10% w/w. (Reprinted with permission from ref. [191] Copyright © 2020, American Chemical Society). Adsorption capacity of cellulose hydrogel (D) prepared by dissolving lemon peel cellulose without mixing raw lemon peel (LPMCCH) and with mixing the lemon peel of 0.2–0.6 g (LPMCC/LPH). Effects of pH (a), temperature (b), ionic type and strength (c), and adsorption/desorption cycles (d). (Reprinted with permission from ref. [192] Copyright © 2021, Elsevier, B.V.).
Figure 6. Properties and efficiencies cellulose membranes prepared by dissolving bamboo cellulose (AC) and lemon peel cellulose mixed with raw lemon peel (D) in 1-buthyl-3-methyl imidazolium chloride (BmimCl) as the casting solution. Dynamic contact angle (A), zeta potential (B), and effectiveness (C) of membranes prepared from bamboo cellulose at the concentration of 3, 5, and 10% w/w. (Reprinted with permission from ref. [191] Copyright © 2020, American Chemical Society). Adsorption capacity of cellulose hydrogel (D) prepared by dissolving lemon peel cellulose without mixing raw lemon peel (LPMCCH) and with mixing the lemon peel of 0.2–0.6 g (LPMCC/LPH). Effects of pH (a), temperature (b), ionic type and strength (c), and adsorption/desorption cycles (d). (Reprinted with permission from ref. [192] Copyright © 2021, Elsevier, B.V.).
Gels 09 00546 g006
Table 1. Recent studies on IL/DMSO solvent systems used for cellulose dissolution.
Table 1. Recent studies on IL/DMSO solvent systems used for cellulose dissolution.
ILConcentration of DMSO in IL and Temperature to Dissolve CelluloseConcentration of CelluloseRef.
C3OMeImAc60% mole, 60 °C12% w/w MCC a[91]
C4MeImAc60% mole, 60 °C16% w/w MCC a
(BmimAc)
C3OMe2ImAc60% mole, 60 °C19% w/w MCC a
C4Me2ImAc60% mole, 60 °C22% w/w MCC a
BmimAc20% w/w, 70 °C8% w/w MCC, Avicel, and α-cellulose b[93]
40% w/w, R.T. c20% w/w Avicel b[94]
40% w/w, R.T. c~10% w/w MCC b[95]
75% w/w, 80 °C14% w/w cellulose powder b[96]
BmimCl50% w/w, 15–75 °C3% w/w cellulose powder b[92]
30% w/w, 120 °C-[97]
EmimAc50% w/w, 60 °C2.5% w/w cotton b[98]
25% w/w, 50 °C10% w/w MCC b[99]
75% w/w, 80 °C14% w/w cellulose powder b[96]
EmimCl75% w/w, 80 °C14% w/w cellulose powder b[96]
a solubility test. b fixed amount. c room temperature (~25 °C).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Taokaew, S. Recent Advances in Cellulose-Based Hydrogels Prepared by Ionic Liquid-Based Processes. Gels 2023, 9, 546. https://doi.org/10.3390/gels9070546

AMA Style

Taokaew S. Recent Advances in Cellulose-Based Hydrogels Prepared by Ionic Liquid-Based Processes. Gels. 2023; 9(7):546. https://doi.org/10.3390/gels9070546

Chicago/Turabian Style

Taokaew, Siriporn. 2023. "Recent Advances in Cellulose-Based Hydrogels Prepared by Ionic Liquid-Based Processes" Gels 9, no. 7: 546. https://doi.org/10.3390/gels9070546

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop