Open Access Article
Md Maksudur Rahmanab,
Hemant Choudhary
ac,
Blake A. Simmons
ad,
John M. Gladden
ab and
Alberto Rodriguez
*ab
aJoint BioEnergy Institute, 5885 Hollis Street, Emeryville, CA 94608, USA. E-mail: alberto@lbl.gov
bBiomaterials and Biomanufacturing, Sandia National Laboratories, 7011 East Avenue, Livermore, CA 94550, USA
cBioresource and Environmental Security, Sandia National Laboratories, 7011 East Avenue, Livermore, CA 94550, USA
dBiological Systems and Engineering, Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, CA 94720, USA
First published on 7th November 2025
Ionic liquids (ILs) are promising solvents for the pretreatment of lignocellulosic biomass due to their ability to disrupt cellulose, hemicellulose, and lignin structures. However, large-scale implementation requires the development of efficient recovery and recycling methods. This review provides a comprehensive analysis of the recyclability potential of ILs used in biomass pretreatment, emphasizing their mechanisms, recent innovations, and ongoing challenges. We begin by discussing the structural diversity and tunability of ILs, which underlie their effectiveness in biomass deconstruction. The distinct roles of IL anions and cations in dissolving specific biomass components are systematically presented and compared. Advances in IL recycling techniques, including antisolvent precipitation methods, membrane separation, and distillation, are critically examined, with attention to how mechanistic insights can inform the design of more efficient and selective recovery strategies. Despite progress, significant challenges remain to scaling up IL-based biomass processing, including high cost, environmental concerns, and impact of biomass-derived impurities (e.g., lignin residues, sugars, proteins) on IL purity and functionality after reuse. We also review the applicability of different ILs based on life cycle assessments and techno-economic analyses. Lastly, we identify critical research gaps and propose future directions, including the design and development of next-generation ILs with improved recyclability, reduced toxicity, and enhanced economic viability for industrial-scale applications.
Green foundation1. We discuss recent advances in the design of ionic liquids with properties that allow for their recycling in a biorefinery setting. We also provide examples of recovery techniques with the potential for reducing the cost and environmental impact of using ionic liquids at large scales.2. Efficient and sustainable biomass pretreatment is one of the biggest bottlenecks to produce bioproducts. Given the growing push towards decarbonizing the chemical and energy sectors, this topic may be relevant to researchers, policymakers and industries aiming to transition towards circular and bio-based economies. 3. Future efforts in this field will focus on developing biomass conversion processes that are cost-effective and scalable. This review contributes to that goal by summarizing the state of the art and providing a systematic assessment of the unique advantages, challenges and opportunities of using recyclable ionic liquids as biomass pretreatment solvents. |
![]() | ||
| Fig. 1 Representation of the structure of lignocellulosic biomass. Panels: (a) cellulose microfiber; (b) macromolecular polymer with linear configuration; (c) hemicellulose; (d) lignin; (e) constitutional units of p-hydroxyphenyl (H), guaiacyl (G), and syringyl (S); (f) prevalent C–C and C–O bonds found in lignin. Adapted from ref. 18. Copyright 2022, Elsevier. | ||
Despite their potential, the widespread industrial implementation of ILs faces significant challenges, including their high production costs primarily due to energy-intensive recovery and purification processes, inhibitory effects on enzymatic processes, potential environmental toxicity, and scalability issues.34,46 Additionally, life cycle assessments (LCA) and techno economic analyses (TEA) have revealed that, despite their advantages, ILs can have a higher eco-toxicity impact than conventional solvents unless efficient recovery and recycling strategies are implemented.46,47 To address these limitations researchers have responded by proposing innovative recovery techniques involving integrated TEA-based and sustainability-focused frameworks. Research on IL-based biorefining processes like the Ionosolv technology underscore their transformative potential while emphasizing the need for continued research to optimize economic and environmental performance.36,48–50
It is worth noting that ILs remained relatively unnoticed until the early 21st century when researchers and chemical companies recognized their exceptional properties compared to conventional solvents. Since then, there has been a growing interest in developing diverse applications for ILs. The number of science citation index (SCI) publications on ILs has risen exponentially, from just a few in 1996 to over 5000 by 2018, surpassing the annual growth rates of many other popular scientific fields.51,52 Additionally, their practical applications are expanding, as highlighted by Morton and Hamer's patent analysis, which shows an increasing variety of uses for ILs.53 The rapid rise in the use of ILs for bioenergy purposes (Fig. 2) reflects their significant promise in advancing green chemistry, driving sustainable technologies, and enabling applications in biomass valorization and beyond.
![]() | ||
| Fig. 2 Trend in the number of publications involving ILs for biomass pretreatment (black line), IL recycling after pretreatment (violet), and review articles on ILs (blue) over the past two decades. | ||
Over the past decade, several review articles have explored the role of ILs in biomass processing, primarily focusing on their role in biomass deconstruction by enabling lignin separation and enhancing enzymatic hydrolysis (Table 1).50,54–58 However, a comprehensive discussion of ionic liquid mechanisms, recyclability, and sustainability has not been published to date. Specifically, there is a critical need to consolidate insights on the dissolution mechanisms of cellulose, hemicellulose, and lignin, the role of anions and cations in biomass deconstruction, and the implications of biomass-derived impurities on the functionality of the recovered ILs. Additionally, the optimization of recycling techniques such as water- and salt-based methods, membrane-based systems, and co-solvent approaches and their economic and environmental implications require a critical assessment to understand the key challenges that remain to be addressed.16,59,60
| Author (s) | Review article | Key issue(s) discussed | Year |
|---|---|---|---|
| Annegret Stark54 | Ionic liquids in the biorefinery: a critical assessment of their potential | Potential of ionic liquids in an integrated biorefinery system for biomass processing, discussing both the benefits and potential risks | 2011 |
| Azmat Mehmood Asim et al.55 | Acidic ionic liquids: promising and cost-effective solvents for processing of lignocellulosic biomass | Cost-effectiveness of acidic ionic liquids, challenges in recycling and environmental impact | 2019 |
| Florence J. V. Gschwend et al.61 | Quantitative glucose release from softwood after pretreatment with low-cost ionic liquids | High glucose yield from softwood using low-cost ionic liquids | 2019 |
| Nana Yamaki et al.62 | Thermal hazard analysis of a biomass pretreatment process using ionic liquids | Thermal safety concerns in using ionic liquids in industrial processes | 2019 |
| Zeba Usmania et al.56 | Ionic liquid-based pretreatment of lignocellulosic biomass for enhanced bioconversion | Enhanced bioconversion using ionic liquid pretreatment, focus on green chemistry | 2020 |
| Niloofar Nasirpour et al.51 | Ionic liquids: promising compounds for sustainable chemical processes and applications | Sustainability of ionic liquids in chemical processes, industrial scale application | 2020 |
| Isa Hasanov et al.57 | The role of ionic liquids in the lignin separation from lignocellulosic biomass | Efficient lignin separation using protic ionic liquids | 2020 |
| Rajagopal Malolan et al.58 | Green ionic liquids and deep eutectic solvents for desulphurization, denitrification, biomass, biodiesel, bioethanol and hydrogen fuels: a review | Green ionic liquids as viable solutions for fuel processing and bioethanol production | 2021 |
| Hakim Hebal et al.63 | Activity and stability of hyperthermostable cellulases and xylanases in ionic liquids | Enzyme stability in ionic liquids for biomass degradation | 2021 |
| Kuan Shiong Khoo et al.50 | How does ionic liquid play a role in sustainability of biomass processing? | Sustainability and techno-economic challenges in using ionic liquids for biomass processing | 2021 |
| Jinxu Zhang et al.64 | Recent developments in ionic liquid pretreatment of lignocellulosic biomass for enhanced bioconversion | Technological hurdles in ionic liquid pretreatment processes with a focus on biocompatibility | 2021 |
| Xiaoqi Lin et al.18 | Review on development of ionic liquids in lignocellulosic biomass refining | Summarizes the role of ionic liquids in pretreatment of lignocellulosic biomass and highlights challenges in achieving sustainable conversion | 2022 |
| Xiaofang Liu et al.65 | The development of novel ionic liquid-based solid catalysts and the conversion of 5-hydroxymethylfurfural from lignocellulosic biomass | Focus on novel IL-based solid catalysts for the conversion of lignocellulosic biomass to platform chemicals like HMF | 2022 |
| Jikai Zhao, Juhee Lee, Donghai Wang16 | A critical review on water overconsumption in lignocellulosic biomass pretreatment for ethanol production through enzymic hydrolysis and fermentation | Water overconsumption during biomass pretreatment for ethanol production, need for water-efficient methods in industrial processing | 2023 |
| Mohammad Eqbalpour et al.66 | A comprehensive review on how ionic liquids enhance the pyrolysis of cellulose, lignin, and lignocellulose toward a circular economy | Role of ILs in enhancing the pyrolysis of cellulose, lignin, and lignocellulose for circular economy and sustainable energy applications | 2023 |
| Nazife Isik Haykir et al.35 | Applications of ionic liquids for the biochemical transformation of lignocellulosic biomass into biofuels and biochemicals: a critical review | Reviews the role of ILs in biofuel and biochemical production, emphasizing technical and economic challenges | 2023 |
| Francieli Colussi et al.10 | Challenges in using ionic liquids for cellulosic ethanol production | Challenges in pretreating lignocellulosic biomass with ionic liquids for cellulosic ethanol production, including cost and process efficiency | 2023 |
| Paul Wolski et al.67 | Factors that influence the activity of biomass-degrading enzymes in the presence of ionic liquids: a review | Examines the interaction of ionic liquids with biomass-degrading enzymes, with a focus on the stability and activity of these enzymes | 2023 |
| Ting He et al.60 | Latest advances in ionic liquids promoted synthesis and application of advanced biomass materials | Discusses advances in using ILs to produce advanced biomass-based materials, highlighting their application as solvents and modifiers | 2023 |
| Ruolin Li59 | Recent advances in biomass pretreatment using biphasic solvent systems | Biphasic solvent systems, including ILs, for efficient fractionation of biomass and enhancing enzymatic hydrolysis | 2023 |
| A. S. Norfarhana et al.68 | Revolutionizing lignocellulosic biomass: a review of harnessing the power of ionic liquids for sustainable utilization and extraction | Focus on ILs for fractionating lignocellulosic biomass, highlighting green and sustainable methods for cellulose, hemicellulose, and lignin extraction | 2024 |
| Kosuke Kuroda69 | Bioethanol fermentation in the presence of ionic liquids: mini review | Challenges of microbial fermentation in the presence of toxic ionic liquids, need for low-toxicity ILs and IL-resistant microorganisms | 2024 |
| Pedro Verdía Barbará et al.70 | Recent advances in the use of ionic liquids and deep eutectic solvents for lignocellulosic biorefineries and biobased chemical and material production | Current state of lignocellulosic biomass processing using ILs and DESs, focusing on pretreatment chemistry, process flows, and product streams, followed by sustainability assessments and key technological challenges | 2025 |
Protic and aprotic ILs are synthesized by substituting heteroatoms in the cation with protons or alkyl groups, respectively. Protic ILs, which are formed via proton transfer from a Brønsted acid to a Brønsted base, are characterized by extensive hydrogen-bonding networks due to the presence of proton donor and acceptor groups. In contrast, aprotic ILs lack this extensive hydrogen-bonding capability, differentiating them from their protic counterparts.78,79 However, the basic structure of ILs involves an asymmetry between the cation and anion, where the cation is generally a large, organic molecule (e.g., imidazolium, pyridinium), and the anion ranges from simple halides (i.e., Cl−, Br−) to complex fluorinated structures like bis(trifluoromethylsulfonyl)imide [Tf2N]−.80,81
This asymmetry reduces the effective coulombic interactions between the ions, lowering the lattice energy and enabling these salts to remain liquid at temperatures below 100 °C, distinguishing them from conventional salts.73 The cationic component often features alkyl chains or aromatic groups that can influence the physicochemical properties of the ILs. For instance, imidazolium-based cations are known for their tunable solubility and thermal stability. The anion, on the other hand, plays a significant role in determining properties like hydrophilicity, polarity, and overall ionic strength.82–84 The structure of ILs is influenced by the electrostatic attractions and steric packing, which contribute to their liquid-like ordering. This ordering is evident in the pair correlation functions observed in molecular dynamics simulations.85 The presence of a pre-peak in the static structure factor indicates a geometric effect from the packing of alkyl chains, highlighting the importance of the molecular structure in determining the properties of ILs.85
The interaction between ions, including the magnitude of attraction and conformational flexibility, play crucial roles in determining the transport properties and stability of ILs86 while computational methods such as molecular dynamics simulations are essential for understanding these interactions and the resulting liquid structures. Studies using molecular dynamics simulations and X-ray scattering have shown that ILs have distinct pair correlation functions, indicating specific spatial arrangements of ions within the liquid.85,87 The structural organization of ILs can change with temperature, as seen in phase transition from solid to liquid phase which is often studied using techniques like X-ray diffraction and Raman spectroscopy.88 However, the choice of ions directly impacts the physical and chemical properties of IL, such as viscosity, conductivity and solubility. However, ILs are classified into three generations which are typically based on the evolution in design, functionality and application focus, as depicted in Fig. 4.
![]() | ||
| Fig. 4 Properties of ILs and examples of corresponding chemical structures employed over the generations. Panels: (a) first generation; (b) second generation; (c) third generation. Adapted from ref. 73. Copyright 2020, Universal Wiser. | ||
First generation ILs have unique physical properties and are focused on simple ionic salts (e.g., chloroaluminate based with simple cations like imidazolium or pyridinium) used mainly in electrochemistry (e.g., electrolytes in batteries) but highly sensitive to moisture, air instability and corrosiveness. Second generation ILs are more stable and less reactive, which have been designed to overcome the limitation of the first generation and expand their application as green solvents.89 The third generation of ILs introduced task specific functionalities tailored for targeted applications (e.g., biomass processing, catalysis, CO2 capture) and enhanced functionalities (i.e., pharmaceutical synthesis, and electrochemical devices).
While third generation ILs offer task specific functionalities, they often rely on synthetic, non-renewable, and sometimes toxic components, which can lead to ecological concerns during disposal or accidental release. Additionally, the limited biodegradability of many existing ILs constrains their applications where environmental safety is paramount, such as biomedicine and environmental remediation.73 Therefore, to address those concerns and fill the shortcomings of previous generations of ILs, significant efforts should be directed towards developing a new class of ILs with enhanced biodegradability, non-toxicity, and renewable sourcing features. This involves employing natural and renewable components such as amino acids, carbohydrates, and organic acids, moving away from synthetic, halogenated compounds that can pose serious environmental and health risks.
![]() | ||
| Fig. 5 Structural feature of cellulose: (a) general form of cellulose (i.e., cellulose microfibril); (b) cross-section view of 36 chain cellulose elementary microfibril; (c) H-bond network between cellulose chains. Adapted from ref. 92. Copyright 2017, Royal Society of Chemistry. | ||
| Techniques | Description | Insights |
|---|---|---|
| Molecular dynamics (MD) simulation | Computational studies analyzing solute–solvent interaction at molecular levels | Visualize anion and cation interactions with cellulose surface |
| Nuclear magnetic resonance (NMR) | Monitors changes in chemical environments during cellulose dissolution | Identifies hydrogen bonding disruption and chemical shifts |
| Fourier transform infrared spectroscopy (FTIR) | Track changes in functional groups and hydrogen bonding networks | Indicates disruption of crystalline hydrogen bonds |
| Scanning electron microscopy (SEM) | Visualize surface morphology of cellulose before and after dissolution | Shows structural changes and fibril disintegration |
| X-ray diffraction (XRD) | Measures crystallinity index changes during pretreatment | Evaluates reduction in cellulose crystallinity |
| Raman spectroscopy | Identifies molecular vibrations and interaction | Highlights structural transformations in cellulose |
| Thermal analysis (DSC, TGA) | Examines thermal stability and phase transition during dissolution | Provides insights into thermal effects on cellulose crystallinity |
To disrupt this robust network of cellulose, several ILs have been used as solvents91–95 to enable homogenous processing under relatively mild conditions (>150 °C). The dissolution of cellulose in ILs primarily hinges on their ability to disrupt the dense hydrogen-bonding that stabilizes the cellulose structure. This disruption is facilitated by the ILs anion, which acts as a strong hydrogen bond acceptor, forming interactions with the hydroxyl groups on cellulose chains. Swatloski et al.38 first demonstrated that 1-butyl-3-methylimidazolium chloride [BMIM]Cl could dissolve cellulose by breaking these hydrogen bonds, highlighting the critical role of small, basic anions such as (Cl−) and (CH3COO−) in the dissolution process. Subsequent research reinforced that the anion's basicity and size are pivotal. For instance, [EMIM][CH3COO] exhibits superior dissolution capability (up to 16 wt%) due to acetate's strong hydrogen bond acceptor nature and the reduced viscosity compared to halide-based ILs.96 Notably, anions such as acetate (CH3COO−) and chloride (Cl−) play a crucial role by forming strong hydrogen bonds with the hydroxyl groups of cellulose, thereby destabilizing the crystalline structure. Concurrently, the cation also contributes by influencing viscosity and steric accessibility; imidazolium-based cations bearing acidic protons enhance solubility by interacting with cellulose's oxygen atoms. For example, 1-ethyl-3-methylimidazolium [EMIM]+ interacts with the hydrophobic surfaces of cellulose fibrils, preventing reaggregation and promoting solvation. This synergistic action facilitates the separation of cellulose chains and enhances enzymatic accessibility.92,93,95
Mechanistically, the dissolution process is believed to occur by the formation of complexes between IL anions and cellulose hydroxyl groups, weakening the extensive hydrogen bonding within and between cellulose chains, followed by polymer chain separation and solvation by the IL medium. Some studies suggest that co-solvents like DMSO can further accelerate dissolution by reducing viscosity and improving mass transfer, although care must be taken to ensure sufficient anion availability.97,98 In terms of structural transitions, ILs can convert crystalline cellulose I into cellulose II or amorphous cellulose, enhancing accessibility for further chemical or enzymatic transformation. Przypis et al.93 highlights that crystalline to amorphous conversion is essential for increasing cellulose reactivity and solubility. Moreover, the ability of ILs to dissolve cellulose without derivatization under mild conditions, positions them as promising solvents.
Despite its potential, the high viscosity of certain ILs, cost of production, and challenges in recyclability limit scalability, necessitating further research to optimize processes and materials. Additionally, developing cost-effective, biodegradable ILs with high recyclability remains a priority.92,93 Addressing these challenges could revolutionize biomass pretreatment by enabling more efficient, sustainable, and economically viable processes. Future research should focus on optimizing solvent systems, understanding solvent–cellulose interactions at the molecular level, and integrating dissolution studies with downstream conversion processes. These approaches would enhance the feasibility of cellulose-based biorefineries, paving the way for widespread adoption of renewable bioresources.92,93
![]() | ||
| Fig. 6 Primary structure of D-xylo–D-glucan. Adapted from the ref. 102. Copyright 2006, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. | ||
Chemically, hemicellulose is more soluble in alkaline solutions than cellulose and can be partially solubilized under mild pretreatment conditions. However, the variability influences the choice of pretreatment methods and conversion strategies in biomass processing. The dissolution of hemicellulose, a fundamental step in converting biomass into value-added products, is governed by the intricate interplay of chemical and physical processes that deconstruct the hemicellulose matrix. To date, various pretreatment methods (see Table 3) have been employed to deconstruct the hemicellulose in biomass.
| Method | Key mechanism | Efficiency (%) | Environmental impact | Ref. |
|---|---|---|---|---|
| Oxidative torrefaction | Thermal decomposition, furan release | Moderate | Moderate | 103 |
| Hydrothermal pretreatment | Hot water hydrolysis | 81.59 | Low | 104 |
| Surfactant-assisted hydrolysis | Surface tension reduction | 54.70 | Low to moderate | 105 |
| Ionic liquid pretreatment | Hydrogen bond disruption | High | Low | 101 |
| Acidic electrolyzed water | Selective cleavage with FeCl3 | 98.64 | Low | 106 |
Oxidative torrefaction thermally decomposes hemicellulose under controlled oxygen conditions, generating volatile compounds like furans and acids,103 while hydrothermal pretreatment employs liquid hot water to hydrolyze hemicellulose, achieving dissolution rates up to 81.59% under optimized conditions.104 Surfactant-assisted acid hydrolysis enhances dissolution by reducing surface tension and promoting ion diffusion, as demonstrated by increased hemicellulose removal from poplar wood chips.105
IL pretreatments disrupt the hydrogen bonding network in hemicellulose through the interaction of IL anions, achieving efficient dissolution with the added benefit of solvent recovery.101 Zhao et al.101 studied the dissolution of hemicellulose in ILs using computational methods, specifically the Conductor-like Screening Model for Real Solvents (COSMO-RS) to accurately predict the solubility of hemicellulose in ILs. Six hemicellulose models were evaluated (see Fig. 7), with the mid-dimer xylan chain (MDXC) emerging as the most suitable due to its alignment with experimental solubility data. A total of 1368 ILs were screened, emphasizing the role of hydrogen bonding, particularly the anion's acceptor capacity, in enhancing hemicellulose solubility. Smaller anions with strong hydrogen-bonding capabilities, such as chloride and acetate, were found to be most effective. Additionally, the study revealed that longer alkyl chains on cations negatively impact dissolution. The s-profile, a COSMO-RS-derived property, was used to analyze molecular interactions, providing insights into solubility patterns and enabling high-throughput screening of ILs. The findings highlight the potential of ILs as green solvents for hemicellulose dissolution, paving the way for improved lignocellulosic biomass processing.
![]() | ||
| Fig. 7 COSMO-RS charge surfaces of six hemicellulose models: Model 1: xylan; Model 2: mid-monomer of xylan chain; Model 3: mid-dimer of xylan chain; Model 4: mid-trimer of xylan chain; Model 5: mid-tetramer of xylan chain; Model 6: mid-pentamer of xylan chain. (Red zones indicate a positive surface charge, yellow and green zones indicate almost neutral charges, navy blue designate negative surface charge.) Adapted from the ref. 101. Copyright 2022, Royal Society of Chemistry. | ||
The MDXC (mid-dimer xylan chain) model displayed balanced charge interactions, making it most representative of hemicellulose behavior in solution, which is crucial as accurate charge distribution predicts the molecular interactions, especially hydrogen bonding, with ILs. Acidic and catalytic pretreatments, such as FeCl3 coupled with acidic electrolyzed water, selectively remove hemicellulose while preserving cellulose and lignin, reaching a remarkable 98.64% removal rate in poplar wood.106 While each method demonstrates distinct advantages, their integration into sustainable biorefinery frameworks remains a critical avenue for advancing hemicellulose valorization while minimizing environmental impacts.
![]() | ||
| Fig. 8 Typical structure and interactions of lignin. Panels: (a) structure of lignin; (b) linkages of lignin. Adapted from ref. 107. Copyright 2017, Hindawi. | ||
Within biomass, lignin forms a structural matrix with polysaccharides, contributing to the mechanical strength of plant tissues and enhancing their resistance to chemical and enzymatic degradation.93,108,109 Recent studies have explored different solvent systems, including ILs, deep eutectic solvents (DESs), and other innovative approaches, to understand and enhance lignin dissolution. These studies reveal that the dissolution process is primarily driven by specific interactions such as hydrogen bonding, π–π stacking, and electrostatic interactions between lignin and the solvent components. DESs, such as those composed of choline chloride and lactic acid, are effective in dissolving lignin due to their ability to form hydrogen bonds with lignin molecules. The addition of co-solvents like ethylene glycol or γ-valerolactone (GVL) enhances these interactions by altering hydrogen bond strength and electrostatic potential, leading to improved lignin separation and dissolution efficiency.
Zhang et al. (2024)110 explores novel ternary DESs combining choline chloride, lactic acid, and additives like ethylene glycol or GVL for lignin dissolution. These DESs showed enhanced delignification efficiency, retaining lignin's structural integrity while reducing molecular weight heterogeneity and improving antioxidant properties. Liu et al. (2020)109 investigates alkali lignin dissolution using lactic acid–choline chloride DES. The mechanism involves lignin depolymerization through β-O-4 bond cleavage, facilitated by lactic acid's acidity and interactions between lignin phenolic groups and choline chloride. They explored the role of temperature in modifying the molecular weight distribution and structure of lignin, offering insights for valuable applications. ILs are known for their ability to dissolve lignin through a combination of hydrogen bonding and π-stacking interactions. The anion–cation pairs in ILs play a crucial role, with anions forming strong hydrogen bonds with lignin, while cations engage in π-stacking interactions.111,112 Zhang et al. (2017)111 employed Density Functional Theory (DFT) to elucidate the molecular mechanisms underlying lignin dissolution in imidazolium-based ILs. They identified hydrogen bonding between anions (chloride and acetate) and lignin as the dominant interaction driving dissolution, with anions preferentially binding to hydroxyl groups on lignin's β-O-4 model structure. Cations play a complementary role through π-stacking interactions with aromatic lignin units. These findings highlight the potential of engineering IL components to optimize specific interactions, paving the way for more effective lignin solubilization strategies.
Wang et al. (2017)112 investigated lignin dissolution in ILs, emphasizing hydrogen bonding by anions (e.g., chloride, acetate) and π-stacking by cations and demonstrate the interplay between lignin structure and IL components, offering insights to optimize ILs for enhanced lignin solubility and sustainable biorefinery processes. In another example, Mohan et al. (2023)113 integrated multiscale molecular simulations and the COSMO-RS model to systematically screen 5670 ILs for lignin dissolution efficiency (see Fig. 9). By analyzing key thermodynamic parameters, such as activity coefficients and excess enthalpy, they discovered ILs containing acetate, glycinate, and lysinate anions paired with tetraalkylammonium and pyridinium cations as optimal.
![]() | ||
| Fig. 9 Graphical representation of the lignin activity coefficient, enthalpy and solubility parameters by the COSMO-RS model. Panels: (a) logarithmic activity coefficients (ln(γ)) of lignin in different ILs at 363.15 K; (b) excess enthalpy, HE (kcal mol−1) of lignin; (c) the COSMO-RS-based predicted Hansen Solubility Parameters of ILs. COSMO quick tool version 1.7 is used to predict the HSP of ILs. Adapted from ref. 113. Copyright 2023, Springer Nature. | ||
Xue et al. (2016)114 studied binary solvent systems, particularly those utilizing GVL combined with co-solvents such as water or ILs, exhibit exceptional efficiency in lignin dissolution. The solubility of lignin in these systems is primarily influenced by the hydrogen bond basicity of the solvent, which varies depending on the type of lignin and solvent employed.114 GVL–water mixtures, for example, demonstrate a remarkable ability to dissolve dealkaline lignin (DAL) through enhanced hydrogen bonding interactions, maintaining lignin's structural integrity while enabling sustainable delignification under mild conditions. Additionally, the inclusion of water in these systems significantly enhances the cleavage of intermolecular linkages within lignin, thereby facilitating its dissolution and depolymerization.115
Cations, though secondary in direct interaction with biomass, modulate the physical properties of ILs, including viscosity and steric effects, which indirectly influence dissolution efficiency. Imidazolium-based cations like [EMIM]+ (1-ethyl-3-methylimidazolium) and [AMIM]+ (1-allyl-3-methylimidazolium) are widely used, as their unsaturated structures reduce steric hindrance, enhancing biomass solubility.92,116 For example, ILs such as [EMIM][CH3COO] effectively disrupt cellulose crystallinity, while [AMIM][HCOO] demonstrates superior biomass solubility due to stronger anion–polymer interactions.116,119 Schutt et al. (2016)120 revealed that cations can influence the activity of enzymes used in biomass processing. As a consequence, modifications to the cation structure, such as adding oxygen atoms, can improve mass transport properties and reduce enzyme deactivation, thereby enhancing the overall efficiency of biomass deconstruction.120
Innovative systems combining multiple ions, such as cholinium and lysinate or acetate mixtures, exhibit synergistic effects that enhance cellulose dissolution.92 It has been revealed that the combination of ions enhance both solubility and enzymatic hydrolysis while maintaining biocompatibility for downstream processing.121
Although the role of anions and cations in biomass deconstruction is well-documented, challenges remain in optimizing these interactions for large scale/industrial applications. The cost and recyclability of ILs, as well as their impact on the properties of recovered materials, are critical factors that need to be addressed for commercial viability.122,123 Additionally, the presence of water in IL systems can reduce their solvating power, although additives like lithium chloride can mitigate this effect.124 These insights highlight the complexity and potential of ILs in biomass processing, suggesting avenues for future research and development.
Despite these advancements, challenges persist, such as scalability and cost-efficiency, highlighting the need for integrated recycling approaches that combine multiple techniques. We envision a simplified framework for selecting a suitable IL recovery method subsequent to biomass pretreatment as a decision-making scenario, based on current knowledge (Fig. 10). These efforts not only support the circular economy but also ensure the sustainable use of ILs in green chemistry and beyond. The following section will systematically articulate recent advancements of recycling techniques applied to ILs.
![]() | ||
| Fig. 10 Main categories of ILs in biomass pretreatment and related recycling techniques. Panels: (a) scenario-1; (b) scenario-2; (c) scenario-3; (d) scenario-4. | ||
![]() | ||
| Fig. 11 IL recovery process comparison, operational setup, and a cyclic process flowchart. Panels: (a) options for antisolvent recovery and associated utilities. Recovery by distillation requires a heat duty, Q, while the electrodialysis process requires electrical work, W. (b) A concept that uses electrodialysis for antisolvent recovery. The sketch illustrates a section of an electrodialysis cell, where a mixture of antisolvent, water and salt enter in the feed channel and salt, and water enters in the draw channel. The two feed channels are separated by cation-exchange membrane and anion-exchange membranes. (c) Flowchart detailing the compositions of an example cyclic process involving antisolvent recycling by electrodialysis. Adapted from ref. 128. Copyright 2024, Elsevier. | ||
ABS employs salts such as Na2CO3 or MgSO4 to create two immiscible aqueous phases, with ILs preferentially partitioning into the IL-rich phase. This method is particularly effective in recovering ILs from complex matrices such as plant extracts, achieving nearly quantitative recovery rates when the salt type and concentration are optimized.129,130 Similarly, ASC utilizes the addition of antisolvents like methanol to selectively precipitate salts from aqueous IL solutions. This approach enables the separation of inorganic salts from saline effluents or IL mixtures, offering a green and cost-effective solution. For instance, ASC has been used to recover sodium sulfate from saline effluents, with a focus on solvent recovery to minimize environmental impact.131
Despite their advantages, these methods face challenges such as membrane fouling in electrodialysis, salt solubility limitations in ABS, and solvent loss in ASC. Hybrid approaches combining these techniques have shown promise in addressing these limitations. For example, integrating ABS with electrodialysis enhances IL recovery efficiency while reducing energy consumption.128,131 Another advanced hybrid design merges aqueous two-phase extraction, membrane separation, and distillation to recover hydrophilic ILs from dilute aqueous solutions, achieving substantial reductions in both total annual costs and energy consumption compared to standalone distillation processes.132 Additionally, incorporating salting-out agents like (NH4)2SO4 significantly boosts process efficiency, achieving energy savings of up to 91% in certain scenarios.132 Table 4 summarizes the key features of these methods, highlighting their energy efficiency, advantages, and challenges. Continued research into the thermodynamic properties of ILs in aqueous systems and the development of advanced materials for membranes and solvents can further optimize these processes, aligning them with green chemistry principles and fostering their broader adoption.
| Method | Energy efficiency | Key advantages | Challenges |
|---|---|---|---|
| Electrodialysis | 60–200 kWh m−3 | High selectivity, low energy use | Membrane fouling and cost |
| ABS | High (near complete) | Simultaneous IL and solute recovery | Salt concentration dependency |
| ASC | Moderate | Selective salt recovery, green process | Solvent loss, antisolvent recycling challenges |
This review underscores the transformative potential of water- and salt-based recovery strategies in advancing sustainable industrial applications. Exploring hybrid methodologies and refining process conditions hold the key to unlocking their full potential.128,129,131 While these methods offer promising solutions for the recovery of ILs and salts, several challenges remain. The selection of appropriate antisolvents and salts is crucial for optimizing recovery efficiency and minimizing environmental impact. Additionally, the integration of these methods into existing industrial processes requires careful consideration of economic and technical feasibility. Further research is needed to address these challenges and develop more robust and scalable recovery systems.
Pervaporation offers exceptional efficiency and selectivity for IL recovery. For instance, the dehydration and recycling of 1-ethyl-3-methylimidazolium acetate [EMIM][CH3COO] from aqueous solutions achieved a remarkable separation factor of 1500, recovering over 99.9 wt% of the IL. Compared to other methods, pervaporation exhibited the highest IL recovery (>99 wt%) and negligible losses (0.02–0.04 wt%) compared to vacuum distillation and electrodialysis, which showed higher IL losses and lower final IL concentrations, as shown in Table 5. These results highlight the superior efficiency and selectivity of pervaporation for IL dehydration. This capability underscores its potential to improve the economics of biorefineries by reducing IL losses and operational costs.135
| Techniques | IL : H2O (w/w) |
t/P (h kPa−1) | Final IL concentration (wt%) (°C) | IL loss (wt%) |
|---|---|---|---|---|
| PV: pervaporation, VD: vacuum distillation, ED: electrodialysis. | ||||
| PV | 20 : 80 |
4/∼12 | >99 (@100) | 0.02–0.04 |
| ED | 10 : 90 |
4/— | 45 (@20) | 7.0 |
| VD | 20 : 80 |
4/10 | 90 (@100) | 0.1 |
| VD | 53 : 47 |
13.8 | 69 (@80) | 0.15 |
| PV | 53 : 47 |
13.8 | 69 (@80) | 0.03 |
Membrane materials, particularly polydimethylsiloxane (PDMS), have demonstrated stable flux and effective separation performance over extended periods. For example, PDMS membranes were successfully used to recover ILs from methanol, showing increased flux with rising temperatures but reduced flux at higher IL concentrations. These findings illustrate the critical influence of operational parameters on process efficiency.136 Nafion™ membranes have also proven effective, especially for water removal from IL solutions, achieving significant fluxes for both water and solvents like N-methyl-2-pyrrolidone (NMP), further highlighting the versatility of pervaporation in handling various IL–solvent systems.137 Huang and Fedkiw (2016)137 explored the application of Nafion membranes for water removal from IL solutions through pervaporation. By optimizing operational parameters, they effectively reduced water content from 1.0 wt% to 0.5 wt%. Among the tested membranes, the XL membrane exhibited the highest performance, achieving a water flux of 10 mg h−1 cm−2 at 100 °C. Their findings revealed a significant increase in water and solvent flux with temperature, with water flux peaking at 7–10 mg h−1 cm−2. However, selectivity slightly declined above 90 °C. Additionally, gas-sweep rates influenced performance, with higher rates improving water flux (up to 10.4 mg h−1 cm−2) but reducing selectivity due to increased solvent permeation, as shown in Fig. 12.
![]() | ||
| Fig. 12 Effect of temperature and gas as-sweep rate on pervaporation. Panels: effect of temperature on pervaporation results with liquid and gas flow rates of 5 and 50 mL min−1, respectively: (a) XL membrane; (b) Gore A membranes; (c) effect of gas-sweep rate on pervaporation rates: Gore A membrane with metal grid at 100 °C. Adapted from ref. 137. Copyright 2016, Taylor & Francis. | ||
To address membrane swelling and stability challenges, support layers were introduced, ensuring mechanical reliability and consistent operation. These results confirm the feasibility of Nafion membranes for efficient IL recovery, combining high performance with robust thermal and mechanical stability.
However, hollow fiber membranes offer unique advantages in pervaporation processes due to their high surface area-to-volume ratio. They have been extensively used for separating water–organic and organic–water mixtures, and even for applications like water desalination.138 An innovative application includes the concentration of glycerol using hollow fiber pervaporation membranes, which significantly reduces energy consumption compared to distillation while preserving product quality by operating at lower temperatures. This capability highlights their energy efficiency and product safety advantages.139 The dynamic cross-flow coating method for hollow fiber membranes creates defect-free selective layers, enhancing pervaporation for alcohol–water separation with a separation factor of 6.4.139 PDMS/PVDF hollow fiber membranes have shown exceptional potential for phenol recovery from coal chemical wastewater. In pilot-scale experiments, these membranes achieved a phenol removal efficiency of 72% under optimal conditions: 70 °C, 150 L h−1 flow rate, 5 kPa membrane pressure, and a 0.3 gas–water ratio, within a 6-hour cycle. Moreover, the system demonstrated long-term operational stability, maintaining over 60% efficiency for 120 hours.140 The use of PDMS/PVDF hollow fiber composite membranes in such processes has shown high removal efficiency and stable operation over extended periods, proving their suitability for industrial wastewater treatment.140,141 Jie et al. (2014)140 demonstrated phenol removal efficiency increases with time and temperature due to enhanced molecular energy, but prolonged high temperatures may compromise membrane integrity. Higher flow rates reduce concentration polarization and improve efficiency, while lower membrane pressures promote phenol permeation at the risk of fouling. The system demonstrated stability over 120 hours with consistent removal efficiency exceeding 60%. These results underscore the system's potential for energy-efficient and scalable phenol recovery in industrial wastewater treatment applications.
However, there are challenges to the widespread adoption of these systems. Table 6 summarizes key insights and applications of pervaporation and hollow fiber membrane systems for ionic liquid recycling. Membrane stability under varying operational conditions, such as temperature and pH, is essential for ensuring long-term performance. The choice of membrane material significantly impacts the durability and efficiency of the process. Moreover, economic viability remains a key consideration, as the initial investment and operational costs associated with these systems must be justified through comprehensive techno-economic analyses to validate their commercial feasibility.
| Aspect | Key observations | Ref. |
|---|---|---|
| Pervaporation efficiency | Highly selective and efficient for IL recovery, achieving a separation factor of 1500 and >99.9 wt% recovery for [EMIM][CH3COO] | 135 |
| PDMS membranes | Stable flux and good separation performance for IL recovery from methanol; flux increases with temperature but decreases with IL concentration | 136 |
| Nafion membranes | Effective for water removal from IL solutions, achieving high flux for water and solvents like NMP | 137 |
| Hollow fiber membranes | High surface area enhances efficiency in water–organic and organic–water separations and desalination | 138 |
| Glycerol concentration | Reduces energy consumption compared to distillation, operating at lower temperatures to avoid glycerol decomposition | 139 |
| Phenol recovery | PDMS/PVDF membranes used in coal chemical wastewater for phenol recycling, achieving high efficiency and stable operation | 140 and 141 |
In comparison to traditional separation methods, membrane-based recycling using pervaporation and hollow fiber systems offers distinct advantages in energy efficiency and selectivity. Addressing challenges such as membrane stability and economic feasibility will be crucial to fully realizing their potential in industrial applications. Continued development of robust and cost-effective membrane materials will be instrumental in advancing these technologies for IL recycling, paving the way for greener and more sustainable industrial processes.
The effectiveness of distillation hinges on optimizing parameters such as temperature, pressure, and IL composition. Protic ionic liquids (PILs), such as ethanolamine acetate [EthA][CH3COO], have demonstrated recovery rates exceeding 85% in vacuum distillation at the lab scale, achieving glucose and xylose yields of 73.6% and 51.4%, respectively, during biomass pretreatment,146 as shown in Fig. 13. While these yields were lower than those from early separation, they were coupled with higher retention of lignin and hemicellulose, enabling potential downstream valorization. The trends in glucose and xylose release during saccharification are also depicted, highlighting the leveling off sugar yields after five days. This indicates the need for further optimization to improve the enzymatic hydrolysis step in the distillation process. These findings highlight distillation's ability to outperform traditional methods like water washing, which are often energy-intensive and generate toxic waste streams.135 Further optimization has been observed in ILs like 1-ethylimidazolium chloride [EIM]Cl, which, when distilled at 200 °C under vacuum, achieved 93% recovery while maintaining structural integrity.119 Distillation also offers the advantage of retaining IL structural integrity, enabling reuse and alignment with circular economy principles. For instance, the neutral compound distillation of [TMGH][CO2ET] resulted in >99% recovery purity with negligible decomposition, emphasizing its robustness under thermal conditions.147 Studies have demonstrated that distillation is particularly effective when integrated with complementary techniques. For example, combining distillation with membrane separation or desorption columns has enhanced recovery efficiency and reduced operational costs.148
![]() | ||
| Fig. 13 Biomass pretreatment with distillable ILs. Panels: (a) process configuration; (b) comparative sugar yields from different methods of processing using ethanolamine acetate; (c) glucose and xylose yield vs. saccharification time for distillation recovery method using ethanolamine acetate. Adapted from ref. 146. Copyright 2024, Elsevier. | ||
Economic and environmental considerations are paramount for distillation's industrial adoption. Innovations such as agitated thin-film evaporators have enhanced energy efficiency, achieving consistent recovery rates above 90% for ILs like [MTBDH][CH3COO].149 Additionally, reducing vacuum requirements and modifying IL compositions such as employing lower-boiling-point alkanolamines have demonstrated potential to decrease energy consumption and improve economic feasibility.150 Pilot-scale studies have validated these findings, with integrated distillation processes reducing reliance on water washing while achieving comparable sugar recovery rates in consolidated biorefinery designs.146
Despite its advantages, distillation faces challenges such as residual IL contamination in biomass and the generation of inhibitory by-products like furfurals and phenolics. Addressing these issues requires further optimization of process conditions and apparatus design. Nevertheless, the scalability and adaptability of distillation, particularly when coupled with advancements in apparatus design and process integration, underscore its viability as a cornerstone technology for IL recycling. Table 7 provides a detailed comparison of the performance of various ILs under different distillation conditions, illustrating the versatility and potential of this method for widespread industrial applications. However, distillation represents a robust and adaptable solution for IL recovery, balancing technical efficiency with environmental sustainability. Continued innovation in apparatus design, IL formulation, and process integration promises to enhance the viability of distillation, solidifying its role in sustainable IL recycling and industrial applications.
| Ils | Recovery method | Conditions | Recovery rate (%) | Observations | Ref. |
|---|---|---|---|---|---|
| [BMIM][CH3COO] | Vacuum distillation | 80 °C, 6 h, reduced pressure | >90 | Minimal lignin accumulation; high stability | 142 and 144 |
| [EMIM][CH3COO] | Molecular distillation | 170 °C, 0.05 mbar, 4 h | >90 | High purity (>95%); effective for cellulose | 143 and 145 |
| [EIM]Cl | Vacuum evaporation | 200 °C, 1 h, vacuum | ∼93 | Maintains acid : base ratio; no structural change |
119 |
| [TMGH][CO2ET] | Neutral compound distillation | 200 °C, 1.0 mmHg | >99 | Regenerated with high purity; negligible decomposition | 147 |
| [MTBDH][CH3COO] | Thin-film evaporator | 61–82 °C, 17–20 mbar | ∼90 | Consistent performance over five cycles | 149 |
| [EthA][CH3COO] | Vacuum distillation | 5–20 kPa, 140 °C, full vacuum | ∼85 | High sugar yields; minor contamination | 146 |
| [EMIM][CH3COO] + methanol/ethanol | Flash distillation | 25–50 kPa, ∼150 °C | ∼95 | Effective recovery; requires vacuum conditions to avoid IL decomposition | 152 |
| Dialkylimidazolium-based ILs | Flash distillation | 30–101 kPa, 120–180 °C | — | Challenges in achieving high purity due to viscosity and thermal sensitivity | 153 |
The development of co-solvent systems integrated with ILs marks a significant innovation in enhancing their recyclability and functional efficiency, addressing both environmental sustainability and industrial scalability. These systems mitigate the intrinsic limitations of ILs, such as high viscosity and restricted miscibility with non-polar reagents, by tailoring their properties for improved performance. For instance, the incorporation of water as a co-solvent in ILs, as demonstrated by Pugin et al. (2004),151 enabled remarkable improvements in catalyst recyclability during enantioselective hydrogenation reactions. Using a multi-phase IL–water system, their method involved monitoring Rh content with ICP-MS to confirm minimal leaching and employing ferrocene-based ligands for high turnover and enantioselectivity. This IL–water combination not only achieved turnover numbers exceeding 10
000 but also minimized catalyst leaching to 0.9 ppm while enhancing enantioselectivity and activity. Such advancements underscore the superiority of IL–water systems over traditional IL–organic solvent combinations, particularly in reactions involving Rh-ferrocenyl-diphosphine catalysts. Tomar and Jain (2022)154 investigates the versatility of co-solvent systems, emphasizing their ability to stabilize ILs and improve recovery rates in enzyme-catalyzed reactions, with experimental methods focusing on optimizing physicochemical conditions to ensure repeatability across cycles, contributing to the broader goals of green chemistry.
The synergy between ILs and co-solvents is particularly evident in applications like N-debenzylation reactions. Choi et al. (2010)155 demonstrated how combining [BMIM][BF4] with methanol not only enhanced reaction yields but also mitigated palladium leaching. Their method involved catalytic hydrogenation at room temperature and atmospheric hydrogen pressure for 16 hours using Pd/C catalysts. Dichloromethane proved effective in minimizing palladium leaching during extraction, enabling reuse of the ionic liquid phase for up to eight cycles with only a slight reduction in yield, followed by a significant decline. However, challenges such as diminishing yields after repeated cycles call for optimization strategies to extend the lifespan of these systems. Corley and Iacono (2019)156 describes a method to recycle the ionic liquid [DMPIm][NTf2] using a combined cation–anion exchange adsorption–desorption process with a 0.1 M NaCl
:
methanol (90
:
10 v/v) eluent. This approach enables simultaneous desorption of ionic liquid ions, separation from neutral impurities, and recovery of high-purity products with ∼60% yield, followed by liquid–liquid extraction. While addressing challenges in recycling ILs with complex anions, the study does not explore scalability for industrial applications or the long-term stability and reusability of adsorbents, limiting its economic and environmental assessment.
In the realm of cellulose chemistry, Gericke et al. (2011)157 showcased the role of polar co-solvents in reducing IL viscosity and enhancing miscibility with cellulose-dissolving ILs, leading to more efficient derivatization processes. Their study systematically evaluated 18 solvents and binary mixtures, employing solvatochromic parameters such as polarity, acidity, and basicity to assess miscibility and reaction outcomes. Despite these benefits, there remains a need for comprehensive evaluations of the long-term stability and reactivity of such systems under varying environmental conditions. Enhancing miscibility and phase behavior, Najdanovic-Visak et al. (2002)158 study on ternary mixtures of ILs with water and ethanol revealed the potential of a 1
:
1 molar ratio in facilitating effective extraction and recovery processes. Accurate phase diagrams were obtained using laser light scattering techniques to detect cloud points, ensuring reliable data collection. However, the lack of detailed thermodynamic data continues to pose a barrier to process optimization.
The recyclability of superbase-derived ionic liquids (SILs) for cellulose processing has been comprehensively evaluated, focusing on [DBUH][CH3OCH2COO], [DBUH][CH3CH2OCH2COO], and [AMIM]Cl. Among these, [DBUH][CH3OCH2COO] exhibited the highest thermal stability and recyclability, sustaining 10 recovery cycles with a recovery yield of 95–97%, compared to 5 cycles for [AMIM]Cl and 4 cycles for [DBUH][CH3CH2OCH2COO]. Degradation mechanisms, depicted in Fig. 14a, highlight structural transitions affecting recyclability, while Fig. 14b demonstrates improved polymerization and crystallinity of regenerated cellulose processed with [DBUH][CH3OCH2COO]. These results establish [DBUH][CH3OCH2COO] as a superior, sustainable, and economically viable solvent for large-scale cellulose processing.159 Another pivotal advancement in IL recyclability involves biphasic systems. Research by Bagherzadeh and Ghazali-Esfahani in (2012)160 explored toluene as a co-solvent with molybdate-anion-based ILs, achieving over 99% conversion in sulfoxide reduction and retaining catalytic activity across eight cycles. The methodology included employing newly synthesized room-temperature ILs containing molybdate anions, with conversions monitored using GC-MS and catalyst stability verified through repeated reaction cycles.
![]() | ||
| Fig. 14 Degradation mechanisms of ILs and their impact on the degree of polymerization and crystallinity of regenerated cellulose. Panels: (a) degradation mechanism of [DBUH][CH3OCH2COO], [AMIM]Cl, and [DBUH][CH3CH2OCH2COO] in the recycling process; (b) degree of polymerization and crystallinity index of RC obtained from fresh and recycled [DBUH][CH3OCH2COO] and [AMIM]Cl. Adapted from ref. 159. Copyright 2023, Elsevier. | ||
While the results are promising, challenges such as substrate diversity and large-scale applicability persist. On a different front, Giacalone and Gruttadauria (2016)161 introduced covalently supported IL phases, where ILs are immobilized on solid supports like silica and polystyrene. This approach not only simplifies separation and reuse but also improves catalytic performance by providing a stable environment for reactions. Methods focused on covalent attachment techniques and post-reaction recovery assessments to validate reusability. Similarly, IL–water mixtures demonstrated enhanced recyclability and catalytic efficiency, outperforming conventional organic solvents in multiple cycles of use. Their approach included systematic variations in water content to optimize catalyst separation and solvent recovery.162
Recent innovations in liquid–liquid extraction processes for IL recycling, such as those proposed by Pan et al. (2024),163 have opened new pathways for industrial applications. By employing hydrophobic ILs as accommodating agents, this method facilitates the effective recovery of water-miscible ILs. Pan's experimental design involved biphasic slug flow and membrane separation techniques, with efficacy demonstrated through platinum nanoparticle synthesis. This technique has shown significant promise, although its environmental and economic impacts warrant further assessment. Deng et al. (2020)164 discussed the integration of ILs and water in biphasic systems for enzymatic bio transformations, where the dual advantages of automatic purification and enzyme reuse significantly reduced waste and improved sustainability. Their method utilized a continuous-flow setup with compartmentalized spaces to streamline purification and recovery processes. Renewable co-solvents have also gained attention, particularly in lignin extraction. Yang et al. (2020)165 identified how these co-solvents, by increasing hydrogen bonding, selectively remove lignin while maintaining the structural integrity of the ILs, thus enhancing recyclability and reducing environmental impact. Their experimental process combined ILs with renewable co-solvents under controlled thermal conditions to maximize lignin dissolution and recovery.
Economic and environmental implications remain at the forefront of discussions surrounding IL recyclability. Tomar and Jain (2022)154 stressed that co-solvent systems reduce the dependency on fresh solvent inputs, aligning with sustainable industrial practices. However, the implementation of these systems on a commercial scale requires careful optimization to balance cost-effectiveness with technical efficacy. Guidelines proposed by Gericke et al. (2011)157 for co-solvent selection-considering polarity, acidity, and basicity-serve as a roadmap for tailoring IL systems to specific applications. Giacalone and Gruttadauria (2016)161 further emphasized that covalently supported IL phases contribute significantly to reducing solvent waste and lifecycle costs. Nonetheless, challenges such as the economic viability of co-solvent mixtures, potential environmental trade-offs, and the technical complexity of recycling processes highlight the need for continued research and innovation.
To address these challenges, future research should focus on developing advanced co-solvent systems with tunable properties to further enhance IL performance and recyclability. Machine learning and computational chemistry could be leveraged to predict optimal co-solvent combinations and reaction conditions, reducing experimental trial and error. Additionally, exploring hybrid co-solvent systems that combine the strengths of multiple solvents could improve stability and reduce viscosity while maintaining recyclability. For large-scale applications, modular reactor designs that integrate continuous flow processes could streamline IL recovery and reduce costs. Furthermore, advancing greener synthesis routes for ILs and co-solvents, along with detailed life-cycle assessments, would ensure their environmental impact is minimized.
A detailed summary of the influencing factors is listed in Table 8. However, the recovery of ILs requires an integrated approach that combines advanced recovery technologies, a deep understanding of ILs chemistry, and environmentally sustainable practices. As the process is influenced by a complex interplay of structural, operational, and economic factors, making it essential to tailor recovery strategies to the specific properties of ILs and the requirements of the system. Innovations in materials and processes, along with system-specific optimizations, hold the potential to significantly improve recovery efficiency and facilitate broader industrial adoption of ILs. Furthermore, cutting-edge research into eco-friendly technologies, such as CO2-induced phase separations, continues to enhance the sustainability of IL applications, enabling high recovery efficiency with minimal environmental impact.
| Influencing factors | Impacts | Examples | Ref |
|---|---|---|---|
| Recovery method | Determines efficiency, purity, and feasibility | Distillation for thermal stability; ATPS for aqueous systems | 142, 156 and 166 |
| Cation structure | Hydrophobicity and tunability influence recovery methods | Longer alkyl chains enhance hydrophobic extraction | 167 and 168 |
| Anion type | Stability and solubility impact separation behavior | [BF4−], [Tf2N−] affect crystallization and extraction efficiency | 166 and 168 |
| Temperature | Reduces viscosity and enhances separation | High temperatures improve distillation and ATPS efficiency | 167 and 168 |
| Pressure | Facilitates specialized recovery techniques | High-pressure CO2-based ATPS for IL separation | 166 |
| System impurities | Complexity necessitates tailored approaches | Ion-exchange for organic impurities; ATPS for aqueous separations | 156 and 167 |
| Economic/environmental | Balances cost and ecological impact | CO2-induced ATPS as a cost-effective and eco-friendly recovery method | 166, 167 and 171 |
The presence of these impurities can significantly reduce the purity and functionality of recycled ILs. For instance, a study demonstrated that impurities such as lignin and sugars could lead to a decrease in the efficiency of IL recycling, necessitating additional purification steps.173,176 Although there has been significant focus on the recovery rate of ILs, there is a lack of comprehensive analysis regarding the purity and functionality of the recovered ILs after multiple recycling runs. This indicates a need for further research to evaluate how the recycling process affects the chemical properties and effectiveness of ILs in enzymatic hydrolysis.177
The efficiency of IL recycling is often measured by the recovery yield and purity of the ILs. Impurities can lower the recovery yield and require more complex separation techniques, such as aqueous two-phase systems or pervaporation, to achieve high purity levels.135,173 Although previous studies have shown varying effects of lignin on pretreatment efficiency, the specific impact of lignin accumulation in the recycled IL solutions on the pretreatment of different biomass types remains unclear and warrants additional research.177
The presence of impurities can increase the cost of recycling the IL due to the need for additional purification steps. This can affect the overall economic viability of using ILs in biomass processing, as the cost of ILs is a significant factor in the process economics. Although some research174,175 demonstrates a significant reduction in the cost of ionic liquid pretreatment, it lacks a comprehensive economic analysis comparing the cost-effectiveness of this method against other pretreatment technologies, which could provide insights into its practical application in large-scale biorefineries.
While impurities pose challenges to the recyclability of ILs, advancements in separation and purification technologies offer potential solutions. Techniques such as phase separation, the use of adjuvants, and optimized recycling processes can help mitigate the impact of impurities and improve the efficiency of IL recycling. These developments are crucial for making IL-based biomass processing more sustainable and economically viable.
![]() | ||
| Fig. 15 Conversion of glucose in IL, lignin conversion and use of ILs as catalyst. Panels: (a) glucose was converted to HMF; (b) HMF condenses with glucose to form sugar-derived humin; (c) possible formation pathway of lignin-derived humin; (d) successive use of IL and catalyst. Conditions: 277 mg of lignin and 342 mg of sugars were added into 5 g of IL and reacted at 170 °C for 2 h. The catalyst was 10 mol% to IL. Adapted from ref. 176. Copyright 2012, American Chemical Society. | ||
It is worth mentioning that most IL pretreatment studies to date have been performed in batch mode with extended residence times. However, evidence from related pretreatment technologies demonstrates that continuous systems can achieve short residence times and scalable throughputs. For instance, continuous hydrothermal pretreatment of wheat straw at 195 °C for 9–12 min resulted in >90% cellulose recovery and ethanol yields approaching 93% of the theoretical maximum, while sorghum bagasse processed at 180 °C for 10 min achieved glucose and xylose yields of 82.6% and 70.8%, respectively.182 Pilot-scale steam pretreatment of wood chips at 215 °C with a residence time of 7 min and a throughput of 39 kg h−1 produced liquors containing >30 g L−1 glucose and solids with up to 88% glucan hydrolysis yield, and continuous steam explosion of wheat straw at 178 °C for 6 min enabled enzymatic recovery of more than half of the glucose and 46% of the initial xylose in the liquor. Likewise, continuous tubular reactors with extruder feeding and residence-time control, and twin-screw extrusion platforms integrating staged autohydrolysis and alkaline delignification at 210–220 °C, have demonstrated residual lignin contents of ∼2% and ∼80% glucose recovery.183
Emerging designs such as continuous ultrasound baths (5 L, 40/80/120 kHz, 260 W) further illustrate mass-transfer intensification strategies transferable to IL pretreatment.184 Collectively, these studies highlight that reducing residence times from hours to minutes is technically feasible, and adapting such strategies to IL systems will require careful control of residence-time distributions in viscous slurries, extruder-based solid handling, ultrasound-assisted mixing, and energy-integrated solvent recovery. Representative continuous pretreatment studies are summarized in Table 9.
| Approach | Feedstock | Conditions | Throughput | Outcomes |
|---|---|---|---|---|
| Review of pretreatment fundamentals | General lignocellulosic materials | Survey of physical (milling, extrusion), chemical (dilute acid, alkali, organosolv), physicochemical (steam explosion, AFEX), and IL pretreatments | — | Identifies scale-up challenges; steam explosion typically 160–240 °C, seconds–minutes; dilute acid ∼1–5 wt% H2SO4, 120–200 °C; AFEX 60–100 °C, 15–30 min; IL pretreatments noted as emerging but limited by viscosity and recovery |
| Continuous tubular reactor (CTR) with residence-time control | Wheat straw, corn stover | Autohydrolysis in CTR; dynamic model with RT control; PI controller implemented | Pilot-scale; continuous flow | RT control validated at 5–20 min target residence times; experimental deviations reduced to <2% with controller; demonstrated stable throughput of lignocellulosic slurry |
| Continuous pretreatment–hydrolysis–fermentation train | Organic residues (e.g., corn stover, food residues) | Hydrothermal pretreatment (170–200 °C, 5–15 min) integrated with enzymatic hydrolysis and microbial fermentation | Conceptual to pilot scale | Demonstrated continuous production of lactic acid and succinic acid; glucose recovery up to 79% (w/w) after wet-disk milling; ethanol fermentation reached 93% of theoretical yield in wheat straw case study |
| Continuous ultrasound bath (intensification module) | General biomass slurries | CAD-designed bath; 5 L capacity, operated at 75–80% fill; ultrasound at 40, 80, 120 kHz; 260 W power | Flow-controlled; sampling port included | Enhanced lignin removal (up to 79–88%) in fiber-based residues; FEA modeling confirmed uniform cavitation field; enables real-time sampling and continuous operation |
In addition to reactor configuration, the high viscosity of ILs and their strong interactions with biomass limit mass transfer, reduce solids loading, and make scale-up challenging.51 The viscosities of ILs are typically 1–3 orders of magnitude higher than conventional organic solvents, with values ranging from 4.8 to 1110 mPa s−1 (ref. 186) depending on the ion pairing, which hinders diffusion-controlled processes such as enzymatic hydrolysis and biomass fractionation.187 These rheological properties restrict mixing and heat transfer, thereby slowing biomass fractionation and contributing to the excessive solvent requirements often reported in batch studies. Several approaches have been implemented to mitigate these effects. For example, co-solvent addition (e.g., water or volatile organics) can significantly reduce viscosity without compromising pretreatment efficiency, while rational design of low-viscosity ILs, including protic and bio-derived ILs, offers another promising pathway.34,188 Process-intensification methods such as ultrasound, microwave irradiation, and mechanical shear can enhance mixing and diffusion, accelerating lignocellulose dissolution. Furthermore, continuous configurations such as twin-screw extrusion inherently improve mass transfer by applying shear and mixing forces, and continuous ultrasound devices with defined flow control demonstrate additional opportunities to intensify diffusivity in viscous media. Collectively, these strategies offer effective means to alleviate viscosity-related mass transfer limitations and facilitate the scalable application of IL pretreatment.
Another significant challenge in current IL pretreatment methods is the requirement for large solvent-to-biomass ratios in batch operations. Such high IL loadings are typically necessary to ensure complete wetting and swelling of biomass fibers, uniform mixing in viscous solutions, and effective lignin disruption. Although this approach is widely utilized in laboratory studies, industrial adoption will require strategies to overcome the challenge of high solvent consumption. Several strategies have been explored to address this limitation, including high-solids loading (≥20–30 wt%) to reduce solvent use, co-solvent systems to lower viscosity and improve mass transfer, and solvent recycling to minimize the need for fresh IL. In addition, continuous platforms such as twin-screw extrusion can reduce solvent requirements by promoting mixing and solvent penetration at lower IL-to-biomass ratios. These advances indicate that solvent demand, though a major bottleneck, can be alleviated through a combination of solvent engineering, process intensification, and continuous reactor design.
The recycling of ILs involves complex processes to separate and purify the solvents from impurities and other substances. However, the recovery and regeneration of ILs are essential for their sustainable use in industrial processes, particularly in refining and chemical production. ILs recovery methods commonly include distillation, liquid–liquid extraction, and stripping, alongside emerging techniques like membrane separation and solubility-switchable ILs. Each method provides unique advantages and limitations. Distillation, a prominent method, involves separating ILs based on the volatility of components. Earle et al. (2006)189 demonstrated the distillation of imidazolium-based ILs at 200–300 °C under vacuum conditions (0.001 mbar), achieving low decomposition rates and high regeneration efficiency. However, the requirement of ultra-low pressures and high energy consumption hinders industrial-scale application.189 Vacuum distillation has shown promise in removing high-boiling-point compounds like dibenzothiophene (DBT) from ILs but struggles with larger molecules such as carbazole (b.p. 354 °C). Techniques involving the distillation of volatile components from non-volatile ILs are better suited for practical applications.190
Liquid–liquid extraction is another practical method, particularly using water or organic solvents. For instance, Gao et al. (2018)191 illustrated the regeneration of hydrophilic ILs (i.e., 1-methyl-3-(4-sulfonic acid butyl)) imidazole p-toluenesulfonic acid [(CH2)4SO3HMIm][Tos] through water extraction, maintaining desulfurization efficiency across five cycles with minimal performance decline (i.e., 43.6% to 41.2%). However, the use of volatile organic compounds (VOCs) as solvents raises environmental and cost concerns. Cross-contamination and additional equipment needs further complicate this process.142 Stripping methods, utilizing gaseous agents like nitrogen or steam, have shown high regeneration efficiencies. Liu et al.192 demonstrated nitrogen stripping for hydroxyl ammonium ILs, achieving over 95% efficiency within five cycles. Steam stripping, as applied by Hardacre et al., effectively removed reaction byproducts but remains challenging for large-scale applications due to equipment and operational complexities.
Emerging techniques like rotary evaporation, crystallization, membrane regeneration, and force field-based methods offer additional pathways for IL recovery. Rotary evaporation, demonstrated by Yao et al.,193 achieved efficient recycling of [C4mim][BF4] with good extraction performance. However, crystallization, while delivering high purity, remains energy-intensive and suitable only for small-scale applications. Membrane regeneration provides high selectivity and lower energy demand but suffers from fouling and low throughput. Advanced innovations such as solubility-switchable ILs (see Fig. 16), described by Kamimura et al.,194 represent promising solutions for scalable IL recovery. These ILs enable efficient phase separation and recovery under controlled conditions but require further development for industrial adoption. Similarly, continuous microfluidic processes, as explored by Pan et al., (2023)195 employs a continuous microfluidic process for the purification of metal-ion-loaded ILs, specifically utilizing inline analytical tools and a modified Nelder-Mead simplex algorithm for statistical optimization to determine the best operating conditions for ion extraction and phase separation. The process includes the extraction of Fe(III) ions from the ionic liquid [BMIM][NTf2] into deionized water, followed by membrane separation of the ionic liquid and aqueous phases. While these processes improve separation efficiency, they remain costly and technically challenging.
![]() | ||
| Fig. 16 Solubility-switchable ILs. Panels: (a) partition ratio of the solubility-switchable ILs in CH2Cl2–water system; (b) relationship between the number of carbon atoms in the cation and partition ratios in water–CH2Cl2 system. Adapted from ref. 194. Copyright 2023, The Chemical Society of Japan and Wiley-VCH GmbH. | ||
Although many laboratory studies have demonstrated recycling of ionic liquids, they typically report a few recovery cycles that correspond to hours rather than months of operation, limiting their industrial relevance. A prominent case study is the ionoSolv process developed by Hallett and colleagues,196–198 which employed distillation and antisolvent precipitation as primary recovery strategies for low-cost protic ionic liquids. At laboratory scale, solvent recycling was demonstrated for up to six consecutive cycles with stable performance. However, at pilot scale, this translated to only a few hours of continuous operation, underscoring the limitations of extrapolating from batch studies. To bridge this gap, the spin-out company Lixea196–198 established a dedicated pilot plant, where solvent recovery and reuse were demonstrated for over one year of continuous operation while maintaining biomass fractionation efficiency and solvent integrity. This example highlights both the promise and the challenges of IL recycling: while recovery by distillation and antisolvent is technically feasible, long-term operation must address impurity accumulation, solvent losses, and energy costs. The ionoSolv case thus provides a critical benchmark, emphasizing that sustainable IL implementation requires validation at pilot and demonstration scale, supported by techno-economic and life cycle assessments.
Despite advancements, challenges remain (i.e., poor electrical conductivity) the degradation of ILs (i.e., corrosive nature of some ILs, leading to degradation of cell components and contamination of the ILs), during electrochemical processes.199 Impurities accumulating during use and their corrosive nature further complicate recycling efforts, necessitating additional purification steps.199–201 These limitations underline the need for integrated approaches combining efficient, scalable regeneration methods with innovations in IL design.
In addition to solvent recovery, energy consumption is a critical parameter influencing both the economic and environmental performance of pretreatment. Conventional processes such as steam explosion are relatively energy efficient, consuming ∼1.3 kg of steam and 0.31–0.47 kWh of electricity per kg dry biomass, equivalent to 2–4 MJ kg−1 of biomass.206 Recent innovations, including fluidized-bed detoxification systems operating at 70 °C and 150 kg h−1 throughput, have demonstrated further reductions in energy demand by minimizing additional steam requirements while simultaneously improving ethanol yields by 14% compared to conventional washing.207 In contrast, batch IL pretreatments often require higher energy inputs due to large solvent loadings and extended heating. A recent techno-economic assessment of butylamine depots identified solvent loading as a major driver of both cost and energy demand, where high ratios (850 g kg−1 slurry) substantially increased heating requirements.208 Lowering the solvent concentration to 5 wt% (59 g kg−1 slurry) reduced sugar production costs by ∼33% and decreased energy input, while solvent recovery efficiencies of 93–99% via thin-film drying and distillation further lowered net energy demand.208 These findings suggest that with optimized solvent loading, efficient recovery, and integration into continuous processes, IL pretreatment could achieve energy performance comparable to or better than established methods such as dilute acid209–211 and AFEX.212,213
Despite these encouraging outcomes, detailed systematic energy analyses of IL pretreatment and solvent recycling are still limited in the literature. Most reported studies are confined to laboratory studies, with little evaluation across different IL families, recovery methods, or process scales. This gap prevents a clear understanding of trade-offs and hinders meaningful comparisons with conventional technologies. Future research should therefore prioritize detailed energy benchmarking of IL processes, explicitly linking solvent recovery efficiency, process configuration, and scale-up with overall energy performance, to establish their true competitiveness against established fractionation methods.
The transition of ILs from laboratory-scale to industrial applications has catalyzed advancements in production, driving down costs per kilogram through economies of scale.203 Such reductions are already enabling the integration of IL-based technologies in alkylation processes, with several companies leading scale-up efforts. However, much of the current research focuses on maximizing IL removal efficiency in applications like desulfurization and denitrogenation. Greater emphasis is needed on optimizing entire processes, conducting comprehensive economic analyses, and enhancing IL regeneration.214 Innovations such as IL immobilization and ultrasound-assisted methods offer promising avenues for overcoming mass transfer limitations, reducing operational costs, and enhancing industrial feasibility.215,216 Research highlights the need for economic optimization to facilitate IL adoption. Ahmed et al. (2020)217 emphasize that the high costs of production and recycling limit the commercialization of IL technologies. Scaling up IL technologies poses further challenges, as large volumes complicate recycling, increasing energy and resource demands.205 Techniques like distillation, extraction, and membrane separation require significant investments in energy and equipment, offsetting potential benefits.156 Additionally, the synthesis of ILs often relies on costly raw materials and intricate processes, making them less competitive than conventional solvents.201,218 Pan et al. (2023)195 further highlight the transition from technical feasibility to commercial viability, emphasizing the importance of scaling production while minimizing associated costs.
Achieving economic viability for IL technologies requires holistic advancements in synthesis, recycling, and process efficiency. For instance, enhancing IL immobilization can minimize material requirements and simplify regeneration processes. Additionally, alternative methods such as ultrasound irradiation show potential to mitigate mass transfer limitations, providing a pathway toward broader industrial adoption. Future research should focus on cost-effective recycling strategies, mass production techniques, and the development of integrated IL-based systems that leverage the unique properties of these solvents while addressing their economic limitations.
However, from an environmental perspective, ILs are not without concerns. Although their negligible vapor pressure reduces air emissions, many exhibit toxicity and poor biodegradability. They can enter the environment through waste streams from refineries, solvents used during IL regeneration, or traces left in treated products. This has led to growing concern about their ecotoxicity and biodegradability, necessitating research into their environmental fate and safe disposal methods.
A number of studies have demonstrated varying degrees of toxicity and biodegradability among ILs. For example, Thuy Pham et al. (2010)219 examined the toxicity of ammonium, pyridinium, and imidazolium-based ILs, concluding that ammonium ILs are the least toxic, while increasing aromatic nitrogen atoms significantly heightens toxicity. In terms of biodegradability, pyridinium-based ILs outperform imidazolium-based ILs, with structural modifications, such as the inclusion of ester groups in the side chains, significantly reducing toxicity and enhancing enzymatic degradation potential.220,221 Moreover, the length of alkyl side chains influences IL behavior, where chains of 6 to 8 carbon atoms are more biodegradable, though longer chains may inadvertently increase toxicity.220
The choice of anions also plays a pivotal role. While some anions, like alkylsulfates, exhibit favorable biodegradability, others, such as fluorinated anions, present challenges due to their hydrolytic instability and harmful decomposition products, like hydrofluoric acid.222,223 The combination of cationic and anionic components must be carefully optimized to maximize efficiency in applications like desulfurization, denitrogenation, and alkylation, while minimizing environmental risks. For example, pyridinium- and imidazolium-based ILs containing anions like tetrafluoroborate [BF4], hexafluorophosphate [PF6], and nitrate [NO3] have shown desulfurization efficiencies above 90% in advanced extraction methods.155 However, nitrate anions pose hazards due to their explosive nature, and tetrachloroferrate [FeCl4] anions have been shown to be toxic to marine organisms.219,224
Addressing these concerns requires robust recycling and regeneration strategies. Innovative techniques such as electrochemical leaching and electrodeposition have been developed to recover valuable materials like platinum from spent fuel cell electrodes, using chloride-based ILs under mild conditions without harmful gas emissions.225 These approaches represent a shift away from traditional pyro-hydrometallurgical processes, which are less environmentally friendly. However, inefficiencies in recycling can lead to significant waste generation, including the loss of ILs and secondary waste streams, which undermine the sustainability of IL-based processes.167,226 Furthermore, regulatory hurdles related to toxicity and poor biodegradability often increase the economic burden of implementing IL technologies.33 Fig. 17 illustrates a structured framework of questions designed to assess the environmental sustainability of solvent synthesis processes. It evaluates key aspects such as the incorporation of renewable feedstocks, energy efficiency, and waste minimization. The framework highlights the critical need to reduce the use of hazardous reagents and byproducts, ensuring that the synthesis process adheres to the principles of green chemistry.
![]() | ||
| Fig. 17 Analytical framework of questions aimed at evaluating the environmental friendliness of a solvent's synthesis process. Panels: (a) questions related to the synthesis steps of a given solvent regarding some environmental issues that allow a general comparison between solvents, adapted from the ref. 33 and 227. Copyright 2022, Elsevier; Copyright 2011, Royal Society of Chemistry. (b) Greenhouse gas emissions resulting from the energy consumed during manufacture (solid bars, assuming 0.042 g CO2 emissions per kJ) and from the eventual oxidation or incineration (hollow bars) of 1 kg of solvent, adapted from ref. 227 and 228. Copyright 2011, Royal Society of Chemistry (c) comparison between [BMIM][BF4] and fossil-based solvents (hexane and THF) and organic-based green solvents (CPME and 2-METHF), in relation to synthesis, recovery, toxicity and biodegradability. Note: The larger the arrow, the greener the solvent, adapted from ref. 33. Copyright 2022, Elsevier. | ||
When applied to ILs, this assessment reveals significant trade-offs between their advantageous properties and the environmental costs associated with their production. Notably, many ILs fail to satisfy these criteria due to the reliance on toxic chemicals, high energy demands, and substantial waste generation, prompting a reconsideration of their classification as environmentally friendly solvents. Despite these challenges, ILs remain a valuable tool in sustainable chemistry. To comprehend their full potential, researchers must prioritize the design of ILs with enhanced biodegradability and lower toxicity, while advancing recycling technologies to reduce environmental impact. By balancing performance, safety, and sustainability, ILs can play a transformative role in greener industrial applications.
A critical comparison (see Table 10) reveals that no single recycling strategy is universally optimal since they all present trade-offs between recovery efficiency, energy consumption, scalability, and environmental impact. Variations in solvent recovery efficiencies and recycling performance across ionic liquids, distillable solvents, and DESs (e.g., imidazolium ILs, cholinium ILs, distillable solvents, DESs, etc.) highlight the need for standardized evaluation protocols to enable consistent and comparable assessments. Imidazolium-based ILs remain effective for biomass fractionation, but their high toxicity and costly recovery restrict broader applicability. Conversely, cholinium lysinate offers clear biocompatibility advantages, although its recovery remains challenging, suggesting that diluted cholinium lysinate formulations might be more practical and sustainable.18,68,81 Distillable solvents such as ethanolamine and other amine-based IL analogues are particularly attractive, as they can be efficiently recovered by distillation and reused, representing a practical pathway toward sustainable recycling.50,121 Looking forward, research should therefore prioritize the development of recyclable protic ILs and distillable solvent systems, alongside benchmarking against DESs and biphasic systems that offer inherent recyclability and reduced environmental burden.51,59
| Recycling strategy | Recovery efficiency | Energy demand | Scalability | Environmental impact |
|---|---|---|---|---|
| Antisolvent | ∼80–95% | High | Limited by waste streams; lab-scale | Large aqueous waste, costly purification |
| Distillation (volatile and amine-based solvent) | >95% (volatile ILs) | Moderate–high | Pilot-scale demonstrated (volatile/protic ILs) | Lower waste, but heat intensive |
| Membrane separation | ∼70–80% | Low–moderate | Pilot-scale; fouling remains a barrier | Reduced waste; membrane disposal issues |
| Co-solvent/hybrid | ∼70–80% | Moderate | Emerging; potential with IL–DES systems | Co-solvent toxicity and separation challenges |
| Deep eutectic solvents | — | Low-moderate | Lab- to pilot-scale; viscosity a challenge | Biodegradable, low-cost, but lower pretreatment efficiency than ILs |
| Biphasic solvents | — | Low | Promising at pilot scale; solvent selection critical | Phase separation aids recovery; solvent cross-contamination risk |
ILs have emerged as versatile compounds with potential applications across chemical and environmental processes, yet their LCA underscores both opportunities and challenges in their sustainability. Although limited studies have studied LCA of ILs,47,229–232 it is imperative to highlight these assessments to foster environmentally responsible processes and broader industrial adoption. Studies, such as Kralisch et al. (2005),233 pioneered LCA analyses of ILs like 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIM][BF4]) for reactions such as 1-octene metathesis, revealing that solvent-free reactions are not always ecologically advantageous. Their comparative analysis of energy requirements between biphasic and homogeneous reactions challenged the assumption that biphasic systems are inherently superior due to recycling ease.
Over the years, the environmental trade-offs of ILs compared to conventional solvents have been scrutinized. For instance, Zhang et al. (2008)234 demonstrated that [BMIM][BF4] poses greater environmental impacts than organic solvents like acetone in synthesizing cyclohexane. Similarly, Amado Alviz and Alvarez (2017)235 found that [BMIM][Br] exhibited higher ecotoxicity in pharmaceutical synthesis compared to toluene. These findings highlight that despite ILs benefits in reducing emissions during use, their synthesis and recovery often offset ecological advantages. ILs roles in carbon capture have also been a focus, with Cuéllar-Franca et al. (2016)236 advocating LCA for assessing ILs like [P66614][124Triz] against monoethanolamine (MEA). Farahipour and Karunanithi (2014)237 revealed that [BMIM][CH3COO] for carbon capture reduced greenhouse gas (GHG) emissions by only 50%, falling short of the 75% reduction with MEA. Likewise, Peterson (2013)230 conducted a cradle-to-grave LCA on 1-hexyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide and trihexyl(tetradecyl)phosphonium 1,2,3-triazolide for CO2 refrigerant systems and highlighted minimal environmental impacts from IL synthesis for CO2 refrigerant systems, emphasizing the need for cradle-to-grave analyses. Righi et al. (2011)238 compared [BMIM]Cl with NMMO/H2O in cellulose dissolution, noting that while effective, IL processes contributed significantly to abiotic depletion and toxicity.
Huebschmann et al. (2011)239 assessed simplified LCA combined with cost analysis for catalytic phenol and benzoyl chloride conversion using [BMIM]Cl, [MIM][BuSO3], and [C18MIM]Br, finding [MIM][BuSO3] ecologically superior due to exothermic synthesis versus [C18MIM]Br's energy-intensive process. While batch synthesis of [BMIM]Cl had low environmental impacts, continuous synthesis proved threefold ecologically advantageous. Mehrkesh and Karunanithi (2013)240 reported higher environmental impacts for IL synthesis compared to TNT production, reinforcing concerns about ILs ecological burden.
Optimization of ILs through molecular design and synthesis route adjustments is essential for sustainable application. For example, Guo et al. (2023)229 identified [BMIM][NTf2] as a cost-effective, environmentally superior IL for methanol/dimethyl carbonate azeotrope separation, outperforming other ILs in thermodynamic efficiency and total annual cost. However, toxicity issues, as seen with [BMIM]Cl and [EMIM][TCB], underscore the need for molecular tailoring and advanced recovery technologies to mitigate impacts. Fig. 18 illustrates simplified synthesis routes, and compares production impacts, highlighting [BMIM][NTf2]'s advantages. Energy-intensive production, as noted by de Jesus and Filho (2022),33 remains a critical barrier, compounded by decomposition losses and high utility consumption during IL use. Recent studies emphasize IL synthesis complexities and their significant environmental impacts. Imidazolium-based ILs for methanol/dimethyl carbonate separation, for example, show high impacts due to energy-intensive production and the use of volatile carbon, nitrogen, sulfur, and halogen compounds.33,229 While ILs reduce CO2 capture emissions, challenges like thermal decomposition losses and additional electricity needs compromise their benefits. As global sensitivity analysis (GSA) analyses by Baaqel et al. (2023)241 demonstrate, identifying environmental hotspots can guide greener designs.
![]() | ||
| Fig. 18 Life cycle analysis of ILs. Panels: (a) life cycle perspective of ILs; adapted from ref. 47. Copyright 2019, Elsevier (b) life cycle trees for IL production; (c) comparison of life cycle impacts of the production of 1 kg of ILs, adapted from ref. 229. Copyright 2023, American Chemical Society. | ||
The potential of ILs in natural gas dehydration and other separations remains tempered by their poor biodegradability and long-term toxicity. Comparisons with volatile organic solvents reveal that while ILs may reduce emissions during use, their overall ecological footprint including high impacts from synthesis and disposal often negates benefits. [BMIM][CH3COO] showed a 50% global warming potential reduction compared to unabated processes but lagged behind MEA's 75% reduction.237 Future pathways, integrating computational modeling and LCA, as suggested by Cuellar-Franca et al. (2016),236 aim to align ILs with green chemistry principles, emphasizing biodegradability and resource efficiency. These efforts are critical in ensuring that IL innovations address environmental hotspots holistically, preventing burden shifts across the life cycle.
Although ILs like [BMIM][NTf2] offer promising pathways for sustainable industrial applications, their environmental trade-offs necessitate further optimization. Through comprehensive LCA and targeted innovations, ILs can achieve a balance between functionality and sustainability, enabling greener chemical processes without unintended ecological costs.
Similarly, Ferrari et al. (2021)250 highlighted the importance of optimizing pretreatment conditions-solid loading and water content had greater impacts on energy consumption than temperature, underscoring the role of process parameters in economic outcomes. Innovations like ensiling biomass, as shown by Magurudeniya et al. (2021),251 have reduced IL consumption by up to 50%, shortened hydrolysis times from 72 to 24 hours without compromising sugar yield, and allowed one-pot processes for biofuel production, leading to reduced production costs and environmental footprints. Similarly, biocompatible ILs like cholinium lysinate ([Ch][Lys]) have enabled one-pot processes that integrate pretreatment, enzymatic hydrolysis, and fermentation without requiring solid–liquid separation or detoxification steps, achieving ethanol MSPs as low as $3 per gallon under ideal conditions.252 A techno-economic analysis indicates that using recyclable ILs like [BMIM][NTf2] can achieve costs competitive with traditional organic solvents like 1-octadecene (ODE).202 However, high IL costs, ranging from $20 to $100 per kg, and cost-intensive recovery processes remain significant challenges.196,253
Industrial applications of IL pretreatment coupled with pyrolysis for co-producing biofuels and chemicals like furfural and levoglucosenone have shown potential to enhance economic returns through coproduct revenues and waste heat recovery, achieving MSPs of $1640 per tonne and $3590 per tonne, respectively.254 Despite these advancements, challenges like IL toxicity, low biodegradability, and high viscosity continue to limit mixing efficiency and sustainability. To address these issues, current research is focusing on developing low-cost, biodegradable ILs such as choline-based derivatives and integrating advanced recovery technologies like distillation–filtration systems, which are critical for improving the economic and environmental feasibility of IL-based biorefineries.249,254 Moving forward, greater attention should be paid towards valorization of lignin and other coproducts to offset production costs and enhance economic feasibility.
1. The molecular mechanisms underpinning biomass dissolution by ILs, particularly the role of specific anion–cation interactions, remain insufficiently underexplored. A deeper mechanistic insight, supported by advanced computational modeling and experimental studies, is critical to guide the rational design of more effective ILs.
2. The recovery and recycling of ILs are often energy-intensive, challenging the economic and environmental sustainability of the process. Existing recycling techniques such as distillation, membrane separation, and antisolvent precipitation methods face limitations in energy use, scalability, and the accumulation of biomass-derived impurities, which degrade IL performance over repeated cycles. This highlights the urgent need for more robust, energy-efficient, and cost-competitive recycling technologies.
3. The presence of impurities/biomass-derived components, including lignin fragments, sugars, and inorganic residues, introduces further complexity, affecting IL functionality, recyclability, and the quality of biomass-derived products. These impurities reduce the functional integrity of ILs during repeated cycles and increase the complexity of the recovery process. Standardized protocols for IL selection and optimization tailored to specific feedstocks and pretreatment conditions are lacking, limiting process reproducibility and scalability a challenge.
4. Despite being promoted as “green solvents”, many ILs tend to inhibit enzyme activity and/or are toxic to microorganisms used in bioconversion processes. Additionally, concerns persist over the biodegradability and the environmental footprint of IL synthesis and disposal. Innovative biocompatible IL formulations and effective detoxification strategies are required to address this challenge.
5. The commercialization of IL-based biomass pretreatment is hindered by high upfront costs, competition with cheaper alternatives, and limited industrial collaboration. Limited data on life cycle impacts and techno-economic feasibility of IL-based processes restrict IL deployment at scale, requiring rigorous data modeling and systems analysis.
Despite these challenges, future advancements can pave the way for the sustainable and large-scale application of ILs in biomass pretreatment. Developing cost-effective and biodegradable ILs through iterative designs that integrate renewable and bio-derived components is a promising direction. Innovations in hybrid recycling techniques, combining methods such as distillation, membrane separation, and antisolvent recovery, could enhance scalability and efficiency. Furthermore, standardizing IL optimization protocols, conducting comprehensive system analyses, and fostering collaboration among academia, industry, and policymakers will be critical for translating IL-based biomass processing from promising research to industrial reality.
| This journal is © The Royal Society of Chemistry 2025 |