Next Article in Journal
Studying the Effect of Shortening Carbon Nanotubes via Ball Milling on Cellulose Acetate Nanocomposite Membranes for Desalination Applications
Next Article in Special Issue
Bioselective PES Membranes Based on Chitosan Functionalization and Virus-Imprinted NanoMIPs for Highly Efficient Separation of Human Pathogenic Viruses from Water
Previous Article in Journal
Removal of MS2 and fr Bacteriophages Using MgAl2O4-Modified, Al2O3-Stabilized Porous Ceramic Granules for Drinking Water Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Green Chemistry and Molecularly Imprinted Membranes

1
Institute on Membrane Technology, CNR-ITM, University of Calabria, Via P. Bucci, 17/C, 87030 Rende, CS, Italy
2
Faculté des Sciences de Monastir, Université de Monastir, Bd. de l’Environnement, Monastir 5019, Tunisia
3
Department of Engineering and of the Environment, University of Calabria, 87030 Rende, CS, Italy
4
College of Chemical Engineering, Nanjing Tech University, Nanjing 211816, China
5
Centre of Excellence in Desalination Technology, King Abdulaziz University, Jeddah 21589, Saudi Arabia
*
Author to whom correspondence should be addressed.
Membranes 2022, 12(5), 472; https://doi.org/10.3390/membranes12050472
Submission received: 6 April 2022 / Revised: 22 April 2022 / Accepted: 24 April 2022 / Published: 27 April 2022
(This article belongs to the Special Issue Advances in Molecularly Imprinted Membranes)

Abstract

:
Technological progress has made chemistry assume a role of primary importance in our daily life. However, the worsening of the level of environmental pollution is increasingly leading to the realization of more eco-friendly chemical processes due to the advent of green chemistry. The challenge of green chemistry is to produce more and better while consuming and rejecting less. It represents a profitable approach to address environmental problems and the new demands of industrial competitiveness. The concept of green chemistry finds application in several material syntheses such as organic, inorganic, and coordination materials and nanomaterials. One of the different goals pursued in the field of materials science is the application of GC for producing sustainable green polymers and membranes. In this context, extremely relevant is the application of green chemistry in the production of imprinted materials by means of its combination with molecular imprinting technology. Referring to this issue, in the present review, the application of the concept of green chemistry in the production of polymeric materials is discussed. In addition, the principles of green molecular imprinting as well as their application in developing greenificated, imprinted polymers and membranes are presented. In particular, green actions (e.g., the use of harmless chemicals, natural polymers, ultrasound-assisted synthesis and extraction, supercritical CO2, etc.) characterizing the imprinting and the post-imprinting process for producing green molecularly imprinted membranes are highlighted.

1. Introduction

Chemistry is at the heart of technological progress that stimulates the production of new essentials of modern life. Due to its presence in all areas of our life, such as the materials and the objects that surround us and that we use every day, (food, drugs, fertilizers), chemistry affects all people. In this context, owing to the development and expansion of chemical industries, large amounts of non-degradable chemicals are present in the environment, leading to pollution and toxicity hazard for human, fauna and flora health. Currently, chemistry is strengthened by the emergence of the green chemistry (GC) concept, which is based on producing more and better, but also by bringing the degree of pollution to the lowest level to ensure both environmental protection and health safety. Moreover, green synthesis in chemistry is today one of the essential aspects to be taken into consideration in the development of new products. For achieving these results, GC concept deals with the efficient use of raw materials, the removal of wastes and the avoidance of using toxic and/or hazardous reagents and solvents in the manufacture and application of chemical products [1,2,3,4,5].
As Kharissova et al. reviewed [1], green chemistry finds application in several material syntheses such as organic, inorganic, and coordination materials and nanomaterials. Based on its 12 principles, it is oriented toward many innovative fields of research. Among them, extremely important is the application of green chemistry in the production of imprinted materials by means of its combination with molecular imprinting technology (MIT).
The latter is a powerful advanced strategic approach leading to the production of polymeric materials endowing specific recognition sites able to selectively interact with targeted analytes called “template molecules” [6,7,8,9,10]. These kinds of interactions imitate the molecular recognition mechanisms typical of living systems such as the interactions of receptor–ligand, antigen–antibody and enzyme–substrate, thus conferring to the imprinted materials’ biomimetic features [6,7,11]. Over the last decade, the growing demand for highly selective separation systems has led to a rapid development of molecular imprinting technology due to the high selective recognition, retention and transport properties exhibited by printed materials. Currently, they find a wide variety of application, such as affinity separation, recovery of bioactive compounds and critical raw materials, sensing of substances in clinical and environmental field, water decontamination, and so on [2,7,9,10,11,12,13]. Molecularly imprinted materials are produced in the form of polymers and membranes. Their peculiarity is that they possess specific recognition sites toward a particular compound of interest (called template) and are able to selectively recognize and separate it from complex mixtures containing other analytes, including their structural analogs.
These specific recognition properties render them highly selective and advantageous with respect to their corresponding non-imprinted materials (that are not selective) for achieving specific detection and separation at the molecular level [7,8,9,10,11,12,13,14,15,16,17].
In the case of molecularly imprinted polymers (MIPs), the recognition sites are usually created during the polymerization process. The synthesis of MIPs entails the polymerization of a functional monomer around the template molecules with the aid of a cross-linker. Subsequently, the template is extracted from the neonatal polymer matrix leading to the formation of recognition sites that exhibit high complementarity to it in shape, size and chemical function [8,9,10,11,12,15,16]. Molecularly imprinted membranes (MIMs) represent a special format of imprinted polymers combining their specific recognition properties with the typical features of membrane science.
Membrane processes are advanced and sustainable technologies that are increasingly replacing traditional separation techniques or integrating with them to achieve better utilization of raw materials, greater separation efficiency at lower costs and high value products. This is pursued in the logic of a circular economy, which aims at resource recycling and waste valorization, also considering environmental and human protection as well as economic and social needs. Today, membrane operations and particularly pressure-driven processes are used successfully in various area such as chemical, pharmaceutical, food, biotechnological, water treatment and much more [18,19,20,21,22,23]. However, an increase in selectivity for achieving high separation levels of tailored compounds from complex mixtures is necessary. From this point of view, the creation of specific recognition sites within a membrane matrix (or on its surface) leads to the production of highly selective membranes such as MIMs. The employment of these smart membranes as such or integrated with traditional membranes is promising for developing sustainable green processes.
MIMs are produced via different routes. For example, composite MIMs are prepared via the phase inversion technique embedding pre-synthesized MIP particles within the membrane matrix or copolymerizing a thin layer of an imprinted polymer with the surface of a pre-existing membrane. The phase inversion technique is also applied for creating the recognition sites directly into the membrane matrix during its formation. In this case, template molecules are added to the cast solution, and after membrane formation, their removal frees the membrane’s recognition sites. In this last method, MIPs are not used, and non-composite imprinted membranes are obtained [7,11,12,13,17,18].
MIMs offer several advantages over MIPs. In particular, even if MIPs exhibit high specificity, they suffer of a low loading capacity and a scarce possibility of working in continuous operation mode. However, due to their high crosslinking status, they are poorly processable. Conversely, the exploitation of both the inherent selectivity conferred by the imprinting procedure and the typical features of membrane-based separation processes (continuous mode operation, easy-scale up, large-scale application, mild operating conditions of pressure and temperature, etc.) allow MIMs to overcome these drawbacks and exhibit superior selectivity and separation efficiency [7,9,11,13,17,24,25,26]. Some examples of MIMs application are the detection and separation of biomolecules, the enantiomeric separation, the recovery of bioactive compounds from different matrices, and the removal of pesticides and dyes from water [7,11,13,18,25,26].
It is undeniable that due to its numerous advantages, molecular imprinting technology is attracting more and more attention. However, until now, despite their high separation performance, MIMs are largely used only at the academic level, while their acceptance on the industrial scale is still in an embryonic stage and needs to be heartened. Some factors hindering their application are the reproducibility, the use of high quantities of solvents and reagents often not quite eco-friendly during their fabrication, the increase in membrane cost, and so on. Some of them might be restrained with actions devoted to increasing the visibility and the industrial trials of these innovative tools. In addition, the integration of other membrane operations, such as nanofiltration, ultrafiltration, reverse osmosis, and membrane distillation with MIMs will stimulate their future large-scale application as well as their market. Finally, in this perspective and in view of the strict regulations aimed at the protection of the environment and the reduction of wastes, the scientific community and manufacturers are increasingly exploring the employment of greener strategies in their production processes while maximizing their efficiency and environmental friendliness.
The sustainable potential of molecular imprinting technology has been recently discussed [27]. This review presents an overview of the basic principles and approaches of GC for producing novel green polymers and membranes. Moreover, it discusses the general aspects concerning the application of the GC concept to the production of MIPs. Finally, it highlights the production of green molecularly imprinted membranes in agreement with the principles of green molecular imprinting.
Considering the advantages of GC in many sectors and its relevance in producing eco-friendly high selective separation tools, we are confident that this review will give great contribute to the current research trend in stimulating the production and the employment of molecularly imprinted membranes while applying the concept of green chemistry both on the research and industrial scale.

2. Green Chemistry

Sustainable development of syntheses, manufacturing materials and separation processes is becoming more and more a priority on the educational, research and industrial levels for obtaining benefits needed by modern society. In this scenario, chemistry plays a central role in improving the economy, the environment and quality of life. The strict environmental regulations make the chemical industry one of the main sectors affected by environmental problems. These rules require the development of new approaches in chemical processes and in the synthesis of chemical compounds that are safer for the environment and human health. It is in this context that a new branch of chemistry dealing with ecological approaches and called “green chemistry” was born. Today, it attracts great attention and appears as a strategic alternative to traditional chemical processes to reduce the environmental problems and afford the new needs of competitiveness. The concepts of GC is based on the identification and development of sustainable pathways in chemical synthesis and processes, emphasizing environmental and ethical objectives [1,2,3,4]. Green chemistry, also known as “benign chemistry” or “sustainable chemistry”, was born in 1990 in the United States of America, by virtue of initiatives aimed at developing chemical products and processes capable of eliminating or reducing the use and production of hazardous substances. Starting from these good purposes, the “twelve principles of green chemistry” were formulated [1,28]. Figure 1 summarizes them.
As is evident, the principles of GC promote the use of renewable materials and chemical processes with less impact on the environment and with reduction of the quantity of material and risks and dangerous reagents, thus contributing to the sustainability of chemicals and manufacturing processes. According with these principles, many research studies have made significant growth in the green chemistry applications field either at the level of syntheses and reactions or at the level of chemical processes using green routes. For example, the use of GC in chemical reactions requires experimental conditions and synthesis protocols, which include green constituents as solvents, reagents and catalysts [1,2,3,4,28]. From this viewpoint, considering that pharmaceutical and fine chemical industries generate the most abundant wastes, many efforts for reducing them began in the third millennium. For example, for reducing the emission of solvents lost in waste during organic syntheses (as in the production of advanced pharmaceutical intermediates), the use of catalytic methods such as catalysis heterogeneous and bio-catalysis has been foreseen [2,29].
In order to make meaningful comparisons concerning the effectiveness of the various synthetic strategies, their sustainability is evaluated by measuring some parameters such as the environmental factor (E-factor) and atom mass economy (AE). The first indicates the mass of total waste produced versus mass of final obtained product. A value in the range of 25–100 kg indicates a high amount of produced waste with a consequent negative environmental impact [2,30].
The atom economy is a number given from the ratio between the formula weight of the obtained product with respect to the total formula weight of the reactants [2,30]. Assuming exact stoichiometric amounts of starting ingredients and a theoretical chemical yield, it is useful for a quick prediction of the waste that will be generated in the process. Other green metrics, as for example the mass of process water and of used solvents with respect to the mass of the final product, the net mass of materials used, the energy consumed and so on, are discussed in the literature [30,31,32,33].
Briefly, some of the advantages offered by greener processes are: the avoidance of unneeded wastes; the possibility of recycling solvents, catalysts, and other reagents; the development of lower-hazard reactions and small quantity of reactants, thus preventing disasters; low energy consumption, prevention of contamination, better product quality, and so on [1,2,3,4]. Currently, green strategies lead in producing inorganic and organic compounds, nanomaterials, composites, aerogels, quantum dots, etc. Furthermore, green approaches can help in improving some traditional materials such as ceramics, polymers, adsorbents, bioplastics and biocomposites [1,2,3,4].
Among various applications of GC, great advantages derive from the employment of water and other green solvents in chemical reactions such as aqueous catalysis [34,35,36], of supercritical fluids (i.e., supercritical CO2) [37,38], ionic liquids (ILs) [39,40,41], deep eutectic solvents (DES) [42,43,44] and fluorous media [45]. Other green approaches are microwave-assisted and ultrasound-assisted processes [46,47,48], hydro/solvo thermal reactions [49,50], magnetic field-assisted synthesis [51], mechanochemistry [52], and UV irradiation [53,54]. All these approaches and the evolution of GC have been accurately reviewed [1,55,56,57,58,59,60,61], highlighting that over the last 20 years, the principles of green chemistry have strengthened the environmental sustainability of chemical processes.

3. Toward Green Polymers and Membranes

One of the different goals pursued in the field of materials science is the application of GC for producing sustainable green polymers and membranes [62,63,64,65,66,67,68,69,70,71,72,73,74,75] as well as inorganic–organic hybrid materials of based on a polymeric matrix holding a small amount of inorganic material (such as carbon-based nanotubes, metal nanoparticles and graphene oxide) [76,77,78,79,80,81]. Green chemistry has been applied for fabricating numerous biopolymers, biopolymer-based membranes [64,66] and different synthetic polymers, such as acrylic-based polymers [82], poly(vinyl) chloride [83], polyurethane [84], and so on. Synthetic processes include the use of biomass-based sources [85] and renewable raw monomers such as triglycerides, terpenes, allylic and olefinic monomers [85,86]. An important aspect dealing with the production of hybrid materials is the compatibility of both organic and inorganic ones, and for achieving this goal, surfactants are often used. In this context, microwave irradiation for liquefying lignin as the starting material for the synthesis of flexible polyurethane was performed [87]. A microwave-assisted method for producing piperazine-containing bisphenol formaldehyde polymer was also applied [88].
An innovative strategic application of GC takes place in the valorization of vegetable wastes for preparing biodegradable bioplastic films. Perrotto et al. [89] produced environmentally friendly and freestanding bioplastic-based films in a simple one steep process (in aqueous solutions of hydrochloric acid), exploiting different waste matrices (carrot, parsley, radicchio and cauliflower). They maintained the color of starting material and antioxidant properties. Moreover, these films exhibited mechanical properties similar to those of traditional synthetic plastics (i.e., polypropylene, polyethylene, polystyrene, poly (methyl methacrylate) (see Figure 2) [89].
The combination of these bioplastics with other polymeric materials produced composite films with reduced oxygen permeability and increased mechanical resistance, suitable for packaging application. One example is the polyvinyl acetate/carrot bioplastic blend [89]. Other cases of food waste valorization are the synthesis of cellulose-based bioplastics from wastes of parsley and spinach stems, rice hulls, and cocoa pod husks by digestion in trifluoroacetic acid (TFA), casting, and subsequent solvent evaporation [90]. Protein-based polymers and starch-based plastics [91,92] are also produced. The successful recycling of non-edible parts of vegetables demonstrated that it is possible to substitute non-renewable plastic resources with renewable biodegradable resources, thus centering on green chemistry and circular economy concepts [93,94].
The preparation strategies of greener membranes fulfilling the principles of GC were classified by Szekely and co-authors on the basis of the priority of their contributions [95]. Figure 3 summarizes them [95].
The first necessities are the use of greener solvents and minimizing the use of toxic chemicals for rendering the process safer. From this viewpoint, the preferred solvents are water, acetone, isopropanol, ionic liquids and supercritical CO2, while dimethylformammide, dimethylacetamide, dioxane, hexane, chloroform and N-methylpyrrolidinone are undesired (see Figure 4) [95].
In producing polyimide P84 membranes, Soroko et al. substituted the toxic dimethylformammide and dioxane with the greener acetone and dimethylsulfoxide [96,97]. In addition, the use of inorganic salts (i.e., calcium chloride, sodium tartrate), non-toxic organic molecules (i.e., citric acid) or UV irradiation for post-membrane preparation crosslinking helps in reducing the waste of toxic reagents [97,98,99,100,101]. Another fruitful strategy is the employment of bio-based solvents, such as glycerol and its derivatives, aqueous solutions of carbohydrates and gluconic acid, lignin-derived solvents, fatty acid methyl esters, and so on [102,103]. The second priority deals with the minimization of waste production, energy consumption and operating costs by reducing the steps of the membrane preparation procedure as much as possible. One possible approach is the combination of crosslinking and coagulation in one step, by adding the cross-linker to the cast solution. Third, the use of renewable and degradable materials as bio-based materials is advisable.
The fourth goes for solubilizing and crosslinking polymers at room temperature for reducing energy consumption. The production of degradable membranes is the final greener aspect that aims at the substitution of the conventional petroleum-based membranes with easy degradable bio-based membranes. Nevertheless, the production of bio-based polymers is not yet enough to satisfy the wide request of membrane manufacturing at the industrial level [66,95].
Cellulose, a polysaccharide produced by plants made up of long macromolecular chains of β-D-glucose, is one of the most employed raw materials for preparing bio-based membranes. It is used for preparing polymeric flat-sheet films and hollow-fibers with green synthetic routes as the employment of green solvents such as, methyl lactate, N-methylmorpholine-N-oxide and ionic liquids [66,95,103,104,105,106,107,108]. Cellulose-based polymers are blended with other polymers, or they are hybridized with inorganic materials [109,110,111]. As it was critically reviewed by Galiano et al., bio-based polymeric membranes have also been prepared using chitosan, hyaluronic acid, poly(isoprene), sodium alginate, and more [66].
The global market trend of green chemistry is increasing in different fields, especially in the pharmaceutical sector as well as agro-food processes and packaging, and it was predicted that its values will reach USD 165 billion by the year 2027 [112]. Nevertheless, the production of green separation materials endowed with high specific and selective separation ability at both the ionic and molecular levels, such as imprinted materials, is still in infancy. Therefore, more efforts at the research level, including development and technological transfer, are necessary for assessing the potential of integrating green chemistry with imprinting technology for a possible application of green highly selective tools at a large scale in the near future.

4. Green Chemistry in the Synthesis of Molecularly Imprinted Polymers: General Aspects

Imprinting technology is a multidisciplinary approach that bio-mimics the interactions of enzyme–substrate and antigen–antibody occurring in living systems in order to produce selective, resistant and reusable imprinted materials (i.e., polymers and membranes). These advanced separation tools exhibit high selective recognition and separation properties toward a specific compound of interest (both called “template” [7,12,13,113,114,115,116]. Currently, molecularly imprinted polymers are produced at the research and industrial levels and are employed in different areas of science and technology [13,114,116,117,118]. Prominent applications include chromatographic separation [119,120], solid-phase extraction and microsolid-phase extraction [121,122,123], sensing [124,125,126], chiral separation [127], drug delivery [128,129], and so on. This high specificity is due to the presence of recognition sites complementary to the template molecules and capable of recognizing them in a selective way from complex mixtures. The selective recognition sites are created during the synthesis of MIPs and make them advantageous with respect to traditional non-imprinted polymers for obtaining tailored separations at the molecular level [8,9,10,11,12,15,16].
Nevertheless, even if the features of these materials are in line with the GC concept, conventionally used reagents and strategies are not accurately green. Therefore, environmental awareness, the objective of an efficient use of raw materials and the simultaneous increased demand for highly selective separation systems, has led to the application of the principles of green chemistry to their production. This was pursued considering the needs of sustainable practices and GC advances, as well discussed by Erythropel et al., who proposed the “green chemisTREE” as a window display for the actions and continued growth of green chemistry [130].
The traditional synthesis of an imprinted polymer involves the polymerization of a functional monomer in the presence of the template, with the aid of a crosslinker and an initiator. During the process, the functional monomer polymerizes around the template, which remains entrapped in the nascent polymer chains. The removal of the template after polymerization reveals specific recognition sites distributed into the neonatal imprinted polymeric matrix. These sites are complementary to the template in terms of chemical function, shape and size, and are able to recognize and separate it from a complex mixture containing other compounds, including its structural homologues and opposite enantiomers [12,13,113,131]. Figure 5 shows a schematic representation of the synthetic procedure of MIPs.
Relevant aspects of the overall synthetic process include:
The necessary chemical complementarity between the template and the functional monomer, which forms pre-polymerization complexes (via covalent or non-covalent binding);
The choice of a reaction solvent non-interfering with the monomer–template interactions for warranting the formation of efficient recognition sites;
The use of the cross-linker for ensuring the formation of a three-dimensional cross-linked network and for stabilizing the recognition sites;
The employment of an appropriate organic solvent (or other methods) for removing the template from the imprinted matrix and for freeing the recognition sites.
From the above, it is clear how all these aspects of MIPs synthesis have an influence on the environment and on social impact. Furthermore, the health risk of imprinters and the negative influence of the poor degradability of MIPs that end up in the environment after their use should be consider. In a recent review [132], Arabi et al. listed the critical sides of the traditional imprinting technology (see Figure 6).
The authors critically evidenced the unsustainable points of the imprinting and post-imprinting steps, including the application and disposal of MIPs. This research group also coined the term “greenification” to present for the first time the fourteen green principles of imprinting technology [132] as a general guide for the development of green MIPs (see Figure 7).
These principles cover different aspects, ranging from the employment of non-toxic (or low-level) reagents and synthetic methods to the fabrication of self-cleaning MIPs in a short time and the optimization with the aid of computational design prioritizing operator safety [132].
For example, traditional largely used functional monomers such as methacrylic acid, acrylic acid, and 4-vynil pyridine, as well as the cross-linkers ethylene glycol dimethacrylate, trimethylolpropane trimethacrylate, and divinylbenzene, are known as toxic chemical compounds, and according to the principles of green imprinting, they begin to be replaced with harmless or environmentally friendly monomers. In this context, room temperature ionic liquids (RTILs) and deep eutectic solvents (DESs) have emerged as green functional monomers and solvents [133,134,135]. Ionic liquids are non-volatile and non-flammable compounds miscible with a wide number of organic solvents. They exhibit low vapor pressure and high boiling point, high stability, ionic conductivity and viscosity. RTILs are able to interact with different organic compounds and bio-macromolecules through the formation of hydrogen bonds, anion-exchange, hydrophobic, electrostaic and π–π interactions. RTIL-base MIP presents excellent recognition properties in an aqueous environment [136,137]. One example is represented by the synthesis of a chlorsulfuron-imprinted MIP using 1-vinyl-3 butyl imidazolium chloride. Binding studies evidenced a higher binding capacity of the MIP (3.88 mg∙g−1) toward the template molecules with respect to the non-imprinted polymer (2.96 mg∙g−1). Furthermore, in competitive adsorption, this MIP showed a high binding selectivity (47.2%) toward the template with respect to its analogs [138]. In a different work, 1-viny-3-carboxybutyl imidazolium bromide resulted in an efficient functional monomer in the synthesis of a polymer imprinted with synephrine [139]. Table 1 lists some ionic liquids used as functional monomers and their relative template.
Importantly, ionic liquids are not only used as functional monomers and porogens, they are also employed as additives, cross-linkers and dummy templates. These aspects are well discussed in different papers [134,136,159,160,161,162]. However, the employment of RILs is restricted by their high cost. In addition, not all of them are assessed to be non-toxic, and in some cases, it is preferred to use their derivatives [163].
Other strategies involve the use of metal ions [164], boronates [165] and bio-based monomers, etc. [132,166,167]. For example, chitosan, and sodium alginate are natural excellent biocompatible, eco-friendly and cost-effective materials simply polymerizing under mild conditions and interact efficiently with different types of templates (i.e., ions, organic molecules and biomolecules [132,167,168]. Furthermore, inorganic salts such as calcium chloride and N, O-bismethacryloyl ethanolamine (NOBE) resulted a valid alternative to move toward greener crosslinking agents [60,169]. Self-initiated polymerization is also a good strategy for avoiding the use of initiators. For example, it has been demonstrated that acrylic monomers such as 2-hydroxyethyl methacrylate, glycidyl acrylate and methacrylic acid are self-initiating polymerizable functional monomers by a simple excitation to a triplet state [170,171].
Considering that some polymerization methods lead to the production of high volumes of disposal and toxic solvents (e.g., chloroform, dichloromethane, N, N-dimethylformamide, hexane), which can contaminate the environment and operators, are studying new green solvents as alternatives to traditional ones [172]. In this perspective, deep eutectic solvents are emerging as a new generation of green solvents [173]. They consist of two components that are a hydrogen donor (e.g., choline chloride) and hydrogen acceptor (e.g., alcohols, amides, amines, urea). They have similar features to ionic liquids, but they are cheaper, safer, highly biodegradable and can be produced by non-ionic-based compounds [173,174,175,176,177,178]. As it was well discussed in recent review papers [60,133,135], they have an excellent recognition ability in aqueous media and are increasingly employed in imprinting technology. Recently, an imprinted polymer was fabricated using a bio-based deep eutectic solvent for the enrichment of organophosphorous in fruits and vegetables [179]. Adsorption experiments carried out with 5 mL of solution containing 5 mg of MIP and a mix of pesticides having each one an initial concentration of 218 mg∙L−1 revealed an excellent adsorption capacity toward all pesticides in a short time (30 s). The highest value (218.62 mg∙g−1) was observed in the case of chlorpyrifos, while the adsorption capacity of the non-imprinted polymers was low (48.58 mg∙g−1). Another example is the synthesis of MIPs used for the recovery of the bioactive compound synaptic acid from agricultural wastes [180]. More in detail, the imprinted polymer was applied for selectively remove sinapic acid from waste rape seed extract after oil manufacture. The maximum adsorption capacity was 121 mg∙g−1, while that of the non-imprinted polymer was 23 mg∙g−1. Selectivity studies carried out in the presence of the competing compounds ferulic acid, cinnamic acid, and vanillic acid showed a selectivity factor synaptic acid/competitor of 20.86, 28.77 and 24.26, respectively. Conversely, the non-imprinted polymer was not selective and exhibited similar adsorption capacity toward all tested compounds [180].
Other strategic approaches devoted to the reduction of conventional solvent consumption and emission involve the combination of green porogenic solvents with traditional functional monomers and cross-linkers or solvent reflux during polymerization [181,182]. The supercritical carbon dioxide (CO2)-assisted synthesis is also a fruitful alternative to the traditional synthesis employing organic solvents. Supercritical CO2, which is obtainable as a high pure subproduct of the industry, combines the properties of gas and liquid states. Owing to its numerous features (apolar, high diffusion coefficient and mass transport capacity, cheap, inert, low viscosity, non-flammable, non-toxic, odorless, recyclable), it represents a sustainable solvent both at the research and industrial levels and is a good porogen of imprinting processes [60,183,184].
Microwave and ultrasound have been present for some years as sustainable strategies used in the imprinting process. More in detail, microwave-assisted synthesis and the ultrasound-assisted synthesis are applied as innovative green approaches during the polymerization step. The use of microwave consistently reduces the polymerization time with respect to traditional heating. This is due to the promotion of high heat transfer into the reaction mixture that facilitated the increase in reaction rate and the decrease in the energy consumption [60,132,185,186]. A reduction of reaction rate is also obtained when employing ultrasound. This is due to the cavitation effect determined by the ultrasonic energy that increases the solubility and the diffusivity of reactants into the polymerization solvent [187,188]. It was demonstrated that MIPs are synthesized with the aid of these green actions, exhibiting similar or higher specific recognition properties of those prepared via traditional routes [189,190,191,192,193,194].
Another extremely important point to consider is the type of chosen template. The use of toxic templates results as hazardous for human health. In some cases, this problem is overcome by the use of non-toxic dummy templates. Bagheri et al. [195] exploited this approach for fabricating dummy molecularly imprinted polymers able to remove acrylamide from biscuit samples. Instead of the toxic acrylamide, the similar propanamide was employed as a dummy template in a synthetic process carried out in a green aqueous environment, thus also avoiding the use of organic solvents [195]. Dummy molecularly imprinted polymers for detecting and quantifying acrylamide in other food matrices have also been synthesized [196]. In a different way, the drug ractopmanine was detected in pig tissues with MIPs synthesized using ritodrine as the dummy template [197]. Another example of a dummy template is the natural isoflavon daidzein in the production of MIPs capable of removing fluoroquinolones from fish samples [198]. Dummy molecularly imprinted resins resulted in efficient solid-phase extraction of plant growth regulators [199].
The template removal from the polymer matrix after both the imprinting process and the subsequent recognition stage is also a crucial issue. First, the traditional process involves the use of large solvent volume. Second, polymer swelling (due to the solvent action) as well as extreme extractive conditions of pH and temperature can alter the created structure of recognition sites, thus negatively affecting their performance. In this scenario, acetic acid and sodium dodecyl sulfate (SDS) are largely used for removing bio-based templates from imprinted polymers even if the surfactant can be adsorbed by the MIP, allowing for the presence of a negative charge [132]. In order to reduce or avoid the volume of the extraction solvent, microwave-assisted extraction, ultrasound-assisted extraction and pressurized hot water extraction were found to be effective in replacing conventional organic solvents [7,132,200]. Lorenzo et al. [201] intensively discussed the mechanism of these strategies adopted for template removal. A good strategy for reducing reactants consumption, waste generation operating cost, and imprinters exposure is multi-template imprinting, which entails the contemporary use of two or more templates aimed at creating their corresponding recognition sites in a unique polymer matrix [202,203,204].
Staying faithful to the principles of greenification, current trends deal with the application of green actions on one or more of the aspects discussed above, also combining eco-friendly and traditionally used chemicals, synthetic routes and post-imprinting phases. Even if greenificated imprinted polymers present the advantages of harmlessness, eco-sustainability and biodegradability, further efforts are still underway to implement production and application. In this context, the roadmap that goes from 2012 to 2030 in Figure 8 envisages the achievement of various objectives [132]. Some of them are the almost total employment of bio-based monomers from renewable resources, solvent-free imprinting, the elimination of the post-imprinting stage, the recovery of wastes from imprinted and their corresponding non-imprinted polymers, and the conversion of wastes in functional materials [132].
In parallel to the research improvement in producing greenificated MIPs, attention was also focalized on the application of the principles of green imprinting technology to the production of green imprinted membranes, which are an advanced form of imprinted polymers, as is discussed in next paragraph.

5. Green Molecularly Imprinted Membranes

In accordance with the principles of green molecular imprinting, one of the biggest challenges for scientists is the use of greener approaches and/or materials for the production of green imprinted membranes, which represent a special format of imprinted polymers.
In general, membrane separation processes entail the separation and concentration of one or more desired compounds, employing a membrane as a separation system. The separation occurs owing to the different permeability of the solutes through the membrane under the application of a suitable driving force (i.e., pressure gradient, concentration gradient, etc.). Some typical features rendering competitive membrane-based operations with respect to traditional techniques are easily scaled-up, stability in a wide pH range, work in mild conditions of pressure and temperature, and have no phase change or requirement of additives (or in minimal part), low energy consumption and environmental impact. They can operate as single units or in a continuous integrated manner. The choice of the membrane is an important aspect to consider, because it represents the key to the separation process. Most relevant parameters determining the separation performance are membrane permeability, selectivity and stability [18,19,20,21,22,23]. From this viewpoint, the idea of introducing tailored specificity into a traditional membrane has opened up new frontiers in the field of membrane operations from microscale to nanoscale applications. This is because imprinted membranes are intelligent tools exploiting contemporary typical features of imprinting and membrane technologies, thus offering several advantages with respect to imprinted polymers and traditional membranes [205,206,207,208]. For example, MIPs suffer from a low load capacity and a poor possibility of working continuously, and due to their high level of crosslinking, they are poorly processable. Conversely, MIMs are able to operate in a continuous mode, can be applied at large-scale and show higher binding capacity and separation efficiency. In addition, they exhibit improved selectivity with respect to the traditional ones but preserve their stability and permeability properties and separate structural homologues and enantiomers between them [206,207,208,209]. Molecularly imprinted membranes are applied in different areas both as an alternative to the traditional separation technologies or integrated with them as well as with non-imprinted membranes for obtaining a high purification level of specific molecules in a sustainable way [7,206,210,211,212,213]. However, despite their many advantages and although their production has grown a bit more in recent years, the number of works dealing with the production of MIMs is still low compared to that of MIPs (see Figure 9).
Therefore, more efforts are necessary for their wide-ranging development, not only in the field of research but also at an industrial level. As summarized in Figure 10, some examples of MIMs application are the separation of macromolecules and drugs and the selective recovery of drugs, bioactive compounds, and herbal ingredients from different matrices [206,213,214,215,216,217,218,219]. Other applications are the clinical monitoring of drugs and toxic compounds [220,221], the drug delivery [222,223,224,225,226], the enantiomeric separation [227,228,229], as well as the detection and removal of contaminants from water and other sources [206,209,230,231,232]. Among all these applications, one example is the production of artemisinin-imprinted composite membranes for the selective separation and purification of the anti-malaria drug artemisinin from ethanol solutions containing its structural homologue artemether. In adsorption experiments, the maximum adsorption capacity of MIMs was 158.85 mg∙g−1, while that of the corresponding non-imprinted membrane was 37.35 mg∙g−1. Furthermore, the imprinted membrane exhibited an adsorption selectivity for artemisinin/artemether of 2.04. In competitive permeation experiments, the permeate flux of artemisinin was 12.5 mg∙cm−2 s−1 × 10−4, while that of artemether was 2.68 mg∙cm−2 s−1 × 10−4 [216]. Another example is the extraction of the herbal active ingredient Ebracteolata B, which exhibits different pharmacological effects such as anti-tubercle and anti-cancer activities, from Euphorbia fischeriana extract [217].
The separation of template molecules is achieved via either their selective retention or facilitated permeation [7,206,209,229,230,233,234].
Molecularly imprinted membranes are prepared in various configurations (flat sheets, hollow fibers, nanofibers) by exploiting different strategies, where the interactions of template-functional monomers and template-recognition sites of the membrane matrix occur via covalent or non-covalent bond (similar to MIPs) [7,13,206,209,213,235,236].
One of the traditional preparation methods of MIMs is the surface imprinting of a pre-existing membrane via the copolymerization of a thin imprinted polymer layer with the surface of a pre-existing membrane (e.g., commercial or previously prepared). This route obtains composite flat-sheets, hollow fibers as well as nanofiber membranes. The application of the phase inversion technique leads to the production of composite MIMs by means of the hybridization of previously synthesized MIP particles with a polymer commonly used for preparing membranes, either via “wet” or “dry” phase inversion (or their combination). These last methods are also useful for preparing non-composite MIMS MIMs using a polymer ad hoc functionalized with chemical functions able to interact with the template molecules. Therefore, in this case, the simultaneous formation of the membrane structure and of its selective recognition sites occur [7,13,206,209,213,229,230,235,236]. As an example of membrane preparation, Donato et al. [229] developed S-naproxen-imprinted membranes via photo-copolymerization of the functional monomer 4-vinylpiridine with the surface of a commercial polypropylene microfiltration membrane. The enantioselective-imprinted membrane exhibited a facilitated permeation of the enantiomer template. At optimized operating conditions (T = 25 °C, pH = 3.4, P = 0.4 bar, ((R,S)-Nap) = 7.0 µg/mL−1), the permselectivity factor S-naproxen/R-naproxen was 1.8. The water permeation flux was typical of ultrafiltration. On the contrary, the pristine commercial membrane and the blank non-imprinted membranes were not selective. In particular, the commercial membrane allowed for the permeation of both enantiomers, while the blank membrane exhibited a low permeation rate owing to the absence of the S-naproxen recognition sites [229]. Composite MIMs prepared via the phase inversion technique hybridizing the poly (vinylidene) fluoride matrix with polymer particles imprinted with 4,4-methylendianiline exhibited high specific retention toward the template with respect to non-imprinted membranes and those prepared with the only commercial poly (vinylidene) fluoride [237]. In permeation experiments performed in isopropanol at pressure of 0.1 bar with an initial feed concentration of 10 mg∙L−1 (in 100 mL), MIMs containing 33 wt.% of MIP particles showed the highest binding capacity (7.5 µmol∙g−1). At the same conditions, the corresponding non-imprinted membrane and the simple PVDF membrane exhibited a binding capacity of 4.4 and 2.0 µmol∙g−1, respectively. The permeability of the poly (vinylidene) fluoride-based membrane was in the nanofiltration range, while that of MIM and non-imprinted membranes was typical of ultrafiltration, indicating that the addition of polymer particles to the poly (vinylidene) fluoride matrix increased its permeability performance. Moreover, the MIM exhibited a selectivity factor of 1.82 toward 4,4-ethylendianiline [237].
Over time, these approaches were improved and strategically combined for producing advanced membranes [238,239,240,241,242,243,244,245,246,247,248].
Figure 11 shows the representation of flat-sheet membranes exhibiting selective binding toward template molecules, thus separating them from competing compounds that accumulate in the permeate stream [7].
The high performance of selective separation of MIMs, combined with the typical characteristics of membrane processes, makes their use rational with the principles of green imprinting technology, even if some characteristics relating to their preparation need to be better addressed with new and greener interventions. Similar to the case of MIPs, this is already occurring in part, as for example via the use of more environmentally friendly functional monomers, solvents and polymers that form the membrane structure. Other ecological actions are the application of new synthetic routes or the dummy template and multi-template imprinting, as well as the template extraction with ultrasound, microwave or supercritical CO2.
For example, the phase inversion technique is useful for preparing MIMs with safe co-polymers such as poly(acrylonitrile-co-acrylic acid) and natural polymers such as chitosan, sodium alginate, cellulose, β-cyclodextrin-based, etc. In particular, natural polymers (and their derivatives) have emerged as high promising materials that form membranes owing to their low cost, eco-friendly features as well as the abundance of active chemical functions (i.e., amino, carboxyl and hydroxyl, groups) with affinity toward many compounds and establishing with them multiple interactions [249,250,251,252]. An important aspect of the employment of these materials is the possibility of avoiding in some cases the polymerization process, while leading the formation of a “polymer–template complex”. This is due to the interactions between the functional groups of the template and the complementary chemical moieties of the polymeric material that form the membrane [250,251,253,254]. This structure is stabilized with the aid of a cross-linker both during or after membrane formation. These types of polymer–template interactions comprise hydrogen bonding, electrostatic and π–π interactions, and van der Waals forces. A problem with these materials is the structural stability in severe conditions; thus, the scientific community is making efforts in the direction of producing more stable innovative bio-based MIMs. Table 2 reports some examples on natural materials used in fabricating innovative MIMs.
In developing green membranes, natural polymers are used as such for direct membrane preparation or as membrane support, coatings or additives for preparing composite membranes via the simultaneous use of traditional polymers (e.g., poly (vinylidene) fluoride, polysulfone, poly (ether sulfone) polyacrilonitrile, etc.). Among natural materials, chitosan possesses both amino and hydroxyl groups, the presence of hydroxyl groups characterizes cellulose, and sodium alginate is a polyelectrolyte rich in carboxyl groups. The macrocyclic β-cyclodextrin consists of an external hydrophilic part and a hydrophobic inner cavity playing a key role in the recognition process. Owing to the biocompatibility, biodegradability and non-toxicity of natural polymers, MIMs prepared with them are suitable for application in the medical, nutraceutical and pharmaceutical fields as well as in water treatment. For example, bacterial cellulose was used for producing both diosgenin-imprinted membranes [250] and quercetin-imprinted [251] membranes, which exhibited a sustained selective release of the template molecules. Recognition sites were created directly into the polymer matrix during the membrane formation step via the phase inversion technique. In a different way, chitosan was used for fabricating a MIM-based sensor able to selectively detect and remove the 4-nitrophenol from drinking water [260]. In batch adsorption studies, this membrane exhibited a maximum adsorption of 723.25 μmol∙g−1 of 4-nitrophenol, while the adsorption capacity exhibited by the corresponding non-imprinted membrane was 517.69 μmol∙g−1. In addition, MIMs were selective with respect to competing phenol, 3-nitrophenol and 4-methoxyphenol, while non-imprinted membranes were not selective. The treatment of real samples (containing 7.19 μmol∙L−1 of this toxic phenolic compound) with MIMs leads to a removal efficiency of 70.6% [260]. More recently, sodium alginate resulted in high efficiency as a polymer, forming enantioselective MIMs with tailored recognition sites specific for d-tryptophan [266]. After their formation, the membranes were crosslinked with calcium chloride by the coordination of two carboxyl groups of the natural polymer and one Ca2+ ion exchanged with Na+. In pressure-driven permeation tests, these innovative smart membranes were able to separate tryptophan isomers from a racemic solution thorough a facilitated permeation of the template enantiomer. At an operating transmembrane pressure of 0.2 MPa, feed concentration of 0.5 mmol∙L−1 and pH above the isoelectric point of tryptophan (5.89), the permeation flux of d-tryptophan was 5.8 × 10−5 mol∙m−2∙h−1, and the permeation enantiomeric excess was about 99% (membrane thickness was 0.02 mm) [266]. Conversely, the non-imprinted membrane showed a permeation flux of almost two times lower (2.91 × 10−5 mol∙m−2∙h−1). Composite MIMs were also prepared via coating the surface of a poly (vinylidene) fluoride membrane with a d-tryptophan-imprinted sodium alginate film, as is shown in Figure 12 [254].
After the coating step and before the template extraction, the new composite MIMs were crosslinked with calcium chloride. Optimized membranes with an imprinted layer thickness of 0.02 mm exhibited a water flux of 6.46 L∙m−2∙h−1, a d-tryptophan flux of about 1.3 L∙m−2∙h−1 and an enantiomeric excess of 99.13 [254]. Electrospun methylene blue-sodium alginate/polyethylene oxide imprinted nanofiber membranes [267] and methyl orange-TiO2/calcium alginate hydrogel as a matrix [268] were also produced. The binding capacity of these membranes toward the template dye was 14.13 and 3186.7 mg∙g−1, respectively. Recently, for producing green tetracycline-imprinted nanocomposite membranes, a biomass-based strategy was developed [273]. In this context, biomass-activated imprinted carbon nanoparticles were embedded into the matrix of porous cellulose acetate/chitosan-blended membranes via phase inversion [273]. These obtained hybrid imprinted membranes exhibited a high permeate rate of the template molecules with respect to the competing structural analog oxytetracycline. The permselectivity factor was 2.4. Other examples of biomaterials are lignin, starch and water-soluble proteins [253,274]. For increasing the number of recognition sites, these materials are also functionalized via esterification, etherification, graft copolymerization, oxidation and Schiff’s base reaction, thus producing biopolymers derivatives [253].
In agreement with the principles of the green molecular imprinting, the employment of greener solvents such as acetone, dimethylsulfoxide, ethanol, isopropanol, ionic liquids, water and deep eutectic solvents for membrane preparation is also continuously increasing as an alternative to the traditional toxic solvents (e.g., dimethylacetammide and dimethylformammide and chloroform). In addition, multi-template imprinting, the use of less toxic functional monomers, initiators and cross-linkers, and more green template extraction strategies is growing [27,275]. A thin polymeric layer imprinted with thymopentin was polymerized on the surface of a regenerated cellulose acetate membrane using the ionic liquid 1-vinyl-3-ethyl acetate imidazolium chloride as a functional monomer [276]. Membranes were tested in solid-phase extraction disks for purifying thymopentin. Scatchard analysis indicated the presence of both high and low affinity binding sites exhibiting a maximum binding capacity equal to 13.07 and 8.04 mg∙g−1, respectively [276]. In another case, the ionic liquid 1-butyl-3-methylimidazolium chloride was used as a co-additive for preparing blended salicylic acid-imprinted cellulose acetate/polyethylene glycol-4000 membranes. The aim was the recovery of this active pharmaceutical ingredient from wastewaters [277]. Membranes exhibited higher binding affinity with respect to those prepared without the ionic liquid. In addition, the selectivity factor for salicylic acid with respect to the competing compounds p-hydroxybenzoic acid and phenol was 5.85 and 5.90, respectively [278]. Fan and coauthors developed a macroporous cryogel-imprinted membrane using (1-vinyl-3-(2-amino-2-oxoethyl) imidazolium chloride as a functional monomer and bovine serum albumin as a model template protein [279], in a phosphate buffer as solvent and porogen. The pre-polymerization mixture was infiltrated between two glass plates and the subsequent polymerization took place at −18 °C, leading to the formation of a macroporous membrane structure having a uniform distribution of pores (see Figure 13).
Permeation tests carried out in a diffusion cell with solutions containing the template and the similar human serum albumin (having each one the concentration of 0.4 mg∙L−1) evidenced a high transport rate of the template molecules: the permeation amount of bovine serum albumin was 2.82 mg∙cm−2, while that of the competing protein was 1.63 mg∙cm−2 [278].
Other syntheses, involving the application of the dummy template strategy, led to the production of innovative MIMs. For example, the separation of the herbal Chinese medicine (anti-malaria) artemisinin from the contaminant artemether was achieved with composite membranes imprinted with the dummy compound artesunate. This was because artemisinin has no chemical groups that can interact with functional monomers [279]. Membrane were prepared via the phase inversion adding pre-synthesized artesunate–MIP particles to a poly (vinylidene) fluoride cast solution.
Figure 14 shows the behavior of the adsorption capacity and the selectivity factor (α) artemisinin/artemether of the best membrane.
In another work, the detection of the mycotoxin citrinin in rice was accomplished using membranes imprinted with the less toxic dummy template 1-napthol [280]. Recently, in an eco-friendly synthesis involving the use of 1-vinylimidazole as a functional monomer and mild photo-co-polymerization conditions on the surface of a nylon-66 membrane, gatifloxacin was used as a dummy template. The prepared composite MIMs allowed for the simultaneous recognition and extraction of the antibiotics enrofloxacin and ciprofloxacin from egg samples [281].
The chemical structure of some investigated compounds and that of their relative dummy templates used in MIMs fabrication are reported in Table 3.
In a different approach, cinnamic and ferulic acids were used simultaneously for fabricating dual-template imprinted membranes exhibiting permeability typical of ultrafiltration and capable of detecting them in cereal samples [284].
In addition to the aspect of template use and extraction discussed in the previous paragraph, (limiting the use of toxic or precious templates as well as solvent consumption for their removal), the dummy template and the multi-template imprinting strategies are useful to control/avoid the template bleeding that sometimes represents a drawback of the subsequent recognition process. This is because possible template traces remaining in the membrane can negatively affect precise analytical determinations if released [176,285,286]. A contribution in this direction also comes from the use of supercritical CO2 both during the polymerization step and during template extraction [287,288,289].
As is evident, the combination of the concept of green chemistry with that of molecular imprinting technology has proven successful in the development of “green intelligent membranes”, which are promising for application in different sectors characterizing our life. However, it is necessary to make further efforts to use them now more than before on a large scale and on the industrial level, even in new integrated processes that require high selective separation efficiency.

6. Conclusions

Over time, human necessities and technological progress have allowed for contemporary increases in chemical processes at the research and industrial levels. However, the consumption of large solvent, toxicity of some used materials, and disposal problems have led to a status quo no longer sustainable both from the points of view of human health and from that of the host planet. The development of more eco-friendly processes has therefore become an emergency, allowing for the advent of green chemistry. According to its 12 principles based on ecological approaches, it plays a key role as a strategic alternative to the traditional chemical processes for reducing environmental problems and coping with new requirements of sustainability and economic affordability. Today, green chemistry finds application in organic and inorganic syntheses, in chemical reactions, and separation processes, as well as in the production of greener polymers and membranes, such as biopolymers and biopolymer-based membranes via the valorization of wastes. In this scenario, the world of molecular imprinting has embraced the concept of green chemistry and the current trend is devoted to the development of eco-friendly processes for producing green molecularly imprinted materials.
In particular, in agreement with the principles of green molecular imprinting and with their high selective separation performance, green MIMs are promising efficient tools for application in different areas. Strategic green actions characterizing their production are focalized on the minimization of waste production and energy and solvents consumption as well as on the use of harmless chemicals. They are realized with the aid of computational design, prioritizing operator security and including the use of greener or natural polymers as membrane-forming material, greener functional monomers, cross-linkers and solvents such as ionic liquids, deep eutectic solvents, acetone, dimethylsulfoxide, ethanol, water and supercritical CO2. The dummy template imprinting and the multi-template imprinting strategies represent other effective approaches that are also effective in limiting the use of toxic or precious templates and in avoiding the template bleeding problem.
The typical characteristics of membrane processes have allowed them to spread widely with great success, predominantly in the case of pressure-driven membrane operations. Conversely, despite the excellent properties of molecularly imprinted membranes, there is still considerable work to accomplish for better exploitation of the combination of green chemistry with imprinting technology for their possible application at a large scale in the near future. From this viewpoint, and taking present the greenificated roadmap from 2012 to 2030 of green imprinting technology, this review represents an opportunity for stimulating the awareness of exploring other green aspects of MIMs production for enhancing their sustainability and environmental remediation. In this perspective, it is legitimate to predict that the production of advanced green imprinted membranes and their integration with traditional membrane operations such as ultrafiltration, nanofiltration, reverse osmosis or membrane distillation will make it possible to market them.

Author Contributions

L.D.: identification of the subject, writing—original draft preparation, writing—review and editing, curation of figures and tables, supervision. I.I.N.: writing—original draft preparation, writing—review and editing, curation of graphical abstract. M.M.: supervision. E.D.: review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This review received no external funding.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kharissova, O.V.; Kharisov, B.I.; González, C.M.O.; Méndez, Y.P.; López, I. Greener synthesis of chemical compounds and materials. R. Soc. Open Sci. 2019, 6, 191–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sheldon, R.A. Metrics of green chemistry and sustainability: Past, present, and future. ACS Sustain. Chem. Eng. 2018, 6, 32–48. [Google Scholar] [CrossRef] [Green Version]
  3. Anastas, P.; Eghbali, N. Green Chemistry: Principles and Practice. Chem. Soc. Rev. 2020, 39, 301–312. [Google Scholar] [CrossRef] [PubMed]
  4. Unterlass, M.M. Green synthesis of inorganic–organic hybrid materials: State of the art and future perspectives. Eur. J. Inorg. Chem. 2016, 2016, 1135–1156. [Google Scholar] [CrossRef]
  5. Sheldon, R.A. Fundamentals of green chemistry: Efficiency in reaction design. Chem. Soc. Rev. 2012, 41, 1437–1451. [Google Scholar] [CrossRef] [Green Version]
  6. Janczura, M.; Lulínski, P.; Sobiech, M. Imprinting technology for effective sorbent fabrication: Current state-of-art and future prospects. Materials 2021, 14, 1850. [Google Scholar] [CrossRef]
  7. Donato, L.; Drioli, E. Imprinted Membranes for Sustainable Separation Processes. Front. Chem. Sci. Eng. 2021, 15, 775–792. [Google Scholar] [CrossRef]
  8. He, S.; Zhang, L.; Bai, S.; Yang, H.; Cui, Z.; Zhang, X.; Li, Y. Advances of molecularly imprinted polymers (MIP) and the application in drug delivery. Eur. Polym. J. 2021, 143, 110–179. [Google Scholar]
  9. Huang, Y.; Wang, R. Review on Fundamentals, Preparations and Applications of Imprinted Polymers. Curr. Org. Chem. 2018, 22, 1600–1618. [Google Scholar] [CrossRef]
  10. Zaidi, S.A. Molecular imprinting polymers and their composites: A promising material for diverse applications. Biomater Sci. 2017, 5, 388–402. [Google Scholar] [CrossRef] [PubMed]
  11. Torres-Cartas, S.; Catalá-Icardo, M.; Meseguer-Lloret, S.; Simó-Alfonso, E.F.; Herrero-Martínez, J.M. Recent advances in molecularly imprinted membranes for sample treatment and separation. Separations 2020, 7, 69. [Google Scholar] [CrossRef]
  12. Chen, L.; Wang, X.; Lu, W.; Wu, X.; Li, J. Molecular imprinting Perspectives and applications. Chem. Soc. Rev. 2016, 45, 2137. [Google Scholar] [CrossRef]
  13. Algieri, C.; Drioli, E.; Ahmed, C.; Iben Nasser, I.; Donato, L. Emerging Tools for Recognition and/or Removal of Dyes from Polluted Sites: Molecularly Imprinted Membranes. J. Membr. Sep. Technol. 2014, 3, 243–266. [Google Scholar]
  14. Culver, H.A.; Steichen, S.D.; Peppas, N.A. A Closer look at the impact of molecular imprinting on adsorption capacity and selectivity for protein templates. Biomacromolecules 2016, 17, 4045–4053. [Google Scholar] [CrossRef]
  15. Yang, S.; Wang, Y.; Jiang, Y.; Li, S.; Liu, W. Molecularly imprinted polymers for the identification and separation of chiral drugs and biomolecules. Polymers 2016, 8, 216. [Google Scholar] [CrossRef]
  16. Irshad, J.M.; Iqbal, N.; Mujahid, A.; Afzal, A.; Hussain, T.; Sharif, A. Molecularly imprinted nanomaterials for sensor applications. Nanomaterials 2013, 3, 615–637. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, R.; Liu, C.; Chen, Z. Imprinted composite membranes. Prog. Chem. 2020, 32, 989–1002. [Google Scholar]
  18. Donato, L. A Spasso con le Membrane. Un Mondo Tutto da Scoprire; CNR Edizioni: Roma, Italy, 2021. [Google Scholar]
  19. Nascimento, T.A.; Fdz-Polanco, F.; Peña, M. Membrane-based technologies for the up-concentration of municipal wastewater: A review of pretreatment intensification. Sep. Purif. Rev. 2020, 49, 1–19. [Google Scholar] [CrossRef]
  20. Piacentini, E.; Mazzei, R.; Drioli, E.; Giorno, L. From biological membranes to artificial biomimetic membranes and systems. In Comprehensive Membrane Science and Engineering, 2nd ed.; Drioli, E., Giorno, L., Fontananova, E., Eds.; Elsevier: Kidlington, UK, 2017; pp. 1–16. [Google Scholar]
  21. Didaskalou, C.; Buyuktiryaki, S.; Kecili, R.; Fonte, C.P.; Szekely, G. Valorisation of agricultural waste with an adsorption/nanofiltration hybrid process: From materials to sustainable process design. Green Chem. 2017, 19, 3116–3125. [Google Scholar] [CrossRef] [Green Version]
  22. Rajesha, K.; Arun, M.I. Handbook of Membrane Separations: Chemical, Pharmaceutical, Food, and Biotechnological Applications, 2nd ed.; CRC Press: London, UK, 2015; pp. 465–481. [Google Scholar]
  23. Kubota, N.; Hashimoto, T.; Mori, Y. Advanced Membrane Technology and Applications, 1st ed.; John Wiley & Sons Inc.: New York, NY, USA, 2008; pp. 101–129. [Google Scholar]
  24. Lu, J.; Qin, Y.; Wu, Y.; Meng, M.; Yan, Y.; Li, C. Recent advances in ion imprinted membranes: Separation and detection via ion-selective recognition. Environ. Sci. Water Res. Technol. 2019, 5, 1626–1653. [Google Scholar] [CrossRef]
  25. Boysen, R.I.; Schwarz, L.J.; Nicolau, D.V.; Hearn, M.T.W. Molecularly imprinted polymer membranes and thin films for the separation and sensing of biomacromolecules. J. Sep. Sci. 2017, 40, 314–335. [Google Scholar] [CrossRef]
  26. Yoshikawa, M.; Tharpa, K.; Dima, S.O. Molecularly Imprinted Membranes: Past, Present, and Future. Chem. Rev. 2016, 116, 11500–11528. [Google Scholar] [CrossRef] [PubMed]
  27. Keçili, R.; Yılmaz, E.; Ersöz, A.; Say, R. Sustainable Nanoscale Engineering: From Materials Design to Chemical Processing, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  28. Anastas, P.T.; Warner, J.C. Principles of green chemistry. In Green Chemistry: Theory and Practice; Anastas, P.T., Warner, J.C., Eds.; Oxford University Press: Oxford, UK, 1998; pp. 29–56. [Google Scholar]
  29. Koenig, S.G. Scalable Green Chemistry: Case Studies from the Pharmaceutical Industry, 1st ed.; Taylor and Francis Group, CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
  30. Dicks, A.P.; Hent, A. Green Chemistry Metrics. A Guide to Determining and Evaluating Process Greenness; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  31. Andraos, J. The Algebra of Organic Synthesis: Green Metrics, Design Strategy, Route Selection and Optimization; Taylor and Francis Group, CRC Press: Boca Raton. FL, USA, 2012. [Google Scholar]
  32. Andraos, J. Unification of reaction metrics for green chemistry: Applications to reaction analysis. Org. Process. Res. Dev. 2005, 9, 149–163. [Google Scholar] [CrossRef]
  33. Lapkin, A.; Constable, D.J.C. Green Chemistry Metrics: Measuring and Monitoring Sustainable Processes; Wiley-Blackwell: Chichester, UK, 2008. [Google Scholar]
  34. Li, C.-J.; Chen, L. Organic chemistry in water. Chem. Soc. Rev. 2006, 35, 68–82. [Google Scholar]
  35. Li, C.-J. Organic Reactions in Aqueous Media with a Focus on Carbon−Carbon Bond Formations:  A Decade Update. Chem. Rev. 2005, 105, 3095–3166. [Google Scholar]
  36. Díaz-Álvarez, A.E.; Francos, J.; Crochet, P.; Cadierno, V. Recent advances in the use of glycerol as Green solvent for synthetic organic chemistry. Curr. Green Chem. 2014, 1, 51–65. [Google Scholar] [CrossRef]
  37. Santosh, R.M.; Ganapati, D.Y. Effect of Supercritical CO2 as Reaction Medium for Selective Hydrogenation of Acetophenone to 1-Phenylethanol. ACS Omega 2018, 3, 7124–7132. [Google Scholar]
  38. Wood, C.D.; Cooper, A.I.; Desimone, J.M. Green synthesis of polymers using supercritical carbon dioxide. Curr. Opin. Solid State Mater. Sci. 2004, 8, 325–331. [Google Scholar] [CrossRef]
  39. Samantha, L.P.; Mega, K.; Douglas, R.M.; Karolina, M.; Jennifer, M.P. Ionic liquids for renewable thermal energy storage—A perspective. Green Chem. 2022, 24, 102–117. [Google Scholar]
  40. Lei, Z.; Chen, B.; Koo, Y.M.; MacFarlane, D.R. Introduction: Ionic Liquids. Chem. Rev. 2017, 117, 6633–6635. [Google Scholar] [CrossRef] [Green Version]
  41. Chatel, G.; MacFarlane, D.R. Ionic Liquids and Ultrasound in Combination: Synergies and Challenges. Chem. Soc. Rev. 2014, 43, 8132–8149. [Google Scholar] [CrossRef]
  42. Alonso, D.A.; Baeza, A.; Chinchilla, R.; Guillena, G.; Pastor, I.M.; Ramón, D.J. Deep Eutectic Solvents: The Organic Reaction Medium of the Century. Eur. J. Org. Chem. 2016, 4, 612–632. [Google Scholar] [CrossRef] [Green Version]
  43. Li, X.; Row, K.H. Development of deep eutectic solvents applied in extraction and separation. J. Sep. Sci. 2016, 39, 3505–3520. [Google Scholar] [CrossRef]
  44. Azizi, N.; Dezfooli, S.; Mahmoudi, M. Greener synthesis of spirooxindole in deep eutectic solvent. J. Mol. Liq. 2014, 194, 62–67. [Google Scholar] [CrossRef]
  45. O’Neal, K.L.; Zhang, H.; Yang, Y.; Hong, L.; Lu, D.; Weber, S.G. Fluorous media for extraction and transport. J. Chromatogr. A 2010, 1217, 2287–2295. [Google Scholar] [CrossRef] [PubMed]
  46. Dąbrowska, S.; Chudoba, T.; Wojnarowicz, J.; Łojkowski, W. Current Trends in the Development of Microwave Reactors for the Synthesis of Nanomaterials in Laboratories and Industries: A Review. Crystals 2018, 8, 379. [Google Scholar] [CrossRef] [Green Version]
  47. Roberts, B.A.; Strauss, C.R. Toward Rapid, “Green”, Predictable Microwave-Assisted Synthesis. Acc. Chem. Res. 2005, 38, 653–661. [Google Scholar] [CrossRef]
  48. Cravotto, G.; Cintas, P. Power Ultrasound in Organic Synthesis: Moving Cavitational Chemistry from Academia to Innovative and Large-Scale Applications. Chem. Soc. Rev. 2006, 35, 180–196. [Google Scholar] [CrossRef] [PubMed]
  49. Ye, N.; Yan, T.; Jiang, Z.; Wu, W.; Fang, T. A review: Conventional and supercritical hydro/solvo thermal synthesis of ultra-fine particles as cathode in lithium battery. Ceram. Int. 2018, 44, 4521–4537. [Google Scholar] [CrossRef]
  50. Darr, J.A.; Zhang, J.; Makwana, N.M.; Weng, X. Continuous hydrothermal synthesis of inorganic nanoparticles: Applications and future directions. Chem. Rev. 2017, 117, 11125–11238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Marcin, K. Magnetic-field-induced synthesis of magnetic wire-like micro- and nanostructures. Nanoscale 2017, 9, 16511–16545. [Google Scholar]
  52. Mateti, S.; Mathesh, M.; Liu, Z.; Tao, T.; Ramireddy, T.; Glushenkov, A.M.; Chen, Y.I. Mechanochemistry: A force in disguise and conditional effects towards chemical reactions. Chem. Commun. 2021, 57, 1080–1092. [Google Scholar] [CrossRef] [PubMed]
  53. Glasing, J.; Champagne, P.; Cunningham, M.F. Current opinion in Green and sustainable chemistry graft modification of chitosan, cellulose and alginate using reversible deactivation radical polymerization (RDRP). Curr. Opin. Green Sustain. Chem. 2016, 2, 15–21. [Google Scholar] [CrossRef]
  54. Ardila Arias, A.N.; Arriola, E.; Reyes Calle, J.; Berrio Mesa, E.; Fuentes Zurita, G. Mineralización de etilenglicol por foto-fenton asistido con ferrioxalato. Rev. Int. Contam. Ambient. 2016, 32, 213–226. [Google Scholar] [CrossRef] [Green Version]
  55. Scala, A.; Neri, G.; Micale, N.; Cordaro, M.; Piperno, A. State of the Art on Green Route Synthesis of Gold/Silver Bimetallic Nanoparticles. Molecules 2022, 27, 1134. [Google Scholar] [CrossRef]
  56. Kurniawan, Y.S.; Priyangga, K.T.A.; Krisbiantoro, P.A.; Imawan, A.C. Green Chemistry Influences in Organic Synthesis: A Review. J. Multidiscip. Appl. Nat. Sci. 2021, 1, 1–12. [Google Scholar] [CrossRef]
  57. Pena-Pereira, F.; Lavilla, I.; Bendicho, C. Greening sample preparation: An overview of cutting-edge contributions. Curr. Opin. Green Sustain. Chem. 2021, 30, 10048. [Google Scholar] [CrossRef]
  58. Aparecida de Marco, B.; Saú Rechelo, B.; Gandolpho Tótoli, E.; Kogawa, A.C.; Nunes Salgado, H.R. Evolution of green chemistry and its multidimensional impacts: A review. Saudi Pharm. J. 2019, 27, 1–8. [Google Scholar] [CrossRef]
  59. Rente, D.; Paiva, A.; Duarte, A.R. The Role of Hydrogen Bond Donor on the Extraction of Phenolic Compounds from Natural Matrices Using Deep Eutectic Systems. Molecules 2021, 26, 2336. [Google Scholar] [CrossRef]
  60. Viveiros, R.; Rebocho, S.; Casimiro, T. Green Strategies for Molecularly Imprinted Polymer Development. Polymers 2018, 10, 306. [Google Scholar] [CrossRef] [Green Version]
  61. Horváth, I.T.; Anastas, P.T. Innovations and Green Chemistry. Chem. Rev. 2007, 107, 2169–2173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Salehi, E.; Khajavian, M.; Sahebjamee, N.; Mahmoudi, M.; Drioli, E.; Matsuura, T. Advances in nanocomposite and nanostructured chitosan membrane adsorbents for environmental remediation: A review. Desalination 2022, 527, 115565. [Google Scholar] [CrossRef]
  63. Zou, D.; Nunes, P.S.; Vankelecom, I.F.J.; Figoli, A.; Lee, Y.M. Recent advances in polymer membranes employing non-toxic solvents and materials. Green Chem. 2021, 23, 9815–9843. [Google Scholar] [CrossRef]
  64. Stefaniak, K.; Masek, A. Green Copolymers Based on Poly(Lactic Acid)—Short Review. Materials 2021, 14, 5254. [Google Scholar] [CrossRef]
  65. Shanxue, J.; Bradley, P.L. Green synthesis of polymeric membranes: Recent advances and future prospects. Curr. Opin. Green Sustain. Chem. 2020, 21, 1–8. [Google Scholar]
  66. Galiano, F.; Briceño, K.; Marino, T.; Molino, A.; Christensen, K.V.; Figoli, A. Advances in biopolymer-based membrane preparation and applications. J. Membr. Sci 2018, 564, 562–586. [Google Scholar] [CrossRef]
  67. Baatout, Z.; Teka, S.; Jaballah, N.; Sakly, N.; Sun, X.; Maurel, F.; Majdoub, M. Water-insoluble cyclodextrin membranes for humidity detection: Green synthesis, characterization and sensing performances. J. Mater. Sci. 2018, 53, 1455–1469. [Google Scholar] [CrossRef]
  68. Dan, L.V.; Miaomiao, Z.; Zhicheng, J.; Shaohua, J.; Qilu, Z.; Ranhua, X.; Chaobo, H. Green Electrospun Nanofibers and Their Application in Air Filtration. Macromol. Mater. Eng. 2018, 303, 1800336. [Google Scholar]
  69. Marino, T.; Blefari, S.; Di Nicolò, E.; Figoli, A. A more sustainable membrane preparation using triethyl phosphate as solvent. Green Processing Synth. 2017, 6, 295–300. [Google Scholar] [CrossRef]
  70. Mohshim, D.F.; Hilmi, M.; Man, Z. Composite blending of ionic liquid–poly(ether sulfone) polymeric membranes: Green materials with potential for carbon dioxide/methane separation. J. Appl. Polym. Sci. 2016, 133, 43999. [Google Scholar] [CrossRef]
  71. Li, S.; Qin, F.; Qin, P.; Karim, M.N.; Tan, T. Preparation of PDMS membrane using water as solvent for pervaporation separation of butanol-water mixture. Green Chem. 2013, 15, 2180–2190. [Google Scholar] [CrossRef]
  72. Worthington, M.J.; Kucera, R.L.; Chalker, J.M. Green chemistry and polymers made from sulfur. Green Chem. 2017, 19, 2748–2761. [Google Scholar] [CrossRef] [Green Version]
  73. Hu, J.; Zhang, C.; Zhang, X.; Chen, L.; Jiang, L.; Meng, Y.; Wang, X. A green approach for preparing anion exchange membrane based on cardo polyetherketone powders. J. Power Sources 2014, 272, 211–217. [Google Scholar] [CrossRef]
  74. Figoli, A.; Marino, T.; Simone, S.; Di Nicolò, E.; Li, X.M.; He, T.; Tornaghi, S.; Drioli, E. Towards non-toxic solvents for membrane preparation: A review. Green Chem. 2014, 16, 4034–4059. [Google Scholar] [CrossRef]
  75. Barroso, T.; Temtem, M.; Casimiro, T.; Aguiar-Ricardo, A. Antifouling performance of poly(acrylonitrile)-based membranes: From green synthesis to application. J. Supercrit. Fluids 2011, 56, 312–321. [Google Scholar] [CrossRef]
  76. Srivastva, A.N.; Saxena, N.; Kumar, M. Green Polymers Decorated with Metal Nanocomposites: Application in Energy Storage, Energy Economy and Environmental Safety. In Metal Nanocomposites for Energy and Environmental Applications; Springer: Singapore, 2022; pp. 269–292. [Google Scholar]
  77. Yang, G.; Kong, H.; Chen, Y.; Liu, B.; Zhu, D.; Guo, L.; Wei, G. Recent advances in the hybridization of cellulose and carbon nanomaterials: Interactions, structural design, functional tailoring, and applications. Carbohydr. Polym 2022, 279, 118947. [Google Scholar]
  78. Tan, N.P.B.; Lee, C.H.; Li, P. Green synthesis of smart metal/polymer nanocomposite particles and their tuneable catalytic activities. Polymers 2016, 8, 105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Bibi, S.; Yasin, T.; Hassan, S.; Riaz, M.; Nawaz, M. Chitosan/CNTs green nanocompositemembrane: Synthesis, swelling and polyaromatic hydrocarbons removal. Mater. Sci. Eng. C 2015, 46, 359–365. [Google Scholar] [CrossRef]
  80. Ghasemlou, M.; Daver, F.; Ivanova, E.P.; Adhikari, B. Synthesis of green hybrid materials using starch and non-isocyanate polyurethanes. Carbohydr. Polym. 2020, 229, 115535. [Google Scholar] [CrossRef] [PubMed]
  81. Chen, G.; Lu, J.; Yu, Y. A novel green synthesis approach for polymer nanocomposites decorated with silver nanoparticles and their antibacterial activity. Analyst 2014, 139, 5793–5799. [Google Scholar] [CrossRef] [PubMed]
  82. Iles, A.; Martin, A.; Rosen, C.M. Undoing chemical industry lock-ins: Polyvinyl chloride and Green chemistry. Int. J. Philos. Chem. 2017, 23, 29–60. [Google Scholar]
  83. Kolya, H.; Jana, S.; Tripathy, T. Green synthesis of graft copolymers based on starch and acrylic monomers by solid phase polymerization technique. Am. J. Polym. Sci. Eng. 2016, 4, 103–110. [Google Scholar]
  84. Levina, M.A.; Miloslavskii, D.G.; Pridatchenko, M.L.; Gorshkov, A.V.; Shashkova, V.T.; Gotlib, E.M.; Tiger, R.P. Green chemistry of polyurethanes: Synthesis, structure, and functionality of triglycerides of soybean oil with epoxy and cyclocarbonate groups—Renewable raw materials for new urethanes. Polym. Sci. Ser. B 2015, 57, 584–592. [Google Scholar] [CrossRef]
  85. Wang, Z.; Ganewatta, M.S.; Tang, C. Sustainable polymers from biomass: Bridging chemistry with materials and processing. Prog. Polym. Sci. 2020, 101, 101197. [Google Scholar] [CrossRef]
  86. Scholten, P.B.V.; Detrembleur, C.; Michael, A.R.M. Plant-Based Nonactivated Olefins: A New Class of Renewable Monomers for Controlled Radical Polymerization. ACS Sustain. Chem. Eng. 2019, 7, 2751–2762. [Google Scholar] [CrossRef]
  87. Cinelli, P.; Anguillesi, I.; Lazzeri, A. Green synthesis of flexible polyurethane foams from liquefied lignin. Eur. Polym. J. 2013, 49, 1174–1184. [Google Scholar] [CrossRef]
  88. Singh, P.; Quraishi, M.A.; Ebenso, E.E. Microwave assisted Green synthesis of Bisphenol polymer containing piperazine as a corrosion inhibitor for mild steel in 1 M HCl. Int. J. Electrochem. Sci. 2013, 8, 10890–10902. [Google Scholar]
  89. Perotto, G.; Ceseracciu, L.; Simonutti, R.; Paul, U.C. Bioplastics from vegetable waste via an eco-friendly water-based process. Green Chem. 2018, 20, 894–902. [Google Scholar] [CrossRef]
  90. Bayer, I.S.; Guzman-Puyol, S.; Heredia-Guerrero, J.A.; Ceseracciu, L.; Pignatelli, F.; Ruffilli, R.; Cingolani, R.; Athanassio, A. Direct Transformation of Edible Vegetable Waste into Bioplastics. Macromolecules 2014, 47, 5135–5143. [Google Scholar] [CrossRef]
  91. Marelli, B.; Patel, N.; Duggan, T.; Perotto, G.; Shirman, E.; Li, C.; Kaplan, D.L.; Omenetto, F.G. Programming function into mechanical forms by directed assembly of silk bulk materials. Proc. Natl. Acad. Sci. USA 2017, 114, 451–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Shen, L.; Worrell, E.; Patel, M. Present and future development in plastics from biomass. Biofuels Bioprod. Biorefining 2010, 4, 25–40. [Google Scholar] [CrossRef]
  93. Kiser, B. Circular economy: Getting the circulation going. Nature 2016, 531, 443–446. [Google Scholar]
  94. Pfaltzgraff, L.A.; De Bruyn, M.; Cooper, E.C.; Budarin, V.; Clark, J.H. Food waste biomass: A resource for high-value chemicals. Green Chem. 2013, 15, 307–314. [Google Scholar] [CrossRef]
  95. Szekely, G.; Jimenez-Solomon, M.F.; Marchetti, P.; Kim, J.F.; Livingston, A.G. Sustainability assessment of organic solvent nanofiltration: From fabrication to application. Green Chem. 2014, 16, 4440–4473. [Google Scholar] [CrossRef] [Green Version]
  96. Soroko, I.; Bhole, Y.; Livingston, A.G. Environmentally friendly route for the preparation of solvent resistant polyimide nanofiltration membranes. Green Chem. 2011, 13, 162–168. [Google Scholar] [CrossRef]
  97. Zhu, Y.; Wang, Z.; Zhang, C.; Wang, J.; Wang, S. A novel membrane prepared from sodium alginate cross-linked with sodium tartrate for CO2 capture. Chin. J. Chem. Eng. 2013, 21, 1098–1105. [Google Scholar] [CrossRef]
  98. Ishikawa, K.; Ueyama, Y.; Mano, T.; Koyama, T.; Suzuki, K.; Matsumura, T. Self-setting barrier membrane for guided tissue regeneration method: Initial evaluation of alginate membrane made with sodium alginate and calcium chloride aqueous solutions. J. Biomed. Mater. Res. 1999, 47, 111–115. [Google Scholar] [CrossRef]
  99. Uranga, J.; Nguyen, B.T.; Si, T.T.; Guerrero, P.; De la Caba, K. The Effect of Cross-Linking with Citric Acid on the Properties of Agar/Fish Gelatin Films. Polymers 2020, 12, 291. [Google Scholar] [CrossRef] [Green Version]
  100. Vanherck, K.; Koeckelberghs, G.; Vankelecom, I.F.J. Crosslinking polyimides for membrane applications: A review. Prog. Polym. Sci. 2013, 38, 874–896. [Google Scholar] [CrossRef]
  101. Gu, Y.; Jérôme, F. Bio-based solvents: An emerging generation of fluids for the design of eco-efficient processes in catalysis and organic chemistry. Chem. Soc. Rev. 2013, 42, 9550. [Google Scholar] [CrossRef] [PubMed]
  102. Ismalaj, E.; Strappaveccia, G.; Ballerini, E.; Elisei, F.; Piermatti, O.; Gelman, D.; Vaccaro, L. γ-Valerolactone as a Renewable Dipolar Aprotic Solvent Deriving from Biomass Degradation for the Hiyama Reaction. ACS Sust. Chem. Eng. 2014, 2, 2461–2464. [Google Scholar]
  103. Cheng, H.N.; Dowd, M.K.; Selling, G.W.; Biswas, A. Synthesis of cellulose acetate from cotton byproducts. Carbohydr. Polym. 2010, 8, 449–452. [Google Scholar] [CrossRef]
  104. Abdul Khalil, H.P.S.; Bhat, A.H.; Ireana Yusra, A.F. Green composites from sustainable cellulose nanofibrils: A review. Carbohydr. Polym. 2012, 87, 963–979. [Google Scholar] [CrossRef]
  105. Medina-Gonzalez, Y.; Aimar, P.; Lahitte, J.F.; Remigy, J.C. Towards green membranes: Preparation of cellulose acetate ultrafiltration membranes using methyl lactate as a biosolvent. Int. J. Sustain. Eng. 2011, 4, 75–83. [Google Scholar] [CrossRef] [Green Version]
  106. Sukma, F.M.; Çulfaz-Emecen, P.Z. Cellulose membranes for organic solvent nanofiltration. J. Membr. Sci. 2018, 545, 329–336. [Google Scholar] [CrossRef]
  107. Zhang, Y.; Shao, H.; Wu, C.; Hu, X. Formation and Characterization of Cellulose Membranes from N-Methylmorpholine-N-oxide Solution. Macromol. Biosci. 2001, 1, 141–148. [Google Scholar] [CrossRef]
  108. Li, X.L.; Zhu, L.P.; Zhu, B.K.; Xu, Y.Y. High-flux and anti-fouling cellulose nanofiltration membranes prepared via phase inversion with ionic liquid as solvent. Sep. Purif. Technol. 2011, 83, 66–73. [Google Scholar] [CrossRef]
  109. Sivakumar, M.; Mohan, D.R.; Rangarajan, R. Studies on cellulose acetate-polysulfone ultrafiltration membranes: II. Effect of additive concentration. J. Membr. Sci. 2006, 268, 208–219. [Google Scholar] [CrossRef]
  110. Uppal, N.; Pappu, A.; Gowri, V.K.S.; Thakur, V.K. Cellulosic fibres-based epoxy composites: From bioresources to a circular economy. Ind. Crop. Prod. 2022, 182, 114895. [Google Scholar] [CrossRef]
  111. He, H.; Cheng, M.; Liang, Y.; Zhu, H.; Sun, Y.; Dong, D.; Wang, S. Intelligent cellulose nanofibers with excellent biocompatibility enable sustained antibacterial and drug release via a pH-responsive mechanism. J. Agric. Food Chem. 2020, 68, 3518–3527. [Google Scholar] [CrossRef] [PubMed]
  112. Green Chemicals Market Size, Trends: Industry Report, 2020–2027. Available online: https://dataintelo.com/report/green-chemicals-market (accessed on 21 February 2022).
  113. Parisi, O.I.; Francomano, F.; Dattilo, M.; Patitucci, F.; Prete, S.; Amone, F.; Puoci, F. The Evolution of Molecular Recognition: From Antibodies to Molecularly Imprinted Polymers (MIPs) as Artificial Counterpart. J. Funct. Biomater. 2022, 13, 12. [Google Scholar] [CrossRef] [PubMed]
  114. Bhogal, S.; Kaur, K.; Mohiuddin, I.; Kumar, S.; Lee, J.; Brown, R.J.; Malik, A.K. Hollow porous molecularly imprinted polymers as emerging adsorbents. Environ. Pollut. 2021, 288, 117775. [Google Scholar] [CrossRef]
  115. Haupt, K.; Medina Rangel, P.X.; Bui, B.T.S. Molecularly Imprinted Polymers: Antibody Mimics for Bioimaging and Therapy. Chem. Rev. 2020, 120, 9554–9582. [Google Scholar] [CrossRef] [PubMed]
  116. Haupt, K. Molecular Imprinting; Springer Science & Business Media: Berlin, Germany, 2012; p. 325. [Google Scholar]
  117. Ylmnaz, E.; Schmidt, R.H.; Mosbach, K. Molecularly Imprinted Materials: Science and Technology, 1st ed.; Taylor & Francis Group: Boca Raton, FL, USA, 2005; pp. 25–57. [Google Scholar]
  118. Wackerlig, J.; Schirhagl, R. Applications of Molecularly Imprinted Polymer Nanoparticles and Their Advances toward Industrial Use: A Review. Anal. Chem. 2016, 88, 250–261. [Google Scholar] [CrossRef] [PubMed]
  119. Huang, C.; Wang, H.; Ma, S.; Bo, C.; Ou, J.; Gong, B. Recent application of molecular imprinting technique in food safety. J. Chromatogr. A 2021, 1657, 462579. [Google Scholar] [CrossRef]
  120. Arabi, M.; Ostovan, A.; Bagheri, A.R.; Guo, X.; Wang, L.; Li, J. Strategies of molecular imprinting-based solid-phase extraction prior to chromatographic analysis. Trends Anal. Chem. 2020, 128, 115923. [Google Scholar] [CrossRef]
  121. Cui, Y.; Kang, W.; Qin, L.; Ma, J.; Liu, X.; Yang, Y. Magnetic surface molecularly imprinted polymer for selective adsorption of quinoline from coking wastewater. Chem. Eng. J. 2020, 397, 125480. [Google Scholar] [CrossRef]
  122. Yang, X.; Zhang, Z.; Li, J.; Chen, X.; Zhang, M.; Luo, L.; Yao, S. Novel molecularly imprinted polymers with carbon nanotube as matrix for selective solid-phase extraction of emodin from kiwi fruit root. Food Chem. 2014, 145, 687–693. [Google Scholar] [CrossRef]
  123. Chen, W.; Ma, Y.; Pan, J.; Meng, Z.; Pan, G.; Sellergren, B. Molecularly imprinted polymers with stimuli-responsive affinity: Progress and perspectives. Polymers 2015, 7, 1689–1715. [Google Scholar] [CrossRef] [Green Version]
  124. Crapnell, R.D.; Hudson, A.; Foster, C.W.; Eersels, K.; Van Grinsven, B.; Cleij, J.C. Recent advances in electrosynthesized molecularly imprinted polymer-sensing platforms for bioanalyte detection. Sensors 2019, 19, 1204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Gui, R.; Jin, H. Recent advances in synthetic methods and applications of photo-luminescent molecularly imprinted polymers. J. Photochem. Photobiol. C Photochem. Rev. 2019, 41, 100315. [Google Scholar] [CrossRef]
  126. Song, X.; Xu, S.; Chen, L.; Wei, Y.; Xiong, H. Recent advances in molecularly imprinted polymers in food analysis. Appl. Polym. Sci. 2014, 131, 40766. [Google Scholar] [CrossRef] [Green Version]
  127. Maier, N.M.; Lindner, W. Chiral recognition applications of molecularly imprinted polymers: A critical review. Anal. Bioanal. Chem. 2007, 389, 377–397. [Google Scholar] [CrossRef]
  128. Tuwahatua, C.A.; Yeung, C.C.; Lam, Y.W.; Arul, V.R.L. The molecularly imprinted polymer essentials: Curation of anticancer, ophthalmic, and projected gene therapy drug delivery systems. J. Control Release 2018, 287, 24–34. [Google Scholar] [CrossRef] [PubMed]
  129. Pijush, K.P.; Alongkot, T.; Roongnapa, S. Biomimetic insulin-imprinted polymer nanoparticles as a potential oral drug delivery system. Acta Pharm. 2017, 67, 149–168. [Google Scholar]
  130. Erythropel, H.C.; Zimmerman, J.B.; de Winter, T.M.; Petitjean, L.; Melnikov, F.; Lam, C.H.; Lounsbury, A.W.; Mellor, K.E.; Janković, N.Z.; Tu, Q.; et al. The Green ChemisTREE: 20 years after taking root with the 12 principles. Green Chem. 2018, 20, 1929–1961. [Google Scholar]
  131. Del Blanco, S.G.; Donato, L.; Drioli, E. Development of molecularly imprinted membranes for selective recognition of primary amines in organic medium. Sep. Purif. Technol. 2012, 87, 40–46. [Google Scholar] [CrossRef]
  132. Arabi, M.; Ostovan, A.; Li, J.; Wang, X.; Zhang, Z.; Choo, J.; Chen, L. Molecular Imprinting: Green Perspectives and Strategies. Adv. Mater. 2021, 33, 2100543. [Google Scholar] [CrossRef]
  133. Wu, X.; Du, J.; Li, M.; Wu, L.; Han, C.; Su, F. Recent advances in green reagents for molecularly imprinted polymers. RSC Adv. 2018, 8, 311–327. [Google Scholar]
  134. Ding, S.; Lyu, Z.; Niu, X.; Zhou, Y.; Liu, D.; Falahati, M.; Du, D.; Lin, Y. Integrating ionic liquids with molecular imprinting technology for biorecognition and biosensing: A review. Biosens. Bioelectron. 2020, 149, 111830. [Google Scholar] [CrossRef] [PubMed]
  135. Chang, X.X.; Mubarak, N.M.; Mazari, S.A.; Jatoi, A.S.; Ahmed, A.; Khalid, M.; Walvekar, R.; Abdullah, E.C.; Karri, R.R.; Siddiqui, M.T.H.; et al. A review on the properties and applications of chitosan, cellulose and deep eutectic solvent in green chemistry. J. Ind. Eng. Chem. 2021, 104, 362–380. [Google Scholar] [CrossRef]
  136. Liu, H.; Jin, P.; Zhu, F.; Nie, L.; Qiu, H. A review on the use of ionic liquids in preparation of molecularly imprinted polymers for applications in solid-phase extraction. Trends Anal. Chem. 2021, 134, 116132. [Google Scholar] [CrossRef]
  137. Booker, K.; Bowyer, M.C.; Lennard, C.J.; Holdsworth, C.I.; McCluskey, A. Molecularly imprinted polymers and room temperature ionic liquids: Impact of template on polymer morphology. Aust. J. Chem. 2007, 60, 51–56. [Google Scholar] [CrossRef]
  138. Guo, L.; Deng, Q.; Fang, G.; Gao, W.; Wang, S. Preparation and evaluation of molecularly imprinted ionic liquids polymer as sorbent for on-line solid-phase extraction of chlorsulfuron in environmental water samples. J. Chromatogr. A 2011, 1218, 6271–6277. [Google Scholar] [CrossRef] [PubMed]
  139. Fan, J.P.; Tian, Z.Y.; Tong, S.; Zhang, X.H.; Xie, Y.L.; Xu, R.; Qin, Y.; Li, L.; Zhu, J.H.; Ouyang, X.K. A novel molecularly imprinted polymer of the specific ionic liquid monomer for selective separation of synephrine from methanol–water media. Food Chem. 2013, 141, 3578–3585. [Google Scholar] [CrossRef]
  140. Yan, H.Y.; Liu, S.T.; Gao, M.M.; Sun, N. Ionic liquids modified dummy molecularly imprinted microspheres as solid phase extraction materials for the determination of clenbuterol and clorprenaline in urine. J. Chromatogr. A 2013, 1294, 10–16. [Google Scholar] [CrossRef]
  141. Qiao, F.; Gao, M.; Yan, H. Molecularly imprinted ionic liquid magnetic microspheres for the rapid isolation of organochlorine pesticides in environmental water. J. Sep. Sci. 2016, 39, 1310–1315. [Google Scholar] [CrossRef]
  142. Lu, X.; Yang, Y.; Zeng, Y.; Li, L.; Wu, X. Rapid and reliable determination of p-Nitroaniline in wastewater by molecularly imprinted fluorescent polymeric ionic liquid microspheres. Biosens. Bioelectron. 2018, 99, 47–55. [Google Scholar] [CrossRef] [PubMed]
  143. Wu, Y.; Wang, Y.; Wang, X.; Wang, C.; Li, C.; Wang, Z. Electrochemical sensing of alpha-fetoprotein based on molecularly imprinted polymerized ionic liquid film on a gold nanoparticle modified electrode surface. Sensors 2019, 19, 3218. [Google Scholar] [CrossRef] [Green Version]
  144. Afzali, Z.; Mohadesi, A.; Ali Karimi, M.; Fathirad, F.A. Highly selective and sensitive electrochemical sensor based on graphene oxide and molecularly imprinted polymer magnetic nanocomposite for patulin determination. Microchem. J. 2022, 177, 107215. [Google Scholar] [CrossRef]
  145. Luo, X.; Dong, R.; Luo, S.; Zhan, Y.; Tu, X.; Yang, L. Preparation of water compatible molecularly imprinted polymers for caffeine with a novel ionic liquid as a functional monomer. J. Appl. Polym. Sci. 2013, 127, 2884–2890. [Google Scholar] [CrossRef]
  146. Xiang, H.; Peng, M.; Li, H.; Peng, S.; Shi, S. High-capacity hollow porous dummy molecular imprinted polymers using ionic liquid as functional monomer for selective recognition of salicylic acid. J. Pharm. Biomed. Anal. 2017, 133, 75–81. [Google Scholar] [CrossRef]
  147. Yuan, Y.; Liang, S.; Yan, H.; Ma, Z.; Liu, Y. Ionic liquid-molecularly imprinted polymers for pipette tip solid-phase extraction of (Z)-3-(chloromethylene)-6 flourothiochroman-4-one in urine. J. Chromatogr. A 2015, 1408, 49–55. [Google Scholar] [CrossRef] [PubMed]
  148. Afzali, M.; Mostafavi, A.; Shamspur, T. Developing a novel sensor based on ionic liquid molecularly imprinted polymer/gold nanoparticles/graphene oxide for the selective determination of an anti-cancer drug imiquimod. Biosens. Bioelectron. 2019, 143, 111620. [Google Scholar] [CrossRef] [PubMed]
  149. Zhu, G.; Li, W.; Wang, L.; Wang, P.; Shi, D.; Wang, J.; Fan, J. Using ionic liquid monomer to improve the selective recognition performance of surface imprinted polymer for sulfamonomethoxine in strong polar medium. J. Chromatogr. A 2019, 1592, 38–46. [Google Scholar] [CrossRef]
  150. Gao, Z.X.Y.; Pan, M.F.; Fang, G.Z.; Jing, W.; He, S.Y.; Wang, S. An ionic liquid modified dummy molecularly imprinted polymer as a solid-phase extraction material for the simultaneous determination of nine organochlorine pesticides in environmental and food samples. Anal. Methods 2013, 5, 6128–6134. [Google Scholar] [CrossRef]
  151. Yuan, S.; Deng, Q.; Fang, G.; Wu, J.; Li, W.; Wang, S. Protein imprinted ionic liquid polymer on the surface of multiwall carbon nanotubes with high binding capacity for lysozyme. J. Chromatogr. B 2014, 960, 239–246. [Google Scholar] [CrossRef] [PubMed]
  152. Xiaopeng, H.; Yide, X.; Yiwei, L.; Yanran, C.; Baizhao, Z. An effective ratiometric electrochemical sensor for highly selective and reproducible detection of ochratoxin A: Use of magnetic field improved molecularly imprinted polymer. Sens. Actuat. 2022, 359, 131582. [Google Scholar]
  153. Zhang, D.; Tang, J.; Liu, H. Rapid determination of lambda-cyhalothrin using a fluorescent probe based on ionic-liquid-sensitized carbon dots coated with molecularly imprinted polymers. Anal. Bioanal. Chem. 2019, 411, 5309–5316. [Google Scholar] [CrossRef] [PubMed]
  154. Sun, Y.; Feng, X.; Hu, J.; Bo, S.; Zhang, J.; Wang, W.; Li, S.; Yang, Y. Preparation of hemoglobin (Hb)-imprinted poly(ionic liquid)s via Hb-catalyzed eATRP on gold nanodendrites. Anal. Bioanal. Chem. 2020, 412, 983–991. [Google Scholar] [CrossRef] [PubMed]
  155. Zhu, G.; Li, W.; Wang, P.; Cheng, G.; Chen, L.; Zhang, K.; Li, X. One-step polymerization of hydrophilic ionic liquid imprinted polymer in water for selective separation and detection of levofloxacin from environmental matrices. J. Sep. Sci. 2019, 43, 639–647. [Google Scholar] [CrossRef]
  156. Fan, J.P.; Yu, J.X.; Yang, X.M.; Zhang, X.H.; Yuan, T.T.; Peng, H.L. Preparation, characterization, and application of multiple stimuli-responsive rattle-type magnetic hollow molecular imprinted poly (ionic liquids) nanospheres (Fe3O4@void@PILMIP) for specific recognition of protein. Chem. Eng. J. 2018, 337, 722–732. [Google Scholar] [CrossRef]
  157. Zhang, X.Y.; Zhang, N.; Du, C.B.; Guan, P.; Gao, X.M.; Wang, C.Y.; Du, Y.F.; Ding, S.C.; Hu, X.L. Preparation of magnetic epitope imprinted polymer microspheres using cyclodextrin-based ionic liquids as functional monomer for highly selective and effective enrichment of cytochrome c. Chem. Eng. J. 2017, 317, 988–998. [Google Scholar] [CrossRef]
  158. Yang, G.M.; Zhao, F.Q. Molecularly imprinted polymer grown on multiwalled carbon nanotube surface for the sensitive electrochemical determination of amoxicillin. Electrochim. Acta 2015, 174, 33–40. [Google Scholar]
  159. Hamdan, S.; Moore, L., Jr.; Lejeune, L.J.; Hasan, F.; Carlisle, T.K.; Bara, J.E.; Gin, D.; LaFrate, A.L.; Noble, R.; Spivak, D.A. Ionic Liquid Cross-linkers for Chiral Imprinted NanoGUMBOS. J. Colloid Interface Sci. 2016, 463, 29–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Li, J.; Hu, X.; Guan, P.; Song, D.; Qian, L.; Du, C.; Song, R.; Wang, C. Preparation of “dummy” l-phenylalanine molecularly imprinted microspheres by using ionic liquid as a template and functional monomer. J. Sep. Sci. 2015, 38, 3279–3287. [Google Scholar] [CrossRef]
  161. Booker, K.; Holdsworth, C.I.; Doherty, C.M.; Hill, A.J.; Bowyerc, M.C.; McCluskey, A. Ionic liquids as porogens for molecularly imprinted polymers: Propranolol, a model study. Org. Biomol. Chem. 2014, 12, 7201–7210. [Google Scholar] [CrossRef] [PubMed]
  162. He, C.; Long, Y.; Pan, J.; Li, K.; Liu, F. Molecularly imprinted silica prepared with immiscible ionic liquid as solvent and porogen for selective recognition of testosterone. Talanta 2008, 74, 1126–1131. [Google Scholar] [CrossRef]
  163. Hulsbosch, J.; De Vos, D.E.; Binnemans, K.; Ameloot, R. Biobased Ionic Liquids: Solvents for a Green Processing Industry. ACS Sustain. Chem. Eng. 2016, 4, 2917–2931. [Google Scholar]
  164. Subat, M.; Borovik, A.S.; König, B. Synthetic Creatinine Receptor:  Imprinting of a Lewis Acidic Zinc(II)cyclen Binding Site to Shape Its Molecular Recognition Selectivity. J. Am. Chem. Soc. 2004, 126, 3185–3190. [Google Scholar] [CrossRef] [PubMed]
  165. Bie, Z.; Chen, Y.; Ye, J.; Wang, S.; Liu, Z. Boronate-Affinity Glycan-Oriented Surface Imprinting: A New Strategy to Mimic Lectins for the Recognition of an Intact Glycoprotein and Its Characteristic Fragments. Angew. Chem. Int. Ed. 2015, 54, 10211–10215. [Google Scholar] [CrossRef]
  166. Ostovan, A.; Ghaedi, M.; Arabi, M.; Yang, Q.; Li, J.; Chen, L. Hydrophilic multi-template molecularly imprinted biopolymers based on a green synthesis strategy for determination of B-family vitamins. Appl. Mater. Interfaces 2018, 10, 4140–4150. [Google Scholar] [CrossRef]
  167. Deng, H.; Wei, Z.; Wang, X. Enhanced adsorption of active brilliant red X-3B dye on chitosan molecularly imprinted polymer functionalized with Ti(IV) as Lewis acid. Carbohydr. Polym. 2017, 157, 1190–1197. [Google Scholar] [CrossRef]
  168. Herrero, E.P.; Del Valle, E.M.M.; Peppas, N.A. Protein Imprinting by Means of Alginate-Based Polymer Microcapsules. Ind. Eng. Chem. Res. 2010, 49, 9811–9814. [Google Scholar] [CrossRef]
  169. Sibrian-Vazquez, M.; Spivak, D.A. Molecular imprinting made easy. J. Am. Chem. Soc. 2004, 126, 7827–7833. [Google Scholar] [CrossRef] [PubMed]
  170. Panagiotopoulou, M.; Beyazit, S.; Nestora, N.; Haupt, K.; Bui, B.T.S. Initiator-free synthesis of molecularly imprinted polymers by polymerization of self-initiated monomers. Polymers 2015, 66, 43–51. [Google Scholar] [CrossRef]
  171. Wang, H.; Brown, H.R. Self-initiated photo-polymerization and photo-grafting of acrylic monomers. Macromol. Rapid Commun. 2004, 25, 1095–1099. [Google Scholar] [CrossRef]
  172. Kerton, F.M.; Marriott, R. Alternative Solvents for Green Chemistry, 2nd ed.; Royal Soc. Chem.: Cambridge, UK; Chicago, IL, USA, 2013; pp. 1–30. [Google Scholar]
  173. Thang, B.; Thang, H.; Row, K.H. Application of deep eutectic solvents in the extraction and separation of target compounds from various samples. J. Sep. Sci. 2015, 38, 1053–1064. [Google Scholar] [CrossRef] [PubMed]
  174. Karimi, M.; Dadfarnia, S.; Shabani, A.M.H.; Tamaddon, F.; Azadi, D. Deep eutectic liquid organic salt as a new solvent for liquid-phase microextraction and its application in ligand less extraction and preconcentraion of lead and cadmium in edible oils. Talanta 2015, 144, 648–654. [Google Scholar] [CrossRef]
  175. Zhang, Q.H.; Vigier, K.D.O.; Royer, S.; Jerome, F. Deep eutectic solvents: Syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 7108–7146. [Google Scholar] [CrossRef] [PubMed]
  176. Madikizela, L.M.; Tavengwa, N.T.; Tutu, H.; Chimuka, L. Green aspects in molecular imprinting technology: From design to environmental applications. Trends Environ. Anal. Chem. 2018, 17, 14–22. [Google Scholar] [CrossRef]
  177. Liu, Y.; Wang, Y.; Dai, Q.; Zhou, Y. Magnetic deep eutectic solvents molecularly imprinted polymers for the selective recognition and separation of protein. Anal. Chem. Acta 2016, 936, 168–178. [Google Scholar] [CrossRef] [PubMed]
  178. Li, G.; Wang, W.; Wang, Q.; Zhu, T. Deep Eutectic Solvents Modified Molecular Imprinted Polymers for Optimized Purification of Chlorogenic Acid from Honeysuckle. J. Chromatogr. Sci. 2016, 54, 271–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Surapong, N.; Santaladchaiyakit, Y.; Burakham, R. A water-compatible magnetic dual-template molecularly imprinted polymer fabricated from a ternary biobased deep eutectic solvent for the selective enrichment of organophosphorus in fruits and vegetables. Food Chem. 2022, 384, 132475. [Google Scholar] [CrossRef] [PubMed]
  180. Sun, Y.; Yao, C.; Zeng, J.; Zhang, Y.; Zhang, Y. Eco-friendly deep eutectic solvents skeleton patterned molecularly imprinted polymers for the separation of sinapic acid from agricultural wastes. Colloids Surf. A Physicochem. Eng. Asp. 2022, 640, 128441. [Google Scholar] [CrossRef]
  181. Li, X.J.; Chen, X.X.; Sun, G.Y.; Zhao, Y.X.; Aisa, H.A. Green synthesis and evaluation of isoquercitrin imprinted polymers for class-selective separation and purification of flavonol glycosides. Anal. Methods 2015, 7, 4717–4724. [Google Scholar] [CrossRef]
  182. Pan, S.D.; Shen, H.Y.; Zhou, L.X.; Chen, X.H.; Zhao, Y.G.; Cai, M.Q.; Jin, M.C. Controlled synthesis of pentachlorophenol-imprinted polymers on the surface of magnetic graphene oxide for highly selective adsorption. J. Mater. Chem. A 2014, 2, 15345–15356. [Google Scholar] [CrossRef]
  183. Boyère, C.; Jérôme, C.; Debuigne, A. Input of supercritical carbon dioxide to polymer synthesis: An overview. Eur. Polym. J. 2014, 61, 45–63. [Google Scholar] [CrossRef]
  184. Da Silva, M.S.; Viveiros, R.; Aguiar-Ricardo, A.; Bonifácio, V.D.B.; Casimiro, T. Supercritical fluid technology as a new strategy for the development of semi-covalent molecularly imprinted materials. RSC Adv. 2012, 2, 5075–5079. [Google Scholar] [CrossRef]
  185. Zhang, Y.; Li, Y.; Hu, Y.; Li, G.; Chen, Y. Preparation of magnetic indole-3-acetic acid imprinted polymer beads with4-vinylpyridine and β-cyclodextrin as binary monomer via microwave heating initiated polymerization and their application to trace analysis of auxins in plant tissues. J. Chromatogr. A 2010, 1217, 7337–7344. [Google Scholar] [CrossRef]
  186. Zhang, Y.; Liu, R.; Hu, Y.; Li, G. Microwave Heating in Preparation of Magnetic Molecularly Imprinted Polymer Beads for Trace Triazines Analysis in Complicated Samples Microwave Heating in Preparation of Magnetic Molecularly Imprinted Polymer Beads for Trace Triazines Analysis in Complicat. Anal. Chem. 2009, 81, 967–976. [Google Scholar] [CrossRef]
  187. Dmitrienko, S.G.; Popov, S.A.; Chumichkina, Y.A.; Zolotov, Y.A. The sorption properties of polymers with molecular imprints of 2,4-dichlorophenoxy acetic acid synthesized by various methods. Russ. J. Phys. Chem. A. 2011, 85, 472–477. [Google Scholar] [CrossRef]
  188. Price, G. Ultrasonically enhanced polymer synthesis. Ultrason. Sonochem. 1996, 3, S229–S238. [Google Scholar] [CrossRef]
  189. Chen, H.; Son, S.; Zhang, F.; Yan, J.; Li, Y.; Ding, H.; Ding, L. Rapid preparation of molecularly imprinted polymers by microwave-assisted emulsion polymerization for the extraction of florfenicol in milk. J. Chromatogr. B 2015, 983, 32–38. [Google Scholar] [CrossRef] [PubMed]
  190. Hou, J.; Li, H.; Wang, L.; Zhang, P.; Zhou, T.; Ding, H.; Ding, L. Rapid microwave-assisted synthesis of molecularly imprinted polymers on carbon quantum dots for fluorescent sensing of tetracycline in milk. Talanta 2016, 146, 34–40. [Google Scholar] [CrossRef] [PubMed]
  191. Asfaram, A.; Ghaedi, M.; Dashtian, K. Ultrasound assisted combined molecularly imprinted polymer for selective extraction of nicotinamide in human urine and milk samples: Spectrophotometric determination and optimization study. Ultrason. Sonochem. 2017, 34, 640–650. [Google Scholar] [CrossRef] [PubMed]
  192. Phutthawong, N.; Pattarawarapan, M. Facile synthesis of magnetic molecularly imprinted polymers for caffeine via ultrasound-assisted precipitation polymerization. Polym. Bull. 2013, 70, 691–705. [Google Scholar] [CrossRef]
  193. Schwarz, L.J.; Potdar, M.K.; Danylec, B.; Boysen, R.I.; Hearn, M.T.W. Microwave-assisted synthesis of resveratrol imprinted polymers with enhanced selectivity. Anal. Methods 2015, 7, 150–154. [Google Scholar] [CrossRef]
  194. Xia, X.; Lai, E.P.C.; Ormeci, B. Ultrasonication-Assisted Synthesis of Molecularly Imprinted Polymer-Encapsulated Magnetic Nanoparticles for Rapid and Selective Removal of 17 β-Estradiol from Aqueous Environment. Polym. Eng. Sci. 2012, 47, 1–9. [Google Scholar]
  195. Bagheri, A.R.; Mehrorang Ghaedi, M.A.; Ostovan, A.; Wang, X.; Li, J. Dummy molecularly imprinted polymers based on a green synthesis strategy for magnetic solid-phase extraction of acrylamide in food samples. Talanta 2019, 195, 390–400. [Google Scholar] [CrossRef]
  196. Ning, F.; Qiu, T.; Wang, Q.; Peng, H.; Li, Y.; Wu, X.; Zhang, Z.; Chen, L.; Xiong, H. Dummy-surface molecularly imprinted polymers on magnetic graphene oxide for rapid and selective quantification of acrylamide in heat-processed (including fried) foods. Food Chem. 2017, 221, 1797–1804. [Google Scholar] [CrossRef] [PubMed]
  197. Xiao, X.; Yan, K.; Xu, X.; Li, G. Rapid analysis of ractopamine in pig tissues by dummy-template imprinted solid-phase extraction coupling with surface-enhanced Raman spectroscopy. Talanta 2015, 138, 40–45. [Google Scholar] [CrossRef] [PubMed]
  198. Sun, X.; Wang, J.; Li, Y.; Yang, J.; Jin, J.; Shah, S.M.; Chen, J. Novel dummy molecularly imprinted polymers for matrix solid-phase dispersion extraction of eight fluoroquinolones from fish samples. J. Chromatogr. A 2014, 1359, 1–7. [Google Scholar] [CrossRef]
  199. Wang, M.; Chang, X.; Wu, X.; Yan, H.; Qiao, F. Water-compatible dummy molecularly imprinted resin prepared in aqueous solution for green miniaturized solid-phase extraction of plant growth regulators. J. Chromatogr. A 2016, 1458, 9–17. [Google Scholar] [CrossRef]
  200. Batlokwa, B.S.; Mokgadi, J.; Nyokong, T.; Torto, N. Optimal template removal from molecularly imprinted polymers by pressurized hot water extraction. Chromatographia 2011, 73, 589–593. [Google Scholar] [CrossRef] [Green Version]
  201. Lorenzo, R.A.; Carro, A.M.; Alvarez-Lorenzo, C.; Concheiro, A. To remove or not to remove? The challenge of extracting the template to make the cavities available in molecularly imprinted polymers (MIPs). Inter. J. Mol. Sci. 2011, 12, 4327–4347. [Google Scholar] [CrossRef] [Green Version]
  202. Xie, Y.; Li, Q.; Qin, L.; Zhou, X.; Fan, Y. Multi-templates surface molecularly imprinted polymer for simultaneous and rapid determination of sulfonamides and quinolones in water: Effect of carbon-carbon double bond. Environ. Sci. Pollut. Res. 2021, 28, 54950–54959. [Google Scholar] [CrossRef] [PubMed]
  203. Lu, W.; Liu, J.; Li, J.; Wang, X.; Lv, M.; Cui, R.; Chen, L. Dual-template molecularly imprinted polymers for dispersive solid-phase extraction of fluoroquinolones in water samples coupled with high performance liquid chromatography. Analyst 2019, 144, 1292–1302. [Google Scholar] [CrossRef] [PubMed]
  204. Lu, W.; Wang, X.; Wu, X.; Liu, D.; Li, J.; Chen, L.; Zhang, X. Multi-template imprinted polymers for simultaneous selective solid-phase extraction of six phenolic compounds in water samples followed by determination using capillary electrophoresis. J. Chromatogr. A 2017, 1483, 30–39. [Google Scholar] [CrossRef] [PubMed]
  205. Fernandes, C.; Tiritan, M.E.; Pinto, M.M.M. Chiral Separation in Preparative Scale: A Brief Overview of Membranes as Tools for Enantiomeric Separation. Symmetry 2017, 9, 206. [Google Scholar] [CrossRef] [Green Version]
  206. Yang, H.; Liu, H.B.; Tang, S.Z.; Qiu, Z.D.; Zhu, H.X.; Song, Z.X. Synthesis, performance, and application of molecularly imprinted membranes: A review. J. Environ. Chem. Eng. 2021, 9, 106352. [Google Scholar] [CrossRef]
  207. Yoshikawa, M.; Hotta, N.; Kyoumura, J.; Osagawa, Y.; Aoki, T. Chiral recognition sites from carbonyldioxyglyceryl moiety by an alternative molecular imprinting. Sensors Actuators B: Chem. 2005, 104, 282–288. [Google Scholar] [CrossRef]
  208. Zaidi, S.A. Recent developments in molecularly imprinted polymer nanofibers and their applications. Anal. Methods 2015, 7, 7406–7415. [Google Scholar] [CrossRef]
  209. Algieri, C.; Drioli, E.; Guzzo, L.; Donato, L. Bio-mimetic sensors based on molecularly imprinted membranes. Sensors 2014, 14, 13863–13912. [Google Scholar] [CrossRef] [Green Version]
  210. Wu, Y.; Luj, J.; Xing, W.; Ma, F.; Gao, J.; Lin, X.; Yu, C.; Yan, M. Double-layer-based molecularly imprinted membranes for template-dependent recognition and separation: An imitated core-shell-based synergistic integration design. Chem. Eng. J. 2020, 397, 125371. [Google Scholar] [CrossRef]
  211. Kenta, S.; Masakazu, Y. Molecularly Imprinted Chitin Nanofiber Membranes: Multi-Stage Cascade Membrane Separation within the Membrane. J. Membr. Sep. Technol. 2016, 5, 103–114. [Google Scholar]
  212. Masakazu, Y. Separation with molecularly imprinted membranes. Kobunshi Ronbunshu 2014, 71, 223–241. [Google Scholar] [CrossRef]
  213. Trotta, F.; Biasizzo, M.; Caldera, F. Molecularly imprinted membranes. Membranes 2012, 2, 440–477. [Google Scholar] [CrossRef] [Green Version]
  214. Yoshikawa, M.; Guiver, M.D.; Robertson, G.P. Surface plasmon resonance studies on molecularly imprinted films. J. Appl. Polym. Sci. 2008, 110, 2826–2832. [Google Scholar] [CrossRef] [Green Version]
  215. Lu, J.; Qin, Y.; Wu, Y.; Meng, M.; Dong, Z.; Yu, C.; Yan, Y.; Li, C.; Nyarko, F.K. Bidirectional molecularly imprinted membranes for selective recognition and separation of pyrimethamine: A double-faced loading strategy. J. Membr. Sci. 2020, 101, 117917. [Google Scholar] [CrossRef]
  216. Cui, J.; Wu, Y.; Meng, M.; Lu, J.; Wang, C.; Zhao, J.; Yan, Y. Bio-inspired synthesis of molecularly imprinted nanocomposite membrane for selective recognition and separation of artemisinin. J. Appl. Polym. Sci. 2016, 133, 43405. [Google Scholar] [CrossRef]
  217. Yukun, M.; Haijun, W.; Mengyan, G. Stainless Steel Wire Mesh Supported Molecularly Imprinted Composite Membranes for Selective Separation of Ebracteolata Compound B from Euphorbia fischeriana. Molecules 2019, 24, 565. [Google Scholar]
  218. Wei, M.H.; Chen, H.Y.; Wang, S.; Jiang, W.Y.; Wang, Y.; Wu, Z.F. Synthesis and characterization of hybrid molecularly imprinted membrane with blending SiO2 nanoparticles for ferulic acid. J. Inorg. Organomet. Polym. 2017, 27, 586–597. [Google Scholar] [CrossRef]
  219. Donato, L.; Tasselli, F.; Drioli, E. Molecularly imprinted membranes with affinity properties for folic acid. Sep. Sci. Technol. 2010, 45, 2273–2279. [Google Scholar] [CrossRef]
  220. Sánchez-González, J.; Odoardi, S.; Bermejo, A.M.; Bermejo-Barrera, P.; Romolo, F.S.; Moreda-Piñeiro, A.; Strano-Rossi, S. HPLC-MS/MS combined with membrane-protected molecularly imprinted polymer micro-solid-phase extraction for synthetic cathinones monitoring in urine. Drug Test. Anal. 2019, 11, 33–44. [Google Scholar] [CrossRef] [PubMed]
  221. Odabas, M.; Uzun, L.; Baydemir, G.; Aksoy, N.H.; Acet, Ö.; Erdönmez, D. Cholesterol imprinted composite membranes for selective cholesterol recognition from intestinal mimicking solution. Colloids Surf. B: Biointerfaces 2018, 163, 266–274. [Google Scholar] [CrossRef] [PubMed]
  222. Silvestri, D.; Barbani, N.; Coluccio, M.L.; Pegoraro, C.; Giusti, P.; Cristallini, C. Poly (ethylene-co-vinyl alcohol) membranes with specific adsorption properties for potential clinical application. Sep. Sci. Technol. 2007, 42, 2829–2847. [Google Scholar] [CrossRef]
  223. Jantarat, C.; Attakitmongkol, K.; Nichsapa, S.; Sirathanarun, P.; Srivaro, S. Molecularly imprinted bacterial cellulose for sustained-release delivery of quercetin. J. Biomater. Sci. Polym. Ed. 2020, 31, 1961–1976. [Google Scholar] [CrossRef] [PubMed]
  224. Bakhshpour, M.; Yavuz, H.; Denizli, A. Controlled release of mitomycin C from PHEMAH–Cu(II) cryogel membranes. Artif. Cells Nanomed. Biotechnol. 2018, 46, 946–954. [Google Scholar] [CrossRef] [Green Version]
  225. Hu, X.; Liu, S.; Zhou, G.; Huang, Y.; Xie, Z.; Jing, X. Electrospinning of polymeric nanofibers for drug delivery applications. J. Control. Release 2014, 185, 12–21. [Google Scholar] [CrossRef]
  226. Suedee, R.; Bodhibukkana, C.; Tangthong, N.; Amnuaikit, C.; Kaewnopparat, S.; Srichana, T. Development of a reservoir-type transdermal enantioselective controlled delivery system for racemic propranolol using a molecularly imprinted polymer composite membrane. J. Control Release 2008, 129, 170–178. [Google Scholar] [CrossRef] [PubMed]
  227. Huang, Q.; Li, H.; Guo, T.; Li, S.; Shen, G.; Ban, C.; Liu, J. Chiral separation of (d,l)-lactic acid through molecularly imprinted cellulose acetate composite membrane. Cellulose 2018, 25, 3435–3448. [Google Scholar] [CrossRef]
  228. Son, S.H.; Jegal, J. Chiral Separation of D,L-Serine Racemate Using a Molecularly Imprinted Polymer Composite Membrane. J. Appl. Polym. Sci. 2007, 104, 1866–1872. [Google Scholar] [CrossRef]
  229. Donato, L.; Figoli, A.; Drioli, E. Novel composite poly (4-vinylpyridine)/polypropylene membranes with recognition properties for (S)-naproxen. J. Pharm. Biomed. Anal. 2005, 37, 1003–1008. [Google Scholar] [CrossRef] [PubMed]
  230. Lah, N.F.C.; Ahmed, A.L.; Low, S.C. Molecular imprinted membrane biosensor for pesticide detection: Perspectives and challenges. Polym. Adv. Technol. 2021, 32, 17–30. [Google Scholar]
  231. Donato, L.; Greco, M.C.; Drioli, E. Preparation of molecularly imprinted membranes and evaluation of their performance in the selective recognition of dimethoate. Desalin. Water Treat. 2011, 30, 171–177. [Google Scholar] [CrossRef]
  232. Yoshimatsu, K.; Ye, L.; Stenlund, P.; Chronakis, I.S. A simple method for preparation of molecularly imprinted nanofiber materials with signal transduction ability. Chem. Commun. 2008, 2022–2024. [Google Scholar] [CrossRef]
  233. Li, J.; Zhang, L.; Fu, C. The recognizing mechanism and selectivity of the molecularly imprinting membrane. In Molecularly Imprinted Catalysts. Principles, Syntheses, and Applications; Li, S., Cao, S., Piletsky, S.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 159–182. [Google Scholar]
  234. Ulbricht, M. Membrane separations using molecularly imprinted polymers. J. Chromatogr. B 2004, 804, 113–125. [Google Scholar] [CrossRef]
  235. Cui, J.; Xie, A.; Liu, Y.; Xue, C.; Pan, J. Fabrication of multi-functional imprinted composite membrane for selective tetracycline and oil-in-water emulsion separation. Compos. Commun. 2021, 28, 100985. [Google Scholar] [CrossRef]
  236. Yoshikawa, M.; Tanioka, A.; Matsumoto, H. Molecularly imprinted nanofiber membranes. Curr. Opin. Chem. Eng. 2011, 1, 18–26. [Google Scholar] [CrossRef]
  237. Donato, L.; Tasselli, F.; De Luca, G.; Del Blanco, S.; Drioli, E. Novel hybrid molecularly imprinted membranes for targeted 4,40-methylendianiline. Sep. Purif. Technol. 2013, 116, 184–191. [Google Scholar] [CrossRef]
  238. Xing, W.; Ma, Z.; Wang, C.; Lu, J.; Gao, J.; Yu, C. Metal-organic framework based molecularly imprinted nanofiber membranes with enhanced selective recognition and separation performance: A multiple strengthening system. Sep. Purif. Technol. 2022, 278, 119624. [Google Scholar] [CrossRef]
  239. Qu, Y.; Qin, L.; Guo, M.; Liu, X.; Yang, Y. Multilayered molecularly imprinted composite membrane based on porous carbon nanospheres/pDA cooperative structure for selective adsorption and separation of phenol. Sep. Purif. Technol. 2022, 280, 119915. [Google Scholar] [CrossRef]
  240. Dong, Z.; Lu, J.; Wu, Y.; Meng, M.; Yu, C.; Chang, S. Antifouling molecularly imprinted membranes for pretreatment of milk samples: Selective separation and detection of lincomycin. Food Chem. 2020, 333, 127477. [Google Scholar] [CrossRef]
  241. Zhang, Y.; Tan, X.; Liu, X.; Li, C.; Zeng, S.; Wang, H.; Zhang, S. Fabrication of multilayered molecularly imprinted membrane for selective recognition and separation of artemisinin. ACS Sustain. Chem. Eng. 2019, 7, 3127–3137. [Google Scholar] [CrossRef]
  242. Meng, M.; Feng, Y.; Liu, Y.; Dai, X.; Pan, J.; Yan, Y. Fabrication of submicrosized imprinted spheres attached polypropylene membrane using ‘‘two-dimensional’’ molecular imprinting method for targeted separation. Adsorpt. Sci. Technol. 2017, 35, 162–177. [Google Scholar] [CrossRef]
  243. Zhang, R.; Guo, X.; Shi, X.; Sun, A.; Wang, L.; Xiao, T. Highly permselective membrane surface modification by cold plasma-induced grafting polymerization of molecularly imprinted polymer for recognition of ptyrethroid insecticides in fish. Anal. Chem. 2014, 86, 11705–11713. [Google Scholar] [CrossRef] [PubMed]
  244. Raharjo, Y.; Fahmi, Z.M.; Wafiroh, S.; Widati, A.A.; Amanda, E.R.; Ismail, A.F. Incorporation of imprinted-zeolite to polyethersulfone/cellulose acetate membrane for creatinine removal in hemodialysis treatment. J. Teknol. 2019, 81, 137–144. [Google Scholar] [CrossRef] [Green Version]
  245. Cheng, H.; Zhu, X.; Yang, S.; Wu, Y.; Cao, Q.; Ding, Z. A pH controllable imprinted composite membrane for selective separation of podophyllotoxin and its analog. J. App. Polym. Sci. 2013, 128, 363–370. [Google Scholar] [CrossRef]
  246. Sueyoshi, Y.; Utsunomiya, A.; Yoshikawa, M.; Robertson, G.P.; Guiver, M.D. Chiral separation with molecularly imprinted polysulfonealdehyde derivatized nanofiber membranes. J. Membr. Sci. 2012, 401–402, 89–96. [Google Scholar]
  247. Barahona, F.; Turiel, E.; Martín-Esteban, A. Supported liquid membrane-protected molecularly imprinted fibre for solid-phase microextraction of thiabendazole. Anal. Chim. Acta. 2011, 694, 83–89. [Google Scholar] [CrossRef] [PubMed]
  248. Amor, F.I.E.H.; Nasser, I.I.; Baatout, Z.; Teka, S.; Jaballah, N.; Donato, L.; Majdoub, M.; Ahmed, C. New Polymer Inclusion Membrane Containing Modified β-Cyclodextrin: Application to Molecular Facilitated Transport. J. App. Chem. 2016, 5, 435–445. [Google Scholar]
  249. Xu, L.; Huang, Y.A.; Zhu, Q.J.; Ye, C. Chitosan in molecularly-imprinted polymers: Current and future prospects. Int. J. Mol. Sci. 2015, 16, 18328–18347. [Google Scholar] [CrossRef]
  250. Dima, S.O.; Dobre, T.; Stoica-Guzun, A.; Oancea, F.; Jinga, S.I.; Nicolae, C.A. Molecularly imprinted bio-membranes based on cellulose nano-fibers for drug release and selective separations. Macromol. Symp. 2016, 359, 124–128. [Google Scholar] [CrossRef]
  251. Donato, L.; Chiappetta, G.; Drioli, E. Surface functionalization of PVDF membrane with a naringin-imprinted polymer layer using photo-polymerization method. Separ. Sci. Technol. 2011, 46, 1555–1562. [Google Scholar] [CrossRef]
  252. Lay, S.; Ni, X.F.; Yu, H.N.; Shen, S.R. State-of-the-art applications of cyclodextrins as functional monomers in molecular imprinting techniques: A review. J. Sep. Sci. 2016, 39, 2321–2331. [Google Scholar] [CrossRef]
  253. Na, L.; Hu, Y. Construction of natural polymeric imprinted materials and their applications in water treatment: A review. J. Hazard. Mater. 2021, 403, 123643. [Google Scholar]
  254. Zhou, Z.; He, L.; Mao, Y.; Chai, W.; Ren, Z. Green preparation and selective permeation of d-Tryptophan imprinted composite membrane for racemic tryptophan. Chem. Eng. J. 2017, 310, 63–71. [Google Scholar] [CrossRef]
  255. Tamahkar, E.; Bakhshpour, M.; Denizli, A. Molecularly imprinted composite bacterial cellulose nanofibers for antibiotic release. J. Biomater. Sci. Polym. Ed. 2019, 30, 450–460. [Google Scholar] [CrossRef]
  256. Piloto, A.M.L.; Ribeiro, D.S.M.; Rodrigues, S.S.M.; Santos, J.L.M.; Sampaio, P.; Sale, G. Imprinted Fluorescent Cellulose Membranes for the On-Site Detection of Myoglobin in Biological Media. ACS Appl. Bio Mater. 2021, 4, 4224–4235. [Google Scholar] [CrossRef]
  257. Zhang, C.; Zhong, S.; Yang, Z. Cellulose acetate-based molecularly imprinted polymeric membrane for separation of vanillin and o-vanillin. Braz. J. Chem. Eng. 2008, 25, 365–373. [Google Scholar] [CrossRef] [Green Version]
  258. Liu, Q.Q.; Zhao, Y.; Pan, J.F.; Bruggen, B.V.D.; Shen, J.N. A novel chitosan base molecularly imprinted membrane for selective separation of chlorogenic acid. Sep. Purif. Technol. 2016, 164, 70–80. [Google Scholar] [CrossRef]
  259. Xiao, X.D.; Li, Z.Q.; Liu, Y.; Jia, L. Preparation of chitosan-based molecularly imprinted material for enantioseparation of racemic mandelic acid in aqueous medium by solid phase extraction. J. Sep. Sci. 2019, 42, 3544–3552. [Google Scholar] [CrossRef] [PubMed]
  260. Di Bello, M.P.; Mergola, L.; Scorrano, S.; Del Sole, R. Towards a new strategy of a chitosan-based molecularly imprinted membrane for removal of 4-nitrophenol in real water samples. Polym. Int. 2017, 66, 1055–1063. [Google Scholar] [CrossRef]
  261. Ma, X.L.; Chen, R.Y.; Zheng, X.; Youn, H.; Chen, Z. Preparation of molecularly imprinted CS membrane for recognizing naringin in aqueous media. Polym. Bull. 2011, 66, 853–863. [Google Scholar] [CrossRef]
  262. Wu, H.; Zhao, Y.Y.; Yu, Y.X.; Jiang, Z.Y. Molecularly Imprinted Chitosan Membrane for Chiral Resolution of Phenylalanine Isomers. J. Funct. Polym. 2007, 20, 262–266. [Google Scholar]
  263. Nong, L.P.; Huang, M.; Zhuang, Y.P. Preparation and Selective Permeation Characterization of L-Tryptophane Molecular Imprinting Chitosan Film. Chem. Res. 2009, 20, 15–18. [Google Scholar]
  264. Zheng, X.F.; Lian, Q.; Yang, H. Synthesis of chitosan–gelatin molecularly imprinted membranes for extraction of L-tyrosine. RSC Adv. 2014, 4, 42478–42485. [Google Scholar] [CrossRef]
  265. Qi, M.; Zhao, K.; Bao, Q.; Pan, P.; Zhao, Y.; Yang, Z.; Wang, H.; Wei, J. Adsorption and Electrochemical Detection of Bovine Serum Albumin Imprinted Calcium Alginate Hydrogel Membrane. Polymers 2019, 11, 622. [Google Scholar] [CrossRef] [Green Version]
  266. Zhou, Z.; Cui, K.; Mao, Y.; Chai, W.; Wang, N.; Ren, Z. Green preparation of d-tryptophan imprinted self-supported membrane for ultrahigh enantioseparation of racemic tryptophan. RSC Adv. 2016, 6, 109992–110000. [Google Scholar] [CrossRef]
  267. Gao, T.; Guan, G.; Wang, X.; Lou, T. Electrospun molecularly imprinted sodium alginate/polyethylene oxide nanofibrous membranes for selective adsorption of methylene blue. Int. J. Biol. Macromol. 2022, 207, 62–71. [Google Scholar] [CrossRef] [PubMed]
  268. Zhao, K.; Feng, L.; Lin, H.; Fu, Y.; Lin, B.; Cui, W.; Wei, J. Adsorption and photocatalytic degradation of methyl orange imprinted composite membranes using TiO2/calcium alginate hydrogel as matrix. Catal. Today 2014, 236, 127–134. [Google Scholar] [CrossRef]
  269. Shin, M.J.; Shin, J.S. A molecularly imprinted polymer undergoing a color change depending on the concentration of bisphenol A. Microchim. Acta 2019, 187, 44. [Google Scholar] [CrossRef] [PubMed]
  270. Wei, S.L.; Liu, W.T.; Huang, X.C.; Ma, J.K. Preparation and application of a magnetic plasticizer as a molecularly imprinted polymer adsorbing material for the determination of phthalic acid esters in aqueous samples. J. Sep. Sci. 2018, 41, 3806. [Google Scholar] [CrossRef]
  271. Lu, J.; Qin, Y.Y.; Wu, Y.L.; Chen, M.N.; Sun, C.; Han, Z.X.; Yan, Y.S.; Li, C.X.; Yan, Y. Mimetic-core-shell design on molecularly imprinted membranes providing an antifouling and high-selective surface. Chem. Eng. J. 2021, 417, 128085. [Google Scholar] [CrossRef]
  272. Liu, Y.; Liu, Y.; Liu, Z.; Hu, X.; Xu, Z. β-cyclodextrin molecularly imprinted solid-phase microextraction coatings for selective recognition of polychlorophenols in water samples. Anal. Bioanal. Chem. 2018, 410, 509–519. [Google Scholar] [CrossRef]
  273. Xing, W.; Wu, Y.; Lu, J.; Lin, X.; Yu, C.; Dong, Z.; Li, C. Biomass-Based Synthesis of Green and Biodegradable Molecularly Imprinted Membranes for Selective Recognition and Separation of Tetracycline. Nano Brief Rep. Rev. 2020, 15, 2050004. [Google Scholar] [CrossRef]
  274. Masakazu, Y.; Kensuke, K.; Akinori, E.; Takashi, A.; Shinichi, S.; Kyoko, H.; Kunihiko, W. Green Polymers from Geobacillus thermodenitrificans DSM465–Candidates for Molecularly Imprinted Materials. Macromol. Biosci. 2006, 6, 210–215. [Google Scholar]
  275. Su, C.; Li, Z.; Zhang, D.; Wang, Z.; Zhou, X. A highly sensitive sensor based on a computer-designed magnetic molecularly imprinted membrane for the determination of acetaminophen. Biosens. Bioelectron. 2020, 148, 111819. [Google Scholar] [CrossRef]
  276. Wang, C.; Hu, X.; Guan, P.; Wu, D.; Qian, L.; Li, J. Separation and purification of thymopentin with molecular imprinting membrane by solid phase extraction disks. J. Pharm. Biomed. Anal. 2015, 102, 137–143. [Google Scholar] [CrossRef]
  277. He, Z.; Meng, M.; Yan, L.; Zhu, W.; Sun, F.; Yan, Y.; Liu, Y.; Liu, S. Fabrication of new cellulose acetate blend imprinted membrane assisted with ionic liquid ([BMIM]Cl) for selective adsorption of salicylic acid from industrial wastewater. Sep. Purif. Technol. 2015, 145, 63–74. [Google Scholar] [CrossRef]
  278. Fan, J.P.; Zhang, F.Y.; Yang, X.M.; Zhang, X.H.; Cao, Y.H.; Peng, H.L. Preparation of a novel supermacroporous molecularly imprinted cryogel membrane with a specific ionic liquid for protein recognition and permselectivity. J. Appl. Polym. Sci. 2018, 135, 46740. [Google Scholar] [CrossRef]
  279. Bai, M.; Qiang, L.Q.; Meng, M.; Li, B.; Wang, S.; Wu, Y. Upper surface imprinted membrane prepared by magnetic guidance phase inversion method for highly efficient and selective separation of artemisinin. Chem. Eng. J. 2021, 405, 126889. [Google Scholar] [CrossRef]
  280. Lee, T.P.; Saad, B.; Nakajima, L.; Takaomi, K. Preparation and Characterization of Hybrid Molecularly Imprinted Polymer Membranes for the Determination of Citrinin in Rice. Sains. Malays. 2019, 48, 1661–1670. [Google Scholar] [CrossRef]
  281. Yuan, Y.; Yuan, X.; Hang, Q.; Zheng, R.; Lin, L.; Zhao, L.; Xiong, Z. Dummy molecularly imprinted membranes based on an eco-friendly synthesis approach for recognition and extraction of enrofloxacin and ciprofloxacin in egg samples. J. Chromatogr. A 2021, 1653, 462411. [Google Scholar] [CrossRef] [PubMed]
  282. Wang, X.J.; Xu, Z.L.; Feng, J.L.; Bing, N.C.; Yang, Z.G. Molecularly imprinted membranes for the recognition of lovastatin acid in aqueous medium by a template analogue imprinting strategy. J. Membr. Sci. 2008, 313, 97–105. [Google Scholar] [CrossRef]
  283. Wan, L.; Gao, H.; Gao, H.; Yan, G.; Wang, F.; Wang, Y. Dummy molecularly imprinted solid phase extraction in a nylon membrane filter for analysis of vardenafil in health care products. Microchem. J. 2021, 165, 106157. [Google Scholar] [CrossRef]
  284. Sergeyeva, T.; Yarynka, D.; Dubey, L.; Dubey, I.; Piletska, E.; Linnik, R.; Antonyuk, M.; Ternovska, T.; Brovko, O.; Piletsky, S.; et al. Sensor Based on Molecularly Imprinted Polymer Membranes and Smartphone for Detection of Fusarium Contamination in Cereals. Sensors 2020, 20, 4304. [Google Scholar] [CrossRef]
  285. Wei, M.; Wang, S.; Jiang, W. Preparation and Characterization of Dual-Template Molecularly Imprinted Membrane with High Flux Based on Blending the Inorganic Nanoparticles. J. Inorg. Organomet. Polym. 2018, 28, 295–307. [Google Scholar] [CrossRef]
  286. Sarafraz-Yazdi, A.; Razavi, N. Application of molecularly imprinted polymers in solid-phase microextraction techniques. TrAC–Trends Anal. Chem. 2015, 73, 81–90. [Google Scholar] [CrossRef]
  287. Zhang, W.; Zhang, Q.; Zhang, X.; Wu, Z.; Li, B.; Dong, X. Preparation and evaluation of molecularly imprinted composite membranes for inducing crystallization of oleanolic acid in supercritical CO2. Anal. Meth. 2016, 8, 5651–5657. [Google Scholar] [CrossRef]
  288. Zhang, W.; Zhang, Q.; Wang, R.; Cui, Y.; Zhang, X.; Hong, L. Preparation of molecularly imprinted composite membranes for inducing bergenin crystallization in supercritical CO2 and adsorption properties. Bull. Korean Chem. Soc. 2012, 33, 703–706. [Google Scholar] [CrossRef] [Green Version]
  289. Da Silva, S.M.; Viveiros, R.; Coelho, M.B.; Aguiar-Ricardo, A.; Casimiro, T. Supercritical CO2-assisted preparation of a PMMA composite membrane for bisphenol A recognition in aqueous environment. Chem. Eng. Sci. 2012, 68, 94–100. [Google Scholar] [CrossRef]
Figure 1. The twelve principles of green chemistry. (Reprinted with permission from Ref [4]. Copyright 2016 John Wiley and Sons).
Figure 1. The twelve principles of green chemistry. (Reprinted with permission from Ref [4]. Copyright 2016 John Wiley and Sons).
Membranes 12 00472 g001
Figure 2. Bioplastic films obtained from vegetable wastes in mild aqueous conditions: (A) carrot bioplastic; (B) parsley bioplasgtic; (C); radicchio bioplastic; (D) cauliflower bioplastic. (Reprinted with permission from Ref. [89]. Copyright 2018 Royal Society of Chemistry).
Figure 2. Bioplastic films obtained from vegetable wastes in mild aqueous conditions: (A) carrot bioplastic; (B) parsley bioplasgtic; (C); radicchio bioplastic; (D) cauliflower bioplastic. (Reprinted with permission from Ref. [89]. Copyright 2018 Royal Society of Chemistry).
Membranes 12 00472 g002
Figure 3. Strategy to develop greener membranes following the principles of green chemistry. The number next to the boxes represent the ranking in order of priority according to their contribution to making a membrane fabrication process greener. (Reprinted with permission from Ref. [95]. Copyright 2014 Royal Society of Chemistry).
Figure 3. Strategy to develop greener membranes following the principles of green chemistry. The number next to the boxes represent the ranking in order of priority according to their contribution to making a membrane fabrication process greener. (Reprinted with permission from Ref. [95]. Copyright 2014 Royal Society of Chemistry).
Membranes 12 00472 g003
Figure 4. Preferred and undesirable solvents for membrane preparation. (Adapted with permission from Ref. [95]. Copyright 2014 Royal Society of Chemistry).
Figure 4. Preferred and undesirable solvents for membrane preparation. (Adapted with permission from Ref. [95]. Copyright 2014 Royal Society of Chemistry).
Membranes 12 00472 g004
Figure 5. Representation of the MIPs synthetic process.
Figure 5. Representation of the MIPs synthetic process.
Membranes 12 00472 g005
Figure 6. Criticism of unsustainable molecular imprinting technology covering imprinting and post-imprinting application and disposal. (Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Figure 6. Criticism of unsustainable molecular imprinting technology covering imprinting and post-imprinting application and disposal. (Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Membranes 12 00472 g006
Figure 7. The fourteen principles of green molecular imprinting expressed as the mnemonic device “GREENIFICATION.” (Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Figure 7. The fourteen principles of green molecular imprinting expressed as the mnemonic device “GREENIFICATION.” (Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Membranes 12 00472 g007
Figure 8. “Greenificated” molecular imprinting technology road map from 2012 to 2030. (Reprinted from ref. [132]. Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Figure 8. “Greenificated” molecular imprinting technology road map from 2012 to 2030. (Reprinted from ref. [132]. Reprinted with permission from Ref. [132]. Copyright 2021 John Wiley and Sons).
Membranes 12 00472 g008
Figure 9. Published papers on MIPs and MIMs from 2000 to 2021 based on web of science core collection. (Reprinted with permission from Ref. [206]. Copyright 2021 Elsevier).
Figure 9. Published papers on MIPs and MIMs from 2000 to 2021 based on web of science core collection. (Reprinted with permission from Ref. [206]. Copyright 2021 Elsevier).
Membranes 12 00472 g009
Figure 10. Application areas of MIMs. (Reprinted with permission from Ref. [206]. Copyright 2021 Elsevier).
Figure 10. Application areas of MIMs. (Reprinted with permission from Ref. [206]. Copyright 2021 Elsevier).
Membranes 12 00472 g010
Figure 11. Representation of a flat-sheet molecularly imprinted membrane able to selectively bind the template molecules. (Reprinted with permission from Ref. [7]. Copyright 2021 Springer Nature).
Figure 11. Representation of a flat-sheet molecularly imprinted membrane able to selectively bind the template molecules. (Reprinted with permission from Ref. [7]. Copyright 2021 Springer Nature).
Membranes 12 00472 g011
Figure 12. Illustration of the preparation process of composite d-tryptophan-sodium alginate/ poly (vinylidene) fluoride imprinted membrane. (Reprinted with permission from Ref. [254]. Copyright 2017 Elsevier).
Figure 12. Illustration of the preparation process of composite d-tryptophan-sodium alginate/ poly (vinylidene) fluoride imprinted membrane. (Reprinted with permission from Ref. [254]. Copyright 2017 Elsevier).
Membranes 12 00472 g012
Figure 13. Pre-polymerization mixture components during the synthesis of the cryogel bovine serum albumin-imprinted membrane fabricated by Fan et al. (a); infiltration of the mixture between two glass plates (b); SEM image (upper) and photo (down) of the obtained free-standing flat-sheet membrane (c); diffusion cell used in permeation tests (d). (Reprinted with permission from Ref. [278]. Copyright 2018 John Wiley and Sons).
Figure 13. Pre-polymerization mixture components during the synthesis of the cryogel bovine serum albumin-imprinted membrane fabricated by Fan et al. (a); infiltration of the mixture between two glass plates (b); SEM image (upper) and photo (down) of the obtained free-standing flat-sheet membrane (c); diffusion cell used in permeation tests (d). (Reprinted with permission from Ref. [278]. Copyright 2018 John Wiley and Sons).
Membranes 12 00472 g013
Figure 14. Adsorption capacity of poly (vinylidene) fluoride artesunate-imprinted (a) and non-imprinted membrane (b) (containing the 16.6 wt.% of MIP particles) toward artemisinin and artemether as well as selectivity factor (α) in time. Artesunate was used as a dummy template. Concentration of each analyte in the feed solution = 200 mg∙L−1; transmembrane pressure = 0.1 MPa; temperature = 25 °C; flow rate = 17 mL∙min−1; active membrane area = 21.23 cm2. (Reprinted with permission from Ref. [279]. Copyright 2021 Elsevier).
Figure 14. Adsorption capacity of poly (vinylidene) fluoride artesunate-imprinted (a) and non-imprinted membrane (b) (containing the 16.6 wt.% of MIP particles) toward artemisinin and artemether as well as selectivity factor (α) in time. Artesunate was used as a dummy template. Concentration of each analyte in the feed solution = 200 mg∙L−1; transmembrane pressure = 0.1 MPa; temperature = 25 °C; flow rate = 17 mL∙min−1; active membrane area = 21.23 cm2. (Reprinted with permission from Ref. [279]. Copyright 2021 Elsevier).
Membranes 12 00472 g014
Table 1. Some examples of ionic liquids used as functional monomers in the synthesis of imprinted polymers.
Table 1. Some examples of ionic liquids used as functional monomers in the synthesis of imprinted polymers.
Functional MonomerSolventTemplateRef.
1-Allyl-3-ethylimidazolium bromide ([AEIM]Br)WaterPhenylephrine
(dummy template of clenbuterol)
[140]
1-allyl-3- ethylimidazolium hexafluorophosphate;Water and chloroform4,4–Dichlorobenzhydrol[141]
3-(anthracen-9-ylmethyl)-1-vinyl1H-imidazol-3-ium chloride;Methanolp-Nitroaniline[142]
1-[3-(N-cystamine)propyl]-3-vinylimidazolium tetrafluoroborate;Watera-Fetoprotein[143]
1-Ethyl- 3-methylimidazolium tetrafluoroborate ([EMIM][BF4]),ethanol/waterPatulin[144]
1-(a-methyl acrylate)-3-methylimidazolium bromide;Methanol and waterCaffeine[145]
1-vinyl-3-methylimidazolium chlorideAcetonitrile and waterBenzoic acid
(dummy template of salicylic acid)
[146]
1-allyl-3-methylimidazolium bromideAcetonitrileBromide (Z)-3-(chloromethylene)-6-flourothiochroman-4-one[147]
1-allyl-3-vinylimidazolium chlorideWater and ethanolImiquimod[148]
1-allyl-3-vinylimidazolium chloride;methanolSulfamonomethoxine[149]
1-(Triethoxysilyl) propyl-3aminopropylimidazole bromideTetrahydrofuran and methanolBisphenol A
(dummy template of organochlorines)
[150]
1-vinyl-3 butyl imidazolium chlorideWaterLysozyme[151]
1-Vinyl-3-ethylimidazolium bromideWaterOchratoxin A[152]
1-Viny-3-carboxybutyl imidazolium bromideMethanol and waterSynephrine[139]
1-vinyl-3 butyl imidazolium tetrafluoroborateMethanol Cyhalothrin[153]
1-vinyl-3-propylimidazole sulfonateWater Hemoglobin [154]
1,6-hexa-3,30 -bis-1-vinylimidazolium bromineWaterLevofloxacin[155]
3-(3-aminopropyl)-1vinylimidazolium chloride WaterBovine serum albumin[156]
Mono-6A-deoxy-6-(1-vinyl imidazolium)-β-cyclodextrin tosylatePhosphate bufferC terminal peptides of cytochrome C[157]
3-Propyl-1-vinyl imidazolium bromide Methanol and waterAmoxicillin [158]
Table 2. Some examples of natural materials used in the production of MIMs.
Table 2. Some examples of natural materials used in the production of MIMs.
Natural MaterialTemplateApplicationRef.
Cellulose
Membranes 12 00472 i001
Diosgenin Sustained release and selective separation [250]
Gentamicin Controlled delivery [255]
Myoglobin Sensing in biological media[256]
QuercetinSustained release[223]
VanillineSelective separation[257]
Chitosan
Membranes 12 00472 i002
Chlorogenic acid Selective separation [258]
L-Mandelic acidEnantioseparation[259]
4-nitrophenolWater treatment[260]
NaringinDebittering[261]
L-PhenylalanineEnantioseparation[262]
L-TryptophanEnantioseparation[263]
L-tyrosineSelective separation [264]
Sodium alginate
Membranes 12 00472 i003
Bovine serum albumin Adsorption and electrochemical detection in aqueous phase[265]
D-TryptophanEnantioseparation[266]
Methylene blueRemoval from water [267]
Methyl orangeRemoval from water[268]
β-cyclodextrin
Membranes 12 00472 i004
Bisphenol ASensing in water[269]
Butyl benzyl phthalate and dibutyl phthalate (dual templates) Sensing in water[270]
CiprofloxacinSelective separation [271]
Triclosan and polychlorophenolsSensing in water[272]
Table 3. Some target molecules and their relative dummy template used in green MIMs production.
Table 3. Some target molecules and their relative dummy template used in green MIMs production.
Target Compound Dummy TemplateApplicationRef.
Artemisinin
Membranes 12 00472 i005
Artesunate
Membranes 12 00472 i006
Separation from similar artemether [279]
Citrinin
Membranes 12 00472 i007
1-Napthol
Membranes 12 00472 i008
Detection in rice[280]
Enrofloxacin
Membranes 12 00472 i009
Gatfloxacin
Membranes 12 00472 i010
Detection and removal from eggs[281]
Lovastatin
Membranes 12 00472 i011
Lovastatin acid
Membranes 12 00472 i012
Separation of statins[282]
Vardefanil
Membranes 12 00472 i013
Sildefanil
Membranes 12 00472 i014
Solid-phase extraction [283]
Zearalenone
Membranes 12 00472 i015
Cyclododecyl-2,4-dihydroxybenzoate
Membranes 12 00472 i016
Detection in cereal samples for inspecting fusarium contamination[284]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Donato, L.; Nasser, I.I.; Majdoub, M.; Drioli, E. Green Chemistry and Molecularly Imprinted Membranes. Membranes 2022, 12, 472. https://doi.org/10.3390/membranes12050472

AMA Style

Donato L, Nasser II, Majdoub M, Drioli E. Green Chemistry and Molecularly Imprinted Membranes. Membranes. 2022; 12(5):472. https://doi.org/10.3390/membranes12050472

Chicago/Turabian Style

Donato, Laura, Imen Iben Nasser, Mustapha Majdoub, and Enrico Drioli. 2022. "Green Chemistry and Molecularly Imprinted Membranes" Membranes 12, no. 5: 472. https://doi.org/10.3390/membranes12050472

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

Article Metrics

Back to TopTop