Molecularly Imprinted Membranes: Past, Present, and Future

Sep 9, 2016 - Biography. Masakazu Yoshikawa received his Bachelor's (1976) and Ph.D. (1982) degrees from Kyoto University under the supervision of Pro...
4 downloads 10 Views 7MB Size
Review pubs.acs.org/CR

Molecularly Imprinted Membranes: Past, Present, and Future Masakazu Yoshikawa,*,† Kalsang Tharpa,‡ and Ştefan-Ovidiu Dima§,∥ †

Department of Biomolecular Engineering, Kyoto Institute of Technology, Matsugasaki, Kyoto 606-8585, Japan Department of Chemistry, University of Mysore, Manasagangotri, Mysore 570 006, India § Faculty of Applied Chemistry and Materials Science, Department of Chemical and Biochemical Engineering, University Politehnica of Bucharest, 1-7 Gheorghe Polizu, 011061 Bucharest, Romania ∥ Bioresources Department, INCDCP-ICECHIM Bucharest, 202 Splaiul Independentei, CP 35-174, 060021 Bucharest, Romania ‡

ABSTRACT: More than 80 years ago, artificial materials with molecular recognition sites emerged. The application of molecular imprinting to membrane separation has been studied since 1962. Especially after 1990, such research has been intensively conducted by membranologists and molecular imprinters to understand the advantages of each technique with the aim of constructing an ideal membrane, which is still an active area of research. The present review aims to be a substantial, comprehensive, authoritative, critical, and general-interest review, placed at the cross section of two broad, interconnected, practical, and extremely dynamic fields, namely, the fields of membrane separation and molecularly imprinted polymers. This review describes the recent discoveries that appeared after repeated and fertile collisions between these two fields in the past three years, to which are added the worthy acknowledgments of pioneering discoveries and a look into the future of molecularly imprinted membranes. The review begins with a general introduction in membrane separation, followed by a short theoretical section regarding the basic principles of mass transport through a membrane. Following these general aspects on membrane separation, two principles of obtaining polymeric materials with molecular recognition properties are reviewed, namely, molecular imprinting and alternative molecular imprinting, followed the methods of obtaining and practical applications for the particular case of molecularly imprinted membranes. The review continues with insights into molecularly imprinted nanofiber membranes as a promising, highly optimized type of membrane that could provide a relatively high throughput without a simultaneous unwanted reduction in permselectivity. Finally, potential applications of molecularly imprinted membranes in a variety of fields are highlighted, and a look into the future of membrane separations is offered.

CONTENTS 1. Introduction 2. Theory for Membrane Separation 2.1. Permeability 2.2. Selectivity 3. Molecular Imprinting 4. Alternative Molecular Imprinting 5. Molecularly Imprinted Membranes 5.1. Membranes Prepared by Conventional Imprinting Method 5.2. MIMs by Alternative Molecular Imprinting 6. Molecularly Imprinted Nanofiber Membranes (MINFMs) 6.1. Reasons for Developing MINFMs 6.2. Molecularly Imprinted Nanofiber Fabrics 6.3. Mass Separation with MINFMs 6.4. Transport Phenomena in MINFMs 6.5. Other Types of MIMs 7. Applications Based on MIMs 7.1. Sensor Applications Based on Supported MIMs 7.2. Sensor Applications Based on Self-Supported MIMs © XXXX American Chemical Society

7.3. Pervaporation, Nanofiltration, Electrodialysis, and Other MIM-Based Applications 8. Perspectives Author Information Corresponding Author Notes Biographies Acknowledgments References

A C C C C D F F H

U U V V V V V V

1. INTRODUCTION Two key processes in any chemical industry that determine both the quality and the cost of products are the synthesis route and the purification step. The synthesis step can be shortened or catalyzed to be made economical, but the purification of a product from reactants, byproducts, catalyst, low-level impurities, and so on involves very general measures associated with almost all processes and often in each reaction step. For instance, the purification of a pharmaceutical product at the

L L M M N R S S

Received: February 8, 2016

T A

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

pervaporation and reverse-osmosis processes. From another perspective, the two terms, solubility and partition, can suggest a form of affinity between the membrane and the target molecule, with dense membranes having a higher affinity but a lower throughput. Second, the diffusion of a target molecule within a membrane is governed by molecular size, shape, and/ or electronic charge. This means that the selectivity interval that a membrane can exhibit based on diffusivity is thought to be intrinsically limited. In contrast to the diffusivity of the substrate (target molecule), affinity, which can also be understood as molecular recognition between the membrane and the substrate, can theoretically range from zero to infinity. Depending on the chemical nature of a given substrate molecule (both targeted and nontargeted molecules), one can design a range of membrane materials to suit an affinity toward the molecule of interest, for instance, by introducing moleculerecognition binding sites into a suitable membrane. In conclusion, membrane-transport phenomena through a dense (nonporous) membrane can be understood by the solution− diffusion mechanism, whereas that through a porous membrane (pore size > 1 nm) occurs by the partition−diffusion mechanism. With this understanding, it is apparent that, to increase the permselectivity and, consequently, the separation performance of any type of membrane, it is essential to create specific binding sites to give the membrane molecular recognition abilities for the target molecule. In this regard, there are various molecular recognition approaches, such as crown ethers,5−7 cyclodextrins,8 and molecular clefts,9 that have been intensively investigated. However, such molecular recognition sites introduced into a membrane do not always show the same recognition performances as when they are free in solution. Apart from the above-mentioned approaches, there is another way to introduce molecular recognition sites into a polymeric membrane: the molecular imprinting of polymers. This approach requires neither rich experience on fine organic synthesis nor a great effort or long research time. Initially, there were a few reports on the introduction of molecular recognition sites within a silicon-based matrix, but this type of matrix showed limited application and many preparation constraints. Starting with the introduction of molecular recognition sites into a polymeric matrix by Wulff and Sarhan,10 this approach became popular and, in time, proved to be the most convenient way of creating molecular recognition sites in polymeric materials.11−24 Another, later approach, called alternative molecular imprinting, revealed an easier path for creating molecular recognition sites in polymeric materials. These two methods, molecular imprinting and alternative molecular imprinting, are discussed in detail in sections 3 and 4 of the present review, followed naturally by an overview of the sysnthesis methods and practical applications of molecularly imprinted membranes (MIMs; section 5) as the main subject of interest of this review. Furthermore, the review continues with a section dedicated to the presumed ultimate technology of the present membrane separation processes, the molecularly imprinted nanofiber membrane (MINFM). Existing and potential applications of molecularly imprinted membranes are further reviewed in section 7, and the review ends with a preliminary discussion of the perspectives on advanced membrane separation (section 8).

part-per-million level from genotoxic or potential genotoxic impurities or from an undesired enantiomeric product increases the overall cost of the product rather dramatically. In general, product purification is carried out by various methods, such as distillation, crystallization, chromatographic separation, and membrane separation. An ideal purification technique should have a single-step purification procedure, with no or minimal use of solvent and minimal use of energy that is stable over a wide range of physicochemical parameters. Such an ideal of purification technology is closely associated with membrane-based separation processes because they can be performed continuously, under mild conditions, and in a majority of cases without phase separation, the only exception being pervaporation.1−4 Present applications of membranebased processes of separation and/or purification include the following: (1) osmotic membrane processes (reverse osmosis, RO; forward osmosis, FO; pressure-enhanced osmosis, PEO; pressure-retarded osmosis, PRO) for seawater desalination for the production of drinking water in underprivileged areas, for the concentration of fruit juices in the food industry, and so on; (2) nanofiltration (NF) membranes used in medicine to separate amino acids and lipids from blood and in the pharmaceuticals industry for the advanced purification of active pharmaceutical ingredients (APIs); (3) ultrafiltration (UF) membranes applied for enzyme recovery, separation/concentration of colloidal particles and macromolecules, and the removal of pathogens from dairy products; (4) microfiltration (MF) membranes that allow the permeation of mono- and multivalent ions and of proteins and sugars, but retain bacteria and suspended solids, used to clear water of pathogens, to reduce water turbidity, and to perform cold sterilization of beverages; (5) ion-exchange membranes used in electrodialysis (ED) to concentrate or remove salt ions and in practical applications in industry for the production of coolingtower water, the conditioning of industrial heat-transfer fluids, the production of irrigation water, and the desalination of seawater, among others; and (6) membranes used in gas−liquid processes such as vapor permeation (VP) or pervaporation (PV) for the purification or removal of organic solvents, the dehydration of (bio)fuels, the breaking of azeotropes, the removal of water during esterification reactions, and the recovery or removal of gases (H2, O2, CO2). In general, mass transfer through a membrane can be understood as two successive processes: The solute’s molecules are first adsorbed into the membrane’s pores and then transported by convection (diffusion and/or advection) through the membrane. In this transport process, the membrane separation performance depends heavily on the factors mentioned above.1−3 First, the incorporation of substrate into the membrane can be due to either solubility or partition. Based on the two main types of membranes, the incorporation of substrate due to solubility is applicable to dense membranes, whereas that of partition is characteristic of porous membrane. Dense membranes can be differentiated from porous membranes by pores of subnanometer size ( 2-Apy > 4-APy, which is opposite to the order of their magnitudes of affinity.83 Acrylic acid (AA) and MAA have frequently been used as functional monomers because these two monomers are hydrogen-bond- and proton-accepting compounds. There are many acidic, basic, and neutral functional monomers, but their diversity is still insufficient, as described in section 1. Likewise, mixtures of theophylline (THO) and caffeine (CAF) are often used in molecular imprinting studies (Scheme 3). For example, Scheme 3. Chemical Structures of CAF and THO

Hong et al. prepared membranes from MAA and EGDMA, which are frequently used as a functional monomer and a crosslinkable monomer, respectively.84 The difference in the membranes was the print molecule, which was THO or CAF in this study. From single-molecule-transport experiments, they found that the substrate that was used as the print molecule was preferentially transported; the ideal permselectivity toward THO for the THO-imprinted membrane was determined to be 2.6, whereas that for the CAF-imprinted membrane was 3.0. As G

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 3. Enantioselective separation with a molecularly imprinted nanoparticle composite membrane. Reproduced with permission from ref 86. Copyright 2002 Elsevier.

Figure 4. Molecularly imprinted polymeric membranes obtained by surface photografting onto porous polypropylene membranes. Reproduced from ref 87. Copyright 2000 American Chemical Society.

Figure 6. Preparation of nanometer thin polymer membranes with uniform nanopores. Reproduced with permission from ref 92. Copyright 2008 Wiley-VCH.

Figure 5. Scheme for obtaining concentration profiles during transport through bifunctional membrane. Reproduced with permission from ref 90. Copyright 2003 Elsevier.

Scheme 4. Chemical Structure of VBGPA

that the capsule released 0.6-nm yellow probes and retained 1.1-nm red and 1.6-nm blue probes. The capsule with GPB as the pore-forming print molecule released 0.6-nm yellow and 1.1-nm red probes and retained just 1.6-nm blue probes, resulting in a blue-colored capsule. Nanocapsules with programmed-sized nanopores were also prepared using 6-O-(4-vinylbenzoyl)-1,2,3,4-tetra-O-acetyl-β-Dglucopyranose (VBGPA) (Scheme 4) as the pore-forming print molecule.93 Dergunov and Pinkhassik applied covalent molecular imprinting in this study,93 whereas noncovalent molecular imprinting was used in the study by Danila et al.92 The study of Dergunov and Pinkhassik93 demonstrated that the chemical environment of programmed-sized nanopores in nanometer-thin organic materials can be controlled. Their technique provides programmed-size pores and programmed chemical environment for membranes. However, it takes too long to realize such membranes in industry.

5.2. MIMs by Alternative Molecular Imprinting

The application of molecularly imprinted polymers obtained by alternative molecular imprinting started with Michaels et al.’s study.37 Furthermore, as described in section 4, applications of molecular imprinting to membrane separations also stemmed from their pioneering study. H

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 7. Candidate materials for alternative molecular imprinting.

to a compound of choice, has evidently not been explored and thus seemed to be a desirable subject for investigation.” In this article,37 they cited Dickey’s article30 from 1955 and concluded as follows: “The model proposed here does not invoke any properties peculiar to ethylene polymers. ... Furthermore, there is no feature of the model that restricts its applicability to the process of pervaporation studied here: Similar behavior may be expected in such processes as gas or vapor transmission and dialysis.” Michaels et al.’s article37 is, as mentioned before, pioneering and worthy of commemoration not only in membrane separation but also in molecular imprinting. As mentioned in section 4, any polymeric materials that can retain molecular memory can be used as materials for alternative molecular imprinting, as exemplified in Figure 7. Another method that can be included in the subdomain of alternative molecular imprinting is the phase-inversion method, which involves the coagulation of a polymeric solution in a nonsolvent. In the case of MIP preparation, the polymeric solution contains the dissolved matrix−print molecule complex. This solution can be cast into various shapes, which is an important technological advantage. Depending on the type of nonsolvent, the phase-inversion method can be wet or dry, that is, wet if the nonsolvent is a liquid, and dry if the nonsolvent is a gas. Through the phase-inversion method, one can obtain MIMs or MIP beads/pearls with various diameters.94−98

Pervaporation separation of xylenes through conditioned polyethylenes was investigated. The polyethylene membrane was conditioned as follows; a high-density polyethylene (HDPE) was first swollen by p-xylene (9.8 g-p-xylene/100 gHDPE) and heated at 100 °C for 24 h. Following heat treatment, the membrane was cooled slowly to ambient temperature and allowed to dry thoroughly under ambient conditions. Single-molecular- transport experiment of p- and oxylenes at 30 °C gave the ideal permselectivity toward p-isomer of 1.8, which showed agreement between calculated selectivity (βp/o) and the permselectivity (αp/o) from pervaporation of binary mixtures. In the introduction of this article, they wrote:37 “Studies by Dickey [ref 30 in this review, incorrectly cited as p 635 in the original article], and more recently by McKee [McKee, R. G., Sc.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, 1957], have demonstrated that polymeric networks formed in the presence of a foreign (nonpolymerizing) compound, when suitably extracted to remove the entrapped solute, exhibit adsorptive selectivity for that compound relative to others of similar molecular structure. These results have been interpreted to mean that ‘holes’ or ‘pockets’ are formed by the foreign solute molecules in this matrix which can accommodate that solute more easily than other, even quite similar molecular species. Application of the principle of tailoring a polymeric network to membrane systems, in an effort to develop transmissive permselectivity I

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 8. Summary of alternative molecularly imprinted polymeric membranes bearing an oligopeptide derivative as a chiral recognition site. Reproduced with permission from ref 16. Copyright 2002 Kluwer Academic Publishers.

9 shows the effect of an applied potential difference on the enantioselective electrodialysis of racemic N-α-acetyltryptophan

As an application of alternative molecular imprinting to membrane separation, chiral separation has been intensively investigated. As summarized in Figure 8, the chiral recognition ability was found to be dependent on the absolute configurations of both the print molecule and the constituent amino acid residues. That is, a membrane consisting of the oligopeptide residue from a D-amino acid and imprinted with a D-amino acid derivative recognized the D-enantiomer in preference to the corresponding L-enantiomer and vice versa.58,70 In general, it is necessary to employ the optically pure print molecule so that one can obtain molecularly imprinted materials for chiral recognition or chiral separation. However, previously reported results58,70 suggested that chiral recognition sites were constructed by using racemic print molecules instead of the corresponding optically pure print molecules. Of course, the efficiency of molecular imprinting that was carried out using racemic print molecules was lower than that obtained using optically pure print molecules. Among racemic print molecules, those with the same configuration as the constituent amino acid residues worked well as print molecules, whereas the antipodes functioned not as print molecules but just as porogens. The membranes that were imprinted with a given amino acid print molecule recognized not only print molecule analogues but also other racemic α-amino acids that had the same absolute configuration as the print molecule.56,57 In terms of enantioselectivity, when a concentration difference was used as the driving force for membrane transport, the transport of the enantiomer that was preferentially incorporated into the membrane was retarded compared to that of its antipode.56,61,62,70 The fact that the permselectivity was opposite to the adsorption selectivity can be attributed to the suppression of the diffusivity of the preferentially adsorbed enantiomer because of its relatively high affinity to the membrane compared to the affinity between the antipode and the membrane. As mentioned before, such a phenomenon has often been observed, especially for optical resolution.99−104 It is interesting and important to selectively transport the enantiomer that is preferentially incorporated into the membrane. Electrodialysis was found to be one way to attain such membrane transport.56,61,62,70 As an example, Figure

Figure 9. Effects of a difference in applied potential on enantioselective electrodialysis and total flux. [The molecular imprinting ratio, (Boc-L-Trp)/(DIDE), was fixed at 0.50; each concentration of racemic Ac-Trp was fixed at 1.0 × 10−3 mol dm−3.] Reproduced with permission from ref 62. Copyright 1999 Elsevier.

(Ac-Trp).62 In the range of applied potential differences of 1.5−2.5 V, the permselectivity toward the L-enantiomer, αL/D, reached 6.0, which was equal to the adsorption selectivity of the studied MIM from tetrapeptide derivatives. In this case, an additional driving force for membrane transport was supplied in the form of a potential difference such that the target molecule that was preferentially incorporated into the membrane could be dissociated from the chiral recognition sites at the same dissociation rate as the antipode. It was revealed that electrodialysis is a suitable method for selectively transporting a target molecule with a charge that is recognized and incorporated into the chiral recognition site constructed by the presence of a print molecule during the membrane preparation process. The simultaneous transport (or extraction) of the D- and Lenantiomers from a racemic mixture is an effective way to J

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

resolve racemic mixtures. Such a dream can be realized by employing two types of MIMs. As shown in Figure 10, two

The dependence of molecular recognition ability on the amino acid residue sequence and content was investigated using tripeptide derivatives consisting of two types of amino acid residues, namely, derivatives of glutamic acid and of phenylalanine.61,63,64 The relationship between the sequence and composition of the eight kinds of tripeptides and their membrane performance was studied. As a result, tripeptide derivatives containing more glutamic acid derivative residues or glutamic acid derivatives as the amino-terminal residue were found to exhibit higher affinity constants. The number of constituent amino acid residues was also expected to affect the molecular recognition abilities of molecularly imprinted oligopeptide membranes. To test this hypothesis, N-α-tert-butoxycarbonyl-L-tryptophan (Boc-L-Trp) was used as the print molecule, and MIMs were constructed from oligopeptide derivatives with between one and eight constituent amino acid residues.68 Also, the molecular recognition sites were constructed from oligopeptide derivatives with three to six constituent amino acid residues. It was revealed that the peptide derivative four amino acid residues, that is, the tetrapeptide derivative, was the best candidate for forming molecular recognition sites. Solvent composition, specifically, the polarity of the solution, in which molecular recognition takes place also greatly affects the membrane performance.65 The effectiveness of alternative molecular imprinting has been confirmed by Kobayashi and co-workers,105−109 the Trotta research group,110,111 Ramamoorthy and Ulbricht,112 Cristallini et al.,113 Jiang et al.,114 Ul-Haq and Park,115 and others. There is also growing interest in membranes from other materials such as zeolites, carbon nanotubes, and aquaporins. In comparison to MIMs, such membranes are different in terms of their preparation, properties, costs of synthesis, and end applications. However, a brief comparison among these membranes is offered in Table 1, although such a comparison does not fall within the scope of the present review. In contrast to existing membranes, MIMs can be tailor-made for specific molecules, which gives rise to both advantages and limitations in their application. Therefore, membranes such as MIMs could be most suitable for sensor preparation or the separation of enantiomers. Detailed applications of MIMs in the field of sensors are described in section 7.

Figure 10. Time−transport curves of Ac-D-Trp and Ac-L-Trp by dualdirection electrodialysis at ΔE = 2.5 V. (Membrane L, membrane from L-amino acid residues and imprinted with Boc-L-Trp; membrane R, membrane from D-amino acid residues and imprinted with Boc-DTrp.) Adapted with permission from ref 70. Copyright 2003 WileyVCH.

MIMs obtained from tetrapeptide derivatives consisting of Damino acid residues and L-amino acid residues separated the enantiomers from their racemic mixture simultaneously.70 This is called “dual-direction electrodialysis”. Chiral separation using dual-direction electrodialysis has also been realized with MIMs from carboxylated polysulfone40 and cellulose acetate.75 In contrast to MIMs from oligopeptide derivatives, as summarized in Figure 11, both the D- and L-enantiomers were found to work well as print molecules in the preparation of imprinted membranes from synthetic polymers,40,41,44,46,49−53 cellulose acetate,75,76 oligopeptide tweezers,72 and tetrapeptide derivatives consisting of glycinyl residues.73

Figure 11. Summary of alternative molecularly imprinted polymeric membranes from entirely nonchiral synthetic polymers, derivative of natural polymer (CA), and oligopeptide tweezers. K

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Table 1. Comparison between MIMs and Membranes Made of Other Materials type ease of synthesis separation mechanism chemical and thermal stability mechanical strength compatibility with solvents regeneration swelling applications cost

MIMs

zeolite membranes

polymeric one-pot synthesis based on shape and arrangement of functional groups stable

inorganic difficult molecular sieving/ partial or capillary condensation high

functionalized fullerene multistep molecular sieving/ functional group affinity high

biological difficult based on gate channel induced by external stimuli low

high limited

poor wide range

poor wide range

depends on support matrix limited

yes yes but low specific to the selected print molecule economical

yes no limited

yes no limited

high

high

yes no water treatment (desalination, etc.) very high

6. MOLECULARLY IMPRINTED NANOFIBER MEMBRANES (MINFMS)

carbon nanotube membranes

aquaporin membranes

nanofiber membrane and its diameter is shown in Figure 13. For a nanofiber membrane with a nanofiber diameter of 100

6.1. Reasons for Developing MINFMs

As indicated in the preceding sections, enhancing permselectivity is relatively easy through molecular imprinting or alternative molecular imprinting techniques. For membrane separation processes, throughput (flux) and permselectivity are two key factors, and in a sense, throughput is more important than permselectivity from the viewpoint of industrial applications. As illustrated in Figure 12, the two key factors in membrane separations, namely, flux and permselectivity, often have a

Figure 13. Relationship between the surface area and diameter of a nanofiber membrane. (Calculation was carried out assuming the density of the polymer to be 1.0 g cm−3.)

nm, the calculated surface area is about 40 m2 g−1, whereas a typical porous polymeric membrane has a surface area of about 0.2 m2 g−1. This estimation shows a 200 times greater surface area for a nanofiber membrane compared to a typical porous membrane cast as a film. If these two types of membranes were molecularly imprinted to increase their affinity toward a specific target molecule, because the binding capacity is directly proportional to the surface area, it is expected that a 200 times higher binding capacity and amount of print molecule would be incorporated into the nanofiber membrane compared to the regular porous film membrane. This statement is synonymous with there being a higher solubility coefficient of the print molecule in the nanofiber membrane than in the corresponding typical film membrane. In terms of the porosities of these two types of membranes, a nanofiber membrane has a porosity of about 80%, whereas a typical film-type membrane has a porosity of just 5−10%.121 Membranes exhibiting high porosities are expected to provide high diffusion coefficients for a given target substrate. As mentioned in sections 1 and 2, the transport of a compound through a membrane is determined by solubility (or partition coefficient) and diffusivity. Nanofiber membranes have both high partition coefficients and high diffusivity, which are transport properties that lead to high membrane performance.

Figure 12. Trade-off relationship in membrane separation.

trade-off relationship. The compromise is given by the fact that enhancement of the flux through the membrane usually leads to a simultaneous reduction in permselectivity and vice versa. It is important and indispensable to enhance both of these key factors so that MIMs will be applicable in various industries. MIMs are required to have high porosity to ensure a high surface area so that they can provide a higher flux and permselectivity. However, it is not easy to enhance the surface areas of MIMs. Electrospray deposition is one suitable method for obtaining MIMs with large surface areas.116−123 First, the difference in surface areas between a typical porous membrane and a nanofiber membrane is quite large, assuming that they are hypothetically fabricated from the same polymeric material with a theoretical volumetric mass density of 1.0 g cm−3. The relationship between the surface area of the L

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

6.2. Molecularly Imprinted Nanofiber Fabrics

ricated to recognize the target organic molecules and to specifically adsorb them. An MINFA aimed at the recognition of metal ions has also been studied. Specifically, an MINFA for lead(II) ions was fabricated from chitosan.133 Studies on MINFAs aimed at the selective adsorption of metal ions are important in relation to environmental protection, the recovery and recycling of useful metal resources, and aqua-metallurgy. MINFAs have also been fabricated by applying method 2. For example, an MINFA for the adsorption of dansyl-Lphenylalanine was obtained from poly(vinyl alcohol) and molecularly imprinted nanoparticles.134 A commercial nanofiber fabric, poly(ethyl acrylate-co-methyl methacrylate-cotrimethyl ammonioethyl methacrylate chloride), was converted into MINFAs for the adsorption of propranolol.135 An MINFA with affinity for rhodamine B was prepared by the encapsulation method as well.136 Polyaniline nanofibers that were not fabricated by electrospray deposition also worked as an MINFA for the recognition of 4-hydroxybenzoic acid.137 The recognition of a target molecule by a nanofiber fabric bearing molecular recognition sites corresponds to the uptake or incorporation of a permeant into a membrane during a membrane transport process. Nanofiber membranes bearing molecular recognition sites have also been fabricated by the simultaneous application of alternative molecular imprinting and electrospray deposition. Such prepared nanofiber materials are called molecularly imprinted nanofiber membranes (MINFMs). The setup for the fabrication of nanofiber materials is simple, and the preparation is relatively easy under ambient conditions.

The first molecularly imprinted nanofiber fabrics (MINFFs) were obtained by applying alternative molecular imprinting methods and electrospraying.124 Poly(allylamine) was used as the functional polymer and was transformed by molecular imprinting into an MINFF that was able to recognize 2,4dichlorophenoxyacetic acid (2,4-D) (Scheme 5). Because Scheme 5. Chemical Structures of PET, Polyallylamine, and 2,4-D

poly(allylamine) itself could not be used in an aqueous environment, poly(ethylene terephthalate) (PET) was used as the polymer matrix. The 2,4-D-imprinted nanofiber membranes exhibited favorable binding properties in aqueous solution. This suggests that electrosprayed nanofiber fabrics have the potential to be applicable in the recognition and separation of target molecules. Molecularly imprinted nanoparticles have been embedded in the electrospraying solution and converted in this way into electrosprayed nanofiber fabrics with a polymer matrix from PET125,126 or polystyrene.127 These studies revealed that molecularly imprinted nanoparticles can be relatively easily encapsulated and electrosprayed as MINFFs. These studies demonstrated that nanofiber fabrics with molecular recognition sites that were introduced by (1) alternative molecular imprinting or by the (2) encapsulation of molecularly imprinted nanoparticles into nanofibers worked well as molecular recognition materials. These types of materials can be used to perform solid-phase extraction and as sensor chips for the detection of a given target molecule. Considering all of these aspects, these nanofiber fabrics with molecular recognition sites can be called, more generally, molecularly imprinted nanofiber adsorbents (MINFAs). These initial studies stimulated many molecular imprinters, and subsequent studies of MINFAs have been reported. An MINFA based on polyimide nanofibers was imprinted with estrone, a natural estrogen suspected to be carcinogenic and environmentally problematic.128 The print molecule of estrone was chemically attached to the functional monomer through a urethane linkage, and it was removed easily after the nanofibrous morphology had been constructed, thus resulting in a covalent type of molecular imprinting. Applying method 1, MINFAs were prepared from polyvinylbutyral and β-cyclodextrin for the selective adsorption of naringin,129 from poly(ether sulfone)130 and aramid131 for the selective removal of bisphenol A [2,2-bis(4-hydroxyphenyl)propane], and also from PET and polyethylenimine for the adsorption of nickel5,10,15,20-tetraphenylporphine.132 These MINFAs were fab-

6.3. Mass Separation with MINFMs

The performances achieved to date by molecularly imprinted membranes obtained using different approaches suggest a clear evolution toward already-established separation membranes used in ultrafiltration, nanofiltration, pervaporation, dialysis, and reverse osmosis. Composite MIMs, micro- and macroporous MIMs, coagulated MIMs (by alternative molecular imprinting), membranes containing MIP nano- and microparticles, and MINFMs are competing for the title of highestperformance MIM, aiming for the highest throughput, the highest selectivity, and the highest stability at the same time. According to Ulbricht,18 two mass separation mechanisms can occur in membranes: One is facilitated permeation, meaning the preferential diffusion of the target molecule as a result of affinity, and the other is retarded permeation due to the affinity binding of the target molecule resulting in the faster elution of competitors of the print molecule. In experimental studies of mass transfer in MINFMs, both continuous and discontinuous, or batch and flow adsorption, are encountered. Moreover, mass separations with MINFMs are currently being studied under both equilibrium and dynamic conditions, with different parameters being evaluated and different membrane performances being obtained. As a first example, the batch adsorption of propanolol (PPL) from complex mixtures with similar substances as competitors, meaning other β-blockers such as atenolol (ATE), metoprolol (MET), and timolol (TIM), was studied on a molecularly imprinted nanofiber composite membrane.135 The adsorption experiments assigned the highest selectivity to the MIP system with propanolol/methyl methacrylate/divinylbenzene (PPL/ MMA/DVB) = 0.8:75:2.5 introduced by electrospinning as 50% MIP particles in Eudragit-RS100 nanofibers. An initial weight ratio higher than 50% did not allow for the M

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

phene sulfone (BTO2), dibenzothiophene sulfone (DBTO2), a n d 4 , 6 - d i m e t h y l d i b e n z o t h i o p h e n e s u l f o n e (4 , 6 DMDBTO2).141 Very interestingly, in this study, the influence of the MIP aspect, prepared in two forms, namely, as nanofibers and as microspheres, indicated the importance of the dimensions of the active molecularly imprinted surface. The nanofiber MIPs gave adsorption capacities almost twice those of the microsphere MIPs, having surface areas that were 30− 60% higher. The surface area was also influenced by the size of the print molecule, with the highest area obtained for nanofiber MIPs being observed for the largest print molecule, meaning DBTO2. However, the binding selectivity coefficient (k) and the adsorption capacity (qe) of the imprinted nanofibers were found to be the highest for the smallest print molecule, meaning BTO2, with the obtained values being 4.9 for k and 8.5 ± 0.6 mg/g for qe. The mass separation studies continued with the determination of the Langmuir and Freundlich adsorption parameters and with kinetic or continuous-flow adsorption studies. The two Langmuir parameters, the adsorption capacity and the adsorption constant, were found to be higher for nanofiber MIPs than for microsphere MIPs. A pseudo-first-order kinetic model fitted the experimental data better than the pseudo-second-order kinetic model, suggesting the physical adsorption of the print molecules on the nanofiber MIPs.

electrospinning of the suspension, whereas higher content of cross-linker (DVB) led to the condensation of the polymer beads introduced in the electrospun nanofibers, which decreased the diffusion of the target molecule into the beads. The highest selectivity obtained for the studied systems was about 1.8, whereas the highest imprinting factor was about 1.4. In a continuous-flow type of experiment, MINFMs were used for mass separation through the molecularly imprinted solidphase extraction (MISPE) technique, aiming to simultaneously extract traces of bisphenol A (BPA) and tebuconazole (TBZ) from vegetables and juices.138 Electrospun PVA nanofibers containing encapsulated molecularly imprinted polymer nanoparticles (MIP-NPs) selective for the two print molecules showed compound recoveries higher than for conventional solid-phase extraction (SPE) based on a hydrophobic/strong cation exchanger column (C18/SCX), of between 78.6% and 88.1% for BPA and between 72.3% and 85.1% for TBZ, as well as better high-performance liquid chromatography (HPLC) separation efficiencies. The mass separation of target molecules from solutions containing analogues with different octanol/ water partition coefficients, namely, hydroquinone, bisphenol C, bisphenol Z, estradial, hexaconazole, penconazole, and flutriafol, was performed using the Scatchard analysis, the dissociation constant (Kd), and the maximum number of binding sites (Bmax), parameters that are frequently used in evaluating mass transfer in MIPs.94,139 The adsorption properties were correlated with the dissociation constants in different solvents, because the binding capacity of the prepared MINFMs in water was similar to that observed in acetonitrile and because Kd in water was higher than that in acetonitrile. For the best MINFM system, the imprinting factor was about 2, with the dissociation constant in acetonitrile being 1.3 times smaller for the imprinted system than for the nonimprinted system. In another recent study,140 polybenzimidazole MINFs packed for the SPE of oxidized organosulfur compounds from diesel products showed high adsorption capacities, removing the sulfonated compounds below the limit of detection, meaning 2.4 mg of sulfur dm−3. The print molecules used, one at a time, were benzothiophene sulfone (BTO2), dibenzothiophene sulfone (DBTO2), and 4,6-dimethyl dibenzothiophene sulfone (4,6-DMDBTO2), and in the selectivity experiments two were used as competitors for the third, which was the print molecule. The separation properties were attributed to the hydrogen-bond interactions between the sulfone oxygen groups and the NH groups of polybenzimidazole and also to the π−π interactions between benzimidazole rings and the aromatic sulfone compounds. The selectivity was expressed as a selectivity coefficient (k), defined as Kd(target molecule)/Kd(interfering molecule), toward the other two interfering compounds and took the values of 40.2 for BTO2 MINFs, 12.9 for DBTO2 MINFs, and 10.9 for 4,6-DMDBTO2 MINFs, suggesting that the selectivity decreased with increasing size of the print molecule as a result of the sterical hindrance, which decreased the accessibility to the binding sites. Recently used natural materials and their derivatives used for designing MIMs are cellulose, cellulose acetate, poly(lactic acid), agarose, cotton, and chitosan. The next paragraph discusses some of the relevant results on MINFMs based on these biopolymers. Chitosan with a 90% deacetylation degree was the biopolymer that stood at the base of bio-MINFs for the selective removal of sulfurous compounds such as benzothio-

6.4. Transport Phenomena in MINFMs

In a recent work,76 cellulose acetate (CA) with 40 mol % acetyl content was used to prepare MINFMs for the print molecules N-α-benzyloxycarbonyl-D-glutamic acid (Z-D-Glu) and N-αbenzyloxycarbonyl-L-glutamic acid (Z-L-Glu) (Scheme 6). Next, the insights into membrane transport phenomena in MINFMs gained from the experimental results and inferences of this study are discussed in detail. First, a scanning electron microscopy (SEM) image of an electrosprayed nanofiber membrane that was obtained by electrospinning in the presence of the print molecule, Z-D-Glu, is shown in Figure 14. Similar morphologies were also observed for both the nanofiber membranes imprinted with the Lenantiomer and the control nanofiber membranes. Based on Scheme 6. Chemical Structures of CA and Z-Glu-OH

N

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

molecules. However, when the MINFMs were constructed using cellulose triacetate (CTA), a polymer with most of its hydroxyl groups acetylated, the results were quite different.77 In the case of CTA, the L-enantiomer of Z-Glu showed chiral selectivity, but it seemed hard for the D-enantiomer print molecule to selectively adsorb the D-enantiomer. Both MINFMs showed that the adsorption selectivity depends on the substrate concentration, which was not the case for the control nanofiber membrane. The adsorption selectivity of the MINFMs increased with decreasing substrate concentration, which indicates the presence of molecular recognition sites formed during the electrospray deposition process in the presence of the print molecule, Z-D-Glu or Z-LGlu. The adsorption isotherms, which provide information concerning the interaction between the substrate and the membrane, were used to confirm the substrate specificity of the membrane. The concentration of Glu in the membrane was determined considering the amount of Glu adsorbed in the membrane and the total volume of the membrane phase, including the membrane and the solution adsorbed in the membrane’s pores. The linear adsorption isotherms passing through the origin for both L-Glu in Figure 16a and D-Glu in Figure 16b show that the L-enantiomer and D-enantiomer adsorbed without any specific interactions on the D-enantiomer-imprinted and Lenantiomer-imprinted nanofiber membranes, respectively. Recalling Figure 15 at this point, it is clear that both the Z-DGlu MINFM and Z-L-Glu MINFM showed enantiomerselective properties toward the D-enantiomer and L-enantiomer, respectively. The nonspecific adsorption of Glu inside the pores of the nanofiber membrane can be described by the equation

Figure 14. SEM image of Z-D-Glu-imprinted nanofiber membrane with a molecular imprinting ratio of 0.50. Reproduced with permission from ref 76. Copyright 2010 Elsevier.

the SEM images, the diameters of the imprinted nanofibers were found to be 200−500 nm, whereas those of the control membrane were in the range of 300−800 nm. Figure 15 shows the adsorption selectivities of the three types of nanofibers, where panels a, c, and e of Figure 15 present the adsorption of Glu by the membranes relative to that of the constitutent repeat unit of CA, whereas panels b, d, and f of Figure 15 show the adsorption selectivities of the nanofiber membranes. The adsorption selectivity, SA(i/j), is defined as SA(i / j) =

(i‐Glu)/(j‐Glu) [i‐Glu]/[j‐Glu]

(9)

where (i-Glu) represents the amount of i-Glu adsorbed by the membrane and [i-Glu] represents the concentration in solution at equilibrium. Both MINFMs showed chiral selectivity, whereas the control nanofiber membrane showed no selectivity. Both Z-D-Glu and Z-L-Glu MINFMs constructed using CA with 40 mol % acetyl content showed chiral selectivity to their respective print

[j‐Glu]M = kA[j‐Glu]

(10)

where j denotes the enantiomer of Glu (L or D) adsorbed nonspecifically, kA is the adsorption constant, [j-Glu]M is the

Figure 15. Effects of the substrate concentration on (a,c,e) Glu adsorption and (b,d,f) adsorption selectivity for (a,b) Z-D-Glu-imprinted, (c,d) control, and (e,f) Z-L-Glu-imprinted membranes. (The two data sets for Glu adsorption in panel c are superimposed and not distinguishable.) Reproduced with permission from ref 76. Copyright 2010 Elsevier. O

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

times higher than that of the typical dense CA membrane and was thus assumed to contain a much higher concentration of molecular recognition sites. This led to the hypothesis that most of the print molecules during electrospray deposition might have reached the grounded aluminum foil without the CA molecules, resulting in a decreased molecular imprinting ratio in comparison to that for the typical CA membrane. This hypothesis seems to be valid, as it was found that the affinity constant (KS) of the typical membrane was 20% lower than that of the MINFMs. From a previous report,69 it is known that the affinity constant increases with decreasing molecular imprinting ratio. In other words, at low molecular imprinting ratio, more functional groups are available to interact with the print molecule, resulting in a higher affinity constant. The selectivity performances of the three nanofiber membranes (Z-D-Glu and Z-L-Glu MINFMs and control membrane) were investigated using a racemic mixture of Glu and a concentration gradient as the driving force for membrane transport. The resulting time−transport plots of the racemic mixture of Glu presented in Figure 17 show that D-Glu was transported through the Z-D-Glu-imprinted nanofiber membrane in preference to its antipode and vice versa and that the control membrane showed hardly any permselectivity. By changing the initial concentration in the feed side and measuring the permselectivity and flux, one can deduce the transport mechanism of a membrane. Figure 18 shows the relationship between the permselectivity and the flux dependence on the feed concentration. In the present section, the specific flux J (mol cm cm−2 h−1), which means the flux per unit membrane thickness, is defined as

Figure 16. Adsorption isotherms of D-Glu and L-Glu in nanofiber membranes imprinted with (a) Z-D-Glu and (b) Z-L-Glu. Reproduced with permission from ref 76. Copyright 2010 Elsevier.

concentration of the j-enantiomer of Glu nonspecifically adsorbed in the membrane, and [j-Glu] is the concentration of the j-enantiomer of Glu in the solution equilibrated with the nanofiber membrane. Now, in the case of adsorption isotherms of D-Glu in Figure 16a and L-Glu in Figure 16b that gave dual adsorption isotherms, a selective adsorption in each MINFM can be seen. Over the substrate concentration of 5.0 × 10−4 mol dm−3, a linearly parallel relationship to nonspecific adsorption was found for each membrane; however, the straight lines with positive intercepts showed selective adsorption in the respective MINFMs. Thus, dual adsorption isotherms aim to include both nonspecific adsorption and selective adsorption on specific recognition sites toward the D-enantiomer (Figure 16a) and the L-enantiomer (Figure 16b). The selective adsorption of Glu inside the imprinted membrane can be evaluated using the equation

J = Qδ /At

where J is the flux (transport rate) of substrate allowed to permeate in a given time t, Q is the amount of transported substrate, and δ is the membrane thickness. As shown in Figure 18, the permselectivity of the MINFMs has an inverse relation with decreasing feed concentration, whereas the control membrane showed a value of unity at any initial feed concentration. Similarly, the slope of the flux through the MINFMs showed an inverse relation with increasing feed concentration until it asymptotically reached a certain value. In contrast, the flux through the control nanofiber membrane was graphically found to increase linearly with the substrate concentration through the origin. Thus, one can conclude that the MINFMs showed enantioselective properties in contrast to the control membrane. The flux of a membrane usually exhibits a trade-off relationship with the permselectivity. However, it was found that the flux values for the molecularly imprinted CA nanofiber membranes were 1−2 orders of magnitude higher than those for molecularly imprinted typical CA membranes while still retaining the permselectivity observed in chiral separations with MINFMs from carboxylated polysulfone,47 polysulfone with aldehyde moiety,50 and cellulose triacetate (CTA).77 The experimental conditions, equations, and results are discussed in the next. The substrate transport through the molecularly imprinted CA nanofiber membranes was based on a concentration gradient, whereas that through the typical CA membranes was based on a potential difference. To compare the fluxes between two such membranes, the most suitable parameter is the molar mobility u (mol cm cm−2 J−1 h−1) of the permeant, Glu.

[i‐Glu]M = kA[i‐Glu] + KS[Site]0 [i‐Glu]/(1 + KS[i‐Glu]) (11)

where [i-Glu]M is the concentration of the i-enantiomer of Glu adsorbed preferentially in the MINFM, [i-Glu] is the concentration of the i-enantiomer of Glu in the solution equilibrated with the nanofiber membrane, KS is the affinity constant between the i-enantiomer and the molecular recognition site, and [Site]0 is the concentration of molecular recognition site in the membrane. The parameters of the adsorption equations (eqs 10 and 11) determined to best fit each adsorption isotherm in Figure 16 are summarize in Table 2, together with those for typical molecularly imprinted CA membranes.75 The concentration of molecular recognition sites, due to Z-DGlu or Z-L-Glu, in the typical molecularly imprinted CA membrane75 was found to be a factor of 2 greater than the concentration obtained by the electrospray deposition process. This was contrary to expectations because of the high surface area of the nanofiber membrane (Figure 13), which was many Table 2. Parameters for Adsorption Isotherms Z-D-Glu-imprinted membranes

Z-L-Glu-imprinted membranes

parameter

MIPMa

MINFMb

MIPMa

MINFMb

kA [Site]0 (mol dm−3) KS (mol−1 dm3)

1.9 × 103 3.4 3.1 × 103

1.5 × 101 7.0 × 10−3 1.6 × 104

2.0 × 103 3.4 3.1 × 103

1.8 × 101 8.0 × 10−3 1.7 × 104

(12)

a

Molecularly imprinted membrane; cited from ref 75. bMolecularly imprinted nanofiber membrane; cited from ref 76. P

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 17. Time−transport curves of racemic Glu through molecularly imprinted and control nanofiber membranes. ([D-Glu]L,0 = [L-Glu]L,0 = 2.50 × 10−4 mol dm−3.) Reproduced with permission from ref 76. Copyright 2010 Elsevier.

Figure 18. Effects of substrate concentration on the transport of racemic Glu through (a) Z-D-Glu-, (b) control, and (c) Z-L-Glu-imprinted nanofiber membranes. Reproduced with permission from ref 76. Copyright 2010 Elsevier.

Table 3. Results of Chiral Separations with Molecularly Imprinted Nanofiber Membranes (MINFMs) Z-D-Glu-imprinted MINFM driving force

value −3

10 Δc (mol dm ) 3

c

ΔEd (V)

αD/L

0.25

1.45

0.50

1.27

1.00

1.11

2.5

2.30

u

control NFM αD/L

a,b

1.96 × (290) 1.04 × (154) 5.49 × (81.2) 6.76 × (1)

−9

10

∼1

10−9

∼1

10−10

∼1

10−12

Z-L-Glu-imprinted MINFM αL/D

a,b

u

−9

1.98 × 10 (293) 2.04 × 10−9 (302) 1.70 × 10−9 (251)

1.44 1.34 1.07 2.30

ua,b 3.81 × (564) 2.49 × (368) 1.32 × (195) 6.76 × (1)

10−9 10−9 10−9 10−12

u = (−J/c)/(dμ/dx) {[(mol cm cm−2 h−1)/(mol cm−3)]/(J mol−1 cm−1) = mol cm cm2 J−1 h−1}. bValues in parentheses are relative values, as the u value for ΔE was set as unity. cConcentration gradient applied as the driving force for membrane transport. dPotential difference applied as the driving force for membrane transport. a

From the Nernst−Planck equation, the flux can be written as142 ⎛ d ln C ZF dφ v dP ⎞ ⎟ + + J = −uCRT ⎜ ⎝ dx RT dx RT dx ⎠

Equation 13 can be further simplified. If the mass transfer through the molecularly imprinted membrane is studied under isothermal and isobaric conditions, the third term in eq 13 can be ignored. If it is further considered that the concentration gradient is the main driving force for mass transport through the membrane, then the flux J can be reduced to the first term of eq 13. Similarly, if the gradient of electrical potential is the main driving force of the mass transfer, then the flux can be calculated with a good approximation using only the second term of eq 13. The molar mobility is a specific parameter equal to the flux per unit membrane thickness, per unit membrane area, per unit concentration, per unit driving force. Table 3 summarizes the

(13)

in which J represents the sum of the D-Glu and L-Glu fluxes through the membrane, C is the initial concentration of each Glu enantiomer, R is the universal gas constant, T is the absolute temperature, dC/dx is the concentration gradient at a given point, z is the valence of the permeant, F is the Faraday constant, dϕ/dx denotes the gradient of electrical potential at that point, v is the partial molar volume of the permeant, and dP/dx is the gradient of pressure at that point. Q

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

morphologies can enhance flux without compromising on permselectivity.

membrane performance for two types of CA membrane, such as MINFMs and typical molecularly imprinted membranes (MIMs). For an easy comparison between the membranes’ performances, Table 3 contains within parentheses the values of relative molar mobility related to that of a typical film-type molecularly imprinted membrane. Yet another example can be seen from nanofibers of polysulfone with alanyl residue as chiral selectors143 and chitin144 exhibiting flux enhancements of 2−3 orders of magnitude without compromising on permselectivity in comparison to the results for dense membranes. Regardless of the mechanism of membrane transport, such as simple diffusion and facilitated transport, a multistage cascade membrane transport is thought to be occurring in nanofiber membranes, as depicted in Figure 19. Multistage cascade

6.5. Other Types of MIMs

The concept of molecular imprinting has found resonance in many other research domains, and because of the specific interactions with each domain, the imprinted materials change accordingly. In this subsection, we highlight the transformations that the molecularly imprinted membranes undergo to better respond to increasingly specific tasks. The materials discussed include composite MIMs, hybrid MIMs, nanocomposite MIMs, hollow-fiber MIMs, (hydro)/(cryo)gel MIMs, flexible MIMs, nanofilm MIMs, biocompatible MIMs, nanofibrous mats used as MIMs, and liquid membranes adapted to work as MIMs. A composite MIM was prepared by the thermally initiated free-radical polymerization of a preorganized mixture of print molecule + functional monomer + cross-linker + initiator on the surface of a microporous ceramic hollow-fiber membrane.145 The optimal preorganization mixture for molecularly imprinting theophylline in a composite MIM was found to be 0.05 mol/L theophylline (print molecule), 0.2 mol/L methacrylic acid (functional monomer), 1 mol/L ethylene glycol dimethacrylate (cross-linker), and 0.02 mol/L 2,2′azobis(isobutyronitrile) (AIBN; initiator) in 20 mL of chloroform. This solution was cast on the surface of an α-Al2O3 ceramic microporous hollow-fiber membrane and thermally polymerized at 60 °C for 24 h. Two- and three-layer MIM coating procedures were performed and compared with the monolayer composite MIM in terms of adsorption capacity. The results showed a maximum binding capacity Q of 64.5 ± 0.3 μmol/g for the two-MIM coating and polymerization steps. Permeation tests performed by cross-flow filtration evidenced a separation factor of 1.98 in favor of theophylline over theobromine for the hollow-fiber MIM, in comparison to a separation factor of 1.08 for the ceramic hollow-fiber membrane alone. From these results, it was concluded that the preferential filtration of the template theophylline can be attributed to the molecular recognition properties gained by molecular imprinting. A drawback of this composite MIM is that the selectivity factor decreased from 2.63 to 2.02 after 2 h and to 1.98 after 3 h. This phenomenon indicates that some of the bound print molecules were no longer desorbing, leading in time to the full occupancy of the imprinting sites and the loss of membrane permselectivity. A second type of hollow-fiber hybrid/composite MIM was obtained by embedding molecularly imprinted polymeric spheres of bisphenol A dimethacrylate as a functional monomer and divinylbenzene as a cross-linker into hollow-fiber membranes obtained by the dry-wet spinning technique.146 A third type of hollow-fiber composite MIM was prepared by embedding a capillary MIP monolith (1 cm in length, 320 or 530 μm in diameter) into a porous hollow-fiber membrane tube, with an important role in this system being played by the thin supported liquid membrane formed at the surface of the open pores. This hollow-fiber-based liquid−solid microextraction system aims to overcome the usual watercompatibility problem in MIP applications and to detect atrazine and other triazine pesticides in water samples.147 Other types of composite MIMs are being built on the same concept of heterogeneous mixing of at least two different compounds, an organic one and an inorganic one. The organic phase is represented by the preorganized imprinting mixture, whereas the inorganic phase is represented by the supporting

Figure 19. Schematic diagrams of membrane transport through a nanofiber membrane: (a) simple diffusion and (b) facilitated transport.

membrane transport can explain why membranes obtained from polysulfones with an N-α-benzoylalanyl moiety or an N-αbenzyloxycarbonylalanyl moiety as a chiral selector showed permselectivity, whereas the corresponding dense membranes hardly gave permselectivity.143 At the same time, the flux values for the nanofiber membranes were 2 orders of magnitude higher than those for the dense membranes. Therefore, one can conclude that multistage cascade membrane transport within a nanofiber membrane leads to an enhancement of the permselectivity and high porosity and, thus, to a higher migration rate of substrate within the membrane. With the performance data for nanofiber membranes discussed so far, it is evident that membranes with nanofiber R

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

matrix on which the polymerization takes place, for example, microporous alumina (Al2O3) membranes with 2−4-μm nominal pore size and 2-mm thickness148,149 or silica (SiO2).150,151 In many other cases, the inorganic phase has been replaced by an organic phase, such as nylon,85,109,152−154 poly(vinylidene fluoride)89,154,155 (example of membrane characteristics: 0.22-μm pore size and 180-μm thickness), polypropylene,87,156,157 regenerated cellulose158,159 (e.g., 0.45μm average pore diameter, 100-μm thickness), and cellulose acetate,75,76,112,160,161 but the mixing of phases remains homogeneous up to micrometer dimensions and heterogeneous at submicronic dimensions, so these membranes are considered to be composites. Calcium alginate,161,162 sodium alginate,163,164 polyamide,85,165 polycarbonate,154 polysulfone,154,166 and polypyrrole167 have also been used as matrices for composite MIMs. If the continuous phase is represented by the membrane and the disperse phase by embedded nanoparticles, then the resultant MIMs are called nanocomposite MIMs or hybrid MIMs. The nanoparticles used to produce hybrid MIMs are extremely diverse, aiming to cover various specific applications. The most frequently encountered nanoparticles in the design of hybrid MIMs are MIP nanoparticles,168−170 carbon nanotubes,161,171,172 silica nanoparticles,173−175 magnetic nanoparticles,146,176−178 bacterial cellulose nanofibers,179−181 TiO2 nanoparticles,182−184 and metallic nanoparticles.185−187 In addition to the mentioned types of molecularly imprinted membranes (MIMs), there are other particular membranes that will be discussed based on their particularities. In this category can be placed imprinted nanofilms and thin films,188−192 flexible MIMs,193−195 (hydro)/(cryo)gel MIMs,196−200 biocompatible MIMs,201−204 nanofiber mats,133,138,205−207 and even supported liquid membrane-protected molecularly imprinted beads.208−211

supporting material involved and its compatibility with the transducer. Researchers have explored and innovated in both types of MIM support systems. Reviews have been written on MIPbased sensors212−218 but not specifically on MIM-based sensors. By “sensor”, we mean an MIM in continuous synchronization with the transducer, not considering a “freestanding MIM” that is not in close proximity with the transducer. In addition, we intend to simplify the classification of numerous articles on MIM-based sensors into two general categories, namely, supported MIMs and self-supported MIMs, thereby making the approach easier for newcomers to this field. A schematic diagram depicting a self-supported MIM, a supported MIM, and a free-standing MIM not in close proximity with the transducer is presented in Figure 20.

7. APPLICATIONS BASED ON MIMS The highly selective interfaces offered by MIMs as an active gate in close proximity with a transducer, through a selective permeate or selective nonpermeate, make MIMs very promising potential sensing materials. One of the key challenges in the development of MIM-based sensors is synchronization at the membrane−transducer interface, which is also the reason for the lack of MIP-based sensors in comparison to applications in other fields. It is certain that, to develop imprinted-polymer-based sensors, the recognizing element should have the properties of a membrane for direct transduction. In this review, we exclude thin-film recognition elements without membrane properties such as that used in quartz-crystal-microbalance- (QCM-) based sensors. Primarily, the efficiency of MIM-based sensors is greatly influenced by the type of membrane support system, namely, self-supported membranes and supported membranes, irrespective of the transduction systems involved. There are advantages and disadvantages to both types of support systems. Whereas supported membranes can be prepared by a wide range of polymerization techniques or monomers finally layering on certain support systems, self-supported MIMs have much narrower options. The twin advantages of selfsupported membranes are high-density imprinted sites coupled with instant transduction, resulting in enhanced sensitivity, permselectivity, and flux. The same advantages can be achieved in supported membrane systems depending on the type of

Figure 20. General classifications of MIM-based sensors.

7.1. Sensor Applications Based on Supported MIMs

Molecularly imprinted membranes standing on an external supporting matrix, such as a filter, tubule, or paste, are categorized as supported MIMs. Supported membrane types were used for the first MIM-based sensors.219 Different techniques for preparing supported MIM-based sensors have been explored. One of the earliest reported MIMs was based on the use of an inorganic material as a membrane support system. It involved the layering of a prepolymerization mixture on a methacrylate-derived silicon wafer, resulting in a 1−3-μm-thick membrane that was used in combination with field-effect devices for the detection of L-phenylalanine anilide, tyrosmnilide, and phenylalaninol.220 The same authors220 also prepared an MIM based on the immobilization of suspended MIP particles (1−25 μm in diameter) in cross-linked agarose with an approximately 0.5-mm-thick membrane. The system was successfully used in a competitive binding assay of morphine and L-phenylalanine anilide, except that the sensor required a long equilibrium time (2 h) and a long response time (20 min). A similar technique was used by Prasad et al.221 for the analysis of phorate in natural water. An imprinted MIP immobilized in a S

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

conductivity, will greatly enhance the application of MIP-coated electrodes. Another desirable option for MIPs as sensors is to use a monomer that is also part of a sensing device. A recent report230 in this direction used pyrogallol doped with alizarin red as the monomer (and also a part of the sensing agent) and doxycycline as the print molecule. The chemistry was based on the enhancement of the chemiluminescence response of oxidized alizarin red by oxidized doxycycline. However, the plots obtained from a selectivity test performed in a mixture containing an analogue of the print molecule also showed a response of equal intensity. Yet another attractive technique is based on the preparation of glassy-membrane-based ion-selective electrodes. Inside a 10 mm long Teflon tube, a polymerization mixture consisting of MAA, EGDMA, and AIBN was thermally polymerized around the print molecule to obtain a 5-mm-diameter glassy membrane.231 The membrane in the end of the tube was directly used as an ion-selective electrode in combination with a Ag/AgCl reference electrode. The resulting sensor had a response time of less than 10 s. Panasyuk-Delaney et al.232 proposed grafting as a technique for preparing self-supported MIMs. The procedure involved pretreatment of the surface membrane or electrode layer with benzophenone, followed by photopolymerization of monomers around the print molecule, resulting in a 10-nm-thick membrane and a 5-min sensor response time. A similar technique was also reported by Delaney et al.233 Use of the inorganic sol−gel polymerization mixture of titanium(IV) butoxide−carboxylate complexes formed on the gate surface of an ion-sensitive field-effect transistor (ISFET) resulting in a membrane with a thickness of ∼90 μm was reported.234 The sensor gave reproducible measurements at ±2 mV at ambient temperature. A novel polyphosphazene polymer was successfully used as the imprinted polymer on the surface of a glassy carbon working electrode.235 The selectivity of the sensor was tested in the presence of a potentially interfering compound at 50 times the concentration of the analyte. Another attractive report on a titanium(IV) butoxide− carboxylate complex-based thin membrane had a very short sensing response time of 45 s for benzylphosphonic acid derivatives and thiolated substrates.236 Li et al. developed a nitrocellulose-based self-supported membrane poly(vinyl alcohol)−Cu2+ imprinted polymer used for the room-temperature phosphometric determination of Cu2+.237 The resulting membrane was transparent and elastic and had a detection limit of 80 fg/spot. The imprinted Cu2+ pore was found to be highly selective through tests involving a series of cations. Using a nanoparticle-modified glassy electrode, a selfassembled MIM was prepared through electropolymerization.238 This technique allows control over the membrane thickness through control of the deposition time. Using this technique, Yuan et al. successfully developed a 70-nm-thick membrane over a Pt-nanoparticle-modified electrode surface for the determination of 17β-estradiol. A comparison between self-supported MIMs and supported MIMs in terms of performance characteristics for sensor applications is needed. The two approaches have distinct application advantages; however, they might differ from analyte to analyte. One of the comparison studies available is the preparation of three different types of MIM sensor based on

poly(vinyl chloride) (PVC) matrix was attached at the end of a pyrex tube for the electrochemical detection of phorate. Using a glass support at the time of polymerization, Piletsky et al.222 developed a supported membrane for the detection of atrazine using a conductometric device setup. The membrane synthesis was based on the radical polymerization of diethylaminoethyl methacrylate and ethylene glycol dimethacrylate. Although the setup was more a detection device than a sensor, its attractive features, including sensitivity, stability (4 months), and rapidity (30 min), make it stand out from other reported MIM-based sensors. Another class of supported membranes involves the use of a gel. This technique consisting of the gelling of an agarose support to act as a membrane cover for finely ground MIP particles was reported.223 The sensor was based on the competitive batch binding assay of 2,4dichlorophenoxy acetic acid against electroactive molecules. This technique has limitations including multistep processes and cracking of particle layer at higher loading (>1 mg). 7.2. Sensor Applications Based on Self-Supported MIMs

MIMs synthesized either from a self-supported prepolymerization mixture or by the adherence of the MIM directly to a transducer are categorized as self-supported MIMs. Because self-supported MIMs have advantages over supported MIMs, researchers have reported many different techniques for preparing self-supported MIMs. The first of its kind was an MIM of about 90-nm thickness on a gold electrode surface prepared by the electropolymerization of phenol in the presence of the print molecule (phenylalanine) with capacitive sensor detection.224 The sensor, which had a response time of 15 min, could analyze phenylalanine within the range of 0.5−8 mg/cm3. Piletsky et al.225 reported a self-assembled monolayer over a gold electrode surface using hexadecyl mercaptan in the presence of cholesterol as the print molecule. The detection of cholesterol, which is an electroinactive compound, was carried out against the electroactive agent Fe(CN)6−3 in the range of (1.5−6.0) × 10−5 mol dm−3 with a performance period of 5 min. A similar approach was employed using chitosan as the self-assembled polymer over a Au electrode surface for the detection of urea.226 Using a fixed electrodeposition time, the authors controlled the thickness of the membrane over the electrode surface. The sensor was used for the analysis of urea in blood samples. Sol−gel spin-coating on a carbon glassy electrode with a thickness of 540 ± 3 nm for use as a sensor for the detection of parathion was also reported as a self-supported MIM.227 The authors used three monomers to facilitate various nonbonding interactions with the print molecule. The sensor showed a linear response over the concentration range from 1.7 × 10−8 to 1.7 × 10−6 mol dm−3. Electrode surface polymerization is among the most attractive techniques for preparing self-supported membranes. One such report involved the use of a gold electrode surface, modified with allyl mercaptan and 1-butanethiol and coated with a polymerization mixture containing the print molecule.228 In the presence of electrolyte, namely, LiCl dissolved in dimethylformamide (DMF), the sensor exhibited a detection range of (1.0 × 10−6)−(1.0 × 10−5) mol dm−3. Recently, a glass electrode with graphene was coated with an MIP through in situ polymerization, resulting in distinguishable cyclic voltammetry curves with changing concentration of analyte.229 Graphene, which has the property of enhancing electrical T

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

8. PERSPECTIVES In membrane separation, as mentioned in section 1, permselectivity and throughput (flux) are two key factors. The enhancement of the former is relatively easy: The incorporation of molecular recognition sites into the membrane by applying standard molecular imprinting or alternative molecular imprinting leads to increased solubility or partition selectivity, and this, as a result, contributes to an increase in permselectivity. The distribution of molecular recognition sites in membranes is important to render a membrane substrate-specific. As shown in Figure 21, there are four different types of distributions of

electrode surface-spin coating, inorganic sol−gel polymerization, and particulate immobilization.239 7.3. Pervaporation, Nanofiltration, Electrodialysis, and Other MIM-Based Applications

As the title of this subsection suggests, the main applications that employ the technology of molecular imprinting of membranes are reverse osmosis, nanofiltration, and electrodialysis. However, reasons and ways to implement MIM technology have been found for other membrane-based separation processes such as pervaporation, ultrafiltration, and membrane chromatography. It is known that, based on the membrane pore size and starting with the smallest pores, the order of this series is pervaporation < reverse osmosis < nanofiltration < ultrafiltration < microfiltration. Pervaporation was the first application imagined and tested for MIMs when, in 1962, Michaels et al. demonstrated the permselectivity of a p-xylene-imprinted polyethylene membrane for the print molecule p-xylene.37 More recent examples studied pervaporation through a composite MIM based on polysulfone−methyl methacrylate240 and the permselectivity mechanism of a cellulose membrane molecularly imprinted with 1,2-dihydroxybenzene.241 The latter study showed that, at low o-xylene concentrations, the permselectivity of the prepared cellulose MIMs depends on both the solubility and diffusivity selectivities, whereas at higher concentrations, the diffusivity selectivity is the dominant factor. Reverse osmosis, a membrane-based separation process known especially as a desalination method to produce drinking water from seawater, uses high pressures, up to 100 bar, to achieve a satisfactory purity. A goal is that other fine purification processes related to the domains of pharmaceuticals, medicine, analytics, and microelectronics will become possible or economically feasible through the molecular imprinting of membranes.85,242,243 Compared with RO, which makes use of nonporous (dense) membranes, nanofiltration is a superior pore-size membranebased separation method, for which molecular imprinting technology promises to allow higher separation properties.244−246 A recent study involved the development of a hybrid, two-step separation process aiming to achieve low genotoxic impurity levels for active pharmaceutical compounds, using the print molecule 1,3-diisopropylurea, a potentially genotoxic impurity.244 The purification process started with an initial step of organic solvent nanofiltration through different commercial NF membranes, up to a concentration of 100 ppm 1,3-diisopropylurea and continued with retention of the print molecule by an MIP/Teflon-membrane hybrid cartridge. The main conclusions were that the NF step was efficient only at higher concentrations (100−1000 ppm), whereas becoming the MIP hybrid cartridge was up to 83% efficient at lower concentrations. Electrodialysis, 4 0 , 5 5 − 5 7 , 6 1 , 6 2 , 7 0 , 7 5 , 8 0 and ultrafiltration100,168,196,197,247,248 are two of the most common separation processes that use membranes, and some of the particular cases that use molecularly imprinted membranes (MIMs) were discussed in more detail within the present review. Particular applications of MIMs are membrane chromatography,249−251 drug delivery,199,210,252 lab-on-a-chip,253,254 bioinspired adhesion membranes,255 enzymatic catalysis,90,256,257 stimuli-responsive membranes,258,259 microbial fuel-cell membranes,260 and even biocompatible materials for artificial organs and prosthetics.261,262

Figure 21. Four types of membranes with molecular recognition sites. Reproduced with permission from ref 266. Copyright 2014 Lifescience Global.

active molecularly imprinted sites in membranes. Figure 21a shows the results for a conventional cast membrane, wherein the molecular recognition sites, although delocalized, are deeply buried in the membrane. In contrast, Figure 21b shows the spread of molecular recognition sites on the surface of the membrane, which will lead to an enhancement of the rate of incorporation of the target molecules into the membrane. In contrast to conventional cast membranes, nanofiber membranes have a much higher availability of molecular recognition sites because of their large surface-area-to-volume ratios, as described in section 6.1 Although availability of molecular recognition sites in MINFMs is high, the accessibility of molecular recognition sites within the nanofibers is difficult for permeants, also known as target molecules. Therefore, the most suitable membrane form is a membrane with its molecular recognition sites localized on the surface of a nanofiber membrane, which is described again in the present section as method 2 of the plausible methods for furthering enhancing the membrane performance of MINFMs. Even though permselectivity is enhanced by the technique of molecular imprinting, a trade-off relationship between flux and permselectivity still remains. Ideally, molecular imprinting is expected to enhance both permselectivity and flux; at worst, the flux should be enhanced without compromising the permselectivity or the membrane performance. To this end, an ideal morphological form of separation membrane, the nanofiber fabric, has emerged. The form of nanofiber fabric provides a large surface area and a high porosity; the former contributes to the enhancement of selectivity through the incorporation of U

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

in the field of molecularly imprinted polymers. He is currently a postdoctoral fellow at the same university, as well as a research scientist third degree (equivalent lecturer) at INCDCP-ICECHIM Institute in Bucharest. His research interests are in the field of molecularly imprinted polymers/membranes and advanced separation and purification methods.

target molecules into a given membrane and the latter contributes to an increase in the diffusivity of the permeant. The flux and permselectivity trade-off in membrane separations is considered unsolved and/or unsolvable, innate problem by many membranologists. Now, however, a possible breakthrough in solving the flux−permselectivity trade-off relationship in membrane separation could be provided by MINFMs that are fabricated by simultaneous electrospray deposition and alternative molecular imprinting. There are four plausible methods for further enhancing the membrane performance of MINFMs: (1) narrowing the diameters of MINFMs, which will lead to an increased surface area (surfaceto-volume ratio), and narrowing the mesh size between nanofibers; (2) localizing the molecular recognition sites on the surfaces of the nanofibers, which can be achieved by applying coaxial two-capillary spinnerets118,119,123,263−266 or postreaction surface modification of the nanofibers; (3) enhancing the concentration of molecular recognition sites by applying a higher molecular imprinting ratio; and (4) employing covalent molecular imprinting during the electrospray deposition process instead of noncovalent molecular imprinting, so that a given print molecule can effectively work as a print molecule. These methods are all modifications of the membrane itself. Membrane separation employing multistage cascade separation267−269 is an alternative for attaining high membrane performance as an improvement of membrane separation process based on the fact that nanofiber membranes give high flux values.143

ACKNOWLEDGMENTS The authors thank Dr. Rita Darkow of Specialty Medicine, Bayer Pharma AG, for stimulating discussions and for translating Russian articles into English. K.T. acknowledges the Central Tibetan Administration of His Holiness the Dalai Lama for a research scholarship. Ş.-O.D. acknowledges access to the scientific literature as a postdoctoral fellow of U.P.B. by means of the Sectoral Operational Programme Human Resources Development 2007−2013 of the Ministry of European Funds through Financial Agreement POSDRU/ 159/1.5/S/134398. REFERENCES (1) Ho, W. S. W.; Sirkar, K. K. Membrane Handbook; Chapman & Hall: New York, 1992; Chapter 1. (2) Mulder, M. Basic Principles of Membrane Technology, 2nd ed.; Kluwer Academic Press: Dordrecht, The Netherlands, 1996; Chapter 1. (3) Baker, R. W. Membrane Technology and Applications, 3rd ed.; Wiley: Chichester, U.K., 2012; Chapter 1. (4) Chromatography can also be continuously operated as simulated moving bed (SMB) chromatography, the operation of which is complicated compared with that of membrane separation. (5) Lehn, J.-M. Supramolecular Chemistry−Scope and Perspective Molecules, Supermolecules, and Molecular Devices. Angew. Chem., Int. Ed. Engl. 1988, 27, 89−112. (6) Cram, D. J. The Design of Molecular Hosts, Guests, and Their Complexes. Angew. Chem., Int. Ed. Engl. 1988, 27, 1009−1020. (7) Pedersen, C. H. The Discovery of Crown Ethers. Angew. Chem., Int. Ed. Engl. 1988, 27, 1021−1027. (8) Bender, M. L.; Komiyama, M. Cyclodextrin Chemistry; Springer: Berlin, 1978. (9) Rebek, J., Jr. Model Studies in Molecular Recognition. Science 1987, 235, 1478−1484. (10) Wulff, G.; Sarhan, A. The Use of Polymers with EnzymeAnalogous Structures for the Resolution of Racemates. Angew. Chem., Int. Ed. Engl. 1972, 11, 341−344. (11) Wulff, G. Molecular Imprinting in Cross-Linked Materials with the Aid of Molecular Templates−A Way towards Artificial Antibodies. Angew. Chem., Int. Ed. Engl. 1995, 34, 1812−1832. (12) Bartsch, R. A., Maeda, M., Eds. Molecular and Ionic Recognition with Imprinted Polymers; ACS Symposium Series; American Chemical Society: Washington, D.C., 1998; Vol. 703. (13) Piletsky, S. A.; Panasyuk, T. L.; Piletskaya, E. V.; Nicholls, I. A.; Ulbricht, M. Receptor and Transport Properties of Imprinted Polymer Membranes − A Review. J. Membr. Sci. 1999, 157, 263−278. (14) Haupt, K.; Mosbach, K. Molecularly Imprinted Polymers and Their Use in Biomimetic Sensors. Chem. Rev. 2000, 100, 2495−2504. (15) Sellergren, B., Ed. Molecularly Imprinted Polymers; Elsevier: Amsterdam, 2000. (16) Yoshikawa, M. Molecularly Imprinted Polymeric Membranes. Bioseparation 2001, 10, 277−286. (17) Komiyama, M.; Takeuchi, T.; Mukawa, T.; Asanuma, H. Molecular Imprinting; Wiley-VCH: Weinheim, Germany, 2003. (18) Ulbricht, M. Membrane Separation Using Molecularly Imprinted Polymers. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2004, 804, 113−125. (19) Cormack, P. A. G.; Elorza, A. Z. Molecularly Imprinted Polymers: Synthesis and Characterization. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2004, 804, 173−182.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected] or [email protected]. Notes

The authors declare no competing financial interest. Biographies Masakazu Yoshikawa received his Bachelor’s (1976) and Ph.D. (1982) degrees from Kyoto University under the supervision of Prof. Takeo Shimidzu. He then joined Prof. Naoya Ogata’s research group at Sophia University in Tokyo as an Assistant Professor. In 1986, he moved to the Department of Chemical Engineering at Kyoto University, where he worked as an Assistant Professor. In 1991, he joined the Department of Polymer Science and Engineering at Kyoto Institute of Technology, where he is currently a Professor. From 1988 to 1990, he was a visiting scientist at National Research Council of Canada and worked with Dr. Takeshi Matsuura. From 2007 to 2009, he was vice president of the Membrane Society of Japan. He is Editorin-Chief of Journal of Membrane and Separation Technology (Lifescience Global, Canada) and Co-Editor-in-Chief of Journal of Chemical Engineering Research Updates. Today, his research interests focus on molecular recognition, molecular imprinting, nanofiber membranes, and membrane materials chemistry. Kalsang Tharpa (Ph.D., Chemistry) is interested in molecular imprinting, sensors, green chemistry, and catalysis. He is a recipient of the Central Tibetan Administration of His Holiness the Dalai Lama’s research scholarship and cofounder of the Tibetan Scientific Society, a science outreach program. He is currently a scientist at SABIC Research & Technology Center, Bangalore, India. Ştefan-Ovidiu Dima received his Ph.D. in chemical engineering in 2013 from University Politehnica of Bucharest, Romania, for a theme V

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(20) Alexander, C.; Andersson, H. S.; Andersson, L. I.; Ansell, R. J.; Kirsch, N.; Nicholls, I. A.; O’Mahony, J.; Whitcombe, M. J. Molecular Imprinting Science and Technology: A Survey of the Literature for the Years up to and Including 2003. J. Mol. Recognit. 2006, 19, 106−180. (21) Maier, N. M.; Lindner, W. Chiral Recognition Applications of Molecularly Imprinted Polymers: A Critical Review. Anal. Bioanal. Chem. 2007, 389, 377−397. (22) Yoshikawa, M.; Tanioka, A.; Matsumoto, H. Molecularly Imprinted Nanofiber Membranes. Curr. Opin. Chem. Eng. 2011, 1, 18−26. (23) Yoshikawa, M.; Higuchi, A. Enantioselective Membranes. In Encyclopedia of Membranes and Technology [Online]; Hoek, E. M. V.; Tarabara, V. V., Eds.; Wiley: New York, 2013. DOI: 10.1002/ 9781118522318.emst131. http://onlinelibrary.wiley.com/doi/10. 1002/9781118522318.emst131/abstract. (24) Whitcombe, M. C.; Kirsch, N.; Nicholls, I. A. Molecular Imprinting Science and Technology: A Survey of the Literature for the Years 2004−2011. J. Mol. Recognit. 2014, 27, 297−491. (25) Breinl, F.; Haurowitz, F. Chemical Examinations on the Precipitate from Haemoglobin and Anti-Haemoglobin Serum and Comments on the Nature of Antibodies. Hoppe-Seyler's Z. Physiol. Chem. 1930, 192, 45−57. (26) Mudd, S. A Hypothetical Mechanism of Antibody Formation. J. Immunol. 1932, 23, 423−427. (27) Pauling, L. A Theory of the Structure and Process of Formation of Antibodies. J. Am. Chem. Soc. 1940, 62, 2643−2657. (28) Polyakov, M. V. Adsorption Properties of Silica Gel and Its Structure. Zhur. Fiz. Khim. 1931, 2, 799−805. (29) Dickey, F. H. The preparation of specific adsorbents. Proc. Natl. Acad. Sci. U. S. A. 1949, 35, 227−229. (30) Dickey, F. H. Specific Adsorption. J. Phys. Chem. 1955, 59, 695− 707. (31) Takagishi, T.; Klotz, I. M. Macromolecule-Small Molecule Interactions; Introduction of Additional Binding Sites in Polyethyleneimine by Disulfide Cross-Linkages. Biopolymers 1972, 11, 483−491. (32) Arshady, R.; Mosbach, K. Synthesis of Substrate-Selective Polymers by Host-Guest Polymerization. Makromol. Chem. 1981, 182, 687−692. (33) Nishide, H.; Deguchi, J.; Tsuchida, E. Selective Adsorption of Metal Ions on Crosslinked Poly(vinylpyridine) Resin Prepared with a Metal Ion as a Template. Chem. Lett. 1976, 5, 169−174. (34) Nishide, H.; Tsuchida, E. Selective Adsorption of Metal Ions on Poly(4-vinylpyridine) Resins in which the Ligand Chain is Immobilized by Crosslinking. Makromol. Chem. 1976, 177, 2295− 2310. (35) Hutchins, R. S.; Bachas, L. G. Nitrate-Selective Electrode Developed by Electrochemically Mediated Imprinting/Doping of Polypyrrole. Anal. Chem. 1995, 67, 1654−1660. (36) Yoshikawa, M.; Izumi, J.; Kitao, T.; Koya, S.; Sakamoto, S. Membranes for Enantioselective Separation: Synthesis and Characterization. In Preprints for the 16th Annual Meeting of the Membrane Society of Japan; Membrane Society of Japan: Tokyo, 1994; p A1-1-1. (37) Michaels, A. S.; Baddour, R. F.; Bixler, H. J.; Choo, C. Y. Conditioned Polyethylene as a Permselective Membrane: Separation of Isomeric Xylenes. Ind. Eng. Chem. Process Des. Dev. 1962, 1, 14−25. (38) Braco, L.; Dabulis, K.; Klibanov, A. M. Production of Abiotic Receptors by Molecular Imprinting of Proteins. Proc. Natl. Acad. Sci. U. S. A. 1990, 87, 274−277. (39) Ståhl, M.; Månsson, M.-T.; Mosbach, K. The Synthesis of a DAmino Acid Ester in an Organic Media with α-Chymotrypsin Modified by a Bio-Imprinting Procedure. Biotechnol. Lett. 1990, 12, 161−166. (40) Yoshikawa, M.; Izumi, J.; Ooi, T.; Kitao, T.; Guiver, M. D.; Robertson, G. P. Carboxylated Polysulfone Membranes Having a Chiral Recognition Site Induced by an Alternative Molecular Imprinting Technique. Polym. Bull. 1998, 40, 517−524. (41) Kondo, Y.; Yoshikawa, M.; Okushita, H. Molecularly Imprinted Polyamides for Chiral Recognition. Polym. Bull. 2000, 44, 517−524.

(42) Yoshikawa, M.; Izumi, J.; Guiver, M. D.; Robertson, G. P. Recognition and Selective Transport of Nucleic Acid Component through Molecularly Imprinted Polymeric Membranes. Macromol. Mater. Eng. 2001, 286, 52−59. (43) Yoshikawa, M.; Asano, Y.; Guiver, M. D. Selective Recognition and Permeation of Bisphenol A with Molecularly Imprinted Polyamide Membranes. Maku 2001, 26, 185−188. (44) Yoshikawa, M.; Hotta, N.; Kyoumura, J.; Osagawa, Y.; Aoki, T. Chiral Recognition Sites from Carbonyldioxyglycerol Moiety by an Alternative Molecular Imprinting. Sens. Actuators, B 2005, 104, 282− 288. (45) Yoshikawa, M.; Guiver, M. D.; Robertson, G. P. Molecularly Imprinted Films from Torlon® Polyamide-Imide. J. Mol. Struct. 2005, 739, 41−46. (46) Yoshikawa, M.; Murakoshi, K.; Kogita, T.; Hanaoka, K.; Guiver, M. D.; Robertson, G. P. Chiral Separation Membranes from Modified Polysulfone Having Myrtenal-Derived Terpenoid Side Groups. Eur. Polym. J. 2006, 42, 2532−2539. (47) Yoshikawa, M.; Nakai, K.; Matsumoto, H.; Tanioka, A.; Guiver, M. D.; Robertson, G. P. Molecularly Imprinted Nanofiber Membranes from Carboxylated Polysulfone by Electrospray Deposition. Macromol. Rapid Commun. 2007, 28, 2100−2105. (48) Yoshikawa, M.; Guiver, M. D.; Robertson, G. P. Surface Plasmon Resonance Studies on Molecularly Imprinted Films. J. Appl. Polym. Sci. 2008, 110, 2826−2832. (49) Hatanaka, M.; Nishioka, Y.; Yoshikawa, M. Polyurea with LLysinyl Residues as Components: Application to Membrane Separation of Enantiomers. Macromol. Chem. Phys. 2011, 212, 1351−1359. (50) Sueyoshi, Y.; Utsunomiya, A.; Yoshikawa, M.; Robertson, G. P.; Guiver, M. D. Chiral Separation with Molecularly Imprinted Polysulfone-Aldehyde Derivatized Nanofiber Membranes. J. Membr. Sci. 2012, 401−402, 89−96. (51) Hatanaka, M.; Nishioka, Y.; Yoshikawa, M. Polyurea Bearing LLysinyl Residue as a Chiral Building Block and Its Application to Optical Resolution. J. Mater. Sci. Res. 2012, 1, 114−122. (52) Hatanaka, M.; Nishioka, Y.; Yoshikawa, M. Chiral Polyurea with L-Lysinyl Residue Aimed for Optical Resolution. J. Membr. Sep. Technol. 2013, 2, 109−119. (53) Hatanaka, M.; Nishioka, Y.; Yoshikawa, M. Chiral Separation with Polyurea Membrane Consisting of L-Lysinyl Residue: Proposal of Facile Method for Prediction of Permselectivity. J. Appl. Polym. Sci. 2013, 128, 123−131. (54) Yoshikawa, M.; Izumi, J.; Kitao, T.; Koya, S.; Sakamoto, S. Molecularly Imprinted Polymeric Membranes for Optical Resolution. J. Membr. Sci. 1995, 108, 171−175. (55) Yoshikawa, M.; Izumi, J.; Kitao, T. Enantioselective Electrodialysis of N-α-Acetyltryptophans through Molecularly Imprinted Polymeric Membranes. Chem. Lett. 1996, 25, 611−612. (56) Yoshikawa, M.; Izumi, J.; Kitao, T.; Sakamoto, S. Molecularly Imprinted Polymeric Membranes Containing DIDE Derivatives for Optical Resolution of Amino Acids. Macromolecules 1996, 29, 8197− 8203. (57) Yoshikawa, M.; Izumi, J.; Kitao, T. Enantioselective Electrodialysis of Amino Acids with Charged Polar Side Chains through Molecularly Imprinted Polymeric Membranes Containing DIDE Derivatives. Polym. J. 1997, 29, 205−210. (58) Yoshikawa, M.; Izumi, J.; Kitao, T.; Sakamoto, S. Alternative Molecularly Imprinted Polymeric Membranes from a Tetrapeptide Residue Consisting of D- or L-Amino Acids. Macromol. Rapid Commun. 1997, 18, 761−767. (59) Yoshikawa, M.; Fujisawa, T.; Izumi, J.; Kitao, T.; Sakamoto, S. Chiral Recognition of N-α-Acetyltryptophans with Molecularly Imprinted Polymeric Membranes Containing DVNE Derivatives. Sen'i Gakkaishi 1998, 54, 77−84. (60) Yoshikawa, M.; Fujisawa, T.; Izumi, J.; Kitao, T.; Sakamoto, S. Molecularly Imprinted Polymeric Membranes Involving Tetrapeptide EQKL Derivatives as Chiral-Recognition Sites toward Amino Acids. Anal. Chim. Acta 1998, 365, 59−67. W

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(61) Yoshikawa, M.; Fujisawa, T.; Izumi, J. Molecularly Imprinted Polymeric Membranes Having EEF Derivatives as a Chiral Recognition Site. Macromol. Chem. Phys. 1999, 200, 1458−1465. (62) Yoshikawa, M.; Izumi, J.; Kitao, T. Alternative Molecular Imprinting, A Facile Way to Introduce Chiral Recognition Sites. React. Funct. Polym. 1999, 42, 93−102. (63) Yoshikawa, M.; Ooi, T.; Izumi, J. Novel Membrane Materials Having EEE Derivatives as a Chiral Recognition Site. Eur. Polym. J. 2001, 37, 335−342. (64) Yoshikawa, M.; Shimada, A.; Izumi, J. Novel Polymeric Membranes Having Chiral Recognition Sites Converted from Tripeptide Derivatives. Analyst 2001, 126, 775−780. (65) Kondo, Y.; Yoshikawa, M. Effect of Solvent Composition on Chiral Recognition Ability of Molecularly Imprinted DIDE Derivatives. Analyst 2001, 126, 781−783. (66) Yoshikawa, M.; Yonetani, K. Molecularly Imprinted Polymeric Membranes with Oligopeptide Tweezers for Optical Resolution. Desalination 2002, 149, 287−292. (67) Yoshikawa, M.; Kondo, Y.; Morita, Y. Relationship between Enantioselectivity of Alternative Molecularly Imprinted Polymeric Membranes and Species of Amino Acid Residues Composing Chiral Recognition Sites. Bioseparation 2001, 10, 323−330. (68) Kondo, Y.; Morita, Y.; Fujimoto, A.; Tounai, M.; Kimura, S.; Yoshikawa, M. Effect of Constituting Amino Acid Residue Numbers on Molecularly Imprinted Chiral Recognition Sites. Chirality 2003, 15, 498−503. (69) Taniwaki, K.; Hyakutake, A.; Aoki, T.; Yoshikawa, M.; Guiver, M. D.; Robertson, G. P. Evaluation of the Recognition Ability of Molecularly Imprinted Materials by Surface Plasmon Resonance (SPR) Spectroscopy. Anal. Chim. Acta 2003, 489, 191−198. (70) Yoshikawa, M.; Izumi, J. Chiral Recognition Sites Converted from Tetrapeptide Derivatives Adopting Racemates as Print Molecules. Macromol. Biosci. 2003, 3, 487−498. (71) Yoshikawa, M.; Nagai, Y.; Moriguchi, K.; Hiraoka, S. Chiral Recognition Ability of Oligopeptide Derivatives Consisting of Glutamyl Residues. J. Appl. Polym. Sci. 2005, 95, 1302−1309. (72) Yoshikawa, M.; Koso, K.; Yonetani, K.; Kitamura, S.; Kimura, S. Optical Resolution of Racemic Amino Acid Derivatives with Molecularly Imprinted Membranes Bearing Oligopeptide Tweezers. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 385−396. (73) Itou, Y.; Nakano, M.; Yoshikawa, M. Optical Resolution of Racemic Amino Acid Derivatives with Molecularly Imprinted Membranes from Tetrapeptide Consisting of Glycinyl Residues. J. Membr. Sci. 2008, 325, 371−375. (74) Izumi, J.; Yoshikawa, M.; Kitao, K. Enantioselective Permeation of Racemic Glutamic Acid through Alternative Molecularly Imprinted Cellulose Acetate Membranes. Maku 1997, 22, 149−154. (75) Yoshikawa, M.; Ooi, T.; Izumi, J. Alternative Molecularly Imprinted Membranes from a Derivative of Natural polymers, Cellulose Acetate. J. Appl. Polym. Sci. 1999, 72, 493−499. (76) Sueyoshi, Y.; Fukushima, C.; Yoshikawa, M. Molecularly Imprinted Nanofiber Membranes from Cellulose Acetate Aimed for Chiral Separation. J. Membr. Sci. 2010, 357, 90−97. (77) Kawasaki, T.; Yoshikawa, M. Nanofiber Membranes from Cellulose Triacetate fro Chiral Separation. Desalin. Water Treat. 2013, 51, 5080−5088. (78) Yoshikawa, M.; Kawamura, K.; Ejima, A.; Aoki, T.; Sakurai, S.; Hayashi, K.; Watanabe, K. Green Polymers from Geobacillus thermodenitrif icans DSM465 − Candidates for Molecularly Imprinted Materials. Macromol. Biosci. 2006, 6, 210−215. (79) Yoshikawa, M.; Kawamura, K.; Watanabe, K. Evaluation of Recognition Ability of Molecularly Imprinted Proteins by Surface Plasmon Resonance (SPR) Spectroscopy. Maku 2007, 32, 40−44. (80) Piletsky, S. A.; Dubey, I. Ya.; Fedoryak, D. M.; Kukhar, V. P. Substrate-Selective Polymeric Membranes. Selective Transfer of Nucleic Acid Components. Biopolim. Kletka 1990, 6, 55−58. (81) Mathew-Krotz, J.; Shea, K. J. Imprinted Polymer Membranes for the Selective Transport of Targeted Neutral Molecules. J. Am. Chem. Soc. 1996, 118, 8154−8155.

(82) Dzgoev, A.; Haupt, K. Enantioselective Molecularly Imprinted Polymer Membranes. Chirality 1999, 11, 465−469. (83) Guo, H.; He, X.; Liang, H. Study of the Binding Characteristics and Transportation Properties of a 4-Aminopyridine Imprinted Polymer Membrane. Fresenius' J. Anal. Chem. 2000, 368, 763−767. (84) Hong, J.-M.; Anderson, P. E.; Qian, J.; Martin, C. R. SelectivePermeable Ultrathin Film Composite Membranes Based on Molecularly-Imprinted Polymers. Chem. Mater. 1998, 10, 1029−1033. (85) 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. (86) Lehmann, M. L.; Brunner, H.; Tovar, G. E. M. Selective Separations and Hydrodynamic Studies: A New Approach Using Molecularly Imprinted Nanosphere Composite Membranes. Desalination 2002, 149, 315−321. (87) Piletsky, S. A.; Matuschewski, H.; Schedler, U.; Wilpert, A.; Piletska, E.; Thiele, T. A.; Ulbricht, M. Surface Functionalization of Porous Polypropylene Membranes with Molecularly Imprinted Polymers by Photograft Copolymerization in Water. Macromolecules 2000, 33, 3092−3098. (88) Wang, H. Y.; Kobayashi, T.; Fujii, N. Surface Molecular Imprinting on Photosensitive Dithiocarbamoyl Polyacrylonitrile Membranes using Photograft Polymerization. J. Chem. Technol. Biotechnol. 1997, 70, 355−362. (89) Hilal, N.; Kochkodan, V. Surface Modified Microfiltration Membranes with Molecularly Recognizing Properties. J. Membr. Sci. 2003, 213, 97−113. (90) Koter, I.; Ceynowa, J. Kinetic Resolution of Chiral Alcohols in Bifunctional Membrane Exhibiting Enzyme Activity and Enantioselective Permeation. J. Mol. Catal. B: Enzym. 2003, 24−25, 17−26. (91) The chiral selectivity of transport through membranes is also represented in terms of the enantiomeric excess (ee) of the permeants. ee is defined as the ratio of the concentration difference to the total concentration of the two enantiomers in the permeant: ee = (|Ci − Cj|)/(Ci + Cj) (92) Danila, D. C.; Banner, L. T.; Karimova, E. J.; Tsurkan, L.; Wang, X.; Pinkhassik, E. Directed Assembly of Sub-Nanometer Thin Organic Materials with Programmed-Size Nanopores. Angew. Chem., Int. Ed. 2008, 47, 7036−7039. (93) Dergunov, S. A.; Pinkhassik, E. Functionalization of Imprinted Nanopores in Nanometer-Thin Organic Materials. Angew. Chem., Int. Ed. 2008, 47, 8264−8267. (94) Dima, Ş.-O.; Sarbu, A.; Dobre, T.; Bradu, C.; Antothe, N.; Radu, A.-L.; Nicolescu, T.-V.; Lungu, A. Molecularly Imprinted Membranes for Selective Separations. Mater. Plast. 2009, 46, 372−378. (95) Dima, Ş.-O.; Dobre, T. Mathematical Modeling and Simulation of Adsorption Process in a Fixed Bed of Molecularly Imprinted Polymeric Pearls. U.P.B. Sci. Bull., Ser. B 2016, 78, 143−154. (96) Dima, Ş.-O.; Meouche, W.; Dobre, T.; Nicolescu, T.-V.; Sarbu, A. Diosgenin-Selective Molecularly Imprinted Pearls Prepared by Wet Phase Inversion. React. Funct. Polym. 2013, 73, 1188−1197. (97) Dima, Ş.-O.; Dobre, T.; Chetraru, O.; Nicolae, C.-A.; Spataru, C.-I.; Sarbu, A. Molecularly Imprinted Poly(acrylonitrile-co-acrylic acid) Matrix with Sclareol. Polym. Eng. Sci. 2014, 54, 1484−1494. (98) Florea, A. M.; Iordache, T. V.; Branger, C.; Ghiurea, M.; Avramescu, S.; Hubca, G.; Sârbu, A. An Innovative Approach to Prepare Hypericin Molecularly Imprinted Pearls Using a ″PhytoTemplate″. Talanta 2016, 148, 37−45. (99) Kakuchi, T.; Takaoka, T.; Yokota, K. Polymeric Chiral Crown Ethers VI. Optical Resolution of α-Amino Acid by Polymers Incorporating 1,3;4,6-Di-O-benzylidene-D-mannitol Residues. Polym. J. 1990, 22, 199−205. (100) Masawaki, T.; Sasai, M.; Tone, S. Optical Resolution of an Amino Acid by an Enantioselective Ultrafiltration Membrane. J. Chem. Eng. Jpn. 1992, 25, 33−39. (101) Aoki, T.; Tomizawa, S.; Oikawa, E. Enantioselective Permeation through Poly{γ-[3-(pentamethyldisiloxantl)propyl]-L-glutamate} Membranes. J. Membr. Sci. 1995, 99, 117−125. X

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(102) Tone, S.; Masawaki, T.; Eguchi, K. The Optical Resolution of Amino Acids by Plasma Polymerized Terpene Membranes. J. Membr. Sci. 1996, 118, 31−40. (103) Wang, H.-D.; Chu, L.-Y.; Song, H.; Yang, J.-P.; Xie, R.; Yang, M. Preparation and Enantiomer Separation Characteristics of Chitosan/β-Cyclodextrin Composite Membranes. J. Membr. Sci. 2007, 297, 262−270. (104) Hazarika, S. Enantioselective Permeation of Racemic Alcohol through Polymeric Membrane. J. Membr. Sci. 2008, 310, 174−183. (105) Kobayashi, T.; Wang, H. Y.; Fujii, N. Molecular Imprinting of Theophylline Acrylonitrile-Acrylic Acid Copolymer Membrane. Chem. Lett. 1995, 24, 927−928. (106) Wang, H. Y.; Kobayashi, T.; Fujii, N. Molecular Imprinting Membranes Prepared by the Phase Inversion Precipitation Technique. Langmuir 1996, 12, 4850−4856. (107) Wang, Y. H.; Kobayashi, T.; Fukaya, T.; Fujii, N. Molecular Imprinting Membranes Prepared by the Phase Inversion Precipitation Technique. 2. Influence of Coagulation Temperature in the Phase Inversion Process on the Encoding in Polymeric Membranes. Langmuir 1997, 13, 5396−5400. (108) Kobayashi, T.; Wang, H. Y.; Fujii, N. Molecular Imprint Membranes of Polyacrylonitrile Copolymers with Different Acrylic Acid Segments. Anal. Chim. Acta 1998, 365, 81−88. (109) Screenivasulu Reddy, P.; Kobayashi, T.; Abe, M.; Fujii, N. Molecular Imprinted Nylon-6 as a Recognition Material of Amino Acids. Eur. Polym. J. 2002, 38, 521−529. (110) Trotta, F.; Drioli, E.; Baggiani, C.; Lacopo, D. Molecular Imprinted Polymeric Membrane for Naringin Recognition. J. Membr. Sci. 2002, 201, 77−84. (111) Trotta, E.; Baggiani, C.; Luda, M. P.; Drioli, E.; Massari, T. A Molecular Imprinted Membrane for Molecular Discrimination of Tetracycline Hydrochloride. J. Membr. Sci. 2005, 254, 13−19. (112) Ramamoorthy, M.; Ulbricht, M. Molecular Imprinting of Cellulose-Sulfonated Polysulfone Blend Membranes for Rhodamine B by Phase Inversion Technique. J. Membr. Sci. 2003, 217, 207−214. (113) Cristallini, C.; Ciardelli, G.; Barbani, N.; Giusti, P. Acrylonitrile-Acrylic Acid Copolymer Membrane Imprinted with Uric Acid for Clinical Uses. Macromol. Biosci. 2004, 4, 31−38. (114) Jiang, Z.; Yu, Y.; Wu, H. Preparation of CS/GPTMS Hybrid Molecularly Imprinted Membrane for Efficient Chiral Resolution of Phenylalanine Isomers. J. Membr. Sci. 2006, 280, 876−882. (115) Ul-Haq, N.; Park, J. K. Chiral Resolution of Phenylalanine by D -Phe Imprinted Membrane Considering Rejection Property. Bioprocess Biosyst. Eng. 2010, 33, 79−86. (116) Morozov, V. N.; Morozova, T. Y.; Kallenbach, N. R. Atomic Force Microscopy of Structures Produced by Electrospraying Polymer Solutions. Int. J. Mass Spectrom. 1998, 178, 143−159. (117) Morozov, V. N.; Morozova, T. Y. Electrospray Deposition as a Method to Fabricate Functionally Active Protein Films. Anal. Chem. 1999, 71, 1415−1420. (118) Ramakrishna, S.; Fujihara, K.; Teo, W.-E.; Lim, T.-C., Ma, Z. An Introduction to Electrospinning and Nanofibers; World Scientific: Singapore, 2005. (119) Greiner, A.; Wendorff, J. H. Electrospinning: A Fascinating Method for the Preparation of Ultrathin Fibers. Angew. Chem., Int. Ed. 2007, 46, 5670−5703. (120) Yoon, K.; Hsiao, B. S.; Chu, B. Functional Nanofibers for Environmental Applications. J. Mater. Chem. 2008, 18, 5326−5334. (121) Chu, B.; Hsiao, B. S. The Role of Polymers in Breakthrough Technologies for Water Purification. J. Polym. Sci., Part B: Polym. Phys. 2009, 47, 2431−2435. (122) Agarwal, S.; Wendorff, J. H.; Greiner, A. Chemistry on Electrospun Polymeric Nanofibers: Merely Routine Chemistry or a Real Challenge? Macromol. Rapid Commun. 2010, 31, 1317−1331. (123) Matsumoto, H.; Tanioka, A. Functionality in Electrospun Nanofibrous Membranes Based on Fiber’s Size, Surface Area, and Molecular Orientation. Membranes 2011, 1, 249−264.

(124) Chronakis, I. S.; Milosevic, B.; Frenot, A.; Ye, L. Generation of Molecular Recognition Sites in Electrospun Polymer Nanofibers via Molecular Imprinting. Macromolecules 2006, 39, 357−361. (125) Chronakis, I. S.; Jakob, A.; Hagström, B.; Ye, L. Encapsulation and Selective Recognition of Molecularly Imprinted Theophylline and 17β-Estradiol Nanoparticles within Electrospun Polymer Nanofibers. Langmuir 2006, 22, 8960−8965. (126) Yoshimatsu, K.; Ye, L.; Lindberg, J.; Chronakis, I. S. Selective Molecular Adsorption Using Electrospun Nanofiber Affinity Membranes. Biosens. Bioelectron. 2008, 23, 1208−1215. (127) 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, 44, 2022− 2024. (128) Kim, W. J.; Chang, J. Y. Molecularly Imprinted Polyimide Nanofiber Prepared by Electrospinning. Mater. Lett. 2011, 65, 1388− 1391. (129) Ma, X.; Chen, Z.; Chen, X.; Chen, R.; Zheng, X. Preparation of Imprinted PVB/b-CD Nanofiber by Electrospinning Technique and Its Selective Binding Abilities for Naringin. Chin. J. Chem. 2011, 29, 1753−1758. (130) Gong, J.; Su, B.; Luo, L.; Zhou, M. Preparation and Recognition Properties of Molecularly Imprinted Polyethersulfone Nanofiber Membranes towards Bisphenol A. Gaofenzi Xuebao 2011, 11, 1266−1272. (131) Wei, Z.; Zhang, Q.; Wang, L.; Peng, M.; Wang, X.; Long, S.; Yang, J. The Preparation and Adsorption Properties of Electrospun Aramid Nanofibers. J. Polym. Sci., Part B: Polym. Phys. 2012, 50, 1414− 1420. (132) Awokoya, K. N.; Moronkola, B. A.; Chigome, S.; Ondigo, D. A.; Tshentu, Z.; Torto, N. Molecularly Imprinted Electrospun Nanofibers for Adsorption of Nickel-5,10,15,20-Tetraphenylporphine (NTPP) in Organic Media. J. Polym. Res. 2013, 20, 148. (133) Li, Y.; Qiu, T.; Xu, X. Preparation of Lead-Ion Imprinted Crosslinked Electron-Spun Chitosan Nanofiber Mats and Application in Lead Removal from Aqueous Solutions. Eur. Polym. J. 2013, 49, 1487−1494. (134) Piperno, S.; Tse Sum Bui, B.; Haupt, K.; Gheber, L. A. Immobilization of Molecularly Imprinted Polymer Nanoparticles in Electrospun Poly(vinyl alcohol) Nanofibers. Langmuir 2011, 27, 1547−1550. (135) Tonglairoum, P.; Chaijaroenluk, W.; Rojanarata, T.; Ngawhirunpat, T.; Akkaramongkolporn, P.; Opanasopit, P. Development and Characterization of Propranol Selective Molecular Imprinted Polymer Composite Electrospun Nanofiber Membrane. AAPS PharmSciTech 2013, 14, 838−846. (136) Liu, H.; Lei, X.; Zhai, Y.; Li, L. Electrospun Nanofiber Membranes Containing Molecularly Imprinted Polymer (MIP) for Rhodamine B (RhB). Adv. Chem. Eng. Sci. 2012, 2, 266−274. (137) Cao, F.; Liao, J.; Yang, K.; Bai, P.; Wei, Q.; Zhao, C. SelfAssembly Molecularly Imprinted Nanofiber for 4-HA Recognition. Anal. Lett. 2010, 43, 2790−2797. (138) Wu, Y.-T.; Zhang, Y.-H.; Zhang, M.; Liu, F.; Wan, Y.-C.; Huang, Z.; Ye, L.; Zhou, Q.; Shi, Y.; Lu, B. Selective and Simultaneous Determination of Trace Bisphenol A and Tebuconazole in Vegetable and Juice Samples by Membrane-Based Molecularly Imprinted SolidPhase Extraction and HPLC. Food Chem. 2014, 164, 527−535. (139) Dima, Ş.-O. Equilibrium and Kinetic Isotherms and Parameters for Molecularly Imprinted with Sclareol Poly(acrylonitrile-co-acrylic acid) Matrix. Polym. Eng. Sci. 2015, 55, 1152−1162. (140) Ogunlaja, A. S.; Sautoy, C.; Torto, N.; Tshentu, Z. R. Design, Fabrication and Evaluation of Intelligent Sulfone-Selective Polybenzimidazole Nanofibers. Talanta 2014, 126, 61−72. (141) Ogunlaja, A. S.; Coombes, M. J.; Torto, N.; Tshentu, Z. R. The Adsorptive Extraction of Oxidized Sulfur-Containing Compounds from Fuels by Using Molecularly Imprinted Chitosan Materials. React. Funct. Polym. 2014, 81, 61−76. (142) Schultz, S. G. Basic Principles of Membrane Transport; Cambridge University Press: Cambridge, U.K., 1980; Chapter 1. Y

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(143) Mizushima, H.; Yoshikawa, M.; Li, N.; Robertson, G. P.; Guiver, M. D. Electrospun Nanofiber Membranes from Polysulfones with Chiral Selector Aimed for Optical Resolution. Eur. Polym. J. 2012, 48, 1717−1725. (144) Shiomi, K.; Yoshikawa, M. Multi-Stage Chiral Separation with Electrospun Chitin Nanofiber Membranes. Sep. Purif. Technol. 2013, 118, 300−304. (145) Ye, Y.-T.; Ma, X. H.; Xu, Z. L.; Zhang, Y. Theophylline Molecular Imprinted Composite Membranes Prepared on a Ceramic Hollow Fiber Substrate. Ind. Eng. Chem. Res. 2014, 53, 346−354. (146) Dai, J.; Zhou, Z.; Zhao, C.; Wei, X.; Dai, X.; Gao, L.; Cao, Z.; Yan, Y. Versatile Method to Obtain Homogeneous Imprinted Polymer Thin Film at Surface of Superparamagnetic Nanoparticles for Tetracycline Binding. Ind. Eng. Chem. Res. 2014, 53, 7157−7166. (147) Chen, J.; Bai, L.; Tian, M.; Zhou, X.; Zhang, Y. Hollow-Fiber Membrane Tube Embedded with a Molecularly Imprinted Monolithic Bar for the Microextraction of Triazine Pesticides. Anal. Methods 2014, 6, 602−608. (148) Meng, M.; Feng, Y.; Liu, Y.; Wang, Y.; Yan, Y. Preparation of Composite-Imprinted Alumina Membrane for Effective Separation of p-Hydroxybenzoic Acid from Its Isomer Using Box-Behnken DesignBased Statistical Modeling. J. Appl. Polym. Sci. 2014, 131, 40621. (149) Liu, Y.; Meng, M. J.; Yao, J. T.; Da, Z. L.; Feng, Y. H.; Yan, Y. S.; Li, C. X. Selective Separation of Phenol from Salicylic Acid Effluent over Molecularly Imprinted Polystyrene Nanospheres Composite Alumina Membranes. Chem. Eng. J. 2016, 286, 622−631. (150) Zhang, Y. Q.; Gao, X. Q.; Xiang, L.; Zhang, Y. H.; Diniz da Costa, J. C. Study on Composite Membranes with High Selective Permeance Properties. J. Membr. Sci. 2010, 346, 318−326. (151) Xi, F. N.; Wu, J. M. Preparation of Macroporous Chitosan Layer Coated on Silica Gel and Its Application to Affinity Chromatography for Trypsin Inhibitor Purification. React. Funct. Polym. 2006, 66, 682−688. (152) Dong, J.; Fan, H.; Sui, D.; Li, L.; Sun, T. Sampling 4Chlorophenol in Water by DGT Technique with Molecularly Imprinted Polymer as Binding Agent and Nylon Membrane as Diffusive Layer. Anal. Chim. Acta 2014, 822, 69−77. (153) Zhang, M.; He, X. W.; Qin, L.; Li, W. Y.; Chen, L. X.; Zhang, Y. K. Preparation of Protein Molecularly Imprinted Composite Membrane by Sol-Gel Coating Method and Its Permeation Mechanism. Chem. J. Chin. Univ. 2008, 29, 498−504. (154) Kochkodan, V.; Hilal, N.; Melnik, V.; Kochkodan, O.; Vasilenko, O. Selective Recognition of Organic Pollutants in Aqueous Solutions with Composite Imprinted Membranes. Adv. Colloid Interface Sci. 2010, 159, 180−188. (155) Scorrano, S.; Mergola, L.; Di Bello, M. P.; Lazzoi, M. R.; Vasapollo, G.; del Sole, R. Molecularly Imprinted Composite Membranes for Selective Detection of 2-Deoxyadenosine in Urine Samples. Int. J. Mol. Sci. 2015, 16, 13746−13759. (156) Schneider, F.; Piletsky, S.; Piletska, E.; Guerreiro, A.; Ulbricht, M. Comparison of Thin-Layer and Bulk MlPs Synthesized by Photoinitiated in situ Crosslinking Polymerization from the Same Reaction Mixtures. J. Appl. Polym. Sci. 2005, 98, 362−372. (157) Kan, B. H.; Feng, L. Z.; Zhao, K. Y.; Wei, J. F.; Zhu, D. W.; Zhang, L. H.; Ren, Q. Preparation and Rebinding Properties of Protein-Imprinted Polysiloxane Using Mesoporous Calcium Silicate Grafted Non-Woven Polypropylene as Matrix. J. Mol. Recognit. 2016, 29, 115−122. (158) Wu, Y. L.; Yan, M.; Yan, Y. S.; Liu, X. L.; Meng, M. J.; Lv, P.; Pan, J. M.; Huo, P. W.; Li, C. X. Fabrication and Evaluation of Artemisinin-Imprinted Composite Membranes by Developing a Surface Functional Monomer-Directing Prepolymerization System. Langmuir 2014, 30, 14789−14796. (159) Wang, C. L.; Hu, X. L.; Guan, P.; Wu, D. F.; Yang, L. F.; Du, C. B. Preparation of Molecularly Imprinted Regenerated Cellulose Composite Membranes by Surface-Initiated Atom Transfer Radical Polymerization Method for Selective Recognition of Lysozyme. Adsorpt. Sci. Technol. 2015, 33, 411−425.

(160) He, Z. H.; Meng, M. J.; Yan, L.; Zhu, W. H.; Sun, F. Q.; Yan, Y. S.; Liu, Y.; Liu, S. J. 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. (161) Zhao, K.; Feng, L.; Lin, H.; Fu, Y.; Lin, B.; Cui, W.; Li, S.; 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. (162) Kan, B. H.; Lin, B. B.; Zhao, K. Y.; Zhang, X. X.; Feng, L. Z.; Wei, J. F.; Fan, Y. C. Imprinting of Bovine Serum Albumin in a Nonwoven Polypropylene Membrane Supported Polyacrylamide/ Calcium Alginate Interpenetrating Polymer Network Hydrogel. RSC Adv. 2014, 4, 55846−55852. (163) Chen, J. H.; Li, G. P.; Liu, Q. L.; Ni, J. C.; Wu, W. B.; Lin, J. M. Cr(III) Ionic Imprinted Polyvinyl Alcohol/Sodium Alginate (PVA/ SA) Porous Composite Membranes for Selective Adsorption of Cr(III) Ions. Chem. Eng. J. 2010, 165, 465−473. (164) Wu, H.; Zhao, Y. Y.; Nie, M. C.; Jiang, Z. Y. Molecularly Imprinted Organic-Inorganic Hybrid Membranes for Selective Separation of Phenylalanine Isomers and Its Analogue. Sep. Purif. Technol. 2009, 68, 97−104. (165) Zhang, W. C.; Zhang, Q.; Zhang, X. Y.; Wang, R. X.; Ge, C. Y. Study on Preparation and Properties of PA6/PSMA-OA Molecularly Imprinted Composite Membranes. Rev. Chim. (Bucharest) 2012, 63, 347−352. (166) Moein, M. M.; Javanbakht, M.; Karimi, M.; Akbari-Adergani, B.; Abdel-Rehim, M. A New Strategy for Surface Modification of Polysulfone Membrane by in situ Imprinted Sol-Gel Method for the Selective Separation and Screening of L-Tyrosine as a Lung Cancer Biomarker. Analyst 2015, 140, 1939−1946. (167) Du, X.; Zhang, H.; Hao, X.; Guan, G.; Abudula, A. Facile Preparation of Ion-Imprinted Composite Film for Selective Electrochemical Removal of Nickel(II) Ions. ACS Appl. Mater. Interfaces 2014, 6, 9543−9549. (168) Nasser, I. I.; Algieri, C.; Garofalo, A.; Drioli, E.; Ahmed, C.; Donato, L. Hybrid Imprinted Membranes for Selective Recognition of Quercetin. Sep. Purif. Technol. 2016, 163, 331−340. (169) Donato, L.; Tasselli, F.; De Luca, G.; Garcia Del Blanco, S.; Drioli, E. Novel Hybrid Molecularly Imprinted Membranes for Targeted 4, 4′-Methylendianiline. Sep. Purif. Technol. 2013, 116, 184−191. (170) Takeda, K.; Kobayashi, T. Hybrid Molecularly Imprinted Membranes for Targeted Bisphenol Derivatives. J. Membr. Sci. 2006, 275, 61−69. (171) dos Santos Moretti, E.; de Fatima Giarola, J.; Kuceki, M.; Prete, M. C.; Pereira, A. C.; Teixeira Tarley, C. R. A Nanocomposite Based on Multi-Walled Carbon Nanotubes Grafted by Molecularly Imprinted Poly(methacrylic acid-hemin) as a Peroxidase-Like Catalyst for Biomimetic Sensing of Acetaminophen. RSC Adv. 2016, 6, 28751− 28760. (172) Luo, Z. M.; Shu, H.; Guo, P. Q.; Zheng, P. L.; Pan, X. Y.; Du, W.; Liu, R. L.; Zeng, A. G.; Chang, C.; Fu, Q. Study of the Allergenic Benzypenicilloyl-HSA and Its Specific Separation from Human Plasma by a Pre-Designed Hybrid Imprinted Membrane. RSC Adv. 2016, 6, 15549−15557. (173) Dinc, M.; Basan, H.; Diemant, T.; Behm, R. J.; Lindén, M.; Mizaikoff, B. Inhibitor-Assisted Synthesis of Silica-Core Microbeads with Pepsin-Imprinted Nanoshells. J. Mater. Chem. B 2016, 4, 4462. (174) Pérez-Quintanilla, D.; Sierra, I. Factors Affecting Hg(II) Adsorption on Hybrid Nanostructured Silicas: Influence of the Synthesis Conditions. J. Porous Mater. 2014, 21, 71−80. (175) Iordache, T. V.; Sarbu, A.; Donescu, D.; Nicolae, C. A.; Jerca, F. A.; Dima, S. O. Selecting the Nature of Imprinted Molecular Organosilica Sieves with Gallic Acid via Thermal Analyses. J. Therm. Anal. Calorim. 2014, 118, 1039−1048. (176) Zhang, M.; Zhao, H. T.; Yang, X.; Zhang, W. T.; Wang, J.; Liu, G. Y.; Zhang, H.; Dong, A. J. Preparation and Characterization of Surface Molecularly Imprinted Film Coated on a Magnetic Nanocore Z

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

for the Fast and Selective Recognition of the New Neonicotinoid Insecticide Paichongding (IPP). RSC Adv. 2016, 6, 3714−3722. (177) Tao, Q. L.; Li, Y.; Shi, Y.; Liu, R. J.; Zhang, Y. W.; Guo, J. Y. Application of Molecular Imprinted Magnetic Fe3O4@SiO2 Nanoparticles for Selective Immobilization of Cellulase. J. Nanosci. Nanotechnol. 2016, 16, 6055−6060. (178) Han, Q.; Wang, X.; Yang, Z. Y.; Zhu, W. Y.; Zhou, X. M.; Jiang, H. J. Fe3O4@rGO Doped Molecularly Imprinted Polymer Membrane Based on Magnetic Field Directed Self-Assembly for the Determination of Amaranth. Talanta 2014, 123, 101−108. (179) 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. (180) Tamahkar, E.; Kutsal, T.; Denizli, A. Surface Imprinted Bacterial Cellulose Nanofibers for Cytochrome c Purification. Process Biochem. 2015, 50, 2289−2297. (181) Piacham, T.; Isarankura-Na-Ayudhya, C.; Prachayasittikul, V. A Simple Method for Creating Molecularly Imprinted Polymer-Coated Bacterial Cellulose Nanofibers. Chem. Pap. 2014, 68, 838−841. (182) Xu, S. F.; Lu, H. Z.; Chen, L. X.; Wang, X. C. Molecularly Imprinted TiO2 Hybridized Magnetic Fe3O4 Nanoparticles for Selective Photocatalytic Degradation and Removal of Estrone. RSC Adv. 2014, 4, 45266−45274. (183) Mizutani, N.; Yang, D. H.; Selyanchyn, R.; Korposh, S.; Lee, S. W.; Kunitake, T. Remarkable Enantioselectivity of Molecularly Imprinted TiO2 Nano-Thin Films. Anal. Chim. Acta 2011, 694, 142−150. (184) Cai, Z. F.; Dai, H. J.; Si, S. H.; Ren, F. L. Molecular Imprinting and Adsorption of Metallothionein on Nanocrystalline Titania Membranes. Appl. Surf. Sci. 2008, 254, 4457−4461. (185) Zhang, J.; Wang, C. Y.; Niu, Y. H.; Li, S. J.; Luo, R. Q. Electrochemical Sensor Based on Molecularly Imprinted Composite Membrane of Poly(O-aminothiophenol) with Gold Nanoparticles for Sensitive Determination of Herbicide Simazine in Environmental Samples. Sens. Actuators, B 2016, DOI: 10.1016/j.snb.2016.02.068. (186) Wu, Y.; Yan, M.; Liu, X. L.; Lv, P.; Cui, J. Y.; Meng, M. J.; Dai, J. D.; Yan, Y. S.; Li, C. X. Accelerating the Design of Multi-Component Nanocomposite Imprinted Membranes by Integrating a Versatile Metal-Organic Methodology with a Mussel-Inspired Secondary Reaction Platform. Green Chem. 2015, 17, 3338−3349. (187) Wang, J. J.; Liu, F. Enhanced and Selective Adsorption of Heavy Metal Ions on Ion-Imprinted Simultaneous Interpenetrating Network Hydrogels. Des. Monomers Polym. 2014, 17, 19−25. (188) Boysen, R. I.; Schwarz, L. J.; Li, S.; Chowdhury, J.; Hearn, M. T. W. Photo-Lithographic Patterning of Biomimetic Molecularly Imprinted Polymer Thin Films onto Silicon Wafers. Microsyst. Technol. 2014, 20, 2037−2043. (189) Dai, J. D.; He, J. S.; Xie, A.; Gao, L.; Pan, J. M.; Chen, X.; Zhou, Z. P.; Wei, X.; Yan, Y. S. Novel Pitaya-Inspired Well-Defined Core-Shell Nanospheres with Ultrathin Surface Imprinted Nanofilm from Magnetic Mesoporous Nanosilica for Highly Efficient Chloramphenicol Removal. Chem. Eng. J. 2016, 284, 812−822. (190) Lazau, C.; Iordache, T. V.; Florea, A. M.; Orha, C.; Bandas, C.; Radu, A. L.; Sarbu, A.; Rotariu, T. Towards Developing an Efficient Sensitive Element for Trinitrotoluene Detection: TiO2 Thin Films Functionalized with Molecularly Imprinted Copolymer Films. Appl. Surf. Sci. 2016, 384, 449. (191) Zarghami, S.; Mohammadi, T.; Kazemimoghadam, M. Diffusive Transport of Cu(II) Ions through Thin Ion Imprinted Polymeric Membranes. Chem. Pap. 2014, 68, 1325−1331. (192) Elmlund, L.; Suriyanarayanan, S.; Wiklander, J.; Aastrup, T.; Nicholls, I. A. Biotin Selective Polymer Nano-Films. J. Nanobiotechnol. 2014, 12, 8. (193) Sergeyeva, T. A.; Piletsky, S. A.; Piletska, E. V.; Brovko, O. O.; Karabanova, L. V.; Sergeeva, L. M.; El'skaya, A. V.; Turner, E. A. P. F. In Situ Formation of Porous Molecularly Imprinted Polymer Membranes. Macromolecules 2003, 36, 7352−7357.

(194) Hillberg, A. L.; Brain, K. R.; Allender, C. J. Design and Evaluation of Thin and Flexible Theophylline Imprinted Polymer Membrane Materials. J. Mol. Recognit. 2009, 22, 223−231. (195) Fan, J.-P.; Li, L.; Tian, Z.-Y.; Xie, C. F.; Song, F.-t.; Zhang, X.H.; Zhu, J.-H. A Novel Free-Standing Flexible Molecularly Imprinted Membrane for Selective Separation of Synephrine in Methanol-Water Media. J. Membr. Sci. 2014, 467, 13−22. (196) Feng, H.; Mao, X.; Chu, B.; Xie, C. Influence of Gelling Properties on Protein Imprinted Agarose Gel Membrane. J. Appl. Polym. Sci. 2014, 131, 40323. (197) Zhao, K.; Lin, B.; Cui, W.; Feng, L.; Chen, T.; Wei, J. Preparation and Adsorption of Bovine Albumin-Imprinted Polyacrylamide Hydrogel Membrane Grafted on Non-Woven Polypropylene. Talanta 2014, 121, 256−262. (198) Akgönüllü, S.; Yavuz, H.; Denizli, A. Preparation of Imprinted Cryogel Cartridge for Chiral Separation of l-Phenylalanine. Artif. Cells, Nanomed., Biotechnol. 2016, 1. (199) Ç etin, K.; Denizli, A. 5-Fluorouracil Delivery from Metal-Ion Mediated Molecularly Imprinted Cryogel Discs. Colloids Surf., B 2015, 126, 401−406. (200) Gun’ko, V. M.; Savina, I. N.; Mikhalovsky, S. V. Morphological, Structural and Adsorption Characterization. Adv. Colloid Interface Sci. 2013, 187−188, 1−46. (201) Cristallini, C.; Bellotti, E.; Spezia, F.; Rosellini, E.; Cascone, M. G.; Barbani, N. A New Strategy to Reduce Amyloid Deposition Using Peptide-Imprinted Membranes. J. Appl. Biomater. Funct. Mater. 2016, 14, 0. (202) Ding, H. Y.; Chen, R. F.; Liu, M. M.; Huang, R.; Du, Y. M.; Huang, C.; Yu, X. Y.; Feng, X. H.; Liu, F. Preparation and Characterization of Biocompatible Molecularly Imprinted Poly(ionic liquid) Films on the Surface of Multi-Walled Carbon Nanotubes. RSC Adv. 2016, 6, 43526−43538. (203) Lopez, T. D.-F.; Francos, M. A. E.; Gonzalez, A. F.; DiazGarcia, M. E.; Badia-Laino, R. Controlled Release of Nafcillin Using Biocompatible ″Dummy″ Molecularly Imprinted Sol-Gel Nanospheres. Curr. Top. Med. Chem. 2015, 15, 262−270. (204) Roy, E.; Maity, S. K.; Patra, S.; Madhuri, R.; Sharma, P. K. A Metronidazole-Probe Sensor Based on Imprinted Biocompatible Nanofilm for Rapid and Sensitive Detection of Anaerobic Protozoan. RSC Adv. 2014, 4, 32881−32893. (205) Li, Y.; Zhang, J.; Xu, C.; Zhou, Y. F. Crosslinked Chitosan Nanofiber Mats Fabricated by One-Step Electrospinning and IonImprinting Methods for Metal Ions Adsorption. Sci. China: Chem. 2016, 59, 95−105. (206) Liu, Y. X.; Cao, B.; An, P.; J ia, J. H.; Luo, C.; Ma, L. J.; Chang, J.; Pan, K. Layer-by-Layer Surface Molecular Imprinting on Polyacrylonitrile Nanofiber Mats. J. Phys. Chem. A 2015, 119, 6661− 6667. (207) Che, A. F.; Yang, Y. F.; Wan, L. S.; Wu, J.; Xu, Z. K. Molecular Imprinting Fibrous Membranes of Poly(acrylonitrile-co-acrylic acid) Prepared by Electrospinning. Chem. Res. Chin. Univ. 2006, 22, 390− 393. (208) Monier, M.; Kenawy, I. M.; Hashem, M. A. Synthesis and Characterization of Selective Thiourea Modified Hg(II) Ion-Imprinted Cellulosic Cotton Fibers. Carbohydr. Polym. 2014, 106, 49−59. (209) Alizadeh, T.; Memarbashi, N. Evaluation of the Facilitated Transport Capabilities of Nano- and Micro-Sized Molecularly Imprinted Polymers (MIPs) in a Bulk Liquid Membrane System. Sep. Purif. Technol. 2012, 90, 83−91. (210) 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. (211) Yoshimi, Y.; Nakayama, T.; Sekine, S. I.; Piletsky, S. A. Star Shaped Molecularly Imprinted Polymer. Polym. Prepr. Jpn. 2006, 55, 5404−5405. (212) Kriz, D.; Ramström, O.; Mosbach, K. Molecular Imprinting New Possibilities for Sensor Technology. Anal. Chem. 1997, 69, 345A−349A. AA

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(213) Hillberg, A. L.; Brain, K. R.; Allender, C. J. Molecular Imprinted Polymer Sensors: Implications for Therapeutics. Adv. Drug Delivery Rev. 2005, 57, 1875−1889. (214) Ansell, R. J.; Kriz, D.; Mosbach, K. Molecularly Imprinted Polymers for Bioanalysis: Chromatography, Binding Assays and Biomimetic Sensors. Curr. Opin. Biotechnol. 1996, 7, 89−94. (215) Trotta, F.; Biasizzo, M.; Caldera, F. Molecularly Imprinted Membranes. Membranes 2012, 2, 440−477. (216) Malitesta, C.; Mazzotta, E.; Picca, R. A.; Poma, A.; Chianella, I.; Piletsky, S. A. MIP Sensors − the Electrochemical Approach. Anal. Bioanal. Chem. 2012, 402, 1827−1846. (217) Algieri, C.; Drioli, E.; Guzzo, L.; Donato, L. Biomimetic Sensors Based on Molecularly Imprinted Membranes. Sensors 2014, 14, 13863−13912. (218) Haupt, K.; Mosbach, K. Plastic Antibodies: Developments and Applications. Trends Biotechnol. 1998, 16, 468−475. (219) Hedborg, E.; Winquist, F.; Lundstrom, I.; Andersson, L. I.; Mosbach, K. Some Studies of Molecularly-Imprinted Polymer Membranes in Combination with Field-Effect Devices. Sens. Actuators, A 1993, 37−38, 796−799. (220) Kriz, D.; Mosbach, K. Competitive Amperometric Morphine Sensor Based on an Agarose Immobilized Molecularly Imprinted Polymer. Anal. Chim. Acta 1995, 300, 71−75. (221) Prasad, K.; Prathish, K. P.; Gladis, J. M.; Naidu, G. R.; Rao, T. P. Design and Development of Imprinted Polymer Inclusion Membrane-Based Field Monitoring Device for Trace Determination of Phorate (O,O-Diethyl S-ethyl thiomethyl phophorodithioate) in Natural Waters. Electroanalysis 2007, 19, 1195−1200. (222) Piletsky, S. A.; Piletskaya, E. V.; Elgersma, A. V.; Yano, K.; Karube, I.; Parhometz, Y. P.; El’skaya, A. V. Atrazine Sensing by Molecularly Imprinted Membranes. Biosens. Bioelectron. 1995, 10, 959−964. (223) Krő ger, S.; Turner, A. P. F.; Mosbach, K.; Haupt, K. Imprinted Polymer-Based Sensor System for Hetbicieds Using Differential-Pulse Voltammetry on Screen-Printed Electrodes. Anal. Chem. 1999, 71, 3698−3702. (224) Panasyuk, T. L.; Mirsky, V. M.; Piletsky, S. A.; Wolfbeis, O. S. Electropolymerized Molecularly Imprinted Polymers as Receptor Layers in Capacitive Chemical Sensors. Anal. Chem. 1999, 71, 4609−4613. (225) Piletsky, S. A.; Piletskaya, E. V.; Sergeyeva, T. A.; Panasyuk, T. L.; El’skaya, A. V. Molecularly Imprinted Self-Assembled Films with Specificity to Cholesterol. Sens. Actuators, B 1999, 60, 216−220. (226) Chen, Y. P.; Liu, B.; Lian, H. T.; Sun, X. Y. Preparation and Application of Urea Electrochemical Sensor Based on Chitosan Molecularly Imprinted Films. Electroanalysis 2011, 23, 1454−1461. (227) Marx, S.; Zaltsman, A.; Turyan, I.; Mandler, D. Parathion Sensor Based on Molecularly Imprinted Sol-Gel Films. Anal. Chem. 2004, 76, 120−126. (228) Shoji, R.; Takeuchi, T.; Kubo, I. Atrazine Sensor Based on Molecularly Imprinted Polymer-Modified Gold Electrode. Anal. Chem. 2003, 75, 4882−4886. (229) Bai, H.; Wang, C.; Chen, J.; Peng, J.; Cao, Q. A Novel Sensitive Electrochemical Sensor Based on in-situ Polymerized Molecularly Imprinted Membranes at Grapheme Modified Electrode for Artemisinin Determination. Biosens. Bioelectron. 2015, 64, 352−358. (230) Li, S.; Li, J.; Lin, Q.; Wei, X. A Molecularly Imprinted Sensor Based on an Electrochemiluminescent Membrane for Ultratrace Doxycycline Determination. Analyst 2015, 140, 4702−4707. (231) D’Agostino, G.; Alberti, G.; Biesuz, R.; Pesavento, M. Potentiometric Sensor for Atrazine Based on a Molecular Imprinted Membrane. Biosens. Bioelectron. 2006, 22, 145−152. (232) Panasyuk-Delaney, T.; Mirsky, V. M.; Ulbricht, M.; Wolfbeis, O. S. Impedometric Herbicide Chemosensors Based on Molecularly Imprinted Polymers. Anal. Chim. Acta 2001, 435, 157−162. (233) Delaney, T. L.; Zimin, D.; Rahm, M.; Weiss, D.; Wolfbeis, O. S.; Mirsky, V. M. Capacitive Detection in Ultrathin Chemosensors Prepared by Molecularly Imprinted Grafting Photopolymerization. Anal. Chem. 2007, 79, 3220−3225.

(234) Zayats, M.; Lahav, M.; Kharitonov, A. B.; Willner, I. Imprinting of Specific Molecular Recognition Sites in Inorganic and Organic Thin Layer Membranes Associated with Ion-Selective Field-Effect Transistor. Tetrahedron 2002, 58, 815−824. (235) Gutierrez-Fernandez, S.; Lobo-Castañoń , M. J.; MirandaOrdieres, A. J.; Tuñoń -Blanco, P.; Carriedo, G. A.; Garcia-Alonso, F. J.; Fidalgo, J. I. Molecularly Imprinted Polyphosphazene Films as Recognition Element in a Voltammetric Rifamycin SV Sensor. Electroanalysis 2001, 13, 1399−1404. (236) Pogorelova, S. P.; Kharitonov, A. B.; Willner, I.; Sukenik, C. N.; Pizem, H.; Bayer, T. Development of Ion-Sensitive Field-Effect Transistor-Based Sensors for Benzylphosphonic Acids and Thiophenols Using Molecularly Imprinted TiO2 Films. Anal. Chim. Acta 2004, 504, 113−122. (237) Li, Z.-M.; Liu, J.-M.; Liu, Z.-B.; Liu, Q.-Y.; Lin, X.; Li, F.-M.; Yang, M.-L.; Zhu, G.-H.; Huang, X.-M. Preparation for Nitrocellulose Membrane-Poly(vinyl alcohol)-Ionic Imprinting and Its Application to Determine Trace Copper by Room Temperature Phosphorimetry. Anal. Chim. Acta 2007, 589, 44−50. (238) Yuan, L.; Zhang, J.; Zhou, P.; Chen, J.; Wang, R.; Wen, T.; Li, Y.; Zhou, X.; Jiang, H. Electrochemical Sensor Based on Molecularly Imprinted Membranes at Platinum Nanoparticles-Modified Electrode for Determination of 17β-Estradiol. Biosens. Bioelectron. 2011, 29, 29− 33. (239) Blanco-López, M. C.; Gutiérrez-Fernández, S.; Lobo-Castañoń , M. J.; Miranda-Ordieres, A. J.; Tuñoń -Blanco, P. Electrochemical Sensing with Electrodes Modified with Molecularly Imprinted Polymer Films. Anal. Bioanal. Chem. 2004, 378, 1922−1928. (240) Yang, Z. G.; Xu, Z. L. Pervaporation of Molecular Imprinted Composite Membranes to Isopropanol/Water System. J. East China Univ. Sci. Technol. 2006, 32, 1373−1377 (abstract only).. (241) Zheng, H. Y.; Yoshikawa, M. Molecularly Imprinted Cellulose Membranes for Pervaporation Separation of Xylene Isomers. J. Membr. Sci. 2015, 478, 148−154. (242) Ulbricht, M. Advanced Functional Polymer Membranes. Polymer 2006, 47, 2217−2262. (243) Jenkins, A. L.; Bae, S. Y. Molecularly Imprinted Polymers for Chemical Agent Detection in Multiple Water Matrices. Anal. Chim. Acta 2005, 542, 32−37. (244) Székely, G.; Bandarra, J.; Heggie, W.; Sellergren, B.; Ferreira, F. C. A Hybrid Approach to Reach Stringent Low Genotoxic Impurity Contents in Active Pharmaceutical Ingredients: Combining Molecularly Imprinted Polymers and Organic Solvent Nanofiltration for Removal of 1,3-Diisopropylurea. Sep. Purif. Technol. 2012, 86, 79−87. (245) Imahashi, M.; Hayashi, K. Concentrating Materials Covered by Molecular Imprinted Nanofiltration Layer with Reconfigurability Prepared by a Surface Sol-Gel Process for Gas-Selective Detection. J. Colloid Interface Sci. 2013, 406, 186−195. (246) De Luca, G.; Donato, L.; Tasselli, F.; Del Blanco, S. G.; Bisignano, F.; Drioli, E. Nanofiltration and Molecularly Imprinted Membranes: A Theoretical Study based on Quantum Mechanics Approach. Procedia Eng. 2012, 44, 1761−1762. (247) Shah, N.; Ha, J. H.; Ul-Islam, M.; Park, J. K. Chiral Separation of Phenylalanine Using D-Phenylalanine Imprinted Membrane Containing L-Phenylalanine Imprinted Submicron/Nanoscale Beads. J. Pharm. Res. (Mohali, India) 2012, 5, 4425−4433. (248) Park, J. K.; Kim, S. J. Separation of Phenylalanine by Ultrafiltration Using D-Phe Imprinted Polyacrylonitrile-Poly(acrylic acid)-Poly(acryl amide) Terpolymer Membrane. Korean J. Chem. Eng. 2004, 21, 994−998. (249) Asliyuce, S.; Uzun, L.; Yousefi Rad, A. Y.; Unal, S.; Say, R.; Denizli, A. Molecular Imprinting Based Composite Cryogel Membranes for Purification of Anti-Hepatitis B Surface Antibody by Fast Protein Liquid Chromatography. J. Chromatogr. B: Anal. Technol. Biomed. Life Sci. 2012, 889−890, 95−102. (250) Maciuk, A.; Moaddel, R.; Haginaka, J.; Wainer, I. W. Screening of Tobacco Smoke Condensate for Nicotinic Acetylcholine Receptor Ligands Using Cellular Membrane Affinity Chromatography Columns AB

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

and Missing Peak Chromatography. J. Pharm. Biomed. Anal. 2008, 48, 238−246. (251) Fujimoto, C. Packing Materials and Separation Efficiencies in Capillary Electrochromatography. TrAC, Trends Anal. Chem. 1999, 18, 291−301. (252) Xi, S.; Hu, X. L.; Liang, L.; Gao, J. Progress in Preparation of New Molecularly Imprinted Polymer Hybrid Membranes. Dev. Appl. Mater. 2010, 25, 67−72. (253) Kubo, I.; Shoji, R.; Fuchiwaki, Y.; Suzuki, H. Atrazine Sensing Chip Based on Molecularly Imprinted Polymer Layer. Electrochemistry 2008, 76, 541−544. (254) Guillon, S.; Lemaire, R.; Linares, A. V.; Haupt, K.; Ayela, C. Single Step Patterning of Molecularly Imprinted Polymers for Large Scale Fabrication of Microbiochips. Lab Chip 2009, 9, 2987−2991. (255) Wu, Y. L.; Liu, X. L.; Meng, M. J.; Lv, P.; Yan, M.; Wei, X.; Li, H. J.; Yan, Y. S.; Li, C. X. Bio-Inspired Adhesion: Fabrication of Molecularly Imprinted Nanocomposite Membranes by Developing a Hybrid Organic-Inorganic Nanoparticles Composite Structure. J. Membr. Sci. 2015, 490, 169−178. (256) Li, S. J.; Zhu, M. Y.; Whitcombe, M. J.; Piletsky, S. A.; Turner, A. P. F. Molecularly Imprinted Polymers for Enzyme-like Catalysis: Principle, Design, and Applications. In Molecularly Imprinted Catalysts: Principles, Syntheses and Applications; Li, S. J., Cao, S. S., Piletsky, S. A., Turner, A. P. F., Eds.; Elsevier: Amsterdam, 2016; pp 1−17. (257) Fang, G. Z.; Lu, J. P.; Pan, M. F.; Li, W.; Ren, L.; Wang, S. Substitution of Antibody with Molecularly Imprinted Film in EnzymeLinked Immunosorbent Assay for Determination of Trace Ractopamine in Urine and Pork Samples. Food Anal. Methods 2011, 4, 590− 597. (258) Minko, S.; Orlov, M.; Tokarev, I.; Katz, E. Stimuli-Responsive Polymer Membrane and Method of Making the Same. U.S. Patent 8,829,062, 2014. (259) Huda, N.; Rashid, M.; Rashid, H. A Review on StimuliResponsive Grafted Membranes Based on Facile Synthesis Process and Extensive Applications. Int. J. Innovation Appl. Stud. 2014, 8, 1296− 1312. (260) Lee, M. H.; Thomas, J. L.; Chen, W. J.; Li, M. H.; Shih, C. P.; Lin, H. Y. Fabrication of Bacteria-Imprinted Polymer Coated Electrodes for Microbial Fuel Cells. ACS Sustainable Chem. Eng. 2015, 3, 1190−1196. (261) Yoshimi, Y.; Sakai, K. Template Recognition and Changes in the Diffusive Permeability of a Thin Layer of Molecularly Imprinted Polymer. Jpn. J. Artif. Organs 2000, 29, 441−445. (262) Yavuz, H.; Andaç, M.; Uzun, L.; Say, R.; Denizli, A. Molecular Recognition Based Iron Removal from Human Plasma with Imprinted Membranes. Int. J. Artif. Organs 2006, 29, 900−911. (263) Sun, Z.; Zussman, E.; Yarin, A. L.; Wendorff, J. H.; Greiner, A. Compound Core-Shell Polymer Nanofibers by Co-Electrospinning. Adv. Mater. 2003, 15, 1929−1932. (264) Li, D.; Xia, Y. Direct Fabrication of Composite and Ceramic Hollow Nanofibers by Electrospinning. Nano Lett. 2004, 4, 933−938. (265) Li, D.; McCann, J. T.; Xia, Y. Use of Electrospinning to Directly Fabricate Hollow Nanofibers with Functionalized Inner and Outer Surfaces. Small 2005, 1, 83−86. (266) Isezaki, J.; Yoshikawa, M. Molecularly Imprinted Nanofiber Membranes: Localization of Molecular Recognition Sites on the Surface of Nanofiber. J. Membr. Sep. Technol. 2014, 3, 119−126. (267) Sun, Y.; Li, L. Y.; Dong, X. Y. Multistage Affinity Cross-Flow Filtration: Process Optimization. Bioprocess Eng. 1997, 16, 229−235. (268) Romero, J.; Zydney, A. L. Staging of Affinity Ultrafiltration Processes for Chiral Separations. J. Membr. Sci. 2002, 209, 107−119. (269) Higuchi, A.; Higuchi, Y.; Furuta, K.; Yoon, B. O.; Hara, M.; Maniwa, S.; Saitoh, M.; Sanui, K. Chiral Separation of Phenylalanine by Ultrafiltration through Immobilized DNA Membranes. J. Membr. Sci. 2003, 221, 207−218.

AC

DOI: 10.1021/acs.chemrev.6b00098 Chem. Rev. XXXX, XXX, XXX−XXX