Next Article in Journal
In Vitro Mechanical Properties of a Novel Graphene-Reinforced PMMA-Based Dental Restorative Material
Next Article in Special Issue
Click Synthesis of Triazole Polymers Based on Lignin-Derived Metabolic Intermediate and Their Strong Adhesive Properties to Cu Plate
Previous Article in Journal
Biopolymer Non-Parametric Analysis: A Degradation Study under Accelerated Destructive Tests
Previous Article in Special Issue
Facile Synthesis of Functionalised Hyperbranched Polymers for Application as Novel, Low Viscosity Lubricant Formulation Components
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Polymeric Membranes for Biomedical Applications

by
Elena Ruxandra Radu
1,2,
Stefan Ioan Voicu
1,2,* and
Vijay Kumar Thakur
3,4,5,*
1
Department of Analytical Chemistry and Environmental Engineering, University Politehnica of Bucharest, 011061 Bucharest, Romania
2
Advanced Polymers Materials Group, University Politehnica of Bucharest, 011061 Bucharest, Romania
3
Biorefining and Advanced Materials Research Center, Scotland’s Rural College (SRUC), Kings Buildings, Edinburgh EH9 3JG, UK
4
School of Engineering, University of Petroleum & Energy Studies (UPES), Dehradun 248007, Uttarakhand, India
5
Centre for Research & Development, Chandigarh University, Mohali 140413, Punjab, India
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(3), 619; https://doi.org/10.3390/polym15030619
Submission received: 17 December 2022 / Revised: 16 January 2023 / Accepted: 21 January 2023 / Published: 25 January 2023
(This article belongs to the Special Issue Feature Papers in Polymer Applications)

Abstract

:
Polymeric membranes are selective materials used in a wide range of applications that require separation processes, from water filtration and purification to industrial separations. Because of these materials’ remarkable properties, namely, selectivity, membranes are also used in a wide range of biomedical applications that require separations. Considering the fact that most organs (apart from the heart and brain) have separation processes associated with the physiological function (kidneys, lungs, intestines, stomach, etc.), technological solutions have been developed to replace the function of these organs with the help of polymer membranes. This review presents the main biomedical applications of polymer membranes, such as hemodialysis (for chronic kidney disease), membrane-based artificial oxygenators (for artificial lung), artificial liver, artificial pancreas, and membranes for osseointegration and drug delivery systems based on membranes.

1. Introduction

The biomedical field is a transdisciplinary field that combines knowledge from medicine, biology, biochemistry, materials science, and biomechanics. A considerable part of biomedical fields is represented by the development of biomedical applications, such as the evolution of biocompatible implants and other medical devices, imaging equipment, regenerative tissue engineering, and drug delivery systems [1]. Biomedical devices, such as bone implants, contact lenses, stents, artificial hearts, tissue adhesives, surgical sutures and dialysis membranes, are obtained by using biomaterials [2,3,4,5]. Biocompatibility is an important property when it comes to the materials used in biomedical applications because it is necessary to use some friendly materials, which are able to not harm the living tissue [6]. Further, the testing of biocompatibility through in vitro and in vivo specific tests is an essential step in developing biomedical devices [7].
In recent years, the use of membranes in various medical applications has been constantly evolving [8,9,10,11,12,13]. Membranes are defined as porous films that act as a separating barrier between two adjacent phases, capable of allowing the transport of substances from one phase to another in a selective manner [14,15]. In general, membranes utilized in the biomedical field can be used in tissue engineering, with a role in purification and in obtaining implants and scaffolds, in obtaining controlled release systems of active substances and in diagnosis, in the form of sensors and various diagnostic tests [16]. Membranes are classified according to their nature into natural membranes and artificial membranes. Synthetic membranes can be further classified according to the material into organic, inorganic or composite membranes [17]. Organic membranes are generally obtained from natural polymers such as cellulose, chitosan, collagen and alginates or from synthetic polymers such as poly (ethylene glycol) (PEG), polyurethane, polysulfone, polylactic acid, poly (acrylic amide), poly (N-vinyl-2-pyrrolidone) [18,19,20,21,22,23,24]. In the biomedical field, membranes are used to detoxify the blood. A well-known example is represented by the membranes used in hemodialysis, in cases of end-stage renal disease, where part of the kidney function is replaced by using a flat membrane or empty fiber dialyzer to remove excess water, salts and excreted metabolic products [24]. However, membranes are often used in other biomedical applications, such as in the developing of artificial organs (artificial liver, oxygenator and pancreas) for increasing the optimal functionalization of physiological functions of the organs [25,26,27]. Further, the application of polymeric membranes in tissue engineering has been highly studied due to their biofunctionality, good mechanical properties and the ability of a possible reparation and regeneration of injured tissues/organs [28,29,30,31,32,33]. Another application of polymeric membranes in the biomedical field is represented by the developing drug delivery system based on membranes [34] or different separation interest molecules such as antibiotics [35,36,37] or proteins [38]. The release of the drug is achieved by the diffusion of the active substance through the polymeric membrane so that the drug release could be controlled and targeted [39,40,41]. This article presents the recent developments of polymeric membranes in biomedical applications (Figure 1).

2. Biomedical Applications of Membranes

2.1. Hemodialysis

Renal failure is one of the major problems suffered by >850 million people worldwide [42]. In chronic kidney failure disease, the kidney functions of the patients are deteriorated, leading to the incapacity of filtering the blood by removing waste products from the body, such as metabolic-resulted toxins with molecular weights more equal to less than 58 kDa [43,44]. Generally, hemodialysis is used in order to ensure a better quality of life for patients suffering from chronic kidney failure [45]. The US Data Renal System (USRDS) reports that almost 80% of hemodialysis patients started dialysis using an indwelling catheter [46]. Further, hemodialysis is defined as an extracorporeal blood purifying method using a semipermeable membrane to conduct blood purification and remove uremic toxins [44,45]. The principle of the hemodialysis membrane is the diffusion of the solvent through a semipermeable membrane [47]. The hemodialysis treatment is used to remove small molecules such as small water-soluble toxins (molecular weight, MW < 500 D), but also a small amount of the middle molecules (MW 500–32,000 Da) from the blood [48]. In hemodialysis treatment, the patient is connected to a dialysis machine and their blood is pumped out via vascular access and filtered using a dialyzer, which is called an artificial kidney. The filtered blood is then pumped back into the patient’s body [49].
The major drawbacks of using semipermeable membranes in hemodialysis are the hemocompatibility through blood exposure to the membrane’s material, which could lead to activation of proinflammatory molecules, and the incapacity of successfully removing some larger toxins molecules [44]. The development of upgraded hemodialysis membranes for increased hemocompatibility and anticoagulant properties was reported [44,50,51,52,53,54]. These membranes were obtained from natural or synthetic polymers, such as polysulfone (PSF) [55,56,57], polyethersulfone (PES) [58,59], polyvinyl alcohol (PVA) [60,61,62], cellulose triacetate (CTA) [63,64,65], polymethylmethacrylate (PMMA) [66,67], polyacrylonitrile (PAN) [53,68], and polyamide (PA) [67].
The first used hemodialysis membranes were made from cellulose-based membranes. Cuprophan® (Wuppertal, Germany) was the first used cellulose-based hemodialysis membrane that was obtained from cotton [48]. Nowadays, cellulose-based hemodialysis membranes are obtained via acrylation with acetate groups resulting in cellulose acetate, cellulose diacetate (CDA), and cellulose triacetate (CTA) with free hydroxyl groups on the surfaces [48,69]. Figure 2 presents the main cellulose-based derivates [65]. The cellulose-based hemodialysis membranes have good toughness, biodegradation, sustainability, and biocompatibility and are cheap in comparison to other polymers [70]. Cellulose acetate (CA) is the most utilized cellulose derivate due to its great solubility in diverse organic solvents and insoluble in water. Dumitriu et al. [63] showed CA functionalization with TiO2 nanoparticles, followed by heparin incorporated via dopamine polymerization. The contact angle results showed that both the neat CA membrane and composite membrane present a hydrophilic nature, but a slight decrease in the contact angle for the composite membrane was observed. Faria et al. [71] studied the potential uremic blood purification with cellulose acetate (CA) functionalized with SiO2 nanoparticles. It was observed that at higher SiO2 content (18%) the hydrophilicity of the membrane increases. The hydraulic permeability increases after the addition of SiO2 nanoparticles, and the rejection of BSA was 99%. Azhar et al. [60] presented the CA-based membrane modified with PVP for better hemocompatibility. The results showed increased hydrophilicity, the pure water flux, BSA rejection, urea, and creatine clearance obtained is 42.4 ± 2 L/m2h, 95 ± 1.023%, 93 ± 1.023% and 89 ± 1.023%.
Polysulfone (PSF) is the most used material in the fabrication of the hemodialysis membrane, with almost 90% of hemodialysis membranes being made of PSF and PSF derivates [72]. PSF is a thermoplastic polymer with great mechanical properties, thermal stability, biocompatibility, and water permeability [47,73]. The PSF membrane showed great stability in extreme basic or acidic environments [74]. Additionally, PSF membranes are resistant to sterilization with steam, ethylene oxide, and gramma radiation [75]. The main disadvantage of PSF is its hydrophobic nature, which favors the molecule adhesion on the membrane surface, inducing blood clot formation [76]. In the past years, researchers investigated the functionalization of PSF in order to increase surface hydrophilicity, resulting in improved hemocompatibility and antifouling properties [54,77]. Polyether sulfone (PES) has similar properties as PSF, with great oxidative, thermal and mechanical properties [78]. The same as PSF, the limitation of PES is represented by the hydrophobic nature, which affects the hemocompatibility of the membrane, even if it is more hydrophilic than PSF due to the presence of the higher atomic weight ratio of sulfone groups [79]. The biocompatibility of neat PES isn’t satisfactory enough and in the past years, researchers obtained different PES-based membranes by adding different additives or by functionalizing the pristine PES. Irfan et al. [80] proposed membrane composites based on PES, multi-wall carbon nanotubes (f-MWCNT) and polyvinylpyrrolidone (PVP) via the phase inversion process for hemodialysis application. The contact angle of the nanocomposite membrane was significantly lower than the PES contact angle (from 88° for PES to 51° for nanocomposite). Further, pure water permeation flux (PWP) rate up to 72.20 L m-2 h-1 exhibited 58.82% reduced protein absorption and better uremic waste clearance of 56.30%, 55.08% and 27.90% of urea, creatinine and lysozyme. Abdelrasoul et al. [52] proposed PES-PVP composite membrane via UV-assisted photochemical synthesis. The addition of PVP increased the hydrophilicity of the composite membrane and showed better resistance to fibrinogen adsorption compared to the pristine PES membrane. Hoseinpour et al. [81] presented carboxymethylcellulose (CMC), and SCMCs were immobilized on the surface of the aminated PES membranes (PES-NH2) via amide bonds. The hydrophilicity of composites was increased in comparison with the hydrophilicity of neat PES due to the surface functional groups of NH2 in PES-NH2, COO− in PES-CMC and COO− and SO3− in PES-SCMC membranes engaging in hydrogen bonding interactions with H2O. The bovine serum albumin (BSA) rejection increases in the case of the composite’s membrane up to 99% in comparison with pristine PES (95%). Xing et al. [82] described the antimicrobial properties of polydopamine (PDA) and PES composites to immobilize antimicrobial metal ions with their functional groups. The hydrophilicity increases after the addition of PDA such that the contact angle decreases from 102° for PES to 40° for the composites.
Generally, the addition of polyvinylpyrrolidone (PVP) is used as an additive in order to increase the hydrophilicity of PES and PSF membranes [83]. In membrane synthesis via phase inversion, PVP plays a major role in the pore formation of the membrane [84]. Mansur et al. [85] presented the effect of the addition of different PVP concentrations (3 wt% and 18 wt%) on the PSF membrane. It was observed that at a higher concentration of PVP, the membrane pores were larger than after the addition of a lower PVP content, leading to increased hydrophilicity of the membrane and also leading to an increased flux and permeability of the membrane, but the BSA rejection was significantly lower at higher concentrations of PVP.
Researchers had proposed surface functionalization with different anticoagulant agents, such as heparin or heparin-like molecules, for better hemocompatibility. Heparin is an anticoagulant agent used in medicine for the prevention of thrombus formation and limits thrombus extension [86]. Heparin has been immobilized onto membranes to increase hemocompatibility through physical absorption or ionic bonding [87,88,89]. Huang et al. [90] presented the chemical binding of heparin onto PSF sheets via a three-step synthesis method (Figure 1). A decrease in the contact angle after functionalization was observed, leading to increased hydrophilicity and hemocompatibility. Ren et al. [91] described the PSF functionalization of heparin via covalent immobilization in order to improve anticoagulant properties. After functionalization, the contact angle decreased from 87° to about 30o, and the coagulation time was prolonged with less fibrin generated in the process of hemodialysis. The limitations of heparin treatment are the expensive cost and potential side effects [92,93].
Different molecules were studied in order to replace the addition of heparin. The heparin-like molecules present an alkyl backbone for synthesized polymers or a uronic backbone for modified biomacromolecules [94]. Zhang et al. [95] used tannic acid as a heparin-like substitute to increase the anticoagulant properties of the PES membrane. The BSA rejection rate and urea clearance rate were 97.1% and 92.0% after the addition of tannic acid. Additionally, the hydrophilicity of the composites reduced platelet adhesion and activation, improving the PES membranes’ hemocompatibility. Tannic acid is a bio-based polyphenol that can be found in all aerial plant tissues [96]. Ma et al. [94] proposed sodium alginate as a heparin-mimicking molecule. The results showed that the presence of sulfonic groups and uronic main chain structure gives excellent anticoagulant activity. Lu et al. [97] proposed heparin-like anticoagulant polypeptides due to the carboxyl and hydroxyl side groups for anticoagulant and thrombolytic therapy. In addition, by simply adjusting the feeding ratio of monomers, anticoagulant activity can be regulated. Song et al. [87] proposed carboxymethyl chitosan nanoparticle (CMCN) and poly (vinyl alcohol) (CMCN/PVA) onto the surface of modified bacterial cellulose sulfate (BCS) membranes via electrospinning as heparin-like substituent membrane. The similarity between heparin and heparin-like membranes (CPBS) is the presence of -SO3, COO−, and -OH groups on the surface of the CPBS membrane. The CPBS membrane hydrophilicity was higher than the BC sulfate membrane, which could provide anticoagulant properties of the heparin-like membrane.

2.2. Drug Delivery Systems

Drug delivery systems are defined as a device or a formulation that is able to deliver an active substance to a target tissue to increase the efficiency of the active substance [98]. These DDS-based membranes have the ability to increase pharmacological activity, thereby reducing the side effects, increasing the solubility of the active substance, protecting the active substance from biodegradation and gradually releasing the active substance [99]. Additionally, to improve the drug release efficiency, different DDS sensitive to both external (magnetic, photothermal and electrical responsive) and internal stimuli (temperature, pH and redox responsive) have been developed. The first attempt to develop DDs was reported in 1950 in agriculture when they tried to develop systems for the controlled release of pesticides from a polymer matrix [100]. After that, this concept of the controlled release of the active substance was borrowed in the biomedical field [101]. In the past years, many biodegradable and bioabsorbable polymers were studied for the production of DDS. The release mechanism from a polymeric matrix is represented through drug diffusion, dissolution, and degradation of the carrier matrix [102]. Moreover, the usage of DDS leads to a decrease in the need for frequent administration of the active substance, leading to improved quality of life. The polymeric membranes used in DDS application require good stability during administration, biocompatibility, biodegradability and absorbability [103,104,105]. Depending on the size of the pores, the membranes can be classified into membranes with micropores (less than 2 nm), membranes with mesopores (between 2 and 50 nm) and membranes with macropores (more than 50 nm) [106]. Apart from the pores’ size, the pores’ shape is very important, so the pores of the membranes can have different shapes, such as cylindrical, conical or irregular. These morphological characteristics are very important when it comes to the development of DDS.
To obtain DDS-based membranes, several techniques are used such as phase separation technique, interfacial polymerization, stretching, ion-track etching, lithography and electrospinning [107,108]. In the ion-track etching method, a membrane can be obtained by irradiating a film with heavy ions forming ion tracks and leading to pore formation [108,109]. The films used for irradiation have been made of polymers, such as polycarbonate (PC), polyethylene terephthalate (PET), polyimide (PI), polypropylene (PP), polyethylene naphthalate (PEN) and also biodegradable polymers, such as polylactic acid (PLA), which are used as a polymer matrix to create track-etched membranes [110,111,112,113,114,115]. Ivanova et al. [115] reported the preparation of a biodegradable membrane based on PLA via the ion-track etching method. The pore size of the obtained biodegradable membrane was reported between 0.6 to 1.5 μm with a shape close to cylindrical. Further, it was observed that an increase in the etching time over 20 min shows a decrease in the roughness of both membrane sides. Lithography is another method for obtaining membranes. Through lithography, an ordered array of nano/micro pores on the surface can be obtained [108]. An example of the use of lithography in the developing of DDS is related by Patil et al. [116], who reported the well-defined pore formation on polyimide (PI), to be the support membrane for grafting poly(acrylic acid) (PAAc) hydrogel via free radical polymerization applying the same excimer laser. This type of DDS is pH-sensitive so at different pH levels (pH 7 and pH 3), it changes its permeability.
The electrospinning method is widely used in obtaining membranes with various applications in the medical field, such as in delivery systems and in tissue engineering, using the application of a high electric field to generate nanofibers from a charged polymer solution or melt [117,118]. It has been observed that if the electrospinning parameters are varied, such as polymer, solvent, polymer solution, processing parameters, and ambient conditions, different morphologies can be obtained in order to improve the mechanical strength and drug loading/releasing performance of membranes [119,120]. To obtain fibers and later membranes by electrospinning, both synthetic polymers can be used, such as polystyrene (PS), poly (vinylchloride) (PVC), and PLA, as well as natural polymers, such as silk fibroin, fibrinogens, dextran, chitin, chitosan, alginate, collagen, and gelatin [120]. Ren et al. [121] reported an electrospun membrane for controlled drug delivery for the acceleration of diabetic wound healing based on PLA and dimethyloxalylglycine (DMOG)-loaded mesoporous silica nanoparticles (DS). All the membranes showed well-organized topological structures, with ellipsoidal-shaped nanopores, which were arranged with the major axis along the fiber direction. Co-electrospinning method, also known as coaxial electrospinning, is a method derived from electrospinning, the difference being that in the case of co-electrospinning, two needles are used instead of one needle, which makes it possible to obtain a core-shell nanofiber [122]. Al-Badani et al. [123] reported a possible polycaprolactone/gelatin membrane as a tunable drug delivery system for bone tissue regeneration via co-electrospinning (Figure 3). It was observed that by adjusting the poly(ε-caprolactone) (PCL) fibers, the degradation of gelatin and the release profile of hydrophilic drugs/proteins could be effectively controlled.
The phase separation technique is one of the simplest methods for obtaining membranes. Phase separation can occur when the thermodynamic quality of a polymer solution usually decreases when the temperature is decreased, and by immersion precipitation [124]. Zeinali et al. [125] reported Poly(butylene succinate) (PBS) membrane via thermally-induced phase separation (TIPS) for Curcumin (CUR) and piperine (PIP) release. Additionally, it was reported that the variation of solvent system, thermal gradient and cooling direction changed the pores’ architecture, which could affect the release profile. The interfacial polymerization method is used for membrane formation, which is one of the most important processes in commercial fabrication [126]. In this method, the polymer solution and the solvent are poured onto a flat surface to form a film. After that, the controlled process of exchanging solvent for a nonsolvent takes place, leading to a membrane formation [127]. Ding et al. [128] described an antibacterial and anti-inflammatory membrane based on porphyrin-covalent organic frameworks (COF) with encapsulated ibuprofen (IBU) via an in situ interfacial polymerization and impregnation approach. The reported membrane displayed high O2 generation and controllable ibuprofen release at body temperature for wound healing.
Generally, the surface modification of membranes is carried out for better biocompatibility and biodegradability [108]. The biocompatibility of the membrane is represented by it is represented by the response to the interaction with biological fluids, without triggering any kind of immune response or infection. The most used techniques for membrane surface modification are blending [129], grafting [130] and plasma initiation [131]. The blending surface modification method is one of the most convenient methods due to reduced cost. Liang et al. [132] proposed the improvement of the blending strategy for membrane modification via surface segregation using surface-tailored amphiphilic nanoparticles. After the functionalization, the membranes achieved a notably increased blending efficiency, resulting in a dramatically enhanced surface hydrophilicity. Parveen et al. [133] reported chitosan/PEG blended PLGA nanoparticles for cancer drug delivery. The functionalization with PLGA decreases the aggregation due to the repulsion of serum proteins, leading to very good stability for days and the lowest percentage of uptake by macrophages. Surface modification of polymeric membranes via grafting has become a very popular method in recent years for preparing a “tailored” membrane surface with desired functions [134]. For example, through this method of functionalization, membranes sensitive to pH variation can be obtained only by grafting different functional groups or such polymers onto the surface of the membrane [135]. Surface modification by grafting can be achieved through several techniques, such as grafting through light [136,137], grating via thermal treatment [138,139], grafting polymerization through plasma irradiation [140,141], atom transfer radical polymerization (ATRP) surface initiated method [75,142,143,144], reversible addition fragmentation chain transfer (RAFT) polymerization [145,146,147] and redox reactions grafting [148].
In the modification of the membrane surface with plasma treatment, plasma is a confined ionized gas obtained from the dissociation of gas-forming plasma after applying an electric field in the reactor [149]. The plasma effect on the membrane occurs only on the surface, not affecting the properties of the bulk [150]. The plasma treatment can introduce chemically reactive functional groups onto polymer substrates to increase biodegradability or biocompatibility [151]. For example, an O2 plasma is able to generate ─OH radicals on the membrane surface, leading to an influence on the hydrophilicity of the membrane surface [152].
Further, the hydrophobic or hydrophilic character of the membrane can have a say in increasing biocompatibility, so hydrophilic membranes are preferred for biomedical applications as they improve the flux through the membrane and fouling properties [131]. The degree of hydrophilicity or wettability of the membrane is one of the most important parameters to control cell behavior through protein adsorption [153]. However, the hydrophilic/hydrophobic character of the active substance is a criterion to be taken into account, so that the functionalization of the membrane surface can improve the interaction between the membrane and the active substance. Hardin et al. [154] shared the functionalization of polyacrylic acid membrane for potential drug delivery application via condensation reaction with N-(3-dimethylaminopropyl)-N-ethylcarbodiimide hydrochloride as our coupling reagent. It was shown that the choice of the hydrophobic group has an effect on the polymer solubilization properties. Sagitha et al. [155] reported the functionalization of the polyurethane membrane with β-Cyclodextrin, leading to increased hydrophilicity and thermal stability. The addition of β-Cyclodextrin provides a higher vapor transmission rate, hemocompatibility and cell viability for functionalized membrane. Moreover, the obtained β-Cyclodextrin- polyurethane membrane was functionalized with nanochitosan for better biocompatibility.
Stimuli-responsive membranes have attracted attention in recent years due to their extraordinary properties such as the ability to change their physicochemical properties in response to changes in their environment [156,157,158]. These stimuli-responsive membranes are able to respond to changes in pH, temperature, light and magnetic fields, being able to release various therapeutic agents in a targeted and gradual manner [157]. For example, pH-responsive membranes are obtained from polymers with ionizable acidic/basic residues which can be employed in drug delivery applications, and this is possible due to the variation of a normal pH in the body and also the variation in pH caused by different conditions, such as inflammation, infection and cancer [159,160]. Notably, pH-responsive membranes are able to target a specific area in the body and release their active agent with an increased therapeutic impact and reducing the side-effects [161]. These pH variations can result in the modification of crosslinking processes so that the protonation or deprotonation of acidic/basic groups can generate distinct interactions between a therapeutic agent and a material, resulting in the release of the active substance [159]. Additionally, a very prominent example of pH-responsive DDS membranes is based on the alginate-chitosan complex, due to advantageous properties such as biocompatibility, biodegradability, pH sensitivity, and mucoadhesive property [162]. Additionally, the protonation of the amino groups leads to the polymeric backbone of chitosan becoming charged, resulting in charge repulsion and swelling [163]. Further, due to their opposite charged polymeric backbones of chitosan and alginate, polyelectrolyte complexes are obtained [164]. Schoeller et al. [163] reported the deposition of chitosan and alginate layers forming polyelectrolyte complex through layer-by-layer assembly on Poly(lactic-co-glycolic acid ) (PLGA) nanofibers to introduce pH-sensitivity for the controlled release of ibuprofen. The results show the inhibition of ibuprofen release at an acidic pH due to the pH-sensitive membrane suggesting that the obtained membrane protects the active substance in the acidic environment of the stomach, leading to an improved release at higher pH values without an early release and reduction in drug concentrations.
Temperature-responsive membranes for DDS are obtained from temperature-sensitive polymers with the property of sol-gel transitions above a certain temperature [165]. The temperature-responsive polymeric membranes are able to change their phase below the lower critical solution temperature (LCST), leading to the hydrogen bonds between the polymer and water molecules, allowing the polymer to swell and release the active substance [166]. The magnetic-responsive membranes are usually filled with magnetic-active nanoparticles of metals, metal oxides, or ferromagnetic materials which under the action of electromagnetic radiation can release the active substance in a targeted manner and enhance drug accumulation at the sites [167,168,169]. Photo-responsive membranes are able to release drugs using light sources, such as ultraviolet (UV), visible, and near-infrared (NIR) light [34]. The permeability of the photo-responsive membrane is influenced by switching irradiation between ultraviolet light (UV) and green light (Vis) [170]. The photo-responsiveness property is ensured by the presence of different photo-reactive groups, such as azobenzene, triphenylmethane and spiropyran groups in the polymer matrices by entrapping, cross-linking, and introducing them as a side chain or part of the main chain of the polymeric matrix [171]. Other physical characteristics, besides stimuli-responsiveness, which are studied and discussed in the DDS field are particle size distribution and polydispersity index (PDI) and length of polymers [172,173]. The PDI is an important physical characteristic to be considered in DDS development due to its influence on stability, drug release profile and drug penetration [172,174]. The length of the polymer can influence the release of the active substance from the DDS-based membrane by increasing the release rate as the polymer length decreases [173].
Further, opsonization is an essential characteristic of biological barriers in order to control drug delivery so that the macrophage cells identify the presence of the membrane-based release system as a foreign body and remove them to protect the body [175,176]. A method to combat opsonization is represented by PEGylation to prevent phagocytosis [176]. PEGylation is used to increase the efficiency of drug and gene delivery and decrease the immunogenicity of the proteins without significantly affecting the activity [177]. The molecular weight (MW) of the PEG link can influence the internalization so that an increase in MW leads to the increased flexibility of the PEG chain, leading to the entanglement of PEG chains and reducing the binding of the ligand to the receiver [178].

2.3. Membrane Oxygenator

Recently, due to the COVID-19 pandemic, the need to develop artificial oxygenators is growing. As the name implies, oxygenators are defined as medical devices for respiratory support. An example of an oxygenator is represented by extracorporeal membrane oxygenation (ECMO), which is used in the treatment of severely ill patients with cardiopulmonary impairment caused by infectious diseases [179]. The general principle of the operation of oxygenators is that when the blood passes through the oxygenator, the oxygen level increases, and the CO2 level decreases so that the non-oxygenated blood is oxygenated (Figure 4) [26,180,181]. The extracorporeal membrane oxygenators are they are made up of a pump that has the role of pumping the blood and the oxygenating membrane that has the role of oxygenating the blood [179,182,183]. The most common causes of complications in patients using an oxygenator are clotting and bleeding [184]. The material from which the oxygenating membrane is obtained must have high permeability, high mechanical strength, free defects, high biocompatibility and hemocompatibility [179,185]. For example, in the case of hemodialysis, membranes are completely sterile and the whole tubes used for extracorporeal circulation of blood are also sterile in order to prevent sepsis. Normally, in order to avoid the formation of clots in the oxygenation installation, they administer anticoagulant drugs, such as heparin. Unfortunately, this method has the disadvantage of the possibility of severe bleeding [186,187,188,189]. Another limitation of the ECMO system is the adherence of various biomolecules on the surface, which can be overcome by incorporating small molecules, such as heparin, in the coating of the ECMO system surface [190]. The anaphylatoxins are attached to the biomaterial surface and have numerous implications in hemostasis. The polymers under the ECMO system have an effect on the CR3 granulocyte receptor, resulting in the adhesion of principal neutrophils and monocytes [190].
Hollow fiber membranes have been used as an oxygenator and are usually obtained via a phase inversion process [191,192]. The commonly used polymers for hollow fiber membranes are hydrophobic polymers, such as polymethylpentene (PMP), polypropylene (PP), PDMS, polysulfone (PSf), polyethersulfone (PES), polyethylene (PE) and polyvinylidene fluoride (PVDF) [26,192,193,194,195,196,197,198]. Wang et al. [194] reported the production of poly (4-methyl-1-pentene)/polypropylene (PMP/PP) thin film composite (TFC) with a PVA/PSS coating was anchored on the membrane surface via crosslinking and PDA binding for membrane oxygenator application. PMP is used in oxygenator membrane production due to increased gas exchange properties, low diffusion resistance and absent plasma leakages. After surface modification of membranes showed excellent hydrophilicity and coagulant factors, adsorption capability, which significantly inhibits the activation of platelets and the adhesion of proteins, thus blocking the blood coagulation due to increased sulfonate groups. Park et al. [199] proposed the fabrication of functionalized poly(vinylidene-co-hexafluoropropylene) or PVDF-co-HFP coated, using Hyflon AD60X to minimize pore wetting and interactions with blood. These functionalized membranes have very low protein adsorption and a high contact angle for both water and blood, leading to improved hemocompatibility.

2.4. Artificial Liver

Liver disease is a worldwide major public health problem, being one of the most life-threatening diseases caused by obesity, non-alcoholic fatty liver disease, high alcohol consumption, hepatitis B or C infection, autoimmune diseases, cholestatic diseases, and iron or copper overload [200]. Chronic liver diseases (CLDs) are more liver-related diseases characterized by a decreased hepatic function as a result of chronic inflammation of the liver, leading to the development of cirrhosis [201]. The only established successful treatment for end-stage liver disease is liver transplantation [202]. The drawback of liver transplantation is organ shortage [203]. As a solution to the transplant limitation, researchers have tried to develop an artificial liver able to provide detoxification, synthesis, biotransformation and excretion functionality as performed by the liver [202]. Various polymeric membranes have been reported in order to develop an artificial liver [25,204,205]. The developed membranes are able to mimic the physiological environment. Further, in the last years, the development of membrane bioreactors was reported, which are able to potentially optimize the treatment during reversible acute liver disease or while waiting for a liver transplant [206]. The hollow fiber membrane bioreactors (HFMBRs) are used to develop liver tissue constructs for bioartificial liver (BAL) or as an in vitro drug discovery and testing platform [207]. Salerno et al. [204] developed biodegradable hollow fiber (HF) membranes of poly(ε-caprolactone) (PCL) with permeability, structural and mechanical properties that supported the cell adhesion and functionality (Figure 5). The endothelial cells were cultured in the lumen of the fibers, and hepatocytes in the shells of the fibers, communicating through their secreted molecules that permeate into the microporous structure of the HFs membrane wall. The sustaining glucose consumption, albumin synthesis, urea production and drug biotransformation function were sustained for 18 days by the hepatic tissue. The biocompatibility of PCL was studied by Slivac et al. [208] where PCL was used as a matrix for electrospun PCL Mats for tissue scaffolds for hepatic cell application. The hepatic cells are grown and attached on obtained scaffolds, yet the PCL scaffolds are able to mimic the bioactivity found in the original tissue matrix. Teotia et al. [25] presented a hollow fiber membrane based on asymmetric porous polysulfones and polysulfone-Tocopheryl polyethylene glycol succinate (PsfTPGS) composite via phase inversion procedure and subsequently surface modified with chitosan using sulfonation with concentrated sulfuric acid. The addition of chitosan confers biocompatibility to the obtained composite membranes due to structural similarity to glycosaminoglycans, a native liver Extracellular matrix (ECM) component leading to supporting cell growth, proliferation and the expression of the liver-specific function.

2.5. Artificial Pancreas

The pancreas is an internal organ located in the abdominal cavity just behind the stomach with a role in the regulation of body metabolism [209]. The function of the pancreas is to control glucose homeostasis through the secretion of endocrine hormones, such as insulin, and to produce exocrine enzymes required for digestion [210]. In type 1 diabetes, patients are unable to produce insulin due to an autoimmune response to the body’s insulin-producing beta cells. As a result, they need insulin treatment [211]. In the case of type 2 diabetes, the organism develops insulin resistance, and the glucose level remains high. The purpose of the artificial pancreas is to protect the pancreatic islets from the immune system’s response, allowing the transfer of insulin, oxygen and other nutrients [212]. The encapsulation of islets could be intravascular or extravascular, depending on the position of the implant. Over the last 40 years, were studied different bio-artificial pancreas (BAP) devices [213]. The semipermeable membranes for intravascular encapsulation are obtained from synthetic polymers, such as polyacrylonitrile-polyvinylchloride (PAN-PVC) copolymer, polycarbonate, ethylene vinyl alcohol copolymer (EVAL) fibers, poly-amino-urethane, nonwoven polytetrafluoroethylene (PTFE) fabric and nylon [214]. These membranes must fulfill several characteristics, so that they can be used to protect and segregate the islet cell from the immune response, allowing the oxygen, glucose, and other nutrients to permeate so that the level of glucose in the blood could be controlled. Further, these membranes must present a high hemocompatibility and biocompatibility, and they must be easy to implant so that, in case of failure, can be replaced very easily [214].

2.6. Osteosynthesis Membrane

Another application of membranes in the medical field is represented by tissue engineering, or more precisely, osteosynthesis and bone regeneration. The research of materials with applicability in bone regeneration started from the need to replace the materials from which metal implants were obtained in order to fix fractures. The major disadvantage of these materials was that after implantation, followed by fracture healing; another surgery was performed to remove the metal implants. Following this implant extraction operation, there is a slight possibility of the apparition of infection, removal problems of jammed implants, implant migration and associated extra health care costs. The solution was to develop new polymeric materials capable of fixing bone defects, as did metal implants, but with the ability to biodegrade at the same time as the actual healing and to facilitate bone healing [215]. Biodegradable materials are defined as materials that, after a period following implantation in the body, are disintegrating [216]. An example of biodegradable materials can be constituted some polymer materials which may disintegrate in vivo via the effect of the biological environment on the integrity of the material, leading to surface erosion or bulk erosion [215]. An osteoconductive material is a material able to serve as a scaffold onto which bone cells, such as osteoblasts and osteoclasts, can attach, migrate, grow and/or divide [217]. The main properties of biodegradable materials for osteosynthesis are excellent mechanical properties, control over degradation time, and biocompatibility [218]. Generally, membranes for osteosynthesis are made of biopolymers from natural sources, such as collagen, chitosan and cellulose, and synthetic sources, such as expanded polytetrafluoroethylene (e-PTFE), poly lactic acid (PLA), and polycaprolactone (PCL) [219,220,221]. Synthetic polymers are more susceptible to an inflammatory response in comparison with natural polymers. Still, an advantage worth mentioning is the capability to control biodegradability, processability, and drug-encapsulating ability [219]. Some of the most used biopolymers in osteosynthesis membranes are chitosan, collagen and cellulose due to increased biocompatibility through a lower immune response and osteoblastic adhesion [222,223,224]. The limitation of collagen is that if collagen is of animal origin, it could transmit disease from animal to human and also, the mechanical properties are lower [225]. In order to increase the mechanical properties, it was reported that collagen crosslinked with glutaraldehyde [226]. Unfortunately, glutaraldehyde shows cytotoxicity due to by-product degradation in later metabolic pathways [227]. As a result, cross-linking materials have been studied, such as EDC(1-ethyl-3-(3-dimethylaminopropyl) carbodiimide)-NHS(N-hydroxysuccinimide), genipin and oligomeric proanthocyanidins (OPCs) [221,226,228]. Liu et al. [229] used genipin, which is a natural cross-linker for collagen type 1 immobilization on Ti surfaces for improvement of the ensuing biological responses.
Further, many researchers are adding hydroxy apatite for an additional increase in osteosynthesis. Zheng et al. [230] presented a multilayer membrane based on collagen/chondroitin sulfate (COL/CS) which was assembled onto an electrospun PCL membrane followed by apatite mineralization (Figure 6). It was shown that the multilayer membrane has favorable mechanical properties, hydrophilicity, biodegradability, and outstanding biocompatibility. It also showed that the multilayer membrane promotes proliferation and osteogenic differentiation due to the addition of the apatite, which is able to promote bone-implant bonding and osteoinduction. The osteoconductivity of the membrane could be used as adjuvant films for osteointegration of a metallic implant, leading to better compatibility of the implant, as the membrane is placed between the implant and the bone. Pandele et al. [20] proposed polylactic acid and micro-structured hydroxyapatite particles in order to obtain composite films. The addition of hydroxyapatite doesn’t affect the degradation temperature, but a decrease in the crystallinity of the composite films was observed.
Cellulose acetate is a cellulose-derivate biopolymer extremely used in membrane development for osteosynthesis membranes due to its biocompatibility, biodegradability and hydrophilicity properties [231,232,233]. Sofi et al. [234] described the development of a scaffold for tissue engineering applications based on cellulose acetate functionalized with hydroxyapatite and silver nanoparticles (Figure 7). The mineralization with hydroxyapatite gives an environment for adhesion, growth, and proliferation of chicken embryo fibroblasts, thus with more hydroxyapatite content which leads to better cell proliferation. The addition of silver nanoparticles confers antimicrobial properties to the composite’s membrane.
Besides functionalization with hydroxyapatite for better osteosynthesis of the membranes, resveratrol is also utilized for a stimulatory effect on bone formation [234,236]. Resveratrol is a polyphenol compound present in fruits and vegetables with great antioxidant properties, antitumoral and antibacterial properties, and also it was reported that it might control cell proliferation [237]. Pandele et al. [238] described the functionalization of cellulose acetate membrane with resveratrol using aminopropyl triethoxysilane (APTS) and glutaraldehyde as linker molecules in order to improve osteointegration (Figure 8). The results showed that the resveratrol functionalization confers osteoblasts viability and differentiation potential in terms of alkaline phosphatase (ALP) activity, which is present in a large amount in the cells of mineralized tissue and bone mineralization and is essential to the formation of hard tissue [239].
A limitation in osteosynthesis application is represented by poor bonding strength between bioceramics and biopolymers due to their dissimilarity in physical and chemical properties, leading to poor bonding strength between the two phases [240]. As a solution to this limitation, the use of coupling agents has been reported with two-parent groups (hydrophilic and hydrophobic) can improve the interface between the polymer and ceramic material, thus improving the properties [241,242,243]. As mentioned earlier, these coupling agents are able to help the interaction between the polymer matrix and the ceramic matrix through two different functional groups, one of which can react with organic molecules, and the other can absorb inorganic surfaces in order to obtain a firm bond [240]. An example of a coupling agent often used is represented by Amino-propyl-triethoxy-silane (APTES) used as a surface modifier for glass and ceramic in order to increase the strength properties of biodegradable composites, as well as to improve the adhesion of the material to tissues [35,244,245]. Biernat et al. [245] reported the functionalization of porous calcium phosphate ceramics (Ca-P)/ poly(L-lactide) (PLA) composites with APTES and alendronate in order to improve the mechanical properties and cytocompatibility. After the functionalization with APTES, an increase in mechanical properties of the composites and an improvement in biocompatibility were observed through Sodium alendronate binds to hydroxyapatites in the bone.

2.7. Membranes for Sensors

Sensors and biosensors based on polymeric membranes used in biomedical applications mainly use three types of membranes – functionalized, molecularly imprinted and composite. The functionalized membranes offer the possibility of immobilization on the surface of the various species that interact with the analyte of interest, the most important being the enzymes for the detection of various organic species [246] or molecules with complexing capacity for the detection of ions (such as crown ethers) [247]. The detection system involves both the use of electrodes and complex structures such as Surface Acoustic Wave (SAW) platforms [248]. Urease immobilized onto poly(vinyl alcohol) was used as an indicator electrode for the urea sensor with a detection limit in the range of 1 × 10−5–5 × 10−4 M and a sensitivity of 19,069 mV/decade [249]. A sensor for triglyceride detection was developed by immobilization of lipase on a polyethyleneimine film (using glutaraldehyde as a linker molecule) deposited on a glassy carbon electrode. The sensor showed a detection limit in the range of 100–500 mg/dL with average recovery values from 95.47 % to 101.05 % [250]. Introduced by Wulff in 1972 [251], molecularly imprinted polymers found their applicability in a wide range of biomedical applications, such as the controlled release of drugs or even sensors and biosensors [252] due to the remarkable properties of molecular or even species with larger dimensions recognition. Thus, molecularly imprinted membranes were used for the detection of cells [253], DNA fragments [252], various saccharides present in the urine [254], proteins [255] or genes [256]. The detection systems used are based either on Quartz Crystal Microbalance (QCM) [257] or the volumetric principle detection [258]. The simultaneous determination of cholesterol and cholestanol was performed with a sensor based on molecularly imprinted polymer membranes obtained from methacrylic acid, ethylene glycolmethacrylate, 2,2-dimetthoxy-2-phenyliacetonephenon. To ensure the electronic conductivity of the entire sensitive structure, the membranes incorporated carbon nanotubes (MWCNTs) and were deposited on the screen-printed carbon electrode with gold nanoparticles [259]. The sensor has been tested both in terms of signal accuracy from simulated solutions and complex matrices, as well as from the point of view of stability in time, at repeated measurements of 8 times a day, maintaining its stability and accuracy for 45 days. Qualitative and quantitative detection of lysozyme was successfully performed using a biosensor from molecularly imprinted membranes embedded with l-cysteine-capped Mn2+-doped ZnS quantum dots [260]. The novelty of the method lies in the possibility of using the optical signal (much more precise than the electronic one), due to the fluorescence effect of quantum dots. The sensor presented linear detection ranged from 1.0 × 10−7 to 1.0 × 10−6 mol L−1 with the detection limit of 10.2 nM. The simultaneous detection of three chemical species - insulin, proinsulin and C-peptides, was achieved by mixing those species with N-methacryloyl-(L) 3-histidine methyl ester, 2-hydroxyethyl methacrylate and ethylene glycol dimethacrylate, followed by the polymerization of the precursors under UV at the surface of the electrode (thick-film boron doped diamond electrode) [261]. The detection limit was in the range of 1–16 pM for insulin, 4–25 pM for proinsulin and 8–88 pM for C-peptide, respectively, in both artificial and real human serum samples. The composite membranes used in the field of biosensors are mainly based on the use as a filler of carbon nanotubes or graphene functionalized with enzyme specific for a certain biological species, the polymer representing usually the matrix that protects the sensitive part and ensures the deposition on different electrodes [262].

3. Conclusions and Future Perspectives

The field of polymer membranes for biomedical applications is one in continuous development, and the challenges that must be solved contribute to the dynamics of this field. The future perspectives are equally addressed to all the applications that membrane materials have in relation to this field. Some of them will be presented below. In the field of hemodialysis, it is unlikely that polysulfone will be replaced too soon. Instead, the process will be able to be combined, in particular with the controlled release of drugs in order to make especially efficient the treatment of diseases derived from chronic renal dysfunction, such as liver cancer, for example. The synthesis of supramolecular architectures that can download the drug under the action of the tumor marker - alpha-fetoprotein. Additionally, in the field of membranes that combine two therapeutic procedures, the development of membranes that treat diabetes, either by releasing insulin or by releasing other active substances used in the treatment of this medical condition, would be equally useful. Another challenge in the field of hemodialysis is the easy development of membranes, especially custom-made for ‘one-day’ hemodialysis processes to remove compounds that have reached the body as a result of various intoxications. The retention of heavy metals [263,264] from the blood or various organic compounds (pesticides, overdoses of medicines or drugs) could be achieved by filtering the blood with the help of the one-day hemodialysis procedure. The future of hemodialysis depends on a combination of the current membranes, as we know them at the moment, and microfluidics. Devices are to be worn permanently by the patient and to filter the blood with membranes that can be changed every two to three days. In the field of artificial liver and artificial pancreas applications, cell viability still presents a problem. The use of hepatocytes of porcine origin or Langerhans cells is limited by the extremely short time in which they maintain their viability inside the membranes used. Finding solutions to condition the membranes or to ‘freeze’ them immediately after obtaining them would make possible the sustainable use of these therapeutic solutions. Related to polymers used for these applications, most probably due to economic and environmental reasons, we will assist in a transition from synthetic polymers to natural ones. Cellulose and its derivatives are the most promising candidates, especially for osseointegration, but also for hemodialysis or drug delivery. The main advantage is given not only by the fact that it is a natural source polymer and presents high biocompatibility but also by the fact that is bioresorbable due to its chemical structure. The only residue resulting from decomposition is glucose.

Author Contributions

Conceptualization, E.R.R., S.I.V. and V.K.T.; data curation, E.R.R., S.I.V. and V.K.T.; writing—original draft preparation, E.R.R., S.I.V. and V.K.T.; writing—review and editing, E.R.R.; supervision, S.I.V.; funding acquisition, S.I.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Ministry of Research, Innovation and Digitization, CNCS/CCCDI – UEFISCDI, project number PN-III-P4-ID-PCE-2020-1154, Hemodialysis combined with stimuli-responsive drug delivery—a new generation of polymeric membranes for advanced biomedical applications within PNCDI III.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qiu, J.; Tanaka, M. Biomedical Applications. In Encyclopedia of Smart Materials; John Wiley and Sons: Hoboken, NJ, USA, 2002. [Google Scholar]
  2. Guo, Z.; Poot, A.A.; Grijpma, D.W. Advanced polymer-based composites and structures for biomedical applications. Eur. Polym. J. 2021, 149, 110388. [Google Scholar] [CrossRef]
  3. Lam, M.T.; Wu, J.C. Biomaterial applications in cardiovascular tissue repair and regeneration. Expert Rev. Cardiovasc. Ther. 2012, 10, 1039–1049. [Google Scholar] [CrossRef]
  4. Xu, J.; Xue, Y.; Hu, G.; Lin, T.; Gou, J.; Yin, T.; He, H.; Zhang, Y.; Tang, X. A comprehensive review on contact lens for ophthalmic drug delivery. J. Control. Release 2018, 281, 97–118. [Google Scholar] [CrossRef] [PubMed]
  5. Ronco, C.; Clark, W.R. Haemodialysis membranes. Nat. Rev. Nephrol. 2018, 14, 394–410. [Google Scholar] [CrossRef] [PubMed]
  6. Saptaji, K.; Gebremariam, M.A.; Azhari, M.A.B.M. Machining of biocompatible materials: A review. Int. J. Adv. Manuf. Technol. 2018, 97, 2255–2292. [Google Scholar] [CrossRef]
  7. Huzum, B.; Puha, B.; Necoara, R.M.; Gheorghevici, S.; Puha, G.; Filip, A.; Sirbu, P.D.; Alexa, O. Biocompatibility assessment of biomaterials used in orthopedic devices: An overview (Review). Exp. Ther. Med. 2021, 22, 1315. [Google Scholar] [CrossRef]
  8. Zhao, D.; Yu, S.; Sun, B.; Gao, S.; Guo, S.; Zhao, K. Biomedical Applications of Chitosan and Its Derivative Nanoparticles. Polymers 2018, 10, 462. [Google Scholar] [CrossRef] [Green Version]
  9. Zagho, M.M.; Hussein, E.A.; Elzatahry, A.A. Recent Overviews in Functional Polymer Composites for Biomedical Applications. Polymers 2018, 10, 739. [Google Scholar] [CrossRef] [Green Version]
  10. Li, H.; Xu, M.; Shi, R.; Zhang, A.; Zhang, J. Advances in Electrostatic Spinning of Polymer Fibers Functionalized with Metal-Based Nanocrystals and Biomedical Applications. Molecules 2022, 27, 5548. [Google Scholar] [CrossRef]
  11. Megahed, M.M.; Abdel Bar, M.M.; Abouelez, E.s.M.; El-Shamy, A.M. Polyamide Coating as a Potential Protective Layer Against Corrosion of Iron Artifacts. Egypt. J. Chem. 2021, 64, 5693–5702. [Google Scholar] [CrossRef]
  12. Abdel-Karim, A.M.; El-Shamy, A.M.; Reda, Y. Corrosion and Stress Corrosion Resistance of Al Zn Alloy 7075 by Nano-Polymeric Coatings. J. Bio- Tribo-Corros. 2022, 8, 57. [Google Scholar] [CrossRef]
  13. Zohdy, K.M.; El-Sherif, R.M.; El-Shamy, A.M. Effect of pH fluctuations on the biodegradability of nanocomposite Mg-alloy in simulated bodily fluids. Chem. Pap. 2022. [Google Scholar] [CrossRef]
  14. Siddique, T.; Dutta, N.K.; Choudhury, N.R. Mixed-Matrix Membrane Fabrication for Water Treatment. Membranes 2021, 11, 557. [Google Scholar] [CrossRef]
  15. Ulbricht, M. Advanced functional polymer membranes. Polymer 2006, 47, 2217–2262. [Google Scholar] [CrossRef] [Green Version]
  16. Woźniak-Budych, M.J. Polymeric membranes for biomedical applications. Phys. Sci. Rev. 2021. [Google Scholar] [CrossRef]
  17. Malik, T.; Razzaq, H.; Razzaque, S.; Nawaz, H.; Siddiqa, A.; Siddiq, M.; Qaisar, S. Design and synthesis of polymeric membranes using water-soluble pore formers: An overview. Polym. Bull. 2019, 76, 4879–4901. [Google Scholar] [CrossRef]
  18. Mbituyimana, B.; Mao, L.; Hu, S.; Ullah, M.W.; Chen, K.; Fu, L.; Zhao, W.; Shi, Z.; Yang, G. Bacterial cellulose/glycolic acid/glycerol composite membrane as a system to deliver glycolic acid for anti-aging treatment. J. Bioresour. Bioprod. 2021, 6, 129–141. [Google Scholar] [CrossRef]
  19. Wsoo, M.A.; Razak, S.I.A.; Bohari, S.P.M.; Shahir, S.; Salihu, R.; Kadir, M.R.A.; Nayan, N.H.M. Vitamin D3-loaded electrospun cellulose acetate/polycaprolactone nanofibers: Characterization, in-vitro drug release and cytotoxicity studies. Int. J. Biol. Macromol. 2021, 181, 82–98. [Google Scholar] [CrossRef]
  20. Pandele, A.M.; Constantinescu, A.; Radu, I.C.; Miculescu, F.; Ioan Voicu, S.; Ciocan, L.T. Synthesis and Characterization of PLA-Micro-structured Hydroxyapatite Composite Films. Materials 2020, 13, 274. [Google Scholar] [CrossRef] [Green Version]
  21. Thakur, V.K.; Voicu, S.I. Recent advances in cellulose and chitosan based membranes for water purification: A concise review. Carbohydr. Polym. 2016, 146, 148–165. [Google Scholar] [CrossRef]
  22. Taemeh, M.A.; Shiravandi, A.; Korayem, M.A.; Daemi, H. Fabrication challenges and trends in biomedical applications of alginate electrospun nanofibers. Carbohydr. Polym. 2020, 228, 115419. [Google Scholar] [CrossRef]
  23. Aydemir Sezer, U.; Ozturk, K.; Aru, B.; Yanıkkaya Demirel, G.; Sezer, S. A design achieved by coaxial electrospinning of polysulfone and sulfonated polysulfone as a core-shell structure to optimize mechanical strength and hemocompatibility. Surf. Interfaces 2018, 10, 176–187. [Google Scholar] [CrossRef]
  24. Seifert, B.; Mihanetzis, G.; Groth, T.; Albrecht, W.; Richau, K.; Missirlis, Y.; Paul, D.; Von Sengbusch, G. Polyetherimide: A New Membrane-Forming Polymer for Biomedical Applications. Artif. Organs 2002, 26, 189–199. [Google Scholar] [CrossRef]
  25. Teotia, R.S.; Kalita, D.; Singh, A.K.; Verma, S.K.; Kadam, S.S.; Bellare, J.R. Bifunctional Polysulfone-Chitosan Composite Hollow Fiber Membrane for Bioartificial Liver. ACS Biomater. Sci. Eng. 2015, 1, 372–381. [Google Scholar] [CrossRef]
  26. He, T.; Yu, S.; He, J.; Chen, D.; Li, J.; Hu, H.; Zhong, X.; Wang, Y.; Wang, Z.; Cui, Z. Membranes for extracorporeal membrane oxygenator (ECMO): History, preparation, modification and mass transfer. Chin. J. Chem. Eng. 2022, 49, 46–75. [Google Scholar] [CrossRef]
  27. Uchiyama, T.; Watanabe, J.; Ishihara, K. Biocompatible polymer alloy membrane for implantable artificial pancreas. J. Membr. Sci. 2002, 208, 39–48. [Google Scholar] [CrossRef]
  28. Becerra, J.; Rodriguez, M.; Leal, D.; Noris-Suarez, K.; Gonzalez, G. Chitosan-collagen-hydroxyapatite membranes for tissue engineering. J. Mater. Sci. Mater. Med. 2022, 33, 18. [Google Scholar] [CrossRef] [PubMed]
  29. Souza, A.P.C.; Neves, J.G.; Navarro da Rocha, D.; Lopes, C.C.; Moraes, Â.M.; Correr-Sobrinho, L.; Correr, A.B. Chitosan/Xanthan membrane containing hydroxyapatite/Graphene oxide nanocomposite for guided bone regeneration. J. Mech. Behav. Biomed. Mater. 2022, 136, 105464. [Google Scholar] [CrossRef]
  30. Tanzli, E.; Ehrmann, A. Electrospun Nanofibrous Membranes for Tissue Engineering and Cell Growth. Appl. Sci. 2021, 11, 6929. [Google Scholar] [CrossRef]
  31. Morelli, S.; Liu, S.-J.; De Bartolo, L. Membrane Systems for Tissue Engineering 2020. Membranes 2021, 11, 763. [Google Scholar] [CrossRef]
  32. Oprea, M.; Voicu, S.I. Recent Advances in Applications of Cellulose Derivatives-Based Composite Membranes with Hydroxyapatite. Materials 2020, 13, 2481. [Google Scholar] [CrossRef]
  33. Oprea, M.; Voicu, S.I. Cellulose Composites with Graphene for Tissue Engineering Applications. Materials 2020, 13, 5347. [Google Scholar] [CrossRef] [PubMed]
  34. Radu, E.R.; Semenescu, A.; Voicu, S.I. Recent Advances in Stimuli-Responsive Doxorubicin Delivery Systems for Liver Cancer Therapy. Polymers 2022, 14, 5249. [Google Scholar] [CrossRef] [PubMed]
  35. Voicu, S.I.; Thakur, V.K. Aminopropyltriethoxysilane as a linker for cellulose-based functional materials: New horizons and future challenges. Curr. Opin. Green Sustain. Chem. 2021, 30, 100480. [Google Scholar] [CrossRef]
  36. Pandele, A.M.; Iovu, H.; Orbeci, C.; Tuncel, C.; Miculescu, F.; Nicolescu, A.; Deleanu, C.; Voicu, S.I. Surface modified cellulose acetate membranes for the reactive retention of tetracycline. Sep. Purif. Technol. 2020, 249, 117145. [Google Scholar] [CrossRef]
  37. Muhulet, A.; Tuncel, C.; Miculescu, F.; Pandele, A.M.; Bobirica, C.; Orbeci, C.; Bobirica, L.; Palla-Papavlu, A.; Voicu, S.I. Synthesis and characterization of polysulfone–TiO2 decorated MWCNT composite membranes by sonochemical method. Appl. Phys. A 2020, 126, 233. [Google Scholar] [CrossRef]
  38. Voicu, Ş.I.; Dobrica, A.; Sava, S.; Ivan, A.; Naftanaila, L. Cationic surfactants-controlled geometry and dimensions of polymeric membrane pores. J. Optoelectron. Adv. Mater. 2012, 14, 923–928. [Google Scholar]
  39. Stamatialis, D.F.; Papenburg, B.J.; Gironés, M.; Saiful, S.; Bettahalli, S.N.M.; Schmitmeier, S.; Wessling, M. Medical applications of membranes: Drug delivery, artificial organs and tissue engineering. J. Membr. Sci. 2008, 308, 1–34. [Google Scholar] [CrossRef] [Green Version]
  40. Van Gheluwe, L.; Chourpa, I.; Gaigne, C.; Munnier, E. Polymer-Based Smart Drug Delivery Systems for Skin Application and Demonstration of Stimuli-Responsiveness. Polymers 2021, 13, 1285. [Google Scholar] [CrossRef]
  41. Abbasnezhad, N.; Kebdani, M.; Shirinbayan, M.; Champmartin, S.; Tcharkhtchi, A.; Kouidri, S.; Bakir, F. Development of a Model Based on Physical Mechanisms for the Explanation of Drug Release: Application to Diclofenac Release from Polyurethane Films. Polymers 2021, 13, 1230. [Google Scholar] [CrossRef]
  42. Jager, K.J.; Kovesdy, C.; Langham, R.; Rosenberg, M.; Jha, V.; Zoccali, C. A single number for advocacy and communication-worldwide more than 850 million individuals have kidney diseases. Kidney Int. 2019, 96, 1048–1050. [Google Scholar] [CrossRef] [PubMed]
  43. Said, N.; Lau, W.J.; Ho, Y.-C.; Lim, S.K.; Zainol Abidin, M.N.; Ismail, A.F. A Review of Commercial Developments and Recent Laboratory Research of Dialyzers and Membranes for Hemodialysis Application. Membranes 2021, 11, 767. [Google Scholar] [CrossRef]
  44. Mollahosseini, A.; Abdelrasoul, A.; Shoker, A. A critical review of recent advances in hemodialysis membranes hemocompatibility and guidelines for future development. Mater. Chem. Phys. 2020, 248, 122911. [Google Scholar] [CrossRef]
  45. Vachharajani, T.J.; Taliercio, J.J.; Anvari, E. New Devices and Technologies for Hemodialysis Vascular Access: A Review. Am. J. Kidney Dis. 2021, 78, 116–124. [Google Scholar] [CrossRef] [PubMed]
  46. Saran, R.; Robinson, B.; Abbott, K.C.; Bragg-Gresham, J.; Chen, X.; Gipson, D.; Gu, H.; Hirth, R.A.; Hutton, D.; Jin, Y.; et al. US Renal Data System 2019 Annual Data Report: Epidemiology of Kidney Disease in the United States. Am. J. Kidney Dis. 2020, 75, A6–A7. [Google Scholar] [CrossRef] [PubMed]
  47. Mohammadi, F.; Mohammadi, F.; Yavari, Z. Characterization of the cylindrical electrospun nanofibrous polysulfone membrane for hemodialysis with modelling approach. Med. Biol. Eng. Comput. 2021, 59, 1629–1641. [Google Scholar] [CrossRef]
  48. Chen, Y.-A.; Ou, S.-M.; Lin, C.-C. Influence of Dialysis Membranes on Clinical Outcomes: From History to Innovation. Membranes 2022, 12, 152. [Google Scholar] [CrossRef]
  49. Swift, O.; Vilar, E.; Farrington, K. Haemodialysis. Medicine 2019, 47, 596–602. [Google Scholar] [CrossRef]
  50. Salimi, E.; Ghaee, A.; Ismail, A.F.; Othman, M.H.D.; Sean, G.P. Current Approaches in Improving Hemocompatibility of Polymeric Membranes for Biomedical Application. Macromol. Mater. Eng. 2016, 301, 771–800. [Google Scholar] [CrossRef]
  51. Song, X.; Ji, H.; Zhao, W.; Sun, S.; Zhao, C. Hemocompatibility enhancement of polyethersulfone membranes: Strategies and challenges. Adv. Membr. 2021, 1, 100013. [Google Scholar] [CrossRef]
  52. Abdelrasoul, A.; Shoker, A. Induced hemocompatibility of polyethersulfone (PES) hemodialysis membrane using polyvinylpyrrolidone: Investigation on human serum fibrinogen adsorption and inflammatory biomarkers released. Chem. Eng. Res. Des. 2022, 177, 615–624. [Google Scholar] [CrossRef]
  53. Zhang, W.; Yue, P.; Zhang, H.; Yang, N.; Li, C.; Li, J.h.; Meng, J.; Zhang, Q. Surface modification of AO-PAN@OHec nanofiber membranes with amino acid for antifouling and hemocompatible properties. Appl. Surf. Sci. 2019, 475, 934–941. [Google Scholar] [CrossRef]
  54. Radu, E.R.; Voicu, S.I. Functionalized Hemodialysis Polysulfone Membranes with Improved Hemocompatibility. Polymers 2022, 14, 1130. [Google Scholar] [CrossRef]
  55. Zhong, D.; Wang, Z.; Zhou, J.; Wang, Y. Additive-free preparation of hemodialysis membranes from block copolymers of polysulfone and polyethylene glycol. J. Membr. Sci. 2021, 618, 118690. [Google Scholar] [CrossRef]
  56. Maggay, I.V.B.; Aini, H.N.; Lagman, M.M.G.; Tang, S.-H.; Aquino, R.R.; Chang, Y.; Venault, A. A Biofouling Resistant Zwitterionic Polysulfone Membrane Prepared by a Dual-Bath Procedure. Membranes 2022, 12, 69. [Google Scholar] [CrossRef]
  57. Abidin, M.N.Z.; Goh, P.S.; Said, N.; Ismail, A.F.; Othman, M.H.D.; Abdullah, M.S.; Ng, B.C.; Hasbullah, H.; Sheikh Abdul Kadir, S.H.; Kamal, F.; et al. Polysulfone/amino-silanized poly(methyl methacrylate) dual layer hollow fiber membrane for uremic toxin separation. Sep. Purif. Technol. 2020, 236, 116216. [Google Scholar] [CrossRef]
  58. Kaleekkal, N.J.; Thanigaivelan, A.; Tarun, M.; Mohan, D. A functional PES membrane for hemodialysis—Preparation, Characterization and Biocompatibility. Chin. J. Chem. Eng. 2015, 23, 1236–1244. [Google Scholar] [CrossRef]
  59. Gores, F.; Montag, P.; Schall, C.; Vienken, J.; Bowry, S.K. Verification of the chemical composition and specifications of haemodialysis membranes by NMR and GPC-FTIR-coupled spectroscopy. Biomaterials 2002, 23, 3131–3140. [Google Scholar] [CrossRef] [PubMed]
  60. Azhar, O.; Jahan, Z.; Sher, F.; Niazi, M.B.K.; Kakar, S.J.; Shahid, M. Cellulose acetate-polyvinyl alcohol blend hemodialysis membranes integrated with dialysis performance and high biocompatibility. Mater. Sci. Eng. C 2021, 126, 112127. [Google Scholar] [CrossRef] [PubMed]
  61. Amri, C.; Mudasir, M.; Siswanta, D.; Roto, R. In vitro hemocompatibility of PVA-alginate ester as a candidate for hemodialysis membrane. Int. J. Biol. Macromol. 2016, 82, 48–53. [Google Scholar] [CrossRef]
  62. Yu, X.; Zhu, Y.; Zhang, T.; Deng, L.; Li, P.; Wang, X.; Hsiao, B.S. Heparinized thin-film composite membranes with sub-micron ridge structure for efficient hemodialysis. J. Membr. Sci. 2020, 599, 117706. [Google Scholar] [CrossRef]
  63. Dumitriu, C.; Voicu, S.I.; Muhulet, A.; Nechifor, G.; Popescu, S.; Ungureanu, C.; Carja, A.; Miculescu, F.; Trusca, R.; Pirvu, C. Production and characterization of cellulose acetate–titanium dioxide nanotubes membrane fraxiparinized through polydopamine for clinical applications. Carbohydr. Polym. 2018, 181, 215–223. [Google Scholar] [CrossRef] [PubMed]
  64. Yamazaki, K.; Matsuda, M.; Yamamoto, K.-i.; Yakushiji, T.; Sakai, K. Internal and surface structure characterization of cellulose triacetate hollow-fiber dialysis membranes. J. Membr. Sci. 2011, 368, 34–40. [Google Scholar] [CrossRef]
  65. Eduok, U.; Abdelrasoul, A.; Shoker, A.; Doan, H. Recent developments, current challenges and future perspectives on cellulosic hemodialysis membranes for highly efficient clearance of uremic toxins. Mater. Today Commun. 2021, 27, 102183. [Google Scholar] [CrossRef]
  66. Maduell, F.; Broseta, J.J.; Rodríguez-Espinosa, D.; Hermida-Lama, E.; Rodas, L.M.; Gómez, M.; Arias-Guillén, M.; Fontseré, N.; Vera, M.; Rico, N. Evaluation and comparison of polysulfone TS-UL and PMMA NF-U dialyzers versus expanded hemodialysis and postdilution hemodiafiltration. Artif. Organs 2021, 45, E317–E323. [Google Scholar] [CrossRef]
  67. Koh, E.; Lee, Y.T. Development of an embossed nanofiber hemodialysis membrane for improving capacity and efficiency via 3D printing and electrospinning technology. Sep. Purif. Technol. 2020, 241, 116657. [Google Scholar] [CrossRef]
  68. Ding, S.; Zhang, T.; Li, P.; Wang, X. Dialysis/adsorption bifunctional thin-film nanofibrous composite membrane for creatinine clearance in portable artificial kidney. J. Membr. Sci. 2021, 636, 119550. [Google Scholar] [CrossRef]
  69. Clark, W.R.; Hamburger, R.J.; Lysaght, M.J. Effect of membrane composition and structure on solute removal and biocompatibility in hemodialysis. Kidney Int 1999, 56, 2005–2015. [Google Scholar] [CrossRef] [Green Version]
  70. Seddik, A.A.; Bashier, A.; Alhadari, A.K.; AlAlawi, F.; Alnour, H.H.; Bin Hussain, A.A.; Frankel, A.; Railey, M.J. Challenges in management of diabetic ketoacidosis in hemodialysis patients, case presentation and review of literature. Diabetes Metab. Syndr. Clin. Res. Rev. 2019, 13, 2481–2487. [Google Scholar] [CrossRef]
  71. Faria, M.; Moreira, C.; Eusébio, T.; Brogueira, P.; de Pinho, M.N. Hybrid flat sheet cellulose acetate/silicon dioxide ultrafiltration membranes for uremic blood purification. Cellulose 2020, 27, 3847–3869. [Google Scholar] [CrossRef]
  72. Bowry, S.K.; Gatti, E.; Vienken, J. Contribution of Polysulfone Membranes to the Success of Convective Dialysis Therapies. Contrib. Nephrol. 2011, 173, 110–118. [Google Scholar]
  73. Serbanescu, O.S.; Voicu, S.I.; Thakur, V.K. Polysulfone functionalized membranes: Properties and challenges. Mater. Today Chem. 2020, 17, 100302. [Google Scholar] [CrossRef]
  74. Voicu, S.I.; Sandru, M. Composite Hybrid Membrane Materials for Artificial Organs. In Handbook of Bioceramics and Biocomposites; Antoniac, I.V., Ed.; Springer International Publishing: Cham, Switzerland, 2016; pp. 407–429. [Google Scholar]
  75. Yue, W.-W.; Li, H.-J.; Xiang, T.; Qin, H.; Sun, S.-D.; Zhao, C.-S. Grafting of zwitterion from polysulfone membrane via surface-initiated ATRP with enhanced antifouling property and biocompatibility. J. Membr. Sci. 2013, 446, 79–91. [Google Scholar] [CrossRef]
  76. Said, N.; Hasbullah, H.; Ismail, A.F.; Othman, M.H.D.; Goh, P.S.; Zainol Abidin, M.N.; Sheikh Abdul Kadir, S.H.; Kamal, F.; Abdullah, M.S.; Ng, B.C. Enhanced hydrophilic polysulfone hollow fiber membranes with addition of iron oxide nanoparticles. Polym. Int. 2017, 66, 1424–1429. [Google Scholar] [CrossRef]
  77. Pandele, A.M.; Oprea, M.; Dutu, A.A.; Miculescu, F.; Voicu, S.I. A Novel Generation of Polysulfone/Crown Ether-Functionalized Reduced Graphene Oxide Membranes with Potential Applications in Hemodialysis. Polymers 2022, 14, 148. [Google Scholar] [CrossRef] [PubMed]
  78. Zhao, C.; Xue, J.; Ran, F.; Sun, S. Modification of polyethersulfone membranes—A review of methods. Prog. Mater. Sci. 2013, 58, 76–150. [Google Scholar] [CrossRef]
  79. Abe, T.; Kato, K.; Fujioka, T.; Akizawa, T. The Blood Compatibilities of Blood Purification Membranes and Other Materials Developed in Japan. Int. J. Biomater. 2011, 2011, 375390. [Google Scholar] [CrossRef] [Green Version]
  80. Irfan, M.; Idris, A.; Yusof, N.M.; Khairuddin, N.F.M.; Akhmal, H. Surface modification and performance enhancement of nano-hybrid f-MWCNT/PVP90/PES hemodialysis membranes. J. Membr. Sci. 2014, 467, 73–84. [Google Scholar] [CrossRef]
  81. Hoseinpour, V.; Ghaee, A.; Vatanpour, V.; Ghaemi, N. Surface modification of PES membrane via aminolysis and immobilization of carboxymethylcellulose and sulphated carboxymethylcellulose for hemodialysis. Carbohydr. Polym. 2018, 188, 37–47. [Google Scholar] [CrossRef] [PubMed]
  82. Xing, J.; Wang, Q.; He, T.; Zhou, Z.; Chen, D.; Yi, X.; Wang, Z.; Wang, R.; Tan, G.; Yu, P.; et al. Polydopamine-Assisted Immobilization of Copper Ions onto Hemodialysis Membranes for Antimicrobial. ACS Appl. Bio Mater. 2018, 1, 1236–1243. [Google Scholar] [CrossRef] [PubMed]
  83. Kourde-Hanafi, Y.; Loulergue, P.; Szymczyk, A.; Van der Bruggen, B.; Nachtnebel, M.; Rabiller-Baudry, M.; Audic, J.-L.; Pölt, P.; Baddari, K. Influence of PVP content on degradation of PES/PVP membranes: Insights from characterization of membranes with controlled composition. J. Membr. Sci. 2017, 533, 261–269. [Google Scholar] [CrossRef]
  84. Tiron, L.; Vlad, M.; Baltă, Ş. Research on hydrophilic nature of polyvinylpyrrolidone on polysulfone membrane filtration. IOP Conf. Ser. Mater. Sci. Eng. 2018, 374, 012059. [Google Scholar] [CrossRef]
  85. Mansur, S.; Hafiz, M.; Othman, M.H.; Ismail, A.; Nidzhom, M.; Zainol Abidin, M.N.; Said, N.; Goh, P.; Hasbullah, H.; Sheikh Abdul Kadir, S.H.; et al. Study on the Effect of PVP Additive on the Performance of PSf/PVP Ultrafiltration Hollow Fiber Membrane. J. Fundam. Appl. Sci. 2018, 14, 343–347. [Google Scholar] [CrossRef]
  86. Reuvers, M. 2.9-Anticoagulant and fibrinolytic drugs. In Drugs During Pregnancy and Lactation, 2nd ed.; Schaefer, C., Peters, P., Miller, R.K., Eds.; Academic Press: Oxford, UK, 2007; pp. 237–253. [Google Scholar]
  87. Song, W.; Zeng, Q.; Yin, X.; Zhu, L.; Gong, T.; Pan, C. Preparation and anticoagulant properties of heparin-like electrospun membranes from carboxymethyl chitosan and bacterial cellulose sulfate. Int. J. Biol. Macromol. 2018, 120, 1396–1405. [Google Scholar] [CrossRef] [PubMed]
  88. Rollason, G.; Sefton, M.V. Inactivation of thrombin in heparin-PVA coated tubes. J. Biomater. Sci. Polym. Ed. 1989, 1, 31–41. [Google Scholar] [CrossRef]
  89. Byun, Y.; Jacobs, H.A.; Kim, S.W. Binding kinetics of thrombin and antithrombin III with immobilized heparin using a spacer. ASAIO J. (Am. Soc. Artif. Organs J.) 1992, 38, M649–M653. [Google Scholar] [CrossRef]
  90. Huang, X.-J.; Guduru, D.; Xu, Z.-K.; Vienken, J.; Groth, T. Immobilization of heparin on polysulfone surface for selective adsorption of low-density lipoprotein (LDL). Acta Biomater. 2010, 6, 1099–1106. [Google Scholar] [CrossRef]
  91. Ren, X.; Xu, L.; Xu, J.; Zhu, P.; Zuo, L.; Wei, S. Immobilized heparin and its anti-coagulation effect on polysulfone membrane surface. J. Biomater. Sci. Polym. Ed. 2013, 24, 1707–1720. [Google Scholar] [CrossRef]
  92. Tu, M.-M.; Xu, J.-J.; Qiu, Y.-R. Surface hemocompatible modification of polysulfone membrane via covalently grafting acrylic acid and sulfonated hydroxypropyl chitosan. RSC Adv. 2019, 9, 6254–6266. [Google Scholar] [CrossRef] [Green Version]
  93. Sakiyama-Elbert, S.E. Incorporation of heparin into biomaterials. Acta Biomater. 2014, 10, 1581–1587. [Google Scholar] [CrossRef] [Green Version]
  94. Ma, L.; Huang, J.; Zhu, X.; Zhu, B.; Wang, L.; Zhao, W.; Qiu, L.; Song, B.; Zhao, C.; Yan, F. In vitro and in vivo anticoagulant activity of heparin-like biomacromolecules and the mechanism analysis for heparin-mimicking activity. Int. J. Biol. Macromol. 2019, 122, 784–792. [Google Scholar] [CrossRef]
  95. Zhang, Z.; Zhao, Y.; Luo, X.; Feng, S.; Wu, L. Preparation of a heparin-like functionalized tannic acid-coated polyethersulfone ultrafiltration membrane for hemodialysis by a simple surface modification method. Appl. Surf. Sci. 2022, 572, 151440. [Google Scholar] [CrossRef]
  96. Robles, H. Tannic Acid. In Encyclopedia of Toxicology, 3rd ed.; Wexler, P., Ed.; Academic Press: Oxford, UK, 2014; pp. 474–475. [Google Scholar]
  97. Lu, D.; Jin, Y.; Wang, X.; Xie, L.; Liu, Q.; Chen, Y.; Wang, H.; Lei, Z. Heparin-like anticoagulant polypeptides with tunable activity: Synthesis, characterization, anticoagulative properties and clot solubilities in vitro. Mater. Sci. Eng. C 2021, 129, 112405. [Google Scholar] [CrossRef]
  98. Rojo, J.; Sousa-Herves, A.; Mascaraque, A. 1.24-Perspectives of Carbohydrates in Drug Discovery. In Comprehensive Medicinal Chemistry III; Chackalamannil, S., Rotella, D., Ward, S.E., Eds.; Elsevier: Oxford, UK, 2017; pp. 577–610. [Google Scholar]
  99. Xie, Y.; Chu, Z.; Jin, W. Beyond separation: Membranes towards medicine. J. Membr. Sci. Lett. 2022, 2, 100020. [Google Scholar] [CrossRef]
  100. Puiggalí-Jou, A.; del Valle, L.J.; Alemán, C. Drug delivery systems based on intrinsically conducting polymers. J. Control Release 2019, 309, 244–264. [Google Scholar] [CrossRef] [PubMed]
  101. Langer, R.S.; Peppas, N.A. Present and future applications of biomaterials in controlled drug delivery systems. Biomaterials 1981, 2, 201–214. [Google Scholar] [CrossRef]
  102. Macha, I.J.; Ben-Nissan, B.; Vilchevskaya, E.N.; Morozova, A.S.; Abali, B.E.; Müller, W.H.; Rickert, W. Drug Delivery From Polymer-Based Nanopharmaceuticals—An Experimental Study Complemented by Simulations of Selected Diffusion Processes. Front. Bioeng. Biotechnol. 2019, 7, 37. [Google Scholar] [CrossRef]
  103. Zhang, Y.; Chan, H.F.; Leong, K.W. Advanced materials and processing for drug delivery: The past and the future. Adv. Drug Deliv. Rev. 2013, 65, 104–120. [Google Scholar] [CrossRef] [Green Version]
  104. Couvreur, P. Nanoparticles in drug delivery: Past, present and future. Adv. Drug Deliv. Rev. 2013, 65, 21–23. [Google Scholar] [CrossRef]
  105. Li, J.; Ding, J.; Liu, T.; Liu, J.; Yan, L.; Chen, X. Poly(lactic acid) Controlled Drug Delivery; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  106. Mays, T. A new classification of pore sizes. Stud. Surf. Sci. Catal.-Stud. Surf. Sci. Catal. 2007, 160, 57–62. [Google Scholar] [CrossRef]
  107. Kabay, N.; Shirazi, M.M.A.; Guler, E.; Bryjak, M. Grand Challenges in Membrane Modules and Processes. Front. Membr. Sci. Technol. 2022, 1, 913597. [Google Scholar] [CrossRef]
  108. Mabrouk, M.; Rajendran, R.; Soliman, I.E.; Ashour, M.M.; Beherei, H.H.; Tohamy, K.M.; Thomas, S.; Kalarikkal, N.; Arthanareeswaran, G.; Das, D.B. Nanoparticle- and Nanoporous-Membrane-Mediated Delivery of Therapeutics. Pharmaceutics 2019, 11, 294. [Google Scholar] [CrossRef] [Green Version]
  109. Rao, V.; Amar, J.V.; Avasthi, D.K.; Narayana Charyulu, R. Etched ion track polymer membranes for sustained drug delivery. Radiat. Meas. 2003, 36, 585–589. [Google Scholar] [CrossRef]
  110. Ferain, E.; Legras, R. Track-etch templates designed for micro- and nanofabrication. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2003, 208, 115–122. [Google Scholar] [CrossRef]
  111. Kozlovskiy, A.; Borgekov, D.; Kenzhina, I.; Zdorovets, M.; Korolkov, I.; Kaniukov, E.; Kutuzau, M.; Shumskaya, A. PET Ion-Track Membranes: Formation Features and Basic Applications. In Proceedings of the International Conference on Nanotechnology and Nanomaterials, Kyiv, Ukraine, 27–30 August 2018; 2019; pp. 461–479. [Google Scholar]
  112. Hanot, H.; Ferain, E. Industrial applications of ion track technology. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2009, 267, 1019–1022. [Google Scholar] [CrossRef] [Green Version]
  113. Liu, F.; Wang, M.; Wang, X.; Wang, P.; Shen, W.; Ding, S.; Wang, Y.J.N. Fabrication and application of nanoporous polymer ion-track membranes. Nanotechnology 2018, 30, 052001. [Google Scholar] [CrossRef]
  114. Baptista, D.; Teixeira, L.M.; Birgani, Z.T.; van Riet, S.; Pasman, T.; Poot, A.; Stamatialis, D.; Rottier, R.; Hiemstra, P.; Habibović, P.J.B. 3D alveolar in vitro model based on epithelialized biomimetically curved culture membranes. Biomaterials 2021, 266, 120436. [Google Scholar] [CrossRef]
  115. Ivanova, N.; Filippova, E.; Tverdokhlebov, S.; Levkovich, N.; Apel, P.Y.J.M.; Technologies, M. Preparation, Structure, and Properties of Track-Etched Membranes Based on Polylactic Acid. Membr. Membr. Technol. 2021, 3, 282–290. [Google Scholar] [CrossRef]
  116. Patil, R.S.; Sancaktar, E. Fabrication of pH-Responsive Polyimide Polyacrylic Acid Smart Gating Membranes: Ultrafast Method Using 248 nm Krypton Fluoride Excimer Laser. ACS Appl. Mater. Interfaces 2021, 13, 24431–24441. [Google Scholar] [CrossRef]
  117. Ahmed, F.E.; Lalia, B.S.; Hashaikeh, R. A review on electrospinning for membrane fabrication: Challenges and applications. Desalination 2015, 356, 15–30. [Google Scholar] [CrossRef]
  118. Luraghi, A.; Peri, F.; Moroni, L. Electrospinning for drug delivery applications: A review. J. Control Release 2021, 334, 463–484. [Google Scholar] [CrossRef] [PubMed]
  119. Mao, Z.; Li, J.; Huang, W.; Jiang, H.; Zimba, B.L.; Chen, L.; Wan, J.; Wu, Q. Preparation of poly (lactic acid)/graphene oxide nanofiber membranes with different structures by electrospinning for drug delivery. RSC Adv. 2018, 8, 16619–16625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and Electrospun Nanofibers: Methods, Materials, and Applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  121. Ren, X.; Han, Y.; Wang, J.; Jiang, Y.; Yi, Z.; Xu, H.; Ke, Q. An aligned porous electrospun fibrous membrane with controlled drug delivery—An efficient strategy to accelerate diabetic wound healing with improved angiogenesis. Acta Biomater. 2018, 70, 140–153. [Google Scholar] [CrossRef]
  122. Campa-Siqueiros, P.I.; Madera-Santana, T.J.; Castillo-Ortega, M.M.; López-Cervantes, J.; Ayala-Zavala, J.F.; Ortiz-Vazquez, E.L. Electrospun and co-electrospun biopolymer nanofibers for skin wounds on diabetic patients: An overview. RSC Adv. 2021, 11, 15340–15350. [Google Scholar] [CrossRef] [PubMed]
  123. Al-Baadani, M.A.; Hii Ru Yie, K.; Al-Bishari, A.M.; Alshobi, B.A.; Zhou, Z.; Fang, K.; Dai, B.; Shen, Y.; Ma, J.; Liu, J.; et al. Co-electrospinning polycaprolactone/gelatin membrane as a tunable drug delivery system for bone tissue regeneration. Mater. Des. 2021, 209, 109962. [Google Scholar] [CrossRef]
  124. van de Witte, P.; Esselbrugge, H.; Peters, A.M.P.; Dijkstra, P.J.; Feijen, J.; Groenewegen, R.J.J.; Smid, J.; Olijslager, J.; Schakenraad, J.M.; Eenink, M.J.D.; et al. Formation of porous membranes for drug delivery systems. J. Control Release 1993, 24, 61–78. [Google Scholar] [CrossRef] [Green Version]
  125. Zeinali, R.; Franco, L.; del Valle, L.J.; Puiggalí, J. Poly(butylene succinate) matrices obtained by thermally-induced phase separation: Pore shape and orientation affect drug release. Polymer 2022, 252, 124916. [Google Scholar] [CrossRef]
  126. Purkait, M.K.; Sinha, M.K.; Mondal, P.; Singh, R. Chapter 1-Introduction to Membranes. In Interface Science and Technology; Purkait, M.K., Sinha, M.K., Mondal, P., Singh, R., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; Volume 25, pp. 1–37. [Google Scholar]
  127. Adamczak, M.; Kamińska, G.; Bohdziewicz, J. Preparation of Polymer Membranes by In Situ Interfacial Polymerization. Int. J. Polym. Sci. 2019, 2019, 6217924. [Google Scholar] [CrossRef]
  128. Ding, L.-G.; Wang, S.; Yao, B.-J.; Li, F.; Li, Y.-A.; Zhao, G.-Y.; Dong, Y.-B. Synergistic Antibacterial and Anti-Inflammatory Effects of a Drug-Loaded Self-Standing Porphyrin-COF Membrane for Efficient Skin Wound Healing. Adv. Healthc. Mater. 2021, 10, 2001821. [Google Scholar] [CrossRef]
  129. LI, W.; Chen, X.; Lu, W.H.; Wang, Z. Preparation and characterization of poly (vinylidene fluoride)/TiO2 hybrid membranes. Front. Environ. Sci. Eng. 2013, 7, 492–502. [Google Scholar] [CrossRef]
  130. Wang, X.-M.; Li, X.-Y.; Shih, K. In situ embedment and growth of anhydrous and hydrated aluminum oxide particles on polyvinylidene fluoride (PVDF) membranes. J. Membr. Sci. 2011, 368, 134–143. [Google Scholar] [CrossRef]
  131. Hariharan, P.; Sundarrajan, S.; Arthanareeswaran, G.; Seshan, S.; Das, D.B.; Ismail, A.F. Advancements in modification of membrane materials over membrane separation for biomedical applications-Review. Environ. Res. 2022, 204, 112045. [Google Scholar] [CrossRef] [PubMed]
  132. Liang, S.; Gao, P.; Gao, X.; Xiao, K.; Huang, X. Improved blending strategy for membrane modification by virtue of surface segregation using surface-tailored amphiphilic nanoparticles. Front. Environ. Sci. Eng. 2016, 10, 9. [Google Scholar] [CrossRef]
  133. Parveen, S.; Sahoo, S.K. Long circulating chitosan/PEG blended PLGA nanoparticle for tumor drug delivery. Eur. J. Pharmacol. 2011, 670, 372–383. [Google Scholar] [CrossRef]
  134. Surface Modification by Graft Polymerization. In Surface Engineering of Polymer Membranes; Xu, Z.; Wan, L.; Huang, X. (Eds.) Springer: Berlin/Heidelberg, Germany, 2009; pp. 80–149. [Google Scholar]
  135. Abbas, M.; Ahmad, N. Surface Modification of RO Membrane by Grafting Hydrophilic Switchable Polymer Brushes. In Proceedings of the Filtech Conference, Cologne, Germany, 22–24 October 2019. [Google Scholar]
  136. Meng, J.; Li, J.; Zhang, Y.; Ma, S. A novel controlled grafting chemistry fully regulated by light for membrane surface hydrophilization and functionalization. J. Membr. Sci. 2014, 455, 405–414. [Google Scholar] [CrossRef]
  137. Jeong, E.; Byun, J.; Bayarkhuu, B.; Hong, S.W. Hydrophilic photocatalytic membrane via grafting conjugated polyelectrolyte for visible-light-driven biofouling control. Appl. Catal. B Environ. 2021, 282, 119587. [Google Scholar] [CrossRef]
  138. Joseph, R.M.; Merrick, M.M.; Liu, R.; Fraser, A.C.; Moon, J.D.; Choudhury, S.R.; Lesko, J.; Freeman, B.D.; Riffle, J.S. Synthesis and characterization of polybenzimidazole membranes for gas separation with improved gas permeability: A grafting and blending approach. J. Membr. Sci. 2018, 564, 587–597. [Google Scholar] [CrossRef]
  139. Wei, C.C.; Li, K. Preparation and Characterization of a Robust and Hydrophobic Ceramic Membrane via an Improved Surface Grafting Technique. Ind. Eng. Chem. Res. 2009, 48, 3446–3452. [Google Scholar] [CrossRef]
  140. Xu, W.-T.; Zhao, Z.-P.; Liu, M.; Chen, K.-C. Morphological and hydrophobic modifications of PVDF flat membrane with silane coupling agent grafting via plasma flow for VMD of ethanol–water mixture. J. Membr. Sci. 2015, 491, 110–120. [Google Scholar] [CrossRef]
  141. Liu, L.; Shen, F.; Chen, X.; Luo, J.; Su, Y.; Wu, H.; Wan, Y. A novel plasma-induced surface hydrophobization strategy for membrane distillation: Etching, dipping and grafting. J. Membr. Sci. 2016, 499, 544–554. [Google Scholar] [CrossRef]
  142. Liu, P.-S.; Chen, Q.; Liu, X.; Yuan, B.; Wu, S.-S.; Shen, J.; Lin, S.-C. Grafting of Zwitterion from Cellulose Membranes via ATRP for Improving Blood Compatibility. Biomacromolecules 2009, 10, 2809–2816. [Google Scholar] [CrossRef]
  143. Zhou, Z.; Zhu, S.; Zhang, D. Grafting of thermo-responsive polymer inside mesoporous silica with large pore size using ATRP and investigation of its use in drug release. J. Mater. Chem. 2007, 17, 2428–2433. [Google Scholar] [CrossRef]
  144. Pauly, A.C.; Schöller, K.; Baumann, L.; Rossi, R.M.; Dustmann, K.; Ziener, U.; de Courten, D.; Wolf, M.; Boesel, L.F.; Scherer, L.J. ATRP-based synthesis and characterization of light-responsive coatings for transdermal delivery systems. Sci. Technol. Adv. Mater. 2015, 16, 034604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Crownover, E.; Duvall, C.L.; Convertine, A.; Hoffman, A.S.; Stayton, P.S. RAFT-synthesized graft copolymers that enhance pH-dependent membrane destabilization and protein circulation times. J. Control. Release 2011, 155, 167–174. [Google Scholar] [CrossRef] [Green Version]
  146. Ying, L.; Yu, W.H.; Kang, E.T.; Neoh, K.G. Functional and Surface-Active Membranes from Poly(vinylidene fluoride)-graft-Poly(acrylic acid) Prepared via RAFT-Mediated Graft Copolymerization. Langmuir 2004, 20, 6032–6040. [Google Scholar] [CrossRef]
  147. Adeli, F.; Abbasi, F.; Babazadeh, M.; Davaran, S. Thermo/pH dual-responsive micelles based on the host–guest interaction between benzimidazole-terminated graft copolymer and β-cyclodextrin-functionalized star block copolymer for smart drug delivery. J. Nanobiotechnology 2022, 20, 91. [Google Scholar] [CrossRef]
  148. Li, X.; Sun, H.-b.; Sun, X. Polysulfone grafted with anthraquinone-hydroanthraquinone redox as a flexible membrane electrode for aqueous batteries. Polymer 2021, 234, 124245. [Google Scholar] [CrossRef]
  149. Luque-Agudo, V.; Hierro-Oliva, M.; Gallardo-Moreno, A.M.; González-Martín, M.L. Effect of plasma treatment on the surface properties of polylactic acid films. Polym. Test. 2021, 96, 107097. [Google Scholar] [CrossRef]
  150. Suganya, A.; Shanmugavelayutham, G.; Rodríguez, C.S. Study on structural, morphological and thermal properties of surface modified polyvinylchloride (PVC) film under air, argon and oxygen discharge plasma. Mater. Res. Express 2016, 3, 095302. [Google Scholar] [CrossRef]
  151. Yoshida, S.; Hagiwara, K.; Hasebe, T.; Hotta, A. Surface modification of polymers by plasma treatments for the enhancement of biocompatibility and controlled drug release. Surf. Coat. Technol. 2013, 233, 99–107. [Google Scholar] [CrossRef]
  152. Madalosso, H.B.; Machado, R.; Hotza, D.; Marangoni, C. Membrane Surface Modification by Electrospinning, Coating, and Plasma for Membrane Distillation Applications: A State-of-the-Art Review. Adv. Eng. Mater. 2021, 23, 2001456. [Google Scholar] [CrossRef]
  153. Polini, A.; Yang, F. 5-Physicochemical characterization of nanofiber composites. In Nanofiber Composites for Biomedical Applications; Ramalingam, M., Ramakrishna, S., Eds.; Woodhead Publishing: Oxford, UK, 2017; pp. 97–115. [Google Scholar]
  154. Hardin, N.Z.; Ravula, T.; Mauro, G.D.; Ramamoorthy, A. Hydrophobic Functionalization of Polyacrylic Acid as a Versatile Platform for the Development of Polymer Lipid Nanodisks. Small 2019, 15, 1804813. [Google Scholar] [CrossRef] [PubMed]
  155. Sagitha, P.; Reshmi, C.R.; Sundaran, S.P.; Binoy, A.; Mishra, N.; Sugith, A. β-Cyclodextrin functionalized polyurethane nano fibrous membranes for drug delivery. J. Drug Deliv. Sci. Technol. 2021, 65, 102759. [Google Scholar] [CrossRef]
  156. Das, S.S.; Bharadwaj, P.; Bilal, M.; Barani, M.; Rahdar, A.; Taboada, P.; Bungau, S.; Kyzas, G.Z. Stimuli-Responsive Polymeric Nanocarriers for Drug Delivery, Imaging, and Theragnosis. Polymers 2020, 12, 1397. [Google Scholar] [CrossRef]
  157. Wandera, D.; Wickramasinghe, S.R.; Husson, S.M. Stimuli-responsive membranes. J. Membr. Sci. 2010, 357, 6–35. [Google Scholar] [CrossRef]
  158. Zhang, S.; Ye, J.; Liu, X.; Wang, G.; Qi, Y.; Wang, T.; Song, Y.; Li, Y.; Ning, G. Dual Stimuli-Responsive smart fibrous membranes for efficient Photothermal/Photodynamic/Chemo-Therapy of Drug-Resistant bacterial infection. Chem. Eng. J. 2022, 432, 134351. [Google Scholar] [CrossRef]
  159. Municoy, S.; Álvarez Echazú, M.I.; Antezana, P.E.; Galdopórpora, J.M.; Olivetti, C.; Mebert, A.M.; Foglia, M.L.; Tuttolomondo, M.V.; Alvarez, G.S.; Hardy, J.G.; et al. Stimuli-Responsive Materials for Tissue Engineering and Drug Delivery. Int. J. Mol. Sci. 2020, 21, 4724. [Google Scholar] [CrossRef]
  160. Kocak, G.; Tuncer, C.; Bütün, V. pH-Responsive polymers. Polym. Chem. 2017, 8, 144–176. [Google Scholar] [CrossRef]
  161. Karimi, M.; Eslami, M.; Sahandi-Zangabad, P.; Mirab, F.; Farajisafiloo, N.; Shafaei, Z.; Ghosh, D.; Bozorgomid, M.; Dashkhaneh, F.; Hamblin, M.R. pH-Sensitive stimulus-responsive nanocarriers for targeted delivery of therapeutic agents. WIREs Nanomed. Nanobiotechnol. 2016, 8, 696–716. [Google Scholar] [CrossRef] [Green Version]
  162. Meng, X.; Li, P.; Wei, Q.; Zhang, H.-X. pH sensitive alginate-chitosan hydrogel beads for carvedilol delivery. Pharm. Dev. Technol. 2011, 16, 22–28. [Google Scholar] [CrossRef]
  163. Schoeller, J.; Itel, F.; Wuertz-Kozak, K.; Gaiser, S.; Luisier, N.; Hegemann, D.; Ferguson, S.J.; Fortunato, G.; Rossi, R.M. pH-Responsive Chitosan/Alginate Polyelectrolyte Complexes on Electrospun PLGA Nanofibers for Controlled Drug Release. Nanomaterials 2021, 11, 1850. [Google Scholar] [CrossRef]
  164. He, C.W.; Parowatkin, M.; Mailänder, V.; Flechtner-Mors, M.; Ziener, U.; Landfester, K.; Crespy, D. Sequence-Controlled Delivery of Peptides from Hierarchically Structured Nanomaterials. ACS Appl. Mater. Interfaces 2017, 9, 3885–3894. [Google Scholar] [CrossRef] [PubMed]
  165. Teotia, A.K.; Sami, H.; Kumar, A. 1-Thermo-responsive polymers: Structure and design of smart materials. In Switchable and Responsive Surfaces and Materials for Biomedical Applications; Zhang, Z., Ed.; Woodhead Publishing: Oxford, UK, 2015; pp. 3–43. [Google Scholar]
  166. Boztepe, C.; Künkül, A.; Yüceer, M. Application of artificial intelligence in modeling of the doxorubicin release behavior of pH and temperature responsive poly(NIPAAm-co-AAc)-PEG IPN hydrogel. J. Drug Deliv. Sci. Technol. 2020, 57, 101603. [Google Scholar] [CrossRef]
  167. Son, J.; Yi, G.; Yoo, J.; Park, C.; Koo, H.; Choi, H.S. Light-responsive nanomedicine for biophotonic imaging and targeted therapy. Adv. Drug Deliv. Rev. 2019, 138, 133–147. [Google Scholar] [CrossRef] [PubMed]
  168. Pantuso, E.; De Filpo, G.; Nicoletta, F.P. Light-Responsive Polymer Membranes. Adv. Opt. Mater. 2019, 7, 1900252. [Google Scholar] [CrossRef]
  169. Tomeh, M.A.; Hadianamrei, R.; Xu, D.; Brown, S.; Zhao, X. Peptide-functionalised magnetic silk nanoparticles produced by a swirl mixer for enhanced anticancer activity of ASC-J9. Colloids Surf. B Biointerfaces 2022, 216, 112549. [Google Scholar] [CrossRef]
  170. Guo, B.; Fan, R.; Shen, S.; Xue, Y.; Zhu, Z.; Xu, R.X. A photo-responsive membrane for tailored drug delivery with spatially and temporally controlled release. J. Mater. Chem. B 2021, 9, 8615–8625. [Google Scholar] [CrossRef]
  171. Nicoletta, F.P.; Cupelli, D.; Formoso, P.; De Filpo, G.; Colella, V.; Gugliuzza, A. Light Responsive Polymer Membranes: A Review. Membranes 2012, 2, 134–197. [Google Scholar] [CrossRef] [Green Version]
  172. Danaei, M.; Dehghankhold, M.; Ataei, S.; Hasanzadeh Davarani, F.; Javanmard, R.; Dokhani, A.; Khorasani, S.; Mozafari, M.R. Impact of Particle Size and Polydispersity Index on the Clinical Applications of Lipidic Nanocarrier Systems. Pharmaceutics 2018, 10, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Khattabi, A.M.; Talib, W.H.; Alqdeimat, D.A. The effect of polymer length on the in vitro characteristics of a drug loaded and targeted silica nanoparticles. Saudi Pharm. J. 2018, 26, 1022–1026. [Google Scholar] [CrossRef] [PubMed]
  174. Zhou, Y.; Fang, A.; Wang, F.; Li, H.; Jin, Q.; Huang, L.; Fu, C.; Zeng, J.; Jin, Z.; Song, X. Core-shell lipid-polymer nanoparticles as a promising ocular drug delivery system to treat glaucoma. Chin. Chem. Lett. 2020, 31, 494–500. [Google Scholar] [CrossRef]
  175. Owens, D.E., 3rd; Peppas, N.A. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int. J. Pharm. 2006, 307, 93–102. [Google Scholar] [CrossRef] [PubMed]
  176. Chaudhari, K.R.; Ukawala, M.; Manjappa, A.S.; Kumar, A.; Mundada, P.K.; Mishra, A.K.; Mathur, R.; Mönkkönen, J.; Murthy, R.S. Opsonization, biodistribution, cellular uptake and apoptosis study of PEGylated PBCA nanoparticle as potential drug delivery carrier. Pharm. Res. 2012, 29, 53–68. [Google Scholar] [CrossRef]
  177. Suk, J.S.; Xu, Q.; Kim, N.; Hanes, J.; Ensign, L.M. PEGylation as a strategy for improving nanoparticle-based drug and gene delivery. Adv. Drug Deliv. Rev. 2016, 99, 28–51. [Google Scholar] [CrossRef] [Green Version]
  178. Shi, L.; Zhang, J.; Zhao, M.; Tang, S.; Cheng, X.; Zhang, W.; Li, W.; Liu, X.; Peng, H.; Wang, Q. Effects of polyethylene glycol on the surface of nanoparticles for targeted drug delivery. Nanoscale 2021, 13, 10748–10764. [Google Scholar] [CrossRef]
  179. He, T.; He, J.; Wang, Z.; Cui, Z. Modification strategies to improve the membrane hemocompatibility in extracorporeal membrane oxygenator (ECMO). Adv. Compos. Hybrid Mater. 2021, 4, 847–864. [Google Scholar] [CrossRef]
  180. Schmidt, M.; Tachon, G.; Devilliers, C.; Muller, G.; Hekimian, G.; Bréchot, N.; Merceron, S.; Luyt, C.E.; Trouillet, J.L.; Chastre, J.; et al. Blood oxygenation and decarboxylation determinants during venovenous ECMO for respiratory failure in adults. Intensive Care Med. 2013, 39, 838–846. [Google Scholar] [CrossRef]
  181. Schlanstein, P.C.; Limper, A.; Hesselmann, F.; Schmitz-Rode, T.; Steinseifer, U.; Arens, J. Experimental method to determine anisotropic permeability of hollow fiber membrane bundles. J. Membr. Sci. 2018, 546, 70–81. [Google Scholar] [CrossRef]
  182. Fujiwara, T.; Nagaoka, E.; Watanabe, T.; Miyagi, N.; Kitao, T.; Sakota, D.; Mamiya, T.; Shinshi, T.; Arai, H.; Takatani, S. New generation extracorporeal membrane oxygenation with MedTech Mag-Lev, a single-use, magnetically levitated, centrifugal blood pump: Preclinical evaluation in calves. Artif. Organs 2013, 37, 447–456. [Google Scholar] [CrossRef]
  183. Lequier, L.; Horton, S.B.; McMullan, D.M.; Bartlett, R.H. Extracorporeal membrane oxygenation circuitry. Pediatr. Crit. Care Med. A J. Soc. Crit. Care Med. World Fed. Pediatr. Intensive Crit. Care Soc. 2013, 14, S7–S12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Abada, E.N.; Feinberg, B.J.; Roy, S. Evaluation of silicon membranes for extracorporeal membrane oxygenation (ECMO). Biomed. Microdevices 2018, 20, 86. [Google Scholar] [CrossRef]
  185. Valencia, E.; Nasr, V.G. Updates in Pediatric Extracorporeal Membrane Oxygenation. J. Cardiothorac. Vasc. Anesth. 2020, 34, 1309–1323. [Google Scholar] [CrossRef]
  186. Zangrillo, A.; Landoni, G.; Biondi-Zoccai, G.; Greco, M.; Greco, T.; Frati, G.; Patroniti, N.; Antonelli, M.; Pesenti, A.; Pappalardo, F. A meta-analysis of complications and mortality of extracorporeal membrane oxygenation. Crit. Care Resusc. J. Australas. Acad. Crit. Care Med. 2013, 15, 172–178. [Google Scholar]
  187. Achneck, H.E.; Sileshi, B.; Parikh, A.; Milano, C.A.; Welsby, I.J.; Lawson, J.H. Pathophysiology of bleeding and clotting in the cardiac surgery patient: From vascular endothelium to circulatory assist device surface. Circulation 2010, 122, 2068–2077. [Google Scholar] [CrossRef] [PubMed]
  188. Combes, A.; Hajage, D.; Capellier, G.; Demoule, A.; Lavoué, S.; Guervilly, C.; Da Silva, D.; Zafrani, L.; Tirot, P.; Veber, B.; et al. Extracorporeal Membrane Oxygenation for Severe Acute Respiratory Distress Syndrome. N. Engl. J. Med. 2018, 378, 1965–1975. [Google Scholar] [CrossRef] [PubMed]
  189. Winnersbach, P.; Hosseinnejad, A.; Breuer, T.; Fechter, T.; Jakob, F.; Schwaneberg, U.; Rossaint, R.; Bleilevens, C.; Singh, S. Endogenous Nitric Oxide-Releasing Microgel Coating Prevents Clot Formation on Oxygenator Fibers Exposed to In Vitro Blood Flow. Membranes 2022, 12, 73. [Google Scholar] [CrossRef]
  190. Doyle, A.; Hunt, B. Current Understanding of How Extracorporeal Membrane Oxygenators Activate Haemostasis and Other Blood Components. Front. Med. 2018, 5, 352. [Google Scholar] [CrossRef]
  191. Ecker, P.; Pekovits, M.; Yorov, T.; Haddadi, B.; Lukitsch, B.; Elenkov, M.; Janeczek, C.; Jordan, C.; Gfoehler, M.; Harasek, M. Microstructured Hollow Fiber Membranes: Potential Fiber Shapes for Extracorporeal Membrane Oxygenators. Membranes 2021, 11, 374. [Google Scholar] [CrossRef]
  192. Teber, O.O.; Altinay, A.D.; Mehrabani, S.A.N.; Tasdemir, R.S.; Zeytuncu, B.; Genceli, E.A.; Dulekgurgen, E.; Pekkan, K.; Koyuncu, İ. Polymeric hollow fiber membrane oxygenators as artificial lungs: A review. Biochem. Eng. J. 2022, 180, 108340. [Google Scholar] [CrossRef]
  193. Bazhenov, S.D.; Bildyukevich, A.V.; Volkov, A.V. Gas-Liquid Hollow Fiber Membrane Contactors for Different Applications. Fibers 2018, 6, 76. [Google Scholar] [CrossRef] [Green Version]
  194. Wang, Y.; Liu, Y.; Han, Q.; Lin, H.; Liu, F. A novel poly (4-methyl-1-pentene)/polypropylene (PMP/PP) thin film composite (TFC) artificial lung membrane for enhanced gas transport and excellent hemo-compatibility. J. Membr. Sci. 2022, 649, 120359. [Google Scholar] [CrossRef] [PubMed]
  195. Narendran, G.; Hoque, S.Z.; Satpathi, N.S.; Nampoothiri, K.N.; Sen, A.K. PDMS membrane-based flexible bi-layer microfluidic device for blood oxygenation. J. Micromechanics Microengineering 2022, 32, 094001. [Google Scholar] [CrossRef]
  196. Dharia, A.; Abada, E.; Feinberg, B.; Yeager, T.; Moses, W.; Park, J.; Blaha, C.; Wright, N.; Padilla, B.; Roy, S. Silicon Micropore-Based Parallel Plate Membrane Oxygenator. Artif. Organs 2018, 42, 166–173. [Google Scholar] [CrossRef] [PubMed]
  197. Zheng, Z.; Wang, W.; Huang, X.; Fan, W.; Li, L. Surface modification of polysulfone hollow fiber membrane for extracorporeal membrane oxygenator using low-temperature plasma treatment. Plasma Process. Polym. 2018, 15, 1700122. [Google Scholar] [CrossRef]
  198. Yi, E.; Kang, H.S.; Lim, S.M.; Heo, H.J.; Han, D.; Kim, J.F.; Park, A.; Choi, D.H.; Park, Y.-I.; Park, H.; et al. Superamphiphobic blood-repellent surface modification of porous fluoropolymer membranes for blood oxygenation applications. J. Membr. Sci. 2022, 648, 120363. [Google Scholar] [CrossRef]
  199. Park, A.; Song, Y.; Yi, E.; Duy Nguyen, B.T.; Han, D.; Sohn, E.; Park, Y.; Jung, J.; Lee, Y.M.; Cho, Y.H.; et al. Blood Oxygenation Using Fluoropolymer-Based Artificial Lung Membranes. ACS Biomater. Sci. Eng. 2020, 6, 6424–6434. [Google Scholar] [CrossRef]
  200. Ginès, P.; Krag, A.; Abraldes, J.G.; Solà, E.; Fabrellas, N.; Kamath, P.S. Liver cirrhosis. Lancet 2021, 398, 1359–1376. [Google Scholar] [CrossRef]
  201. Hope, A.A.; Morrison, R.S. Chapter 53-What Is the Clinical Course of Advanced Liver Disease and What Symptoms Are Associated With It? In Evidence-Based Practice in Palliative Medicine; Goldstein, N.E., Morrison, R.S., Eds.; W.B. Saunders: Philadelphia, PA, USA, 2013; pp. 300–307. [Google Scholar]
  202. Khakpour, S.; Ahmed, H.M.M.; De Bartolo, L. Membrane-Based Bioartificial Liver Devices. In Biomedical Membranes and (Bio)Artificial Organs; World Scientific Series in Membrane Science and Technology: Biological and Biomimetic Applications, Energy and the Environment; World Scientific: Singapore, 2017; Volume 2, pp. 149–178. [Google Scholar]
  203. Jadlowiec, C.C.; Taner, T. Liver transplantation: Current status and challenges. World J. Gastroenterol. 2016, 22, 4438–4445. [Google Scholar] [CrossRef]
  204. Salerno, S.; Tasselli, F.; Drioli, E.; De Bartolo, L. Poly(ε-Caprolactone) Hollow Fiber Membranes for the Biofabrication of a Vascularized Human Liver Tissue. Membranes 2020, 10, 112. [Google Scholar] [CrossRef]
  205. Morelli, S.; Piscioneri, A.; Salerno, S.; De Bartolo, L. Hollow Fiber and Nanofiber Membranes in Bioartificial Liver and Neuronal Tissue Engineering. Cells Tissues Organs 2022, 211, 46–75. [Google Scholar] [CrossRef]
  206. Verma, S.K.; Modi, A.; Bellare, J. Hollow fibre membrane bioreactor functionalized with GO-cryogel 3D matrix promotes liver cell anchoring and their functional maintenance. Med. Devices Sens. 2021, 4, e10128. [Google Scholar] [CrossRef]
  207. Khakpour, S.; Di Renzo, A.; Curcio, E.; Di Maio, F.P.; Giorno, L.; De Bartolo, L. Oxygen transport in hollow fibre membrane bioreactors for hepatic 3D cell culture: A parametric study. J. Membr. Sci. 2017, 544, 312–322. [Google Scholar] [CrossRef]
  208. Slivac, I.; Zdraveva, E.; Ivančić, F.; Žunar, B.; Holjevac Grgurić, T.; Gaurina Srček, V.; Svetec, I.-K.; Dolenec, T.; Bajsić, E.G.; Tominac Trcin, M.; et al. Bioactivity Comparison of Electrospun PCL Mats and Liver Extracellular Matrix as Scaffolds for HepG2 Cells. Polymers 2021, 13, 279. [Google Scholar] [CrossRef] [PubMed]
  209. Jellali, R.; Essaouiba, A.; Leclerc, E.; Legallais, C. Chapter 4-Membrane bioreactors for bio-artificial pancreas. In Current Trends and Future Developments on (Bio-) Membranes; Basile, A., Annesini, M.C., Piemonte, V., Charcosset, C., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 77–108. [Google Scholar]
  210. Cade, J.E.; Hanison, J. The pancreas. Anaesth. Intensive Care Med. 2017, 18, 527–531. [Google Scholar] [CrossRef]
  211. Daneman, D. Type 1 diabetes. Lancet 2006, 367, 847–858. [Google Scholar] [CrossRef]
  212. Hwang, P.T.; Shah, D.K.; Garcia, J.A.; Bae, C.Y.; Lim, D.J.; Huiszoon, R.C.; Alexander, G.C.; Jun, H.W. Progress and challenges of the bioartificial pancreas. Nano Converg. 2016, 3, 28. [Google Scholar] [CrossRef] [Green Version]
  213. Silva, A.I.; Mateus, M. Development of a polysulfone hollow fiber vascular bio-artificial pancreas device for in vitro studies. J. Biotechnol. 2009, 139, 236–249. [Google Scholar] [CrossRef]
  214. Duy Nguyen, B.T.; Nguyen Thi, H.Y.; Nguyen Thi, B.P.; Kang, D.-K.; Kim, J.F. The Roles of Membrane Technology in Artificial Organs: Current Challenges and Perspectives. Membranes 2021, 11, 239. [Google Scholar] [CrossRef]
  215. Eglin, D.; Alini, M. Degradable polymeric materials for osteosynthesis: Tutorial. Eur. Cells Mater. 2008, 16, 80–91. [Google Scholar] [CrossRef]
  216. Godavitarne, C.; Robertson, A.; Peters, J.; Rogers, B. Biodegradable materials. Orthop. Trauma 2017, 31, 316–320. [Google Scholar] [CrossRef]
  217. Wypych, G. 6-Functional Fillers-Applications. In Functional Fillers; Wypych, G., Ed.; ChemTec Publishing: Scarborough, ON, Canada, 2018; pp. 153–179. [Google Scholar]
  218. Cifuentes, S.C.; Benavente, R.; Lieblich, M.; González-Carrasco, J.L. Biodegradable and bioabsorbable materials for osteosynthesis applications: State-of-the-art and future perspectives. Handb. Compos. Renew. Mater. 2017, 5, 109–144. [Google Scholar]
  219. Shiohara, A.; Prieto-Simon, B.; Voelcker, N.H. Porous polymeric membranes: Fabrication techniques and biomedical applications. J. Mater. Chem. B 2021, 9, 2129–2154. [Google Scholar] [CrossRef] [PubMed]
  220. Marques, M.S.; Zepon, K.M.; Petronilho, F.C.; Soldi, V.; Kanis, L.A. Characterization of membranes based on cellulose acetate butyrate/poly(caprolactone)triol/doxycycline and their potential for guided bone regeneration application. Mater. Sci. Eng. C 2017, 76, 365–373. [Google Scholar] [CrossRef] [PubMed]
  221. Wang, J.; Wang, L.; Zhou, Z.; Lai, H.; Xu, P.; Liao, L.; Wei, J. Biodegradable Polymer Membranes Applied in Guided Bone/Tissue Regeneration: A Review. Polymers 2016, 8, 115. [Google Scholar] [CrossRef]
  222. Sbricoli, L.; Guazzo, R.; Annunziata, M.; Gobbato, L.; Bressan, E.; Nastri, L. Selection of Collagen Membranes for Bone Regeneration: A Literature Review. Materials 2020, 13, 786. [Google Scholar] [CrossRef] [Green Version]
  223. Neacsu, P.; Staras, A.I.; Voicu, S.I.; Ionascu, I.; Soare, T.; Uzun, S.; Cojocaru, V.D.; Pandele, A.M.; Croitoru, S.M.; Miculescu, F.; et al. Characterization and In Vitro and In Vivo Assessment of a Novel Cellulose Acetate-Coated Mg-Based Alloy for Orthopedic Applications. Materials 2017, 10, 686. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Acevedo, C.A.; Olguín, Y.; Briceño, M.; Forero, J.C.; Osses, N.; Díaz-Calderón, P.; Jaques, A.; Ortiz, R. Design of a biodegradable UV-irradiated gelatin-chitosan/nanocomposed membrane with osteogenic ability for application in bone regeneration. Mater. Sci. Eng. C 2019, 99, 875–886. [Google Scholar] [CrossRef]
  225. Döri, F.; Huszár, T.; Nikolidakis, D.; Arweiler, N.B.; Gera, I.; Sculean, A. Effect of platelet-rich plasma on the healing of intra-bony defects treated with a natural bone mineral and a collagen membrane. J. Clin. Periodontol. 2007, 34, 254–261. [Google Scholar] [CrossRef]
  226. Nair, M.; Best, S.M.; Cameron, R.E. Crosslinking Collagen Constructs: Achieving Cellular Selectivity Through Modifications of Physical and Chemical Properties. Appl. Sci. 2020, 10, 6911. [Google Scholar] [CrossRef]
  227. Takigawa, T.; Endo, Y. Effects of Glutaraldehyde Exposure on Human Health. J. Occup. Health 2006, 48, 75–87. [Google Scholar] [CrossRef]
  228. Yu, Y.; Xu, S.; Li, S.; Pan, H. Genipin-cross-linked hydrogels based on biomaterials for drug delivery: A review. Biomater. Sci. 2021, 9, 1583–1597. [Google Scholar] [CrossRef] [PubMed]
  229. Liu, C.-F.; Chang, K.-C.; Sun, Y.-S.; Nguyen, D.T.; Huang, H.-H. Immobilizing type I collagen via natural cross-linker genipin to enhance the osteogenic responses to titanium implant surface. J. Mater. Res. Technol. 2021, 15, 885–900. [Google Scholar] [CrossRef]
  230. Zheng, J.; Rahman, N.; Li, L.; Zhang, J.; Tan, H.; Xue, Y.; Zhao, Y.; Zhai, J.; Zhao, N.; Xu, F.; et al. Biofunctionalization of electrospun fiber membranes by LbL-collagen/chondroitin sulfate nanocoating followed by mineralization for bone regeneration. Mater. Sci. Eng. C 2021, 128, 112295. [Google Scholar] [CrossRef] [PubMed]
  231. Ahmed, M.K.; Menazea, A.A.; Abdelghany, A.M. Blend biopolymeric nanofibrous scaffolds of cellulose acetate/ε-polycaprolactone containing metallic nanoparticles prepared by laser ablation for wound disinfection applications. Int. J. Biol. Macromol. 2020, 155, 636–644. [Google Scholar] [CrossRef]
  232. Ioniță, M.; Crică, L.E.; Voicu, S.I.; Dinescu, S.; Miculescu, F.; Costache, M.; Iovu, H. Synergistic effect of carbon nanotubes and graphene for high performance cellulose acetate membranes in biomedical applications. Carbohydr. Polym. 2018, 183, 50–61. [Google Scholar] [CrossRef]
  233. Oprea, M.; Voicu, S.I. Recent advances in composites based on cellulose derivatives for biomedical applications. Carbohydr. Polym. 2020, 247, 116683. [Google Scholar] [CrossRef]
  234. Mizutani, K.; Ikeda, K.; Kawai, Y.; Yamori, Y. Resveratrol Stimulates the Proliferation and Differentiation of Osteoblastic MC3T3-E1 Cells. Biochem. Biophys. Res. Commun. 1998, 253, 859–863. [Google Scholar] [CrossRef]
  235. Sofi, H.S.; Akram, T.; Shabir, N.; Vasita, R.; Jadhav, A.H.; Sheikh, F.A. Regenerated cellulose nanofibers from cellulose acetate: Incorporating hydroxyapatite (HAp) and silver (Ag) nanoparticles (NPs), as a scaffold for tissue engineering applications. Mater. Sci. Eng. C 2021, 118, 111547. [Google Scholar] [CrossRef]
  236. Dai, Z.; Li, Y.; Quarles, L.D.; Song, T.; Pan, W.; Zhou, H.; Xiao, Z. Resveratrol enhances proliferation and osteoblastic differentiation in human mesenchymal stem cells via ER-dependent ERK1/2 activation. Phytomedicine 2007, 14, 806–814. [Google Scholar] [CrossRef]
  237. Risuleo, G. Chapter 33-Resveratrol: Multiple Activities on the Biological Functionality of the Cell. In Nutraceuticals; Gupta, R.C., Ed.; Academic Press: Boston, MA, USA, 2016; pp. 453–464. [Google Scholar]
  238. Pandele, A.M.; Neacsu, P.; Cimpean, A.; Staras, A.I.; Miculescu, F.; Iordache, A.; Voicu, S.I.; Thakur, V.K.; Toader, O.D. Cellulose acetate membranes functionalized with resveratrol by covalent immobilization for improved osseointegration. Appl. Surf. Sci. 2018, 438, 2–13. [Google Scholar] [CrossRef] [Green Version]
  239. Vimalraj, S. Alkaline phosphatase: Structure, expression and its function in bone mineralization. Gene 2020, 754, 144855. [Google Scholar] [CrossRef] [PubMed]
  240. Shuai, C.; Yu, L.; Feng, P.; Gao, C.; Peng, S. Interfacial reinforcement in bioceramic/biopolymer composite bone scaffold: The role of coupling agent. Colloids Surf. B Biointerfaces 2020, 193, 111083. [Google Scholar] [CrossRef] [PubMed]
  241. Manh Vu, C.; Bach, Q.-V.; Duong, L.X.; Thai, N.V.; Thao, V.D.; Duc, P.T.; Nguyen, D.D.; Hoang, T.; Nguyen Van, T.J. Silane coupling agent with amine group grafted nano/micro-glass fiber as novel toughener for epoxy resin: Fabrication and mechanical properties. Compos. Interfaces 2020, 27, 1085–1100. [Google Scholar] [CrossRef]
  242. Pramanik, N.; Mohapatra, S.; Pramanik, P.; Bhargava, P. Processing and Properties of Nano-Hydroxyapatite(n-HAp)/Poly(Ethylene-Co-Acrylic Acid)(EAA) Composite Using a Phosphonic Acid Coupling Agent for Orthopedic Applications. J. Am. Ceram. Soc. 2007, 90, 369–375. [Google Scholar] [CrossRef]
  243. Mallakpour, S.; Khani, M.; Mallakpour, F.; Fathi, M. Polyethylene-based nanocomposite: Structure and properties of poly(vinyl alcohol)/organofunctionalized Mg-doped fluorapatite hybrid. Int. J. Polym. Anal. Charact. 2017, 22, 237–246. [Google Scholar] [CrossRef]
  244. Hasan, M.S.; Ahmed, I.; Parsons, A.; Walker, G.; Scotchford, C. Cytocompatibility and Mechanical Properties of Short Phosphate Glass Fibre Reinforced Polylactic Acid (PLA) Composites: Effect of Coupling Agent Mediated Interface. J. Funct. Biomater. 2012, 3, 706–725. [Google Scholar] [CrossRef] [Green Version]
  245. Biernat, M.; Szwed-Georgiou, A.; Rudnicka, K.; Płociński, P.; Pagacz, J.; Tymowicz-Grzyb, P.; Woźniak, A.; Włodarczyk, M.; Urbaniak, M.M.; Krupa, A.; et al. Dual Modification of Porous Ca-P/PLA Composites with APTES and Alendronate Improves Their Mechanical Strength and Cytobiocompatibility towards Human Osteoblasts. Int. J. Mol. Sci. 2022, 23, 4315. [Google Scholar] [CrossRef]
  246. Himuro, Y.; Takai, M.; Ishihara, K. Poly(vinylferrocene-co-2-hydroxyethyl methacrylate) mediator as immobilized enzyme membrane for the fabrication of amperometric glucose sensor. Sens. Actuators B Chem. 2009, 136, 122–127. [Google Scholar] [CrossRef]
  247. Oprea, M.; Pandele, A.M.; Nicoara, A.I.; Nicolescu, A.; Deleanu, C.; Voicu, S.I. Crown ether-functionalized cellulose acetate membranes with potential applications in osseointegration. Int. J. Biol. Macromol. 2023, 230, 123162. [Google Scholar] [CrossRef]
  248. Palla-Papavlu, A.; Voicu, S.I.; Dinescu, M. Sensitive Materials and Coating Technologies for Surface Acoustic Wave Sensors. Chemosensors 2021, 9, 105. [Google Scholar] [CrossRef]
  249. Abd Hakim, S.; Rianna, M.; Rais, A. Manufacture and characterization of indicator electrodes from PPy + H2SO4 and PPy + Sulfonic acid as a urea sensor using urease enzyme immobilization technique in PVA. Mater. Sci. Energy Technol. 2023, 6, 89–93. [Google Scholar] [CrossRef]
  250. Thakkar, J.B.; Aghera, D.J.; Trivedi, B.; Prabha, C.R. Design and characterization of a biosensor with lipase immobilized nanoparticles in polymer film for the detection of triglycerides. Int. J. Biol. Macromol. 2023, 229, 136–145. [Google Scholar] [CrossRef] [PubMed]
  251. Wulff, G.; Sarhan, A. Über die Anwendung von enzymanalog gebauten Polymeren zur Racemattrennung. Angew. Chem. 1972, 84, 364. [Google Scholar] [CrossRef]
  252. Nawaz, N.; Abu Bakar, N.K.; Muhammad Ekramul Mahmud, H.N.; Jamaludin, N.S. Molecularly imprinted polymers-based DNA biosensors. Anal. Biochem. 2021, 630, 114328. [Google Scholar] [CrossRef]
  253. Pan, J.; Chen, W.; Ma, Y.; Pan, G. Molecularly imprinted polymers as receptor mimics for selective cell recognition. Chem. Soc. Rev. 2018, 47, 5574–5587. [Google Scholar] [CrossRef]
  254. Okutucu, B.; Onal, S. Molecularly imprinted polymers for separation of various sugars from human urine. Talanta 2011, 87, 74–79. [Google Scholar] [CrossRef]
  255. Culver, H.R.; Peppas, N.A. Protein-Imprinted Polymers: The Shape of Things to Come? Chem. Mater. 2017, 29, 5753–5761. [Google Scholar] [CrossRef] [PubMed]
  256. Babamiri, B.; Salimi, A.; Hallaj, R. A molecularly imprinted electrochemiluminescence sensor for ultrasensitive HIV-1 gene detection using EuS nanocrystals as luminophore. Biosens. Bioelectron. 2018, 117, 332–339. [Google Scholar] [CrossRef]
  257. Kadirsoy, S.; Atar, N.; Yola, M.L. Molecularly imprinted QCM sensor based on delaminated MXene for chlorpyrifos detection and QCM sensor validation. New J. Chem. 2020, 44, 6524–6532. [Google Scholar] [CrossRef]
  258. Beytur, M.; Kardaş, F.; Akyıldırım, O.; Özkan, A.; Bankoğlu, B.; Yüksek, H.; Yola, M.L.; Atar, N. A highly selective and sensitive voltammetric sensor with molecularly imprinted polymer based silver@gold nanoparticles/ionic liquid modified glassy carbon electrode for determination of ceftizoxime. J. Mol. Liq. 2018, 251, 212–217. [Google Scholar] [CrossRef]
  259. Jalalvand, A.R. Fabrication of a novel molecularly imprinted biosensor assisted by multi-way calibration for simultaneous determination of cholesterol and cholestanol in serum samples. Chemom. Intell. Lab. Syst. 2022, 226, 104587. [Google Scholar] [CrossRef]
  260. Zhang, X.; Yang, S.; Jiang, R.; Sun, L.; Pang, S.; Luo, A. Fluorescent molecularly imprinted membranes as biosensor for the detection of target protein. Sens. Actuators B Chem. 2018, 254, 1078–1086. [Google Scholar] [CrossRef]
  261. Jalalvand, A.R.; Pinto, L. A novel triple templates molecularly imprinted biosensor assisted by second-order calibration methods based on generation of second-order hydrodynamic linear sweep voltammetric data for simultaneous biosensing of insulin, proinsulin and C-peptide: Application to comparing PARAFAC2 and PARASIAS. Chemom. Intell. Lab. Syst. 2023, 233, 104746. [Google Scholar] [CrossRef]
  262. Ioniţă, M.; Vlăsceanu, G.M.; Watzlawek, A.A.; Voicu, S.I.; Burns, J.S.; Iovu, H. Graphene and functionalized graphene: Extraordinary prospects for nanobiocomposite materials. Compos. Part B Eng. 2017, 121, 34–57. [Google Scholar] [CrossRef]
  263. Serbanescu, O.S.; Pandele, A.M.; Miculescu, F.; Voicu, S.I. Synthesis and Characterization of Cellulose Acetate Membranes with Self-Indicating Properties by Changing the Membrane Surface Color for Separation of Gd(III). Coatings 2020, 10, 468. [Google Scholar] [CrossRef]
  264. Serbanescu, O.S.; Pandele, A.M.; Oprea, M.; Semenescu, A.; Thakur, V.K.; Voicu, S.I. Crown Ether-Immobilized Cellulose Acetate Membranes for the Retention of Gd (III). Polymers 2021, 13, 3978. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Polymeric membrane applications in the biomedical field.
Figure 1. Polymeric membrane applications in the biomedical field.
Polymers 15 00619 g001
Figure 2. Molecular structures of cellulose and some cellulose derivatives (reproduced with permission after [65]).
Figure 2. Molecular structures of cellulose and some cellulose derivatives (reproduced with permission after [65]).
Polymers 15 00619 g002
Figure 3. (A) Schematic illustration of the PCL/Gel membrane prepared by co-electrospinning technique; (B) SEM images of different PCL/Gel co-electrospinning membranes at various magnifications; and (C) the statistics of fiber size in different co-electrospinning membranes [123].
Figure 3. (A) Schematic illustration of the PCL/Gel membrane prepared by co-electrospinning technique; (B) SEM images of different PCL/Gel co-electrospinning membranes at various magnifications; and (C) the statistics of fiber size in different co-electrospinning membranes [123].
Polymers 15 00619 g003
Figure 4. (a) ECMO system composition diagram; (b) Schematic diagram of common membrane oxygenators; (c) hollow fiber membrane filament for gas and blood exchange; (d) Schematic diagram of blood oxygen exchange principle. (reproduced with permission after [26]).
Figure 4. (a) ECMO system composition diagram; (b) Schematic diagram of common membrane oxygenators; (c) hollow fiber membrane filament for gas and blood exchange; (d) Schematic diagram of blood oxygen exchange principle. (reproduced with permission after [26]).
Polymers 15 00619 g004
Figure 5. Poly(ε-caprolactone) (PCL) hollow fiber (HF) membrane bioreactor and scheme of the 3D human hepatic tissue realized by culturing human hepatocytes over and between PCL HF membranes parallel assembled at a distance of 250 µm, and endothelial cells compartmentalized in the lumen of the fibers. The cells were in communication through the porous wall of the membranes (reproduced with permission after [204]).
Figure 5. Poly(ε-caprolactone) (PCL) hollow fiber (HF) membrane bioreactor and scheme of the 3D human hepatic tissue realized by culturing human hepatocytes over and between PCL HF membranes parallel assembled at a distance of 250 µm, and endothelial cells compartmentalized in the lumen of the fibers. The cells were in communication through the porous wall of the membranes (reproduced with permission after [204]).
Polymers 15 00619 g005
Figure 6. Schematic of the biofunctional bone scaffold fabrication with LBL self-assembly on electrospun fiber membranes followed by apatite deposition (reproduced with permission after [230]).
Figure 6. Schematic of the biofunctional bone scaffold fabrication with LBL self-assembly on electrospun fiber membranes followed by apatite deposition (reproduced with permission after [230]).
Polymers 15 00619 g006
Figure 7. Schematic representation of the fabrication of regenerated cellulose nanofiber mats containing HAp and Ag NPs. Sequential steps are shown to describe the various process involved in the fabrication process (reproduced with permission after [235]).
Figure 7. Schematic representation of the fabrication of regenerated cellulose nanofiber mats containing HAp and Ag NPs. Sequential steps are shown to describe the various process involved in the fabrication process (reproduced with permission after [235]).
Polymers 15 00619 g007
Figure 8. Schematic representation of the reaction sequence for the derivatization of cellulose acetate membranes with resveratrol (reproduced with permission after [238]).
Figure 8. Schematic representation of the reaction sequence for the derivatization of cellulose acetate membranes with resveratrol (reproduced with permission after [238]).
Polymers 15 00619 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Radu, E.R.; Voicu, S.I.; Thakur, V.K. Polymeric Membranes for Biomedical Applications. Polymers 2023, 15, 619. https://doi.org/10.3390/polym15030619

AMA Style

Radu ER, Voicu SI, Thakur VK. Polymeric Membranes for Biomedical Applications. Polymers. 2023; 15(3):619. https://doi.org/10.3390/polym15030619

Chicago/Turabian Style

Radu, Elena Ruxandra, Stefan Ioan Voicu, and Vijay Kumar Thakur. 2023. "Polymeric Membranes for Biomedical Applications" Polymers 15, no. 3: 619. https://doi.org/10.3390/polym15030619

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

Article Metrics

Back to TopTop