Next Article in Journal
Removal Performance and Mechanism of Emerging Pollutant Chloroquine Phosphate from Water by Iron and Magnesium Co-Modified Rape Straw Biochar
Previous Article in Journal
Study of the Sensing Kinetics of G Protein-Coupled Estrogen Receptor Sensors for Common Estrogens and Estrogen Analogs
Previous Article in Special Issue
Fabrication of Effective Nanohybrids Based on Organic Species, Polyvinyl Alcohol and Carbon Nanotubes in Addition to Nanolayers for Removing Heavy Metals from Water under Severe Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrospun Nanofiber Membrane: An Efficient and Environmentally Friendly Material for the Removal of Metals and Dyes

1
College of Science, Gansu Agricultural University, Lanzhou 730070, China
2
Research Center of Gansu Military and Civilian Integration Advanced Structural Materials, Key Laboratory of Eco-Environment-Related Polymer Materials, Ministry of Education of China, Key Laboratory of Polymer Materials of Gansu Province, College of Chemistry and Chemical Engineering, Northwest Normal University, Lanzhou 730070, China
3
College of Food Science and Engineering, Gansu Agricultural University, Lanzhou 730070, China
4
State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(8), 3288; https://doi.org/10.3390/molecules28083288
Submission received: 2 March 2023 / Revised: 27 March 2023 / Accepted: 31 March 2023 / Published: 7 April 2023
(This article belongs to the Special Issue Polymer-Based Hybrid Nanomaterials for Water and Air Remediation)

Abstract

:
With the rapid development of nanotechnology, electrospun nanofiber membranes (ENM) application and preparation methods have attracted attention. With many advantages such as high specific surface area, obvious interconnected structure, and high porosity, ENM has been widely used in many fields, especially in water treatment, with more advantages. ENM solves the shortcomings of traditional means, such as low efficiency, high energy consumption, and difficulty in recycling, and it is suitable for recycling and treatment of industrial wastewater. This review begins with a description of electrospinning technology, describing the structure, preparation methods, and factors of common ENMs. At the same time, the removal of heavy metal ions and dyes by ENMs is introduced. The mechanism of ENM adsorption on heavy metal ions and dyes is chelation or electrostatic attraction, which has excellent adsorption and filtration ability for heavy metal ions and dyes, and the adsorption capacity of ENMs for heavy metal ions and dyes can be improved by increasing the metal chelation sites. Therefore, this technology and mechanism can be exploited to develop new, better, and more effective separation methods for the removal of harmful pollutants to cope with the gradually increasing water scarcity and pollution. Finally, it is hoped that this review will provide some guidance and direction for research on wastewater treatment and industrial production.

1. Introduction

In recent decades, mining, agriculture, chemical, and commodity production have severely degraded water resources. Water pollution comes mainly from the discharge of wastewater containing heavy metal ions from the mining and smelting industries, mainly from the extensive use of pesticides in industry and from the discharge of wastewater. Irreversible damage is caused to the environment. Heavy metal ions have become one of the many serious environmental problems. Heavy metal ions are commonly used in metal, leather tanning, metal manufacturing, and nuclear power plants [1,2,3,4,5]. Heavy metals are very difficult to biodegrade; instead, they can be concentrated up to several thousand times in the amplification of the food chain and end up in the human body. Heavy metals can interact strongly with proteins and enzymes in the human body, rendering them inactive. They can also accumulate in the body’s organs, leading to chronic poisoning. Mercury, arsenic, cadmium, hexavalent chromium (Cr), and other heavy metal ions can cause teratogenicity and are highly toxic. Even at low doses, they can pose a serious risk to human health because they accumulate in the body [6,7,8,9].
Dyes have been used for many years in the printing, leather, textile, and food industries [10,11,12,13,14]. Both natural and synthetic dyes can cause serious environmental pollution, and dye wastewater is a major pathway for many water pollution sources. Dyes in wastewater have complex and stable structures, large molecular weights, low biodegradability, and even low concentrations of dyes can have a significant impact on human health. Therefore, the removal of free dyes before the wastewater is discharged into natural waters is important for environmental protection [15,16]. Most dyes are highly toxic and readily soluble in water. Among them, aromatic amines, which are dye intermediates, have been classified as suspected carcinogens by many local authorities and are considered to be the most likely carcinogens in nature.
In the past, various methods have been used to remove heavy metal ions and organic dyes from water, such as membrane separation, reduction precipitation, ion exchange, coagulation–flocculation, adsorption, etc. [17,18,19,20]. Most removal methods are subject to stringent operating conditions and may also produce secondary contaminants with low efficiency. Among these technologies, adsorption is a superior physical and chemical method for the removal of organic dyes and heavy metal ions. To ensure the better adsorption of heavy metals and dyes, adsorbents should be prepared with a large surface area, high reusability, and strong interaction with heavy metal ions and organic dyes. For example, metal–organic skeleton compounds (MOFs) have been applied in adsorption due to their porosity properties, tunable structure, large specific surface area, and easy modification. Madhu N. Nimbalkar synthesized UiO-66-free carboxylic metal–organic skeleton using 1,2,4,5 phenyl tetracarboxylic acid, which has a high adsorption capacity for Pb2+ ions [21]. At the same time, graphene-based compounds have also been widely used in wastewater treatment. For example, regarding a GO/PVDF nanofiber membrane prepared by Fang-fang Ma, the adsorption capacity of MB was 621.1 mg g−1 [22]. At the same time, CNT [23] and MXene [24] have also been widely used in wastewater treatment. However, the preparation process of the above-mentioned powder adsorbents is complicated, the absorbents are difficult to recover, and the process involves easy to produce by-products, which cause secondary pollution. ENM has the advantages of a large specific surface area, high porosity, easy preparation, and no secondary contamination, which makes it very suitable for wastewater treatment [25].
Electrospinning technology mainly uses high power to stretch the polymer solution into a Taylor cone, and when the electric field force overcomes the surface tension, a charged jet can be injected from the top of the cone [26], which due to solvent evaporation and instability of the charged jet allows the solid nanofibers to be deposited randomly on the collector, forming a non-woven nanofiber membrane. Nanofiber membranes have the advantages of large volume ratios and ideal porosity and can be used for energy storage [27], water treatment [28], air filtration [29], ion exchange membranes [30], etc. The electrospinning technique has been widely used for the preparation of nanofiber membranes due to its advantages of mass producibility, speed, and simplicity [31]. Nanofiber membranes are now widely used for the treatment of pollutants in aqueous solutions due to their high specific surface area, easy surface modification, and porous structure [32,33]. This review references the work of researchers over the past five years, and a small amount of classic literature over a slightly longer period, on the removal of metals and dyes from wastewater by ENMs, which focuses on the structure, properties, influencing factors, and some of the working mechanisms of ENMs. The references are from a wide range of sources but mainly focus on nanofiber membrane materials and industrial applications and hopefully will provide some meaningful ideas for colleagues.

2. Preparation and Structure of Nanofiber Membrane

2.1. Electrospinning Technology

ENMs are used to produce fibrous membranes with quasi-3D diameters in the sub-micron or nano dimensions. Electrospinning is an aspecific type of fiber fabrication process that allows a polymer solution or melt to be jet-spun in a strong electric field, where the feedstock is transferred through the tip of a needle to a conical spherical tip (“Taylor cone”) and from the top of the cone, polymer filaments with nanoscale dimensions are produced [34,35], resulting in nanofiber membranes. The single-jet electrospinning device, often used in the laboratory, is shown in Figure 1 and consists of four components: a high-voltage power supply, a rotating liquid thruster, an ejector, and a grounded metal collector to collect the nanofibers [36,37]. The high-voltage power supply is available in the range of 0–50 kV, and the injector is primarily used to control the rate of injection of the rotating solution. In particular, small droplets can form on the spinnerets due to the polymer’s surface tension, and when the droplets are charged under the high voltage, the applied electric field is increased, and the counteracting electric field forces are correspondingly increased. When the electric field strength increases to a certain point, the electric field force is greater than the extension of the surface tension, and the deflected tip forms a Taylor cone and flies from the spinneret to the receiving hole as a jet stream [38].

2.2. Factors Affecting Electrospinning

Several factors can affect the outcome of electrospinning in the production of ENMs. The first is the effect caused by voltage, as high voltage helps to stretch the fiber and increase the surface charge density of the jet and reduce the fiber diameter; however, voltages that are too high can lead to gelation, which can reduce the efficiency of fiber collection [26]. Voltage also affects the diameter of the ENM, and according to the literature [39], as shown in Figure 2a, as the applied voltage increases, the pore diameter of the ENM first decreases and then increases. This means that if the applied voltage is too high or too low, it will affect the increase in the pore diameter of the obtained ENM and the stability of the injection. Secondly, the injection speed is also an important parameter that affects the electrospinning process. If the speed is too slow, the electrospinning time will be prolonged, but if the injection speed is too fast, it will lead to the formation of larger fiber diameters or large droplets in the nanofiber film, which will hinder the effectiveness of electrospinning. As shown in Figure 2b, the pore size of the ENM gradually decreases when the flow rate is increased. In addition, when the flow rate is too high or too low, the pore size distribution is less uniform, and at the same time, beads cannot be formed when the flow rate is too high.
Again, the distance between the needle and the receiver also affects electrospinning. If the distance between them is too short, solvent evaporation will be adversely affected, leading to bonding between fibers, increasing the fiber diameter, and reducing the fiber performance. If the spinning distance is too long, the spinning nozzle will gain more opportunities to split, and the diameter of the fibers obtained will be greatly reduced. As shown in Figure 3, when the distance is too short (100 mm), the fibers will melt, possibly due to a lack of solidification time before the fibers are ejected onto the collector [40], and when the distance increases, the diameter of the ENM increases and the fiber morphology shows a small non-beaded fiber mat. If the distance is too long (300 mm), the diameter increases due to the weak electrostatic force, and beads may form, resulting in the solvent being trapped within the fibers [41]. At the same time, relative humidity and temperature also affect electrospinning [42]. Lower relative humidity will accelerate the evaporation rate of the solvent in the jet, which helps to form finer fibers. The temperature has two opposite effects on average fiber diameter. Higher temperatures accelerate the evaporation rate of the solvent. Lower temperatures will reduce the viscosity of the solution and promote the formation of finer fibers. Finally, electrospinning technology is also affected by the concentration of the solution, as too high a concentration of the solution will prevent jetting, but too low a concentration of the solution will eject small droplets, and electrospinning can only be performed when the concentration of the polymer is within a suitable range.

2.3. Nanofiber Membrane Structure

In electrostatic technology, the structure of the ENM can be transformed by adjusting the voltage parameters or other conditions to meet the requirements of the experiment. In this section, several methods of preparing nanofibers with complex internal structures are presented [43,44,45,46,47,48,49]. Figure 4 is a schematic diagram of the structure of a common nanofiber membrane.

2.3.1. Core–Shell Structure Electrospray Nanofiber Membrane

Compared with ordinary electrospinning nanofibers, nanofibers with “core–shell” structures can synergistically exploit the advantages of both “core” and “shell” structures and exhibit better performance. According to the report by Jian Qiang Wang [50], the specific method for the preparation of a polyacrylonitrile/polypyrrole (PAN/PPy) core–shell structure is provided. Figure 5 is an SEM photograph of the prepared ENM, with Ppy(polypyrrole) well deposited on the front side of the ENM. Moreover, Xue Wu Huang reported [51] that flexible ultra-thin “core–shell” structures, and hydrophilic and conductive polymer nanofiber composites with unique PDA/ACNT “shell” and polymer nanofiber “core” structures were prepared by decorating polyurethane (PU) nanofibers with acidified carbon nanotubes (ACNTs) and then modified with polydopamine (PDA), which showed great potential in water purification. As reported by Fang-fang Ma [52], Ppy particles were deposited on polyvinylidene fluoride (PVDF) nanofibers using polydopamine (PDA) to produce new composite nanofibers with a similar structure of the “core–shell”. The morphological characteristics confirmed the electrospun PVDF/PDA nanofibers were uniformly encapsulated by Ppy particles, forming a typical core–shell structure.
In the core–shell structure of an ENM prepared by electrospinning, compounds such as PAN, PVDF, etc., are modified or coated on the surface of the ENM to form a functionalized nanofiber membrane. Under normal circumstances, the core part is mainly the supporting part, while the shell part is the functionalized part, and in most cases, the function of the ENM is added or changed, mainly by modifying the “shell”. In the ENM used for water purification (WP), the properties of the “shell” can be modified to improve the purification function of the ENM, providing a new idea for future wastewater treatment work, but in this regard, attention must be paid to the effect of non-uniform deposition of compounds on the surface of the nanofiber membrane.

2.3.2. Porous Nanofiber Membrane

According to a report by Faraz Khan Mahar et al. [53], PAN/poly(methyl methacrylate) (PMMA) nanofiber membranes were first prepared and then carbonized to convert the obtained nanofibers into p-CNF, which was then prepared to obtain porous nanofiber membranes. On the other hand, Zhang-Qi Feng et al. prepared porous polyacrylonitrile/graphene oxide nanofiber membranes [54] by adding PAN/GO solution to the N, N-dimethylformamide (DMF)/H2O solvent mixture to prepare porous polyacrylonitrile (PAN)/graphene oxide (GO) ENM by a simple one-step electrospinning method, the SEM of which is shown in Figure 6. From the current publicly reported research work, the prepared porous ENM has the following point: (1) the porous exterior can increase the specific surface area of the ENM, which can further improve the adsorption capacity of the ENM; (2) more importantly, the porous structure can well expose the compound loaded in the nanofiber to the water environment, so that the adsorption performance of the nanofiber membrane can be better reflected [55,56].

3. Applications of ENM Wastewater Treatment

A variety of ENMs are used in wastewater treatment. For example, nanofiltration membranes [57], nanofiltration ultrafiltration membranes [58], and microfiltration membranes [59] are used to remove wastewater. However, the above methods are accompanied by membrane pollution and high energy consumption. Nanofiber membranes as adsorbents have the advantages of no energy consumption and less membrane pollution; therefore, they are used for wastewater treatment. Here, two-dimensional materials (GO, CNC, etc.) have hydroxyl, carboxyl, amino, and other functional groups, which have high adsorption performance for heavy metal ions and dyes. For example, Senelisile M. summarized the application of nanocellulose-based nanofiber membranes containing carbon nanomaterials for dye wastewater treatment. Because carbon nanomaterials have favorable physical and chemical properties, such as strength, stiffness, green color, renewability, and sustainability, they are doped in the nanofiber membranes to improve the adsorption capacity of the membranes. At the same time, MOF-modified membranes showed excellent selectivity and permeability and minimal structural defects on the matrix [60]. In addition, MOFs can be used as dopants in heavy metal adsorption because of their easier controllability of pore size and structure. In this review, two-dimensional materials are added to nanofiber membranes to increase the adsorption sites on the membrane surface and improve the removal ability of the membrane to heavy metal ions and dyes in wastewater.

3.1. Heavy Metal Ion Removal

3.1.1. Heavy Metal Ion Adsorption

It is well known that some specific functional groups can form stable chemical bonds with heavy metal ions for wastewater treatment [13,55,61,62,63]. Amine, hydroxyl, and carboxyl groups can form complexes with metal ions for the adsorption of metal ions. For example, Sana Jamshidifard et al. [64] investigated the incorporation of UiO-66-NH2 MOF into PAN/CS and successfully prepared nanofiber membranes. The membrane showed strong adsorption performance for Pb (II), Cd (II), and Cr (VI) reaching 441.2, 415.6, and 372.6 mg·g−1, respectively. As shown in Figure 7, good adsorption performance was maintained after five adsorption cycles. Based on the adsorption kinetics, it was shown that the complexation of amine plays a signification role in the metal ion adsorption process, indicating that chemisorption was the main factor in the adsorption of the ENM. According to E. Salehi [65], polyethylene glycol (PEG) and amino-modified multi-walled carbon nanotubes (MWCNT-NH2) are used to modify the thin adsorption of chitosan/polyvinyl alcohol (CS/PVA) membranes, where copper ions and amine groups can form complexes for copper ion adsorption. The idea of the study was consistent with that reported by Sheng Deng et al. [66], where the polyether-acrylamide-modified multi-walled carbon nanotubes were co-spun with polyacrylonitrile trimers to prepare a composite nanofiber membrane with excellent adsorption of Cu2+ and Pb2+. The main mechanism is that the PEI molecule attached to the MWCNT contains primary, secondary, and tertiary amine groups in a ratio of about 1:2:130, and the adsorption of metal ions is due to the action of the hydroxyl group and amine group on the nanocomposite membrane [67]. The ion exchange between the metal ions and the hydroxyl ions reacted on Co2+ sites. According to Shuping Wu et al. [68], CS/PVP/PVA nanofiber membranes were synthesized by electrospinning; the maximum monolayer adsorption capacities of Cu (II), Ni (II), Cd (II), and Pb (II) on nanofiber membranes are 34.79, 25.24, 18.07, and 16.05 mg g−1, respectively. This is mainly due to the formation of chemical bonds between adsorbent and adsorbent, i.e., the chelation of amino and hydroxyl groups with heavy metal ions.
Chitosan (CS) has attracted great interest as an excellent adsorbent material with non-toxic and non-hazardous properties and has been extensively studied in the past. Chitosan (CS) is a natural polymer that can remove metal ions due to the presence of the amino group of the 2-amino-2-deoxy-d-glucose (glucosamine) unit [69,70]; mainly due to the possession of many amino groups and hydroxyl groups, functional groups can form stable complexes with heavy metal ions during adsorption [71,72]. According to Lei Li et al. [73], a layered CS nanofiber layer with an average diameter of 75 nm was successfully prepared by electrospinning using 5 wt % chitosan as the spinning solution, and then the nanofibers were crosslinked to remove Cr (VI) from water by static adsorption, and the maximum adsorption capacity of the nanofibers was 131.58 mg·g−1, which was twice that of chitosan powder. Lei Li also prepared a renewable spiral wound assembly of electrospun chitosan nanofiber membrane with a high affinity for the removal of Cr ions [74].
Electrostatic attraction is an important mechanism for the adsorption of heavy metal ions by NEM, and the pH of the solution in this process is the most important factor affecting the adsorption of heavy metal ions. The porous PAN/GO nanofibers with abundant nanopores [54,75] reached 382.5 ± 6.2 mg·g−1 for Cr (VI) adsorption, as shown in Figure 8a. The pH of the solution had a significant effect on the adsorption of ENM. The pH value of 3 reached the maximum adsorption, which gradually decreased as the pH value of the solution increased. The reasons for the change in adsorption capacity are as follows: the metal ion of Cr (VI) in the solution usually exists in three forms of chromate, namely, CrO42−, Cr2O72−, and HCrO4. As shown in Figure 8c, HCrO4 is the predominant form of Cr (VI) in an aqueous solution where the pH decreases and the deprotonated hydroxy nanofibers of PAN and GO sheets form a positively charged surface, which adsorbs the negatively charged HCrO4 by electrostatic attraction. However, as the zeta potential of PAN/GO nanofibers decreases with increasing pH, the adsorption capacity also decreases, so it can be concluded that the negative surface of porous PAN/GO nanofibers decreases the adsorption capacity with increasing pH due to electrostatic forces. In addition, according to Palaniswamy Suresh Kumar et al. [15], most heavy metal ions (such as Pb2+, Cd2+, and Cu2+) are in the divalent state in water at low pH values. Due to the low pH value, the mixed CNFs/TiO2-PAN was positively charged, resulting in a strong electrostatic repulsion between metal ions. As the pH value increases, the electrostatic attraction gradually increases, and the adsorption of metal ions increases. The adsorption capacity of ENM for heavy metal ions is listed in Table 1, and it can be seen that the pH of the solution has a significant effect on the adsorption of the ENM. Meanwhile, the adsorption capacity of other adsorbents for heavy metal ions is listed, as shown in Table 2.
Table 1. The adsorption performance of other nanofiber adsorbents for heavy metal ions.
Table 1. The adsorption performance of other nanofiber adsorbents for heavy metal ions.
Nanofiber AdsorbentPolymerSurface Area (m2·g−1)Heavy Metal IonsMaximum Adsorption Capacity (mg·g−1)pHT (°C)Sample Volume (mL)Kinetics ModelIsotherms ModelRef.
PDA/MnO2/PANPAN66Pb2+185.19625/Pseudo-second-order modelLangmuir isotherm model[76]
PVA/PEIPVA PEI/Cr6+15042510Pseudo-first-order
model
Langmuir isotherm model[77]
PAN/Fe2O3@Fe2O3PAN/Pb2+57/2550Pseudo-first-order modelLangmuir isotherm model[78]
PAN-CNT/TiO2-NH2PAN/Cr6+714220100Pseudo-first-order modelFreundlich isotherm model[79]
PA6@Mg (OH)2PA6/Cr6+294.632540Pseudo-second-order modelFreundlich
isotherm model
[80]
PVA@SiO2PVA370Cu2+489.12625100/Redlich-Peterson
isotherm model
[81]
CS-PGMA-PEICS/Cr6+138.9622510Pseudo-second-order modelLangmuir isotherm model[82]
EDTA-EDA-PANPAN/Cr6+6632525Pseudo-second-orderFreundlich isotherm model[83]
m-PEI/PVDFCCN/Cr6+109325/Pseudo-second-orderFreundlich isotherm model[84]
HMO-PANPAN/Pb2+19472520Pseudo- second- order modelFreundlich isotherm model[85]
CA/Fe3O4CA/Pb2+4462750Pseudo- second- order modelFreundlich isotherm model[86]
PA6/Fe3O4/o-MWCNTsPA6/Pb2+4962550//[87]
Lys-CNFsCNFs220Pb2+27062550Pseudo-second-order modelLangmuir isotherm model[88]
Thiol-functionalized celluloseCS/Pb2+2242550Pseudo-second-order
model
Langmuir isotherm model[89]
CS-DTPA/PEOCS/Pb2+142525/Pseudo-second-order modelLangmuir isotherm model[90]
Hal/Fe3O4/PEO/CSCS38Pb2+67725/Pseudo-second-order modelLangmuir isotherm model[91]
Palygorskite/chitinChNFs/Pb2+53.772550Pseudo-second-order modelFreundlich isotherm model[92]
MgAl-EDTA-LDH@PANPAN/Cu2+120.7525/Pseudo-second-order modelLangmuir isotherm model[93]
Note: “/” indicates no relevant data mentioned in the reference.
Table 2. The adsorption performance of other adsorbents for heavy metal ions.
Table 2. The adsorption performance of other adsorbents for heavy metal ions.
AdsorbentSurface
Area
Adsorption of Metal IonsMaximum Adsorption Capacity
(mg·g−1)
pHT
(°C)
Sample VolumeKinetics
Model
Isotherms
Model
Ref.
rGO/PEI-KOH/Cr6+3982252Pseudo-second-order modelLangmuir isotherm model[76]
Fe3O4@Arg-PPy NC22Cr6+322225/Pseudo-second-order model/[77]
PPy-rGO/Fe3O480Cr6+22633040Pseudo-second-order modelLangmuir isotherm model[79]
Hierarchical MnO2 microspheres252Pb2+139330100/Freundlich isotherm model[82]
NTA-β-CD-CS/Hg2+178.3625/Pseudo-second-orderLangmuir
isotherm model
[83]
C-phenylcalix pyrogallolarene/Cu2+8525///[94]
Sepiolite@polyethyleneimine/SA/Pb2+10945.52510Pseudo-second-order modelLangmuir model[95]
SA@PEI-CDs/Pb2+38042510Pseudo-second-order modelFreundlich isotherm model[96]
PAAO cryogels/Pb2+45052575Pseudo-second-order modelFreundlich isotherm model[97]
CS/PVP/PVA2.12Pb2+16/25/Pseudo-second-order modelLangmuir
isotherm model
[68]
Cellulose/chitosan/alginic acid hydrogels/Cu2+760/25/Pseudo-first-order modelFreundlich isotherm mode[98]
Nanocellulose/sodium alginate/carboxymethyl-chitosan284Pb2+47252550Pseudo-second-order modelLangmuir model[99]
Note: “/” indicates no relevant data mentioned in the reference.

3.1.2. Heavy Metal Ion Filtration

Membrane adsorption is mainly divided into static adsorption and dynamic adsorption, where dynamic adsorption of membranes is also an important process of membrane adsorption. According to the reports of Sana Jamshidifard et al. [64], the effect of the PAN/chitosan/UiO-66-NH2 MOF nanofiber layer on flux and metal ion removal was investigated. As shown in Figure 9a, the permeate flux decreased, and the metal ion removal increased when the thickness of the nanofiber layer was increased. As the thickness of the ENM layer is increased, the permeate flux of metal ions also decreases, mainly because there is a less aqueous solution passing through the membrane, thus reducing the diffusion of the aqueous solution through the membrane. In addition, the increase in metal ion removal with increasing ENM is due to the chemical reaction between the functional groups of the ENM (such as amines) and metal ions, which increases the available active sites for metal ion adsorption and can adsorb large amounts of metals on the surface of the ENM. In the work shown in Figure 9c,d, the performance of the ENM to remove heavy metals was investigated for 24 h. The water flux and metal ion removal rate remained constant for 18 h, after which the number of metal ions and the removal rate decreased significantly with adsorption saturation. Shahnaz Koush kbaghi et al. [100] prepared aminated Fe3O4 nanoparticles filled with chitosan/PVA/PES bilayer nanofiber membranes used for the treatment and adsorption of Pb (II) and Cr (VI) ions from aqueous solutions, and the results showed that the maximum adsorption capacities for Cr (VI) and Pb (II) ions reached 509.7 and 525.8 mg·g−1, respectively. During the treatment process, increasing the thickness of the ENM layer resulted in a slight decrease in water flux and an increase in metal ion recovery. The removal effect of metal ions increased with the increase in the initial concentration of heavy metal ions, while the water flux did not change significantly. Figure 9e,f show the adsorption results of ENM during water treatment repeated for three cycles, which demonstrates the good utility of the synthesized ENM as an adsorbent membrane in industry. In addition, according to Mohammad Pishnamazi et al. [101], UiO-66-NH2 and ZIF-8 metal–organic framework nanoparticles (NMOFs) were incorporated into polyvinylidene fluoride (PVDF) for the separation of Cr(VI) ions by ultrafiltration membranes to investigate the synthesized PVDF/NMOFs monolayer and PVDF/chitosan/NMOFs bilayer nanofiber membranes. The results showed that the maximum flux of PVDF/chitosan nanofiber membrane containing 20 wt % UiO-66-NH2 was 470 L·m−2·h−1, and the Cr (VI) rejection rate was 95.6% and the adsorption capacity of the prepared nanofiber membrane for Cr(VI) was 602.3 mg·g−1 after five adsorption–desorption cycles. This indicates that PVDF/chitosan/UiO-66-NH2 nanofibers have a high potential in membrane separation and Cr (VI) ion adsorption. In conclusion, the thickness of the ENM nanofiber layer affects the adsorption and permeation flux of Cr (VI). As the thickness of the ENM layer increases, the chelating sites for metal ions increase, and a large number of metal ions can be removed simultaneously, and the removal rate of metal ions increases.

3.1.3. Summary of Heavy Metal Removal

The main reason for the adsorption of heavy metal ions by ENM is the chelation reaction between amino, hydroxyl, and other functional groups and metal ions. According to the result of adsorption kinetics, the most important reason influencing adsorption is chemisorption. According to this line of research, the adsorption capacity of ENM for heavy metal ions can be improved by increasing the metal chelating sites [102,103], which we believe can be implemented in the following ways: (1) doping more amino compounds into the electrospinning solution to improve the removal efficiency, but the amount of the doped compounds should be appropriate, as too much compound will lead to unsuccessful preparation of ENM or aggregation on ENM, thus reducing the adsorption efficiency of heavy metal ions; (2) surface chemical treatment of the surface of carbon nanotubes or graphene-based compounds. The surface chemical treatment uses chemical methods to introduce amino groups and other groups on the surface of the compound, mainly to increase the number of amino groups so that more amino groups react with metal ions and improve the removal efficiency of ENM. (3) ENM was prepared by using natural polymers, such as chitosan, which is an excellent environmentally friendly wastewater treatment material because it is non-toxic and non-polluting. (4) The prepared nanofiber membranes were chemically modified by introducing other groups and amino groups on their surface to promote the removal efficiency and hydrophilic appearance of ENM.
The adsorption of heavy metal ions by ENM is currently a new way to remove heavy metal ions from water, but the desorption process of heavy metal ions and its influencing factors are also worthy of great attention because it helps to realize the recycling of heavy metal ions. The desorption process of heavy metal ions by changing the temperature and adjusting the pH value of the solution provides a new idea for the recycling of heavy metal ions in the environment, where the question worthy of our consideration is how to improve the desorption recovery of nanofiber membranes.
Overall, the preparation of ENM provides a technical option for the traditional heavy metal ion adsorption, but it must be pointed out that the preparation of ENM has a direct impact on the removal rate of metal ions, and researchers need to conduct in-depth studies to improve it. Certainly, dynamic adsorption is an important factor among many issues to be considered to improve the adsorption efficiency, the recyclability of ENM used as adsorbents, and the stability of ligands on the fiber membrane surface; the adsorption isotherms, kinetics, and thermodynamic properties of nanofibers also deserve further discussion.

3.2. Dye Removal

3.2.1. Adsorption of Dyes

Dyes and pigments are widely used in textile, leather, paper, cosmetics, plastics, printing, and other industries. These industries discharge non-biodegradable and complex organic pollutants into the water environment, posing a major threat to human health and aquatic organisms. Among them, Congo red (CR) is a synthetic antifouling agent, containing toxic aniline (aromatic amine) compounds that are suspected human carcinogens and mutagens [13,14,104]. As reported by S. Patel [105], polyacrylonitrile (PAN) nanofiber membranes synthesized by electrospinning were chemically modified with different amino (-NH2) functional groups on the surface and used as a novel nanosorbent for the removal of anionic Congo red dyes from aqueous media. They designed three chemical modus used to synthesize three different -NH2 functionalized nanofiber membranes, named PAN-NH2, PAN-CONH2, and PAN-EDA, whose structures are shown in Figure 10a, and the three different single bond NH2-functionalized PAN ENMs obtained were applied as adsorbents for the removal of toxic anionic CR dyes from aqueous media; the experimental results are shown in Figure 10b. It has been shown that the pH of the solution affects the removal of ENM, and at lower pH, the amino-functional groups present on the exterior of the functionalized nanofibers protonate, increasing the affinity for the removal of anionic CR dyes by creating a positively charged surface. However, at higher pH, the removal efficiency decreases due to the deprotonation of the amine group [106], which increases the negative charge on the surface of the functionalized nanofibers, causing electrostatic repulsion. According to F.F. Ma et al. [22], polyvinylidene fluoride (PVDF)/graphene oxide (GO) composite films were prepared by depositing graphene oxide on a polyvinylidene fluoride film using ultrasound, and the adsorption capacity of PVDF/GO composite films on methylene blue (MB) was investigated. The results showed that the adsorption ability of MB achieved 621.1 mg·g−1. As reported by F. F. Ma et al. [52], polydopamine (PDA) was used to modify electrospun PVDF nanofibers, and Ppy particles were deposited on the successfully modified nanofiber membranes to obtain composite nanofibers with a “core–shell” structure, and because it contains many nitrogen-containing groups, the surface roughness of the nanofibers was increased, and the hydrophilicity of the ENM appearance was greatly improved. They use Congo red (CR) and methylene blue (MB) to study the adsorption ability of ENM, and the results showed that the ENM had excellent adsorption capacity for both anionic and cationic dyes, with the maximum adsorption capacities of CR and MB being 384.6 and 370.4 mg·g−1, respectively. Z.M. Shourijeh prepared porous aminated PAN/PVDF composite nanofibers for the removal of Red 23 [107], and the dye removal capacity of the nanofibers was improved by extraction with NaHCO3 and modified diethylene triamine, and the removal of Red 23 was more than 95%.
The removal efficiency of ENMs is closely related to the pH of the solution, mainly because pH affects the surface charge of the ENM, the degree of ionization of the dye, and the functional groups on the active site of the ENM [108]. Decreasing the pH increases the number of protons, which leads to the protonation of amino groups on the surface of the nanofiber membrane and the formation of NH3+ groups, and it improves the electrostatic attraction between the negatively charged dye and the positively charged ENM [109]. According to W.J. Chen et al. [12], a new economical multifunctional cellulose acetate was obtained by carboxymethylation, deacetylation, and a polydopamine (PDA) coating to remove both anionic and cationic dyes, and a schematic diagram of the preparation is shown in Figure 10c. The adsorption performance of the PDA@DCA-COOH ENM was determined to be 67.31 mg·g−1 for CR and 69.89 mg·g−1 for MB. The results show that the adsorption of the dyes by the ENM is related to the charge on the membrane surface and the pH of the solution. For MB, it is a typical cationic dye, and the outer layer of the ENM becomes negatively charged, increasing the electrostatic attraction between the ENM and MB, so the dye will be adsorbed under alkaline conditions, and therefore the amount of MB removed increases with the increase in pH. For Congo red, which is a common anionic dye, the adsorbent has a moderate removal rate for it under acidic conditions. At lower pH, the positively charged protonated amino groups (single bond NH3+) on PDA@DCA-COOH increase, and the electrostatic attraction between the adsorbent and CR increases. Therefore, PDA@DCA-COOH has excellent recovery and adsorption performance for both anionic and cationic dyes, and it is expected to be a new compatible and efficient adsorbent in wastewater treatment.
We have compiled some research works on nanofiber membranes and listed the data for dye adsorption in wastewater treatment in Table 3 and Table 4 for comparison, which we hope will be a reference for readers.
Table 3. The adsorption performance of other nanofiber adsorbents for dyes.
Table 3. The adsorption performance of other nanofiber adsorbents for dyes.
Nanofiber AdsorbentPolymerSurface Area (m2·g−1)DyesMaximum Adsorption Capacity
(mg·g−1)
pHT (°C)Sample VolumeKinetics
Model
Isotherms
Model
Ref.
PPy/PANiPANi/CR250625150Pseudo-second-order modelLangmuir isotherm model[110]
Gelatin/calcium AlginateSA/MB204642525Pseudo-second-order modeLangmuir isotherm model[16]
GO/CSCS/MB58452550Pseudo-second-order
model
Freundlich isotherm mode[111]
Carboxylated Mn2O3Carboxylated148MB1175925400Pseudo-second-order modelLangmuir isotherm model[112]
Functionalized PANPAN/MG200825/Pseudo-second-order kinetic model.Langmuir isotherm
model
[113]
PVP/aluminaPVP417MO3515.52530Pseudo-second-order modelLangmuir isotherm model[114]
PAN-CNTPAN/MG881025/Pseudo-second-order model./[115]
Sodium alginate/polyvinyl alcoholPVA/MB9/25///[116]
Vinyl-modified mesoporous poly(acrylic acid)/SiO2PAA523MG220 3050Pseudo-second-order modelFreundlich isotherm model[117]
NiFe LDH/PAN/GOPAN/RB6.1962540Pseudo-first-order modelLangmuir isotherm model[42]
Mesoporous carbonPVP1642MC5673/10Pseudo-second-order kinetic modelLangmuir adsorption isotherm
model
[118]
ZnO-HT-PAN_HPAN/RB267/2515Pseudo-first-order modelLangmuir isotherm
model
[119]
PVDF/PDAPVDF/MB173/2550Pseudo-second-order modelLangmuir isotherm model[120]
Note: “/” indicates no relevant data mentioned in the reference.
Table 4. The adsorption performance of other adsorbents for dyes.
Table 4. The adsorption performance of other adsorbents for dyes.
AdsorbentSurface Area (m2·g−1)DyesMaximum Adsorption Capacity (m2·g−1)pHT (°C)Sample VolumeKinetics ModelIsotherms ModelRef.
NTA-β-CD-CS/MB16262525Pseudo-second-order modeLangmuir isotherm model[89]
T-QT/CS68.4MB917925/Pseudo-second-order modelLangmuir isotherm model[121]
CS2/CMC2-PEG/CR1053/2520Pseudo-second-order modelFreundlich isotherm model[122]
Pectin/graphene oxide aerogel/RhB419/25100Pseudo-first-order modelFreundlich isotherm model[123]
Benzenesulfonyl hydrazone modified guar gum19CR1065102510Pseudo-second-order modelLangmuir isotherm model[124]
Barberry stem powder/RR 19527///Pseudo-first-order modelLangmuir and Freundlich isotherms model[125]
α-Fe2O3 nanoparticles165RR 19520///Pseudo-second-order modelLangmuir isotherm model[126]
Ca-alginate/citric acid (CA)-sawdust/UiO-66-NH2 hydrogel beads15MB2562510Pseudo-second-order modelFreundlich isotherm model[127]
Note: “/” indicates no relevant data mentioned in the reference.

3.2.2. Summary of Dye Removal

In summary, the adsorption of dyes on nanofiber membranes is mainly physical, with electrostatic attraction being the dominant force. This is because the pH of the solution affects the degree of ionization of the dye and the charge on the surface of the ENM. In our aforementioned work, hydrophilic groups were added to the surface of the ENM by modular assembly or chemical modification to increase the hydrophilicity of the ENM. The protonated amino groups exhibit strong electrostatic attraction to anionic dyes. This affinity improves the efficiency of dye adsorption. By increasing the number of functional groups (such as hydroxyl or carboxyl) [128,129], the prepared nanofiber membrane can absorb anionic and cationic dyes simultaneously, further expanding the practicality of the nanofiber membrane. According to the experiment of Fang-fang Ma et al. [52], the prepared PVDF@PDA@PPY composite nanofiber membranes can adsorb not only cationic dyes and anionic dyes but also adsorb heavy metal chromium ions. However, how to further improve the adsorption efficiency of using ENM in practical applications is a question we need to consider in the future.

3.3. Factors Affecting Performance

In addition to the pH of the solution affecting the nanofibrous membrane, the adsorbent loading also affects the adsorption of the nanofibrous membrane. For example, according to Katherine E. Greenstein et al., who prepared nanofiber membranes with α-Fe2O3 [78], it was found experimentally that as the content of α-Fe2O3 increased, the removal efficiency of the nanofiber membranes for heavy metal ions increased, mainly attributed to the increase in adsorption sites, resulting in more heavy metal ions being adsorbed onto the surface of the nanofiber membranes. In addition, the initial concentration of the solution affects the adsorption of heavy metal ions by the membrane. According to Zhang-Qi Feng et al. [54], the adsorption of chromium ions by the prepared porous PAN/GO nanofiber membranes gradually increased in capacity with increasing initial concentration, but as the adsorption sites saturated, the adsorption gradually reached saturation with increasing initial concentration of the solution. Coexisting examples also affect the adsorption of heavy metal ions by membranes [77]. Of the commonly co-occurring anions, only sulphate and nitrate reduced the removal of Cr(VI) by the nanofibers but did not affect the structure of the adsorbed Cr. Copper can compete with Cr(VI) for complexation with amine groups, thus attenuating the reduction of Cr(VI). In addition to this, factors such as contact time and temperature of the solution can also affect the adsorption of heavy metal ions by the nanofiber membrane.
The contact time of the dye also affects the adsorption of the dye by the nanofiber membrane. According to Fang-fang Ma et al. [22], the adsorption of MB by the prepared GO/PVDF increased with time, but the adsorption capacity of the adsorption gradually reached a plateau when the adsorption sites on the membrane surface reached saturation. Additionally, the loading amount affects the adsorption of dyes by nanofiber membranes. For example, the PVDF/PDA composite membrane [120] prepared by Fang-fang Ma et al. showed the best adsorption of MB at a loading concentration of 30 for PDA mainly because at too high a loading, too much loading on the membrane surface tends to collect on the surface of the membrane, resulting in a lower adsorption capacity. In addition to this, temperature, initial concentration, etc., also affect the adsorption of dyes by the membrane.

3.4. The More Advantageous ENMs

The high specific surface area and adjustable functionality of nanofiber membranes make them much more effective than conventional membranes for surface adsorption treatment of wastewater, with high porosity (typically around 80%) and fully interconnected pores with controlled pore size distribution from micron to submicron, which also makes them very suitable for a wide range of filtration applications [130,131]. The mechanisms of metal ion and dye removal using electrostatic spinning membrane technology mainly include physical adsorption and chemisorption.
Physical adsorption is mainly caused by the electrostatic or intermolecular forces between the adsorbent and the metal ions. For the adsorption of ionic pollutants, physical adsorption, and ion exchange adsorption methods are mainly used. Physical adsorption relies mainly on electrostatic and intermolecular forces between the adsorbate and the anion. Ion exchange relies on the adsorbent releasing ions while adsorbing the same number of ions, thus achieving the remobilization of the contaminant in the water. Common materials such as GO, polydopamine (PDA), β-cyclodextrin (β-CD), and magnetic nanoparticles have been widely used to remove metals and dyes from water [132,133,134].
Chemisorption is an effective method for the adsorption of nonionic dye contaminants. It relies mainly on the formation of strong chemical bonds or surface complexes with metals or dyes by functional groups on the adsorbent material. Polydopamine (PDA), polyvinyl fluoride (PVDF), GO [22,52,135], etc., can form complexes with metal ions and dyes through their special structural properties, thus increasing the adsorption capacity.
In summary, the specific surface area and chelating ability of the adsorbent are the main determinants of physical and chemical adsorption, respectively; therefore, loading physical adsorbent materials onto the surface of ENM to effectively increase the specific surface area of the adsorbent material can achieve efficient adsorption. Compared with physical adsorption, chemisorption mainly relies on the formation of stable chemical bonds or redox reactions between adsorbent materials and ions to achieve water treatment, and according to this principle, excellent adsorbent materials can be obtained by grafting functional groups directly on the surface of spun-bond polymers. From the current research, among them, PDA and GO are used for ENMs to improve their adsorption performance, which is a more mature and effective method.

4. Summary and Outlook

Water pollution caused by heavy metals and dyes has become one of the major environmental problems of the 21st century. In wastewater treatment technology, ENMs have attracted great interest due to their good permeability, stability, easy handling, and low energy consumption [75,136,137]. In the past, innovative developments in electrospinning equipment and technology have enabled the functionalization of nanofibers for electrospray and multi-jet electrospinning [138]. At the same time, electrospinning technology is widely used in the preparation of ENMs due to their high porosity, ease of preparation, and high specific surface area [139]. Nanofiber membranes are also well used in industrial production for water treatment [11,140]. In this review, we report the latest research on the removal of heavy metals and dyes from water by nanocomposite fiber membranes. Furthermore, new materials and theories for the removal of heavy metal ions and dyes by ENM adsorption are analyzed, showing that these electrospun nanofiber membranes exhibit excellent performance in the treatment of heavy metal ions and dyes in water, proving their potential application in wastewater treatment.
Electrostatic spinning offers a simple and versatile method for the preparation of membranes for water treatment, and despite considerable progress, there is still a need to improve efficiency and flux rates by further optimizing the composition, structure, and physicochemical properties of nanofiber membranes, and future work should focus on a comprehensive understanding of the active adsorption sites exposed to the fiber surface, the mechanisms of desorption of ions/molecules, and the transport kinetics in the interconnected pores [141]. Despite these efforts, there are challenges to the further expansion and growth of ENM applications. Firstly, due to the scaling behavior of the ENM after cycling, the actual flux will be much lower than the theoretical value; secondly, in the presence of multiple metal ions or high flux, functionalized or deposited nanoparticles on the surface of the nanofiber membrane may be detached from the nanofibers, and this possible transformation alters the water flux and the adsorption capacity of the ENM for heavy metal ions and dyes, resulting in a reduced removal capacity; finally, the synergy between the membrane material and the surface roughness has to be considered, as it may also affect the adsorption capacity of the coating for dyes and heavy metal ions. While surface functionalization of membranes can significantly improve their filtration performance, it typically involves multiple steps, making it less suitable for industrial product development. It should be used to achieve large-scale production by further simplifying the single manufacturing process. In addition, as the manufacture of some fibrous membranes requires the use of hazardous organic solvents as solvents for the working fluid, further methods for the environmentally friendly post-treatment of fibrous membranes should be sought. Furthermore, to make full use of clean energy and reduce costs, new fiber membranes should be developed that can make full use of sunlight for rapid water treatment (e.g., decontamination, disinfection, and desalination) to directly produce drinking water with high throughput and low cost, which is particularly important for developing countries.
The key issue is that the practical application of ENMs in wastewater treatment needs to be further promoted; at present, these membranes have made great progress in laboratory-scale applications [142], but their mass production in industrial applications is still limited. Although there are still many problems with ENM, such as the need to further improve mechanical properties and production efficiency, it is hoped that in the future, these difficulties will eventually be overcome with the advancement of nanofiber technology and the increased requirements for water environmental management.

Author Contributions

Conceptualization, L.L. and R.G.; methodology, L.L.; software, R.G.; validation, L.L., S.Z., W.G., R.G., and L.Z.; formal analysis, L.L.; investigation, L.L.; resources, L.L.; data curation, L.L.; writing—original draft preparation, L.L., W.G., S.Z., and R.G.; writing—review and editing, L.L., W.G., R.G., and L.Z.; revision, L.L., W.G. and R.G.; visualization, L.L.; supervision, L.L.; project administration, L.L. and S.Z.; funding acquisition, L.L. and R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of Gansu Province (Grant No. 21JR7RA806); the State Key Laboratory of Cryospheric Science, Northwest Institute of Eco-Environment and Resources, Chinese Academy Sciences (Grant No. SKLCS-OP-2021-03); the Education Department of Gansu Province: Higher Education Innovation Fund (Grant No. 2022A-062); and the Northwest Normal University Research Project, China (Grant No. NWNU-LKQN2022-07).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, R.; Liang, R.; Dai, T.; Chen, J.; Shuai, X.; Liu, C. Pectin-based adsorbents for heavy metal ions: A review. Trends Food Sci. Technol. 2019, 91, 319–329. [Google Scholar] [CrossRef]
  2. Babu Poudel, M.; Shin, M.; Joo Kim, H. Interface engineering of MIL-88 derived MnFe-LDH and MnFe2O3 on three-dimensional carbon nanofibers for the efficient adsorption of Cr(VI), Pb(II), and As(III) ions. Sep. Purif. Technol. 2022, 287, 120463. [Google Scholar] [CrossRef]
  3. Hmtshirazi, R.; Mohammadi, T.; Asadi, A.A.; Tofighy, M.A. Electrospun nanofiber affinity membranes for water treatment applications: A review. J. Water Process Eng. 2022, 47, 102795. [Google Scholar] [CrossRef]
  4. Jo, H.; Le, T.-H.; Lee, H.; Lee, J.; Kim, M.; Lee, S.; Chang, M.; Yoon, H. Macrocyclic ligand-embedded graphene-in-polymer nanofiber membranes for lithium ion recovery. Chem. Eng. J. 2023, 452, 139274. [Google Scholar] [CrossRef]
  5. Lam, T.N.; Ma, C.Y.; Hsiao, P.H.; Ko, W.C.; Huang, Y.J.; Lee, S.Y.; Jain, J.; Huang, E.W. Tunable Mechanical and Electrical Properties of Coaxial Electrospun Composite Nanofibers of P(VDF-TrFE) and P(VDF-TrFE-CTFE). Int. J. Mol. Sci. 2021, 22, 4639. [Google Scholar] [CrossRef]
  6. Clementino, M.; Shi, X.; Zhang, Z. Oxidative Stress and Metabolic Reprogramming in Cr(VI) Carcinogenesis. Curr. Opin. Toxicol. 2018, 8, 20–27. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, Q.Y.; Murphy, A.; Sun, H.; Costa, M. Molecular and epigenetic mechanisms of Cr(VI)-induced carcinogenesis. Toxicol. Appl. Pharmacol. 2019, 377, 114636. [Google Scholar] [CrossRef]
  8. Ge, X.; He, J.; Wang, L.; Zhao, L.; Wang, Y.; Wu, G.; Liu, W.; Shu, Y.; Gong, W.; Ma, X.L.; et al. Epigenetic alterations of CXCL5 in Cr(VI)-induced carcinogenesis. Sci. Total Environ. 2022, 838 Pt 1, 155713. [Google Scholar] [CrossRef]
  9. Kumar Shetty, M.; Karthik, K.V.; Patil, J.H.; Murthy Shekhar, S.; Desai, S.M.; Hiremath, P.G.; Ravishankar, R. Sorption studies of Cr (VI) ions from synthetic waste water using chitosan embedded in calcium alginate beads. Mater. Today Proc. 2023, 76, 1–7. [Google Scholar] [CrossRef]
  10. Sanchez, L.M.; Espinosa, E.; Mendoza Zelis, P.; Morcillo Martin, R.; De Haro Niza, J.; Rodriguez, A. Cellulose nanofibers/PVA blend polymeric beads containing in-situ prepared magnetic nanorods as dye pollutants adsorbents. Int. J. Biol. Macromol. 2022, 209 Pt A, 1211–1221. [Google Scholar] [CrossRef]
  11. Bai, X.; Fu, R.; Huang, W.; Zhao, Y.; Liu, B.; Lin, S.; Yan, B.; Yang, Q.; Chen, S. Efficient and recyclable ultra-thin diameter polyacrylonitrile nanofiber membrane: Selective adsorption of cationic dyes. Sep. Purif. Technol. 2022, 302, 122109. [Google Scholar] [CrossRef]
  12. Chen, W.; Ma, H.; Xing, B. Electrospinning of multifunctional cellulose acetate membrane and its adsorption properties for ionic dyes. Int. J. Biol. Macromol. 2020, 158, 1342–1351. [Google Scholar] [CrossRef] [PubMed]
  13. Georgouvelas, D.; Abdelhamid, H.N.; Li, J.; Edlund, U.; Mathew, A.P. All-cellulose functional membranes for water treatment: Adsorption of metal ions and catalytic decolorization of dyes. Carbohydr. Polym. 2021, 264, 118044. [Google Scholar] [CrossRef] [PubMed]
  14. Kadirvelu, K.; Fathima, N.N. Keratin functionalized electrospun PEI/PAN microfiltration system as a simple and sustainable approach for anionic dye removal. J. Environ. Chem. Eng. 2022, 10, 107791. [Google Scholar] [CrossRef]
  15. Kumar, P.S.; Venkatesh, K.; Gui, E.L.; Jayaraman, S.; Singh, G.; Arthanareeswaran, G. Electrospun carbon nanofibers/TiO2-PAN hybrid membranes for effective removal of metal ions and cationic dye. Environ. Nanotechnol. Monit. Manag. 2018, 10, 366–376. [Google Scholar] [CrossRef]
  16. Ma, Y.; Qi, P.; Ju, J.; Wang, Q.; Hao, L.; Wang, R.; Sui, K.; Tan, Y. Gelatin/alginate composite nanofiber membranes for effective and even adsorption of cationic dyes. Compos. Part B Eng. 2019, 162, 671–677. [Google Scholar] [CrossRef]
  17. Chauke, V.P.; Maity, A.; Chetty, A. High-performance towards removal of toxic hexavalent chromium from aqueous solution using graphene oxide-alpha cyclodextrin-polypyrrole nanocomposites. J. Mol. Liq. 2015, 211, 71–77. [Google Scholar] [CrossRef]
  18. O’Connell, D.W.; Birkinshaw, C.; O’Dwyer, T.F. Heavy metal adsorbents prepared from the modification of cellulose: A review. Bioresour. Technol. 2008, 99, 6709–6724. [Google Scholar] [CrossRef]
  19. Gradinaru, L.M.; Bercea, M.; Vlad, S.; Barbalata Mandru, M.; Drobota, M.; Aflori, M.; Ciobanu, R.C. Preparation and characterization of electrospun magnetic poly(ether urethane) nanocomposite mats: Relationships between the viscosity of the polymer solutions and the electrospinning ability. Polymer 2022, 256, 125186. [Google Scholar] [CrossRef]
  20. Zhao, X.; Wang, X.; Lou, T. Simultaneous adsorption for cationic and anionic dyes using chitosan/electrospun sodium alginate nanofiber composite sponges. Carbohydr. Polym. 2022, 276, 118728. [Google Scholar] [CrossRef]
  21. Nimbalkar, M.N.; Bhat, B.R. Simultaneous adsorption of methylene blue and heavy metals from water using Zr-MOF having free carboxylic group. J. Environ. Chem. Eng. 2021, 9, 106216. [Google Scholar] [CrossRef]
  22. Ma, F.-F.; Zhang, D.; Huang, T.; Zhang, N.; Wang, Y. Ultrasonication-assisted deposition of graphene oxide on electrospun poly(vinylidene fluoride) membrane and the adsorption behavior. Chem. Eng. J. 2019, 358, 1065–1073. [Google Scholar] [CrossRef]
  23. Hayati, B.; Maleki, A.; Najafi, F.; Gharibi, F.; McKay, G.; Gupta, V.K.; Harikaranahalli Puttaiah, S.; Marzban, N. Heavy metal adsorption using PAMAM/CNT nanocomposite from aqueous solution in batch and continuous fixed bed systems. Chem. Eng. J. 2018, 346, 258–270. [Google Scholar] [CrossRef]
  24. Cheng, Y.; Li, M.; Song, Y. Theoretical study of M2CO2 MXenes stability and adsorption properties for heavy metals ions removal from water. Comput. Mater. Sci. 2023, 220, 112042. [Google Scholar] [CrossRef]
  25. Huang, X.; Wang, R.; Jiao, T.; Zou, G.; Zhan, F.; Yin, J.; Zhang, L.; Zhou, J.; Peng, Q. Facile Preparation of Hierarchical AgNP-Loaded MXene/Fe(3)O(4)/Polymer Nanocomposites by Electrospinning with Enhanced Catalytic Performance for Wastewater Treatment. ACS Omega 2019, 4, 1897–1906. [Google Scholar] [CrossRef] [Green Version]
  26. 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]
  27. Kadam, V.V.; Wang, L.; Padhye, R. Electrospun nanofiber materials to filter air pollutants—A review. J. Ind. Text. 2016, 47, 2253–2280. [Google Scholar] [CrossRef]
  28. Zhang, L.; Wei, T.; Jiang, Z.; Liu, C.; Jiang, H.; Chang, J.; Sheng, L.; Zhou, Q.; Yuan, L.; Fan, Z. Electrostatic interaction in electrospun nanofibers: Double-layer carbon protection of CoFe2O4 nanosheets enabling ultralong-life and ultrahigh-rate lithium ion storage. Nano Energy 2018, 48, 238–247. [Google Scholar] [CrossRef]
  29. Phan, D.N.; Khan, M.Q.; Nguyen, N.T.; Phan, T.T.; Ullah, A.; Khatri, M.; Kien, N.N.; Kim, I.S. A review on the fabrication of several carbohydrate polymers into nanofibrous structures using electrospinning for removal of metal ions and dyes. Carbohydr. Polym. 2021, 252, 117175. [Google Scholar] [CrossRef] [PubMed]
  30. Akampumuza, O.; Gao, H.; Zhang, H.; Wu, D.; Qin, X.-H. Raising Nanofiber Output: The Progress, Mechanisms, Challenges, and Reasons for the Pursuit. Macromol. Mater. Eng. 2018, 303, 1700269. [Google Scholar] [CrossRef]
  31. Hou, Z.; Wen, Z.; Wang, D.; Wang, J.; Philippe François-Xavier, C.; Wintgens, T. Bipolar jet electrospinning bi-functional nanofibrous membrane for simultaneous and sequential filtration of Cd2+ and BPA from water: Competition and synergistic effect. Chem. Eng. J. 2018, 332, 118–130. [Google Scholar] [CrossRef]
  32. Yoon, K.; Kim, K.; Wang, X.; Fang, D.; Hsiao, B.S.; Chu, B. High flux ultrafiltration membranes based on electrospun nanofibrous PAN scaffolds and chitosan coating. Polymer 2006, 47, 2434–2441. [Google Scholar] [CrossRef]
  33. Bilad, M.R.; Westbroek, P.; Vankelecom, I.F.J. Assessment and optimization of electrospun nanofiber-membranes in a membrane bioreactor (MBR). J. Membr. Sci. 2011, 380, 181–191. [Google Scholar] [CrossRef]
  34. Cui, J.; Li, F.; Wang, Y.; Zhang, Q.; Ma, W.; Huang, C. Electrospun nanofiber membranes for wastewater treatment applications. Sep. Purif. Technol. 2020, 250, 117116. [Google Scholar] [CrossRef]
  35. Mishra, R.K.; Mishra, P.; Verma, K.; Mondal, A.; Chaudhary, R.G.; Abolhasani, M.M.; Loganathan, S. Electrospinning production of nanofibrous membranes. Environ. Chem. Lett. 2018, 17, 767–800. [Google Scholar] [CrossRef]
  36. Collins, R.T.; Jones, J.J.; Harris, M.T.; Basaran, O.A. Electrohydrodynamic tip streaming and emission of charged drops from liquid cones. Nat. Phys. 2007, 4, 149–154. [Google Scholar] [CrossRef]
  37. Wang, S.H.; Chang, F.C. Cu(II) and Au(III) recovery with electrospun lignosulfonate CO(2)-activated carbon fiber. Int. J. Biol. Macromol. 2022, 203, 505–514. [Google Scholar] [CrossRef]
  38. Zhang, L.; Aboagye, A.; Kelkar, A.; Lai, C.; Fong, H. A review: Carbon nanofibers from electrospun polyacrylonitrile and their applications. J. Mater. Sci. 2013, 49, 463–480. [Google Scholar] [CrossRef]
  39. Cheng, T.; Li, S.; Xu, L.; Ahmed, A. Controllable preparation and formation mechanism of nanofiber membranes with large pore sizes using a modified electrospinning. Mater. Des. 2019, 178, 107867. [Google Scholar] [CrossRef]
  40. Aijaz, M.O.; Karim, M.R.; Alharbi, H.F.; Alharthi, N.H. Novel optimised highly aligned electrospun PEI-PAN nanofiber mats with excellent wettability. Polymer 2019, 180, 121665. [Google Scholar] [CrossRef]
  41. Akiyama, G.; Matsuda, R.; Sato, H.; Hori, A.; Takata, M.; Kitagawa, S. Effect of functional groups in MIL-101 on water sorption behavior. Microporous Mesoporous Mater. 2012, 157, 89–93. [Google Scholar] [CrossRef]
  42. Khalili, R.; Sabzehmeidani, M.M.; Parvinnia, M.; Ghaedi, M. Removal of hexavalent chromium ions and mixture dyes by electrospun PAN/graphene oxide nanofiber decorated with bimetallic nickel–iron LDH. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100750. [Google Scholar] [CrossRef]
  43. Sabetzadeh, N.; Gharehaghaji, A.A.; Javanbakht, M. Porous PAN micro/nanofiber separators for enhanced lithium-ion battery performance. Solid State Ion. 2018, 325, 251–257. [Google Scholar] [CrossRef]
  44. He, M.; Jiang, H.; Wang, R.; Xie, Y.; Zhao, C. Fabrication of metronidazole loaded poly (epsilon-caprolactone)/zein core/shell nanofiber membranes via coaxial electrospinning for guided tissue regeneration. J. Colloid Interface Sci. 2017, 490, 270–278. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, M.; Ma, W.; Wu, S.; Tang, G.; Cui, J.; Zhang, Q.; Chen, F.; Xiong, R.; Huang, C. Electrospun frogspawn structured membrane for gravity-driven oil-water separation. J. Colloid Interface Sci. 2019, 547, 136–144. [Google Scholar] [CrossRef]
  46. Chen, Z.; Zhang, J.; Guo, Y.; Zhang, H.; Cao, J.; Xu, Q.; Wang, S.; Wang, B.; Liu, Z. Effects of various factors on the modification of carbon nanotubes with polyvinyl alcohol in supercritical CO2 and their application in electrospun fibers. Chem. Res. Chin. Univ. 2014, 30, 690–697. [Google Scholar] [CrossRef]
  47. Wang, R.; Hu, Q.-H.; Wang, Q.-Y.; Xiang, Y.-L.; Huang, S.-H.; Liu, Y.-Z.; Li, S.-Y.; Chen, Q.-L.; Zhou, Q.-H. Efficiently selective removal of Pb(II) by magnetic ion-imprinted membrane based on polyacrylonitrile electro-spun nanofibers. Sep. Purif. Technol. 2022, 284, 120280. [Google Scholar] [CrossRef]
  48. Wang, S.G.; Jiang, X.; Chen, P.C.; Yu, A.G.; Huang, X.J. Preparation of coaxial-electrospun poly[bis(p-methylphenoxy)]phosphazene nanofiber membrane for enzyme immobilization. Int. J. Mol. Sci. 2012, 13, 14136–14148. [Google Scholar] [CrossRef] [Green Version]
  49. Huang, Y.J.; Chen, Y.F.; Hsiao, P.H.; Lam, T.N.; Ko, W.C.; Luo, M.Y.; Chuang, W.T.; Su, C.J.; Chang, J.H.; Chung, C.F.; et al. In-Situ Synchrotron SAXS and WAXS Investigation on the Deformation of Single and Coaxial Electrospun P(VDF-TrFE)-Based Nanofibers. Int. J. Mol. Sci. 2021, 22, 12669. [Google Scholar] [CrossRef]
  50. Wang, J.; Pan, K.; He, Q.; Cao, B. Polyacrylonitrile/polypyrrole core/shell nanofiber mat for the removal of hexavalent chromium from aqueous solution. J. Hazard. Mater. 2013, 244–245, 121–129. [Google Scholar] [CrossRef] [PubMed]
  51. Huang, X.; Zhang, S.; Xiao, W.; Luo, J.; Li, B.; Wang, L.; Xue, H.; Gao, J. Flexible PDA@ACNTs decorated polymer nanofiber composite with superhydrophilicity and underwater superoleophobicity for efficient separation of oil-in-water emulsion. J. Membr. Sci. 2020, 614, 118500. [Google Scholar] [CrossRef]
  52. Ma, F.-f.; Zhang, D.; Zhang, N.; Huang, T.; Wang, Y. Polydopamine-assisted deposition of polypyrrole on electrospun poly(vinylidene fluoride) nanofibers for bidirectional removal of cation and anion dyes. Chem. Eng. J. 2018, 354, 432–444. [Google Scholar] [CrossRef]
  53. Mahar, F.K.; He, L.; Wei, K.; Mehdi, M.; Zhu, M.; Gu, J.; Zhang, K.; Khatri, Z.; Kim, I. Rapid adsorption of lead ions using porous carbon nanofibers. Chemosphere 2019, 225, 360–367. [Google Scholar] [CrossRef] [PubMed]
  54. Feng, Z.-Q.; Yuan, X.; Wang, T. Porous polyacrylonitrile/graphene oxide nanofibers designed for high efficient adsorption of chromium ions (VI) in aqueous solution. Chem. Eng. J. 2020, 392, 123730. [Google Scholar] [CrossRef]
  55. Huang, Z.-M.; Zhang, Y.Z.; Kotaki, M.; Ramakrishna, S. A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci. Technol. 2003, 63, 2223–2253. [Google Scholar] [CrossRef]
  56. Poudel, M.B.; Awasthi, G.P.; Kim, H.J. Novel insight into the adsorption of Cr(VI) and Pb(II) ions by MOF derived Co-Al layered double hydroxide @hematite nanorods on 3D porous carbon nanofiber network. Chem. Eng. J. 2021, 417, 129312. [Google Scholar] [CrossRef]
  57. Yang, L.; Liu, X.; Zhang, X.; Chen, T.; Ye, Z.; Rahaman, M.S. High performance nanocomposite nanofiltration membranes with polydopamine-modified cellulose nanocrystals for efficient dye/salt separation. Desalination 2022, 521, 115385. [Google Scholar] [CrossRef]
  58. Wang, Z.; Sahadevan, R.; Crandall, C.; Menkhaus, T.J.; Fong, H. Hot-pressed PAN/PVDF hybrid electrospun nanofiber membranes for ultrafiltration. J. Membr. Sci. 2020, 611, 118327. [Google Scholar] [CrossRef]
  59. Arribas, P.; García-Payo, M.C.; Khayet, M.; Gil, L. Heat-treated optimized polysulfone electrospun nanofibrous membranes for high performance wastewater microfiltration. Sep. Purif. Technol. 2019, 226, 323–336. [Google Scholar] [CrossRef]
  60. Li, B.; Xu, C.; Yu, D.; Qi, Z.; Wang, Y.; Peng, Y. Enhanced phosphate remediation of contaminated natural water by magnetic zeolitic imidazolate framework-8@engineering nanomaterials (ZIF8@ENMs). J. Colloid Interface Sci. 2022, 613, 71–83. [Google Scholar] [CrossRef]
  61. Li, L.; Liao, Q.; Hou, B.; He, C.; Liu, J.; Li, B.; Yu, M.; Liu, Y.; Lai, B.; Yang, B. Synchronous reduction and removal of hexavalent chromium from wastewater by modified magnetic chitosan beads. Sep. Purif. Technol. 2023, 304, 122363. [Google Scholar] [CrossRef]
  62. Sahoo, S.K.; Panigrahi, G.K.; Sahoo, J.K.; Pradhan, A.K.; Purohit, A.K.; Dhal, J.P. Electrospun magnetic polyacrylonitrile-GO hybrid nanofibers for removing Cr(VI) from water. J. Mol. Liq. 2021, 326, 115364. [Google Scholar] [CrossRef]
  63. Bao, S.; Zhang, C.; Luo, S.; Jiang, L.; Li, Q.; Kong, Y.; Cao, J. HMGA2 mediates Cr (VI)-induced metabolic reprogramming through binding to mitochondrial D-Loop region. Ecotoxicol. Environ. Saf. 2022, 244, 114085. [Google Scholar] [CrossRef] [PubMed]
  64. Jamshidifard, S.; Koushkbaghi, S.; Hosseini, S.; Rezaei, S.; Karamipour, A.; Jafari Rad, A.; Irani, M. Incorporation of UiO-66-NH2 MOF into the PAN/chitosan nanofibers for adsorption and membrane filtration of Pb(II), Cd(II) and Cr(VI) ions from aqueous solutions. J. Hazard. Mater. 2019, 368, 10–20. [Google Scholar] [CrossRef] [PubMed]
  65. Salehi, E.; Madaeni, S.S.; Rajabi, L.; Derakhshan, A.A.; Daraei, S.; Vatanpour, V. Static and dynamic adsorption of copper ions on chitosan/polyvinyl alcohol thin adsorptive membranes: Combined effect of polyethylene glycol and aminated multi-walled carbon nanotubes. Chem. Eng. J. 2013, 215–216, 791–801. [Google Scholar] [CrossRef]
  66. Deng, S.; Liu, X.; Liao, J.; Lin, H.; Liu, F. PEI modified multiwalled carbon nanotube as a novel additive in PAN nanofiber membrane for enhanced removal of heavy metal ions. Chem. Eng. J. 2019, 375, 122086. [Google Scholar] [CrossRef]
  67. Chen, Y.; Pan, B.; Li, H.; Zhang, W.; Lv, L.; Wu, J. Selective Removal of Cu(II) Ions by Using Cation-exchange Resin-Supported Polyethyleneimine (PEI) Nanoclusters. Environ. Sci. Technol. 2010, 44, 3508–3513. [Google Scholar] [CrossRef]
  68. Wu, S.; Li, K.; Shi, W.; Cai, J. Preparation and performance evaluation of chitosan/polyvinylpyrrolidone/polyvinyl alcohol electrospun nanofiber membrane for heavy metal ions and organic pollutants removal. Int. J. Biol. Macromol. 2022, 210, 76–84. [Google Scholar] [CrossRef]
  69. Zhao, R.; Wang, Y.; Li, X.; Sun, B.; Jiang, Z.; Wang, C. Water-insoluble sericin/beta-cyclodextrin/PVA composite electrospun nanofibers as effective adsorbents towards methylene blue. Colloids Surf. B Biointerfaces 2015, 136, 375–382. [Google Scholar] [CrossRef]
  70. Sağ, Y.s.; Aktay, Y. Kinetic studies on sorption of Cr(VI) and Cu(II) ions by chitin, chitosan and Rhizopus arrhizus. Biochem. Eng. J. 2002, 12, 142–152. [Google Scholar] [CrossRef]
  71. Habiba, U.; Afifi, A.M.; Salleh, A.; Ang, B.C. Chitosan/(polyvinyl alcohol)/zeolite electrospun composite nanofibrous membrane for adsorption of Cr(6+), Fe(3+) and Ni(2). J. Hazard. Mater. 2017, 322 Pt A, 182–194. [Google Scholar] [CrossRef]
  72. Luo, Y.; Hu, Z.; Lei, X.; Wang, Y.; Guo, X. Fluorescent magnetic chitosan-based hydrogel incorporating Amino-Functionalized Fe3O4 and cellulose nanofibers modified with carbon dots for adsorption and detection of Cr (VI). Colloids Surf. A Physicochem. Eng. Asp. 2023, 658, 130673. [Google Scholar] [CrossRef]
  73. Li, L.; Li, Y.; Cao, L.; Yang, C. Enhanced chromium (VI) adsorption using nanosized chitosan fibers tailored by electrospinning. Carbohydr. Polym. 2015, 125, 206–213. [Google Scholar] [CrossRef] [PubMed]
  74. Li, L.; Zhang, J.; Li, Y.; Yang, C. Removal of Cr (VI) with a spiral wound chitosan nanofiber membrane module via dead-end filtration. J. Membr. Sci. 2017, 544, 333–341. [Google Scholar] [CrossRef]
  75. Tamayo Marin, J.A.; Londono, S.R.; Delgado, J.; Navia Porras, D.P.; Valencia Zapata, M.E.; Mina Hernandez, J.H.; Valencia, C.H.; Grande Tovar, C.D. Biocompatible and Antimicrobial Electrospun Membranes Based on Nanocomposites of Chitosan/Poly (Vinyl Alcohol)/Graphene Oxide. Int. J. Mol. Sci. 2019, 20, 2987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Li, Y.; Zhao, R.; Chao, S.; Sun, B.; Wang, C.; Li, X. Polydopamine coating assisted synthesis of MnO2 loaded inorganic/organic composite electrospun fiber adsorbent for efficient removal of Pb2+ from water. Chem. Eng. J. 2018, 344, 277–289. [Google Scholar] [CrossRef]
  77. Zhang, S.; Shi, Q.; Korfiatis, G.; Christodoulatos, C.; Wang, H.; Meng, X. Chromate removal by electrospun PVA/PEI nanofibers: Adsorption, reduction, and effects of co-existing ions. Chem. Eng. J. 2020, 387, 124179. [Google Scholar] [CrossRef]
  78. Greenstein, K.E.; Myung, N.V.; Parkin, G.F.; Cwiertny, D.M. Performance comparison of hematite (alpha-Fe(2)O(3))-polymer composite and core-shell nanofibers as point-of-use filtration platforms for metal sequestration. Water Res. 2019, 148, 492–503. [Google Scholar] [CrossRef]
  79. Mohamed, A.; Nasser, W.S.; Osman, T.A.; Toprak, M.S.; Muhammed, M.; Uheida, A. Removal of chromium (VI) from aqueous solutions using surface modified composite nanofibers. J. Colloid. Interface Sci. 2017, 505, 682–691. [Google Scholar] [CrossRef]
  80. Jia, B.-B.; Wang, J.-N.; Wu, J.; Li, C.-J. “Flower-Like” PA6@Mg(OH)2 electrospun nanofibers with Cr (VI)-removal capacity. Chem. Eng. J. 2014, 254, 98–105. [Google Scholar] [CrossRef]
  81. Wu, S.; Li, F.; Wang, H.; Fu, L.; Zhang, B.; Li, G. Effects of poly (vinyl alcohol) (PVA) content on preparation of novel thiol-functionalized mesoporous PVA/SiO2 composite nanofiber membranes and their application for adsorption of heavy metal ions from aqueous solution. Polymer 2010, 51, 6203–6211. [Google Scholar] [CrossRef]
  82. Yang, D.; Li, L.; Chen, B.; Shi, S.; Nie, J.; Ma, G. Functionalized chitosan electrospun nanofiber membranes for heavy-metal removal. Polymer 2019, 163, 74–85. [Google Scholar] [CrossRef]
  83. Luo, X.; Lei, X.; Cai, N.; Xie, X.; Xue, Y.; Yu, F. Removal of Heavy Metal Ions from Water by Magnetic Cellulose-Based Beads with Embedded Chemically Modified Magnetite Nanoparticles and Activated Carbon. ACS Sustain. Chem. Eng. 2016, 4, 3960–3969. [Google Scholar] [CrossRef]
  84. Ma, Y.; Zhang, B.; Ma, H.; Yu, M.; Li, L.; Li, J. Electrospun nanofibrous polyethylenimine mat: A potential adsorbent for the removal of chromate and arsenate from drinking water. RSC Adv. 2016, 6, 30739–30746. [Google Scholar] [CrossRef]
  85. Yadav, P.; Farnood, R.; Kumar, V. HMO-incorporated electrospun nanofiber recyclable membranes: Characterization and adsorptive performance for Pb(II) and As(V). J. Environ. Chem. Eng. 2021, 9, 106507. [Google Scholar] [CrossRef]
  86. Shalaby, T.I.; El-Kady, M.F.; Zaki, A.E.H.M.; El-Kholy, S.M. Preparation and application of magnetite nanoparticles immobilized on cellulose acetate nanofibers for lead removal from polluted water. Water Supply 2016, 17, 176–187. [Google Scholar] [CrossRef]
  87. Bassyouni, D.; Mohamed, M.; El-Ashtoukhy, E.-S.; El-Latif, M.A.; Zaatout, A.; Hamad, H. Fabrication and characterization of electrospun Fe3O4/o-MWCNTs/polyamide 6 hybrid nanofibrous membrane composite as an efficient and recoverable adsorbent for removal of Pb (II). Microchem. J. 2019, 149, 103998. [Google Scholar] [CrossRef]
  88. Liang, C.; Zhao, L.; Qiao, L.; Du, K. Proteinaceous porous nanofiber membrane-type adsorbent derived from amyloid lysozyme protofilaments for highly efficient lead(II) biologic scavenging. J. Hazard. Mater. 2022, 425, 127886. [Google Scholar] [CrossRef]
  89. Choi, H.Y.; Bae, J.H.; Hasegawa, Y.; An, S.; Kim, I.S.; Lee, H.; Kim, M. Thiol-functionalized cellulose nanofiber membranes for the effective adsorption of heavy metal ions in water. Carbohydr. Polym. 2020, 234, 115881. [Google Scholar] [CrossRef]
  90. Surgutskaia, N.S.; Martino, A.D.; Zednik, J.; Ozaltin, K.; Lovecka, L.; Bergerova, E.D.; Kimmer, D.; Svoboda, J.; Sedlarik, V. Efficient Cu2+, Pb2+ and Ni2+ ion removal from wastewater using electrospun DTPA-modified chitosan/polyethylene oxide nanofibers. Sep. Purif. Technol. 2020, 247, 116914. [Google Scholar] [CrossRef]
  91. Li, L.; Wang, F.; Lv, Y.; Liu, J.; Zhang, D.; Shao, Z. Halloysite nanotubes and Fe3O4 nanoparticles enhanced adsorption removal of heavy metal using electrospun membranes. Appl. Clay Sci. 2018, 161, 225–234. [Google Scholar] [CrossRef]
  92. Mamah, S.C.; Goh, P.S.; Ismail, A.F.; Amin, M.A.M.; Ahmad, N.A.; Suzaimi, N.D.; Raji, Y.O. Facile preparation of palygorskite/chitin nanofibers hybrids nanomaterial with remarkable adsorption capacity. Mater. Sci. Eng. B 2020, 262, 114725. [Google Scholar] [CrossRef]
  93. Chen, H.; Lin, J.; Zhang, N.; Chen, L.; Zhong, S.; Wang, Y.; Zhang, W.; Ling, Q. Preparation of MgAl-EDTA-LDH based electrospun nanofiber membrane and its adsorption properties of copper(II) from wastewater. J. Hazard. Mater. 2018, 345, 1–9. [Google Scholar] [CrossRef] [PubMed]
  94. Jumina; Priastomo, Y.; Setiawan, H.R.; Mutmainah; Kurniawan, Y.S.; Ohto, K. Simultaneous removal of lead(II), chromium(III), and copper(II) heavy metal ions through an adsorption process using C-phenylcalix[4]pyrogallolarene material. J. Environ. Chem. Eng. 2020, 8, 103971. [Google Scholar] [CrossRef]
  95. Wu, X.; Song, Y.; Yin, P.; Xu, Q.; Yang, Z.; Xu, Y.; Liu, X.; Wang, Y.; Sun, W.; Cai, H. Fabrication of the composite sepiolite@polyethyleneimine/sodium alginate and its excellent adsorption performance for heavy metal ions. Appl. Clay Sci. 2022, 228, 106647. [Google Scholar] [CrossRef]
  96. Zhang, S.; Arkin, K.; Zheng, Y.; Ma, J.; Bei, Y.; Liu, D.; Shang, Q. Preparation of a composite material based on self-assembly of biomass carbon dots and sodium alginate hydrogel and its green, efficient and visual adsorption performance for Pb2+. J. Environ. Chem. Eng. 2022, 10, 106921. [Google Scholar] [CrossRef]
  97. Chen, J.; Liao, C.; Guo, X.-X.; Hou, S.-C.; He, W.-D. PAAO cryogels from amidoximated P(acrylic acid-co-acrylonitrile) for the adsorption of lead ion. Eur. Polym. J. 2022, 171, 111192. [Google Scholar] [CrossRef]
  98. Zhang, R.; Liu, B.; Ma, J.; Zhu, R. Preparation and characterization of carboxymethyl cellulose/chitosan/alginic acid hydrogels with adjustable pore structure for adsorption of heavy metal ions. Eur. Polym. J. 2022, 179, 111577. [Google Scholar] [CrossRef]
  99. Li, W.; Zhang, L.; Hu, D.; Yang, R.; Zhang, J.; Guan, Y.; Lv, F.; Gao, H. A mesoporous nanocellulose/sodium alginate/carboxymethyl-chitosan gel beads for efficient adsorption of Cu2+ and Pb2+. Int. J. Biol. Macromol. 2021, 187, 922–930. [Google Scholar] [CrossRef]
  100. Koushkbaghi, S.; Zakialamdari, A.; Pishnamazi, M.; Ramandi, H.F.; Aliabadi, M.; Irani, M. Aminated-Fe3O4 nanoparticles filled chitosan/PVA/PES dual layers nanofibrous membrane for the removal of Cr(VI) and Pb(II) ions from aqueous solutions in adsorption and membrane processes. Chem. Eng. J. 2018, 337, 169–182. [Google Scholar] [CrossRef]
  101. Mahmoodi, N.M.; Mokhtari-Shourijeh, Z.; Ghane-Karade, A. Dye removal from wastewater by the cross-linked blend nanofiber and homogenous surface diffusion modeling. Environ. Prog. Sustain. Energy 2017, 36, 1634–1642. [Google Scholar] [CrossRef]
  102. Miao, S.; Guo, J.; Deng, Z.; Yu, J.; Dai, Y. Adsorption and reduction of Cr(VI) in water by iron-based metal-organic frameworks (Fe-MOFs) composite electrospun nanofibrous membranes. J. Clean. Prod. 2022, 370, 133566. [Google Scholar] [CrossRef]
  103. Zhang, X.; Ru, Z.; Sun, Y.; Zhang, M.; Wang, J.; Ge, M.; Liu, H.; Wu, S.; Cao, C.; Ren, X.; et al. Recent advances in applications for air pollutants purification and perspectives of electrospun nanofibers. J. Clean. Prod. 2022, 378, 134567. [Google Scholar] [CrossRef]
  104. Mahapatra, A.; Mishra, B.G.; Hota, G. Adsorptive removal of Congo red dye from wastewater by mixed iron oxide–alumina nanocomposites. Ceram. Int. 2013, 39, 5443–5451. [Google Scholar] [CrossRef]
  105. Patel, S.; Hota, G. Synthesis of novel surface functionalized electrospun PAN nanofibers matrix for efficient adsorption of anionic CR dye from water. J. Environ. Chem. Eng. 2018, 6, 5301–5310. [Google Scholar] [CrossRef]
  106. Kumar, R.; Ansari, M.O.; Parveen, N.; Barakat, M.A.; Cho, M.H. Simple route for the generation of differently functionalized PVC@graphene–polyaniline fiber bundles for the removal of Congo red from wastewater. RSC Adv. 2015, 5, 61486–61494. [Google Scholar] [CrossRef]
  107. Mokhtari-Shourijeh, Z.; Langari, S.; Montazerghaem, L.; Mahmoodi, N.M. Synthesis of porous aminated PAN/PVDF composite nanofibers by electrospinning: Characterization and Direct Red 23 removal. J. Environ. Chem. Eng. 2020, 8, 103876. [Google Scholar] [CrossRef]
  108. Mahmoodi, N.M.; Mokhtari-Shourijeh, Z. Preparation of PVA-chitosan blend nanofiber and its dye removal ability from colored wastewater. Fibers Polym. 2015, 16, 1861–1869. [Google Scholar] [CrossRef]
  109. Mahmoodi, N.M.; Mokhtari-Shourijeh, Z. Modified poly(vinyl alcohol)-triethylenetetramine nanofiber by glutaraldehyde: Preparation and dye removal ability from wastewater. Desalination Water Treat. 2015, 57, 20076–20083. [Google Scholar] [CrossRef]
  110. Chafai, H.; Laabd, M.; Elbariji, S.; Bazzaoui, M.; Albourine, A. Study of congo red adsorption on the polyaniline and polypyrrole. J. Dispers. Sci. Technol. 2016, 38, 832–836. [Google Scholar] [CrossRef]
  111. Yu, R.; Shi, Y.; Yang, D.; Liu, Y.; Qu, J.; Yu, Z.-Z. Graphene oxide/chitosan aerogel microspheres with honeycomb cobweb and radially oriented microchannel structures for broad-spectrum and rapid adsorption of water contaminants. ACS Appl. Mater. Interfaces 2017, 9, 21809–21819. [Google Scholar] [CrossRef]
  112. Berenjian, A.; Maleknia, L.; Chizari Fard, G.; Almasian, A. Mesoporous carboxylated Mn2O3 nanofibers: Synthesis, characterization and dye removal property. J. Taiwan Inst. Chem. Eng. 2018, 86, 57–72. [Google Scholar] [CrossRef]
  113. Patel, S.; Hota, G. Adsorptive removal of malachite green dye by functionalized electrospun PAN nanofibers membrane. Fibers Polym. 2014, 15, 2272–2282. [Google Scholar] [CrossRef]
  114. Shen, J.; Li, Z.; Wu, Y.-n.; Zhang, B.; Li, F. Dendrimer-based preparation of mesoporous alumina nanofibers by electrospinning and their application in dye adsorption. Chem. Eng. J. 2015, 264, 48–55. [Google Scholar] [CrossRef]
  115. Yar, A.; Parlayici, e. Carbon nanotubes/polyacrylonitrile composite nanofiber mats for highly efficient dye adsorption. Colloids Surf. A Physicochem. Eng. Asp. 2022, 651, 129703. [Google Scholar] [CrossRef]
  116. Jiang, X.; Sang, C.; Wang, J.; Guo, J. Preparation of sodium alginate/polyvinyl alcohol composite nanofiber membranes for adsorption of dyes. Text. Res. J. 2021, 92, 3154–3163. [Google Scholar] [CrossRef]
  117. Xu, R.; Jia, M.; Zhang, Y.; Li, F. Sorption of malachite green on vinyl-modified mesoporous poly(acrylic acid)/SiO2 composite nanofiber membranes. Microporous Mesoporous Mater. 2012, 149, 111–118. [Google Scholar] [CrossRef]
  118. Teng, M.; Qiao, J.; Li, F.; Bera, P.K. Electrospun mesoporous carbon nanofibers produced from phenolic resin and their use in the adsorption of large dye molecules. Carbon 2012, 50, 2877–2886. [Google Scholar] [CrossRef]
  119. Phan, D.-N.; Rebia, R.A.; Saito, Y.; Kharaghani, D.; Khatri, M.; Tanaka, T.; Lee, H.; Kim, I.-S. Zinc oxide nanoparticles attached to polyacrylonitrile nanofibers with hinokitiol as gluing agent for synergistic antibacterial activities and effective dye removal. J. Ind. Eng. Chem. 2020, 85, 258–268. [Google Scholar] [CrossRef]
  120. Ma, F.-f.; Zhang, N.; Wei, X.; Yang, J.-h.; Wang, Y.; Zhou, Z.-w. Blend-electrospun poly(vinylidene fluoride)/polydopamine membranes: Self-polymerization of dopamine and the excellent adsorption/separation abilities. J. Mater. Chem. A 2017, 5, 14430–14443. [Google Scholar] [CrossRef]
  121. Zhong, Z.-R.; Jiang, H.-L.; Shi, N.; Lv, H.-W.; Liu, Z.-J.; He, F.-A. A novel tetrafluoroterephthalonitrile-crosslinked quercetin/chitosan adsorbent and its adsorption properties for dyes. J. Mol. Struct. 2023, 1282, 135150. [Google Scholar] [CrossRef]
  122. Zhu, H.; Chen, S.; Duan, H.; He, J.; Luo, Y. Removal of anionic and cationic dyes using porous chitosan/carboxymethyl cellulose-PEG hydrogels: Optimization, adsorption kinetics, isotherm and thermodynamics studies. Int. J. Biol. Macromol. 2023, 231, 123213. [Google Scholar] [CrossRef] [PubMed]
  123. He, Z.; Qin, M.; Han, C.; Bai, X.; Wu, Y.; Yao, D.; Zheng, Y. Pectin/graphene oxide aerogel with bamboo-like structure for enhanced dyes adsorption. Colloids Surf. A Physicochem. Eng. Asp. 2022, 652, 129837. [Google Scholar] [CrossRef]
  124. Wen, Y.; Xie, Z.; Xue, S.; Long, J.; Shi, W.; Liu, Y. Preparation of benzenesulfonyl hydrazone modified guar gum and its adsorption properties for dyes and phytotoxicity assays. Int. J. Biol. Macromol. 2023, 234, 123700. [Google Scholar] [CrossRef]
  125. Kamranifar, M.; Khodadadi, M.; Samiei, V.; Dehdashti, B.; Noori Sepehr, M.; Rafati, L.; Nasseh, N. Comparison the removal of reactive red 195 dye using powder and ash of barberry stem as a low cost adsorbent from aqueous solutions: Isotherm and kinetic study. J. Mol. Liq. 2018, 255, 572–577. [Google Scholar] [CrossRef]
  126. Nassar, M.Y.; Ahmed, I.S.; Mohamed, T.Y.; Khatab, M. A controlled, template-free, and hydrothermal synthesis route to sphere-like 伪-Fe2O3 nanostructures for textile dye removal. RSC Adv. 2016, 6, 20001–20013. [Google Scholar] [CrossRef]
  127. Mallakpour, S.; Sirous, F.; Dinari, M. Comparative study for removal of cationic and anionic dyes using alginate-based hydrogels filled with citric acid-sawdust/UiO-66-NH2 hybrid. Int. J. Biol. Macromol. 2023, 238, 124034. [Google Scholar] [CrossRef]
  128. Duman, O.; Polat, T.G.; Tunç, S. Development of poly(vinyl alcohol)/β-cyclodextrin/P(MVE-MA) composite nanofibers as effective and selective adsorbent and filtration material for the removal and separation of cationic dyes from water. J. Environ. Manag. 2022, 322, 116130. [Google Scholar] [CrossRef]
  129. Wang, X.; Dong, J.; Gong, C.; Zhang, S.; Yang, J.; Zhang, A.; Feng, Z. Bendable poly(vinylidene fluoride)/polydopamine/β-cyclodextrin composite electrospun membranes for highly efficient and bidirectional adsorption of cation and anion dyes from aqueous media. Compos. Sci. Technol. 2022, 219, 109256. [Google Scholar] [CrossRef]
  130. Wang, X.; Hsiao, B.S. Electrospun nanofiber membranes. Curr. Opin. Chem. Eng. 2016, 12, 62–81. [Google Scholar] [CrossRef] [Green Version]
  131. Tijing, L.D.; Woo, Y.C.; Shim, W.-G.; He, T.; Choi, J.-S.; Kim, S.-H.; Shon, H.K. Superhydrophobic nanofiber membrane containing carbon nanotubes for high-performance direct contact membrane distillation. J. Membr. Sci. 2016, 502, 158–170. [Google Scholar] [CrossRef]
  132. Qi, Y.; Yang, M.; Xu, W.; He, S.; Men, Y. Natural polysaccharides-modified graphene oxide for adsorption of organic dyes from aqueous solutions. J. Colloid Interface Sci. 2017, 486, 84–96. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, S.; Zhang, Y.; Bi, G.; Liu, J.; Wang, Z.; Xu, Q.; Xu, H.; Li, X. Mussel-inspired polydopamine biopolymer decorated with magnetic nanoparticles for multiple pollutants removal. J. Hazard. Mater. 2014, 270, 27–34. [Google Scholar] [CrossRef]
  134. Yu, L.; Xue, W.; Cui, L.; Xing, W.; Cao, X.; Li, H. Use of hydroxypropyl-β-cyclodextrin/polyethylene glycol 400, modified Fe3O4 nanoparticles for congo red removal. Int. J. Biol. Macromol. 2014, 64, 233–239. [Google Scholar] [CrossRef]
  135. Zhao, R.; Li, Y.; Sun, B.; Chao, S.; Li, X.; Wang, C.; Zhu, G. Highly flexible magnesium silicate nanofibrous membranes for effective removal of methylene blue from aqueous solution. Chem. Eng. J. 2019, 359, 1603–1616. [Google Scholar] [CrossRef]
  136. Anusiya, G.; Jaiganesh, R. A review on fabrication methods of nanofibers and a special focus on application of cellulose nanofibers. Carbohydr. Polym. Technol. Appl. 2022, 4, 100262. [Google Scholar] [CrossRef]
  137. Yang, M.; Lotfikatouli, S.; Chen, Y.; Li, T.; Ma, H.; Mao, X.; Hsiao, B.S. Nanostructured all-cellulose membranes for efficient ultrafiltration of wastewater. J. Membr. Sci. 2022, 650, 120422. [Google Scholar] [CrossRef]
  138. Herath, A.; Salehi, M.; Jansone-Popova, S. Production of polyacrylonitrile/ionic covalent organic framework hybrid nanofibers for effective removal of chromium(VI) from water. J. Hazard. Mater. 2022, 427, 128167. [Google Scholar] [CrossRef]
  139. Lee, S.Y.; Jeong, Y.J.; Park, W.H. Poly(vinyl alcohol) nanofibrous membranes via green electrospinning and tannin coating for selective removal of Pb(II) ion. Chemosphere 2022, 307 Pt 1, 135719. [Google Scholar] [CrossRef]
  140. Yue, Y.; Liu, Y.; Zhang, W.; Guo, J.; Gong, Y.; Yu, Y. Amidoxime functionalized low-cost cellulose-based adsorbent derived from waste cigarette filters for efficient heavy metal removal. J. Environ. Chem. Eng. 2022, 10, 107846. [Google Scholar] [CrossRef]
  141. Kandel, D.R.; Kim, H.-J.; Lim, J.-M.; Poudel, M.B.; Cho, M.; Kim, H.-W.; Oh, B.-T.; Nah, C.; Lee, S.H.; Dahal, B.; et al. Cold plasma-assisted regeneration of biochar for dye adsorption. Chemosphere 2022, 309, 136638. [Google Scholar] [CrossRef]
  142. Himmler, M.; Schubert, D.W.; Dahne, L.; Egri, G.; Fuchsluger, T.A. Electrospun PCL Scaffolds as Drug Carrier for Corneal Wound Dressing Using Layer-by-Layer Coating of Hyaluronic Acid and Heparin. Int. J. Mol. Sci. 2022, 23, 2765. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Preparation and structure of ENMs.
Figure 1. Preparation and structure of ENMs.
Molecules 28 03288 g001
Figure 2. (a) SEM and TEM images of NFM at different voltages (the inset is the hole diameter distribution map). (b) SEM and TEM images of ENM at different flow rates (the inset is the pore diameter distribution map). Reprinted from ref. [39].
Figure 2. (a) SEM and TEM images of NFM at different voltages (the inset is the hole diameter distribution map). (b) SEM and TEM images of ENM at different flow rates (the inset is the pore diameter distribution map). Reprinted from ref. [39].
Molecules 28 03288 g002
Figure 3. The effect of the distance from the tip to the collector on the ENM (ae) at distances of 100 mm, 150 mm, 200 mm, 250 mm, and 300 mm, respectively. The bar scale in (ae) is 10 μm and in (a′e′) is 100 nm. Reprinted with permission from ref. [40]. Copyright 2019 Elsevier.
Figure 3. The effect of the distance from the tip to the collector on the ENM (ae) at distances of 100 mm, 150 mm, 200 mm, 250 mm, and 300 mm, respectively. The bar scale in (ae) is 10 μm and in (a′e′) is 100 nm. Reprinted with permission from ref. [40]. Copyright 2019 Elsevier.
Molecules 28 03288 g003
Figure 4. SEM images of NEM with different morphologies, (a) Core-shell structure; (b) Hollow structure; Reprinted with permission from ref. [44] Copyright 2017 Elsevier. (c) Porous structure; (d) Aligned structure; Reprinted with permission from ref. [34] Copyright 2020 Elsevier; (e,f) porous PAN micro/nanofiber membrane Reprinted with permission from ref. [43]. Copyright 2018 Elsevier.
Figure 4. SEM images of NEM with different morphologies, (a) Core-shell structure; (b) Hollow structure; Reprinted with permission from ref. [44] Copyright 2017 Elsevier. (c) Porous structure; (d) Aligned structure; Reprinted with permission from ref. [34] Copyright 2020 Elsevier; (e,f) porous PAN micro/nanofiber membrane Reprinted with permission from ref. [43]. Copyright 2018 Elsevier.
Molecules 28 03288 g004
Figure 5. (a) Pure PAN nanofiber; (b) PAN/Ppy nanofiber; (c) PAN/Ppy nanofiber; (d) SEM and TEM image of PAN/Ppy nanofiber membrane. Reprinted with permission from ref. [50]. Copyright 2013 Elsevier.
Figure 5. (a) Pure PAN nanofiber; (b) PAN/Ppy nanofiber; (c) PAN/Ppy nanofiber; (d) SEM and TEM image of PAN/Ppy nanofiber membrane. Reprinted with permission from ref. [50]. Copyright 2013 Elsevier.
Molecules 28 03288 g005
Figure 6. (a,b) SEM photos of CNF; (c,d) SEM photos of p-CNF; (e,f) the diameter distribution diagrams of CNF and p-CNF, respectively. Reprinted with permission from ref. [53]. Copyright 2019 Elsevier; (g) PAN; (h) porous PAN; (i) SEM and TEM images of porous PAN/GO nanofibers; (il) porous structure on the shell and core of porous PAN/GO nanofibers (inset: cross section membrane performance). Reprinted with permission from ref. [54]. Copyright 2020 Elsevier.
Figure 6. (a,b) SEM photos of CNF; (c,d) SEM photos of p-CNF; (e,f) the diameter distribution diagrams of CNF and p-CNF, respectively. Reprinted with permission from ref. [53]. Copyright 2019 Elsevier; (g) PAN; (h) porous PAN; (i) SEM and TEM images of porous PAN/GO nanofibers; (il) porous structure on the shell and core of porous PAN/GO nanofibers (inset: cross section membrane performance). Reprinted with permission from ref. [54]. Copyright 2020 Elsevier.
Molecules 28 03288 g006
Figure 7. Effect of initial concentration of (a) Pb (II), (b) Cd (II), and (c) Cr (VI) ions on the adsorption of metal ions using PAN/chitosan/UiO-66-NH2 nanofibrous adsorbent. Reprinted with permission from ref. [64]. Copyright 2019 Elsevier.
Figure 7. Effect of initial concentration of (a) Pb (II), (b) Cd (II), and (c) Cr (VI) ions on the adsorption of metal ions using PAN/chitosan/UiO-66-NH2 nanofibrous adsorbent. Reprinted with permission from ref. [64]. Copyright 2019 Elsevier.
Molecules 28 03288 g007
Figure 8. (a) The effect of the pH value of the Cr (VI) aqueous solution on the adsorption capacity of PAN/GO porous nanofibers; (b) zeta potential of porous PAN/GO nanofibers in aqueous solutions with different pH values; (c) The form of Cr (VI) at pH from stepwise from 1–10. Reprinted with permission from ref. [54]. Copyright 2020 Elsevier.
Figure 8. (a) The effect of the pH value of the Cr (VI) aqueous solution on the adsorption capacity of PAN/GO porous nanofibers; (b) zeta potential of porous PAN/GO nanofibers in aqueous solutions with different pH values; (c) The form of Cr (VI) at pH from stepwise from 1–10. Reprinted with permission from ref. [54]. Copyright 2020 Elsevier.
Molecules 28 03288 g008
Figure 9. The effect of PAN/chitosan/UiO-66-NH2 nanofiber layer thickness on the removal of (a) flux and (b) metal ions using PVDF/PAN/chitosan/UiO-66NH2 nanofiber membrane, (c) permeation flux, and (d) removal efficiency of metal ions in the nanofiber membrane within 24 h. Reprinted with permission from ref. [64]. Copyright 2019 Elsevier; (e) regenerating metal on the PES/PVA/chitosan/A-Fe3O4-2 nanofiber membrane ion recovery rate; and (f) the influence of water flux. Reprinted with permission from ref. [100]. Copyright 2018 Elsevier.
Figure 9. The effect of PAN/chitosan/UiO-66-NH2 nanofiber layer thickness on the removal of (a) flux and (b) metal ions using PVDF/PAN/chitosan/UiO-66NH2 nanofiber membrane, (c) permeation flux, and (d) removal efficiency of metal ions in the nanofiber membrane within 24 h. Reprinted with permission from ref. [64]. Copyright 2019 Elsevier; (e) regenerating metal on the PES/PVA/chitosan/A-Fe3O4-2 nanofiber membrane ion recovery rate; and (f) the influence of water flux. Reprinted with permission from ref. [100]. Copyright 2018 Elsevier.
Molecules 28 03288 g009
Figure 10. (a) Schematic diagram of functionalized PAN nanofibers; (b) adsorption of functionalized nanofiber membranes. Reprinted with permission from ref. [105]. Copyright 2018 Elsevier; (c) synthesis of PDA@DCA–COOH film. Reprinted with permission from ref. [12]. Copyright 2020 Elsevier.
Figure 10. (a) Schematic diagram of functionalized PAN nanofibers; (b) adsorption of functionalized nanofiber membranes. Reprinted with permission from ref. [105]. Copyright 2018 Elsevier; (c) synthesis of PDA@DCA–COOH film. Reprinted with permission from ref. [12]. Copyright 2020 Elsevier.
Molecules 28 03288 g010
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

Li, L.; Guo, W.; Zhang, S.; Guo, R.; Zhang, L. Electrospun Nanofiber Membrane: An Efficient and Environmentally Friendly Material for the Removal of Metals and Dyes. Molecules 2023, 28, 3288. https://doi.org/10.3390/molecules28083288

AMA Style

Li L, Guo W, Zhang S, Guo R, Zhang L. Electrospun Nanofiber Membrane: An Efficient and Environmentally Friendly Material for the Removal of Metals and Dyes. Molecules. 2023; 28(8):3288. https://doi.org/10.3390/molecules28083288

Chicago/Turabian Style

Li, Li, Wei Guo, Shenggui Zhang, Ruibin Guo, and Li Zhang. 2023. "Electrospun Nanofiber Membrane: An Efficient and Environmentally Friendly Material for the Removal of Metals and Dyes" Molecules 28, no. 8: 3288. https://doi.org/10.3390/molecules28083288

Article Metrics

Back to TopTop