Next Article in Journal
Investigation into Effects of Coating on Stress Corrosion of Cable Bolts in Deep Underground Environments
Previous Article in Journal
Assessment of Digital Image Correlation Effectiveness and Quality in Determination of Surface Strains of Hybrid Steel/Composite Structures
Previous Article in Special Issue
Gas-Phase Fabrication and Photocatalytic Activity of TiO2 and TiO2–CuO Nanoparticulate Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Metal Ion Recovery Based on Membrane Technology and Adsorption Synergy

School of Environmental Science and Engineering, Shandong University, Qingdao 266237, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(14), 3562; https://doi.org/10.3390/ma17143562
Submission received: 3 June 2024 / Revised: 8 July 2024 / Accepted: 15 July 2024 / Published: 18 July 2024
(This article belongs to the Special Issue Functional Nanoparticle Materials: From Synthesis to Application)

Abstract

:
The development of modern industry will generate more and more waste containing metal ions. It is necessary to take appropriate measures to recover these ions, whether from the perspective of environmental protection or improving economic benefits. So far, scientists have studied many methods for recovering metal ions. Among these methods, adsorption and membrane separation have received widespread attention due to their own characteristics. Combining adsorption and membrane separation methods can better leverage their respective advantages to improve the ability of recovering metal ions. This review, therefore, focuses on the synergistic recovery of metal ions by adsorption and membrane separation methods. This article first briefly explains the theoretical principles of membrane separation and adsorption synergy, and then focuses on several technologies that have received attention in different chapters. In these chapters, membrane technology is briefly introduced, followed by the situation and progress of synergistic application with adsorption technology. Then, the article compares and elaborates on the advantages and disadvantages of the above technologies, and finally summarizes and looks forward to these technologies being used to solve the difficulties and challenges in industrial application.

1. Introduction

Metal recovery is significant due to its environmental benefits and its contribution to competitive economic systems [1]. Heavy industries such as mechanical manufacturing, metallurgy, and chemical engineering typically generate a large amount of dust and wastewater containing heavy metals. These substances, including cadmium (Cd), chromium (Cr), lead (Pb), mercury (Hg), etc., enter the ecosystem through the atmosphere or water cycle, even posing a threat to animals, plants and humans through the food chain [2]. On the other hand, China requires a large amount of mineral resources every year to meet industrial demand. For example, China imported 4.415 million tons of lithium ore in 2023, according to news reports. Coupled with enormous amounts of scrap metals awaiting recycling, increasing mineral imports does not achieve carbon neutrality goals. Therefore, how to deal with heavy metal pollution and improve metal utilization is currently a hot topic in the environmental field.
Various methods are employed for the recovery of metal ions, categorized into physical methods like adsorption [3], membrane separation [4] and ion exchange [5]; chemical methods such as oxidation-reduction [6], electrodialysis [7] and chemical precipitation [8]; and biological methods including plant absorption, microbial treatment [9] and algae treatment [10].
Among the above methods, adsorption is widely used due to its high efficiency, simple operation, low cost, good stability, and diverse selection of adsorbents [4,11,12,13]. The key is to choose appropriate adsorption materials. The adsorption materials used for metal ion recovery mainly include activated carbon [14], zeolite [15] and graphene [16]. Also, membrane materials have been extensively employed due to their diverse types, high performance and selectivity, simple modification, mild operation and low chemical consumption [11,17,18,19]. However, the two methods have their own disadvantages, so some scientists combine membrane technology with adsorption technology and synergistically apply it to metal ion recovery. This not only overcomes the limitations of a single technology, but also leverages their respective advantages to achieve synergistic effects of technology, and improves the efficiency and economy of metal ion recovery. An overall analysis of the literature in the Scopus database, especially for the past three years, reveals that the synergy of the two technologies has generated much interest in metal recovery in the past decade. Statistics are as shown in Figure 1.
This review first analyzes the mechanism of membrane technology and adsorption synergistic recovery of metal ions from a theoretical perspective. Then this paper discusses the concepts, classifications, and application progress of several membrane technologies separately. Next, the article compares the membrane technologies discussed and analyze their advantages and limitations. Finally, a summary of this type of membrane technology is provided, and its potential applications in the industrial field are discussed.

2. Theory of Membrane and Adsorption Synergy Technology

Membrane technologies are widely used in metal recovery because of its easy operation and high efficiency. In theory, a membrane can be seen as a barrier that blocks the passage of certain metal components (i.e., retentate) while allowing other components (i.e., permeate) to pass through. Therefore, the removal mechanism of membrane separation mainly involves size exclusion, as well as Donnan exclusion (charge–charge repulsion) effect and adsorption capacity for specific pollutants [4].
In addition to size exclusion, adsorption phenomena also can play a part in metal-polluted water treatments and recovery by membranes. Adsorptive membranes are a modification of existing membranes (e.g., ultrafiltration (UF) membranes) by adding additives to improve membrane properties such as hydrophilicity, porosity, mechanical strength and chemical stability [20,21]. For instance, the UF membrane can be added to some functional groups or compounds that can coordinate with metals (-NH2, -COOH, MOF, etc.) to achieve a membrane’s adsorption abilities [22,23,24]. Therefore, the mechanism of membrane adsorption is like traditional adsorption, where the adsorbent transfers substances from the liquid phase to the solid phase by physical or chemical processes [22]. Many studies have also evaluated membrane adsorption behavior and tried to find the adsorption process and mechanism of membranes through isothermal and kinetic analysis methods [22,25,26,27].
Abdullah et al. [20] reported that although the membrane pore size is smaller after adding adsorptive nanomaterials, the adsorption process is the main removal mechanism, instead of size exclusion. As metal ions are still smaller than the pore size of modified membranes, ions can permeate membranes. This also means that adsorption performance is important for ion removal. The adsorption of metal ions onto mixed matrix membranes (MMMs) may involve four possible steps in the study: (1) bulk diffusion of ions into membrane pores and finger-like layer; (2) inner sphere surface complexation on the outer sites of adsorbents; (3) transfer of ions across the adsorbent interface; and (4) equilibrium stage. Therefore, the rate-controlling step for the sorption of metal ions is chemisorption on the external active sites of adsorbents, while the rate-limiting step is the transfer of ions across the adsorbent interface [20], as shown in Figure 2.

3. Metal Recovery by Different Membrane and Adsorption Synergy Processes

3.1. Ultrafiltration Membranes

Ultrafiltration (UF) membranes work at a low trans-membrane pressure to remove dissolved and colloidal materials. Because the pore sizes of UF membranes are larger than metal ions dissolved in aqueous solutions in the form of ions or complexes, these ions can pass through UF membranes easily [8]. As mentioned above, certain functional groups or compounds could be added to UF membranes to increase their adsorption efficiency. UF can be divided into two sub-UF processes: polymer-enhanced UF (PEUF) and micellar-enhanced UF (MEUF).
PEUF was initially formed by the complexation of polysaccharides extracted from biomass with metal ions [28]. The mechanism of PEUF is to form a complex between polymer ligands and metal ions, making the metal ion larger than the membrane pores and achieving a separation process [29]; as illustrated in Figure 3a [30]. In fact, the selection of the polymer ligands in PEUF is very important to improve the combination between metal ions and polymers. Polymer reagents need to meet the following requirements: high affinity towards target metal ions; high molecular mass; high water wettability; chemical and mechanical stability; and low toxicity and costs [31]. If the above standards cannot be met, metal ions will penetrate the membrane, leading to a decrease in recovery rate.
Similar to PEUF, the MEUF process is also used to prevent metal ions from passing through UF membranes, but this process is achieved by forming micelles between metal ions and surfactants; as depicted in Figure 3b [32]. That is to say, cationic or anionic surfactants should be added according to the situation of the water body when designing MEUF. When the concentration of surfactant exceeds the critical micelle concentration (CMC), surfactant monomers aggregate with metal ions to form amphiphilic transparent micelles, and the hydrodynamic diameter will be larger than the pores of the UF membrane. In addition, some parameters should be considered when performing MEUF; such as pressure, pH, concentration and dosage of active agents, and the metals to be removed [4].
Through the above introduction and analysis, it should be noted that UF membranes themselves have a certain ability to trap metal ions through polymerization or ion interactions, but this ability is not strong. Membrane modification can be achieved by inoculating adsorbent materials on the surface of the membrane, which allows the UF membrane to play its own role while also improving rejection efficiency through surface adsorption. This is the fundamental essence of the so-called synergistic effect.
Some recent studies conducted on UF for the removal of metal ions are presented in Table 1. From the table, it can be seen that various polymers or surfactants can functionalize UF membranes and improve their ability to retain metals. For instance, chitosan—the most important derivative of the natural amino polysaccharide chitin—is receiving increasing attention due to its high binding ability, wide availability and unique functions [33,34,35]. Indiasih et al. combined chitosan as a polymer with PES membrane and studied the effect of the ratio of chitosan to hexavalent chromium Cr(VI) on its rejection rate and permeation flux [36]. The results showed that when chitosan with a mass ratio of 12.5 of chitosan/Cr(VI) was added, the rejection rate of Cr(VI) rapidly increased from 28% to 75%, while the permeation flux only decreased by less than a quarter. The authors believe that the mechanism of increased rejection is that the amino groups of the chitosan used can achieve complete protonation at pH 5. The protonated amine group combines with Cr(VI) to form a chitosan-Cr(VI) complex, as shown in Figure 4c, which achieves adsorption of Cr(VI) in the form of the complex. The number of protonated amine groups also increases with the increase of chitosan mass, thereby enhancing the formation of complex macromolecules. Combined with the separation effect of the membrane, the interception of metal ions is completed. In addition, the color change between the feed solution and the permeate is shown in Figure 4b. This experiment demonstrates the excellent ability of chitosan. Some experts have also carried out functionalization modifications on the basis of chitosan to enhance the ability of UF membranes to recover metals. Nikhil Chandra et al. studied the functionalization of chitosan/poly (styrene sulfonate) (CHS/PSS) polyelectrolyte multilayer membranes (PEM) with tannic acid (TA) and campesterol (CHT) under UF conditions, and investigated the ability of these membranes to recover lithium [37]. The results showed that TA and CHT were immobilized on PEM membranes and achieved a high recovery rate of over 94% and excellent permeation flux. Through a series of experiments, it was also found that the bare membrane itself repels cations due to the surface charge of the membrane. The addition of TA or CHT causes the surface of the membrane to form a porous structure, increasing the surface area of the membrane. At the same time, it also allows metal ions to interact with carboxyl groups in TA or CHT to form complexes. These processes allow metal ions to be trapped inside the membrane. In addition, high cycling performance and lifespan also demonstrate the potential application of functionalized PEM.
Apart from chitosan, metal organic frameworks (MOFs) are widely used to modify UF membranes to strengthen the ability of metal recovery. Yin et al. functionalized Zr-MOFs with amino groups and combined them with ceramic membrane ultrafiltration (CUF) to remove Pb(II) from wastewater [42]. Some results are shown in Figure 5. The membrane separation process of the aqueous solutions was achieved at a high trans-membrane pressure of 0.23 MPa and a low cross-flow velocity of 3.0 m·s−1. MOFs nanomolecules on the CUF adsorbed Pb(II) during this process, achieving the synergistic application of the two technologies. It was found that the removal rate of lead by the composite membrane exceeded 60%, and the adsorption capacity also reached 1795.3 mg/g. This successfully verified the adsorption mechanism of the coordination interaction between the amide group (-NH2) and Pb(II). Additionally, after cleaning the membrane, its recovery efficiency approached 100%. Senol-Arslan et al. used a solvothermal method to prepare Ni-Zn-MOF powder and prepared Ni-Zn-MOF-embedding membranes with different concentrations [43]. Similar to reference [42], the Ni-Zn-MOF membrane not only exhibits ion interactions on the membrane surface, but also possesses the adsorption capacity of MOF, enabling higher rejection rates during the separation process. The influence of temperature, pH and other parameters on the adsorption capacity of Pb(II) was studied through a series of characterizations. It was found that the MOF-modified PSF membrane had a better Pb(II) rejection rate than the pure membrane, and reached 98% under optimal conditions.
Adsorptive ultrafiltration mixed matrix membranes (MMMs) are a new strategy to remove metal cations, which achieve ultrafiltration and adsorption functions in one unit. Adsorptive ultrafiltration MMMs can be divided into four categories, according to the types of adsorbents: inorganic filler, organic filler, biomaterial and mixed filler membrane [47]. Kumar et al. prepared a novel polyethylene tetrazole co-polyacrylonitrile (PVT-co-PAN) membrane using a nonsolvent induced phase separation (NIPS) method [26]. After adding PVT fragments, the membrane became more negatively charged and hydrophilic, and PVT fragments are the main binding sites for adsorbing Cu(II) ions, with an adsorption capacity of 44.3 mg/g; higher than other membranes reported in the literature. Usman Farid et al. [48] performed Ar/O2 plasma treatment on multi-walled carbon nanotubes (MWCNTs) [49]. The experimental results showed that the composite membrane with functionalized MWCNTs maintains an adsorption and removal efficiency of over 90% for Zn2+. In addition, selective adsorption of zinc can also be achieved in the presence of common coexisting ions such as Na+, Fe2+, Mn2+, Mg2+ and Ca2+. Importantly, the adsorption capacity of the CNT membrane for zinc can be easily regenerated under acidic conditions.

3.2. Nanofiltration Membranes

Nanofiltration (NF) is a technology between UF and RO and its pore size is in the range of 1–10 nm [50]. The advantages of NF membranes are their low energy demand, high efficiency for metal removal, easy operation and low pressure requirement [8]. NF membranes are generally made of aliphatic amine monomers, which carry positive or negative charges on the membrane. Due to the electrostatic interaction between metal ions and the membrane, coupled with the loose structure and small pore size of NF membranes, the separation and recovery of metal ions can be well achieved [4,51]. In addition to commercial NF membranes, existing substances or processes can be used to improve NF membranes, achieving the goal of enhancing their ability to adsorb metal ions. Some recent studies about NF membranes are shown in Table 2.
Xu et al. [63] developed a positively charged nanofiltration membrane with a dense polyamide (PA) layer and surfactant functional layer (SFL); namely, the CTAB/PIP/TMC TFC membrane. Hexadecyltrimethylammonium bromide (CTAB) endowed it with a superior positively charged surface, resulting in a separation efficiency of over 98% for Zn2+ and Ni2+, over 96–98% for Cu2+ and over 85% for Pb2+. It also found that CTAB molecules form multilayers through electrostatic adsorption and hydrophobic interactions, changing the surface properties such as pore size, surface charge, and chemical composition of the membranes. Most metal ions are blocked due to charge repulsion. Even if a small number of metal ions reach the PA layer, the carboxyl groups on the surface of the PA layer will combine with the metal ions to complete the adsorption process. Abedi et al. synthesized l-cysteine functionalized cellulose nanocrystals (CysCNCs) using the iodine catalysis method, and dispersed the CNCs into trimeric chloride in n-hexane for heavy metal removal experiments [49]. After evaluating the membrane performance of thin film composite materials (TFC); and CNC, ACNC and CysCNC TFN membranes, it was found that the CysCNC-TFN0.2 membrane exhibited the best heavy metal removal performance, of over 85%. The analysis and results are shown in Figure 6. The reason for achieving such high removal efficiency is that NF membranes mainly rely on electrostatic repulsion and size repulsion to separate metals. The thiol (-SH) functional groups of CysCNCs introduce additional sites to enhance the repulsion of heavy metals through complexation or adsorption [64]. The coordination interaction between Cu-S and Pb-S (chelation effect) can lead to higher repulsion of Cu2+ and Pb2+ metal cations. They further speculate that, as part of the Lewis soft base family, sulfur-containing species have highly polarized donor centers that can strongly react with the orbitals of soft Lewis acids (such as heavy metal ions). Furthermore, the improvement of TFN membrane performance was mainly due to the functionality of nanoparticles and the uniform distribution of particles dispersed in the polyamide layer through solution.
Roy et al. used simple impregnation and cross-flow filtration methods to modify polyethylene imine (PEI) and EDTA, and prepared new amine functionalized composite membranes P-60S and P-60S-EDTA on a tubular ceramic substrate [65]. The experimental results showed that P-60S-EDTA has superior permeability and rejection efficiency, with rejection rates of over 90% for Cu(II), As(V) and Cr(VI). The membrane operated well in both ideal and actual environments. This is because, in the case of the P-60S-EDTA composite membrane, the reduction of pore size and the presence of surface functional groups on the membrane are conducive to the interception of spatial effects. Under low pressure and pH 3 conditions, the chelation of HCrO4 and H2AsO4 ions with protonated surface amino groups is the main mechanism for removing As(V) and Cr(VI). However, as the operating pressure increases, small anions will pass through the composite membrane, leading to a decrease in rejection rate and an increase in the number of discharged anions.
Arouche et al. used carbon nanotubes and boron nitride nanotubes as nanofiltration membranes, and tested the membrane’s rejection effect on heavy metal ions through electric field force testing [66]. The study showed that both nanotubes had the best filtration effect when the electric field strength was 10−8 a.u. The kinetic energy and in situ temperature required for both types of nanotube filtration are the minimum values in 10−8 a.u. Due to differences in hydrophobicity and thermal stability, the effect of boron nitride nanotubes (90–98%) was better than that of carbon’s (80–90%). Meanwhile, Arouche et al. also found that as the electric field (EF) strength increases, the increase in evanescent effect will affect the treatment effect. Based on this, they believe that nanofiltration will have better results at lower EF strengths. An additional inference is that the filtering effect may be better without adding any EF.
Fonseka et al. [67] modified the chromium-based MOF with N-(phosphonylmethyl) iminodiacetic acid (PMIDA) and amine grafted mesoporous silica (SBA15) materials for selective recovery of rare earth elements (REE) and copper, respectively. Two adsorbents were used sequentially to selectively adsorb REE (91%) and Cu (90%) from the concentrated feed with pH regulation. The main driving force for adsorbing REE is the formation of coordination complexes with carboxylate and phosphine ions on the MOF; and the adsorption of Cu is due to the strong chelation bond between Cu and amine ligands. Morgante et al. [59] incorporated the active layer of NH2-MIL-101 (Al) and ZnO nanoparticles into the chitosan matrix, and combined it with the ultrafiltration matrix to prepare a nanofiltration membrane. The characterization analysis and results are shown in Figure 7 and Figure 8. The experiment found that the membrane containing 35 wt% ZnO had the highest rejection rates for MgCl2 (90.10%) and CaCl2 (86.49%) at a solution concentration of 1000 ppm, and the selectivity for these two substances was higher than that of commercial membranes (NF90 and NF270).

3.3. Electrospun Nanofibrous Membranes

The electrospun nanofiber membrane (ENFM) is essentially a polymer-based membrane [77]. Polymer-based membranes can be prepared using various methods, including stretching, phase separation, electrospinning, evaporation, and etching [78]. Among these methods, electrospinning is a simple, reliable, universal, and sustainable technology. The significant characteristics of spinning nanofibers are high surface-to-volume ratio, high porous structure and excellent connectivity, sufficient flexibility, extensive material collection, and the ability to prepare different structures by adding different materials through electrospinning [79,80]. According to Uddin et al.’s article, there are six preparation methods for electrospinning [80]. For convenience, these membranes made from electrospun fibers by these methods are usually referred to as “electrospun nanofiber membranes (ENFMs)”. Since ENFMs exhibit excellent adsorption performance because of their high specific surface area and high porosity, ENFM is a good choice for removing heavy metals from aqueous solutions [77]. Adsorbents require active sites so that metals can be removed through ion interactions or chelation. At present, researchers use CS, cyclodextrin (CD), polyacrylic acid (PAA), polyethyleneimine (PEI), polyindole (PIN) and polyaniline as nanofiber membranes to remove some metal ions [81]. However, most polymers have poor spinnability and stability in aqueous media. Therefore, by combining different polymers with various nanoparticles to improve their performance and stability, electrospun nanofiber composite membranes have been developed [81].
Liu et al. [82] prepared electrospun nanofiber composite PAN–PVA membranes and nanofiber PAN and PVA membranes, fabricated through a two-nozzle electrospinning process. Different modification methods were used to functionalize the PAN components of PAN–PVA nanofiber composite membranes: one was to use amino groups as adsorption sites to attract negatively charged heavy metal ions (such as Cr(VI)), as shown in reaction (I) in Figure 9a. The other method was to hydrolyze PAN with strong alkali to form sodium polyacrylate (as shown in reaction (II) in Figure 9a), and the resulting carboxylate ions coordinated with positively charged heavy metal ions, such as Cd(II). Using GA as a crosslinking agent to chemically crosslink PVA nanofiber components and modifying PAN, subsequently, the composite membrane was used for adsorption experiments of Cr(VI) and Cd(II); this achieved high results (66.5 mg/(g membrane) against Cr(VI) and 33.6 mg/(g membrane) against Cd(II)).
Miao et al. [83] first synthesized iron-based MOFs (Fe-MOFs), and then co-electrospun polyacrylonitrile (PAN) to prepare composite electrospun nanofiber membranes (PAN/MOFs ENFMs). In the co-electrospinning process, ENFMs mainly act as carriers, and MOF particles are the main adsorbents. Fe-MOFs particles are fixed in situ in ENFMs, with MOFs particles partially on the fiber surface and partially encapsulated within the fiber, forming complex porous structures. The results showed that PAN/MOFs ENFMs had a maximum theoretical adsorption capacity of 127.70 mg/g for Cr(VI) at pH 4.0, and partially adsorbed Cr(VI) could be reduced to less toxic Cr(III). The above results are shown in Figure 10. It was also found that the adsorption of Cr(VI) is a spontaneous process dominated by physical adsorption. In Miao’s article, the maximum theoretical adsorption capacity is determined based on the saturation adsorption capacity of the theoretical isotherm, and its value is calculated using the Sips model. The formula is as follows [83]:
qe = qm × (a × Ce)1/n/(1 + (a × Ce)1/n)
a: Sips affinity coefficient; 1/n: Sips affinity coefficient.
Additionally, Mohammed et al. [84] developed a PAN/PEI nanofiber composite membrane (NCM) based on ZIF-8 nanoparticles, and then the adsorption of Cr(VI) in water by NCM was tested. The results showed that the addition of PEI endows the surface of NCM with superhydrophilicity, while the adsorption capacity of NCM is 403.5 mg/g at a ZIF-8 concentration of 5 wt%. Through FTIR and XRD spectra, it was found that the adsorption process of Cr(VI) mainly involves the formation of hydroxyl (-OH), amino (-NH2), amide (-CO-NH) and other structures between PEI and PAN. ZIF-8 first adsorbs Cr(VI), and then these functional groups convert Cr(VI) to Cr(III) through redox reactions. Subsequent analysis also indicated that PAN/PEI@ZIF-8NCM has high selectivity and stability for wastewater treatment.
In addition, Yin et al. [85] used a dual nozzle electrospinning method to prepare electrospun sulfonated polyether sulfone (SPES) nanofiber membranes, and tested the fiber membrane’s adsorption capacity for heavy metal ions by adjusting the fiber diameter. In the ATR-FTIR spectrum of SPES, two new peaks were generated, representing the generation of –SO3H, which is served as adsorption sites for positively charged substances (e.g., Pb(II)). It was found that the rejection rate of Pb(II) by ultrafine nanofiber SPES membrane was 96.2% (the initial concentration of Pb(II) is 1.0 ppm), and the sustainability of the membrane was demonstrated in cyclic experiments. Li et al. [86] used HPEI grafted graphene oxide (GO) and polyacrylonitrile (PAN) to synthesize PAN/GO-HPEI electrospun nanocomposite fiber membrane through electrospinning. Then, HPEI was grafted through chemical grafting reaction to obtain HPEI-g-PAN/GO-HPEI electrospun nanocomposite fiber membrane, which was used as an adsorbent to adsorb Au(III) in aqueous solution. Due to the presence of amino groups, HPEI-g-PAN/GO-HPEI exhibited efficient adsorption of Au(III), with a maximum adsorption capacity of 2893.33 mg/g at the center point. This experiment also found that adsorbed Au(III) can be reduced to Au(I) and irregular thin and granular Au elements attached to the adsorbent.

3.4. Ion-Imprinted Membranes

Ion-imprinting technology (IIT) is derived from molecular imprinting technology, and the principles of these two technologies are similar; except that the template is replaced with ions. Although Dickey [87] proposed the concept of “specific adsorption” in 1949, it was not until 1976 that the application of ion imprinted polymers (IIPs) was first realized [88]. Since the emergence of IIPs, they have attracted attention due to their excellent chemical properties and specific selectivity. Template ions, functional monomers, crosslinking agents, initiators, and solvents are required to synthesize IIP; with the target generally being template ions. The preparation and identification process of IIP is shown in Figure 11.
The ion imprinting membrane (IIM) is a combination of ion imprinting technology and membrane separation technology, which refers to a special functional membrane prepared by direct or modified synthesis between IIPs and membranes. IIM can achieve specific selectivity of IIT and high efficiency and stability of membrane separation, thereby achieving selective separation in complex solution systems. Compared with IIP, IIM can achieve the separation and concentration of metal ions without grinding, and demonstrates high recovery efficiency and low diffusion resistance [89]. In addition, Lu et al. [90] provided a more comprehensive introduction to the origin and development, mechanism, preparation, performance evaluation, application of IIM, etc.
Yang et al. [91] used 12-crown-4 ether (12C4) as a selective metal complexing agent for lithium ions and prepared lithium-ion imprinted membranes (LIIMs) through hydrolysis polymerization; the preparation process is shown in Figure 12h. Observing the SEM images, it can be seen that Figure 12d,e shows that LIIM has a rougher surface than Figure 12a,b. Many microspheres were loaded on the surface of lithium-ion imprinting membranes. It was also found that LIIM exhibited rapid adsorption kinetics and adsorption capacity (132.00 mg/g), and Langmuir isotherm adsorption results confirmed that the imprinting sites were uniform. Perm selectivity experiments and adsorption regeneration cycles demonstrated the excellent selectivity and stability of LIIM. He et al. [92] prepared a multi-layer interlayer (SiO2@SP-PDA) and even coat on the substrate surface (PVDF membrane). By modifying SiO2@SP-PDA with polydopamine (PDA), the aggregation of the ion imprinting layer was reduced while increasing the adsorption capacity to 231.77 mg/g. When analyzing the mechanism of selective adsorption, it was found that the peaks of C-O-C/C-O and O-C=O both shifted, and the intensity of the peaks significantly decreased, indicating that the O-containing functional groups of methacrylic acid (MAA) and 12C4 participated in the chelation of Li+. The changes in adsorption peaks observed in ATR-FTIR spectra before and after Li+ adsorption also confirm this viewpoint. Therefore, the adsorption of Li+ mainly depends on the chelation effect and appropriate spatial structure. Qu et al. [93] prepared a TiO2/PVDF composite LIIM using 12-crown-4 ether as a functional monomer. LIIMs showed high affinity for Li+ (799.82 mg/g, 308 K) in simulated leaching solution (SLS), and also showed high selectivity for Li+ in lithium and other ion systems. Through SEM, FT-IR and XPS analysis, it was found that the adsorption mechanism of LIIMs is the coordination between the active adsorption sites on the crown ether surface and Li+, and oxygen atoms provide lone electron pairs for Li+ to form coordination bonds.
Except for LIIMs, other metal ions can be used to make IIMs. For example, Jin et al. [94] developed an IIM to adsorb Ag+ ions in aqueous solutions. The dynamic adsorption results showed that the prepared IIM exhibited excellent performance (29.98 mg/g) when applied to feed solutions with 12 mg/L and pH = 6. The reason for achieving such results is, on the one hand, the FTIR spectrum showed that N-benzoyl-N’, N’—dibutylthiourea (BDBTU) had successfully adhered to the membrane; on the other hand, XPS and EDS spectra confirmed that the formation of chemical bonds between Ag+ and the -C(=O)-NH-C(=S)- group of BDBTU is believed to occur through electron transfer from the C=O and C=S groups of IIM to Ag+. This reaction enhances the adsorption capacity of Ag+ in the solution. Zhang et al. [95] developed an IIM by self-assembling pyridine monomers in restricted reduced graphene oxide (r-GO) nanochannels for Nd extraction. The membrane exhibited excellent Nd(III) adsorption capacity of 20.6 mg/g in a solution containing seven rare earth elements (REE). The ion radius, average bond length between Nd(III) and pyridine nitrogen, and the coordination of Nd(III) leading to an increase in nitrogen binding energy are the main factors for Nd(III) targeted binding to bipyridine nanocages. Zhang et al. [96] used chitosan to prepare two types of Nd-IIMs. The kinetic experiments on Nd(III) showed that both composite membranes reached 80% of their maximum adsorption capacity within 200 min. Even after 10 cycles of adsorption and desorption, the imprinted composite membrane still maintains a removal rate of about 80%. Mokhtar et al. [97] prepared an ion imprinted PVDF/diVB-TETA-Cu membrane, and a series of results indicate that the prepared membrane is suitable for the selective separation process of Cu2+ ions. In the selective experiment, it is believed that the selective adsorption of copper mainly involves the following steps: (1) the concentration difference between the feed side and the receiving side, which causes Cu2+ and Ni2+ to move from the solution to the membrane interface; (2) the chelation reaction between the active site and Cu2+, which is due to the higher affinity of TETA for Cu(II) and the higher imprinting effect of the membrane on Cu(II); (3) desorption into receiving compartment. Wang et al. [98] has prepared an ion imprinting multilayer membrane for selective separation of copper ions, which alternately assembles polyacrylic acid and chitosan to form a multi-layer structure. When the assembly layers are 3.5 layers of polyacrylic acid and 4 layers of chitosan, the adsorption capacity of Cu(II) is the highest (48.0 mg/g).

3.5. MXenes

MXenes refer to transition metal carbides and nitrides, which have become research hotspots in many fields as an emerging two-dimensional material [99]. The general formula of MXenes is Mn+1XnTx (n = 1, 2, or 3), where M represents transition metals (such as Mo, Cr, Nb, V, Zr, and Sc), X stands for nitrogen or carbon, and T refers to surface halogen or sulfur group element capped groups (such as -F, -OH, -O) [100]. MXene is mainly generated by removing the MAX phase of interlayer third or fourth group elements. MAX phase also refers to layered transition metal carbides and nitrides, which form Mn+1AXn (n = 1, 2, 3, and 4) with post transition metals of the III–VI group (such as Ni, Fe, Zn, and Cu) and A (such as Si, Al) [101]. Due to the difference in n, the allocation form of MAX phase is also different [102]. So far, many MXenes have been theoretically predicted, laying a solid foundation for extensive research on MXenes, as shown in Figure 13 and Figure 14.
Currently, many studies have reported examples of MXene adsorbing metal ions (e.g., Pb, Cu, Cr, U). Powdered activated carbon (PAC) served as the control material for MXene (Ti3C2Tx), and a series of methods were used to focus on the adsorption of Pb(II) by MXene. Jun et al. [103] found that the main adsorption mechanism of Pb(II) is electrostatic attraction in the experimental results of solutions with different pH levels, ion strengths, and humic acid concentrations. Furthermore, FT-IR and XPS revealed that the adsorption mechanism of Pb(II) is ion exchange and the formation of inner spherical complexes. In addition, the value of MXene in removing heavy metals was discovered through adsorption/desorption cycles. Zhang et al. [104] prepared Ti3C2Tx MXene nanosheets with amino groups through alkaline grafting modification based on the popular silane coupling agent (APTES). The prepared alk-MXene-NH2 increased the content of grafted organic functional groups, significantly improved interlayer spacing and chemical state, and exhibited strong adsorption capacity (384.63 mg/g). Biobased materials can also be used to improve the adsorption capacity of MXene due to their low toxicity and high biocompatibility. As a case example, enzymatic hydrolysis of lignin (EHL), lignosulfonate, and chitosan modified Ti2CTx nanosheets were selected, and the differences in the removal of Pb (II) ions by various surfactants were analyzed [105]. The results showed that EHL functionalized Ti2CTx exhibited excellent capacity for Pb(II) (232.9 mg/g) and prevented the re stacking of Ti2CTx nanosheets. Therefore, enhancing the performance of activation sites is an effective strategy for enhancing the removal of Pb ions on MXenes. These above results are shown in Figure 15.
MXene can also be used to adsorb other heavy metals and radioactive elements. Shahzad et al. [106] synthesized 2D Ti3CNTx and Ti3C2Tx MXene nanosheets by peeling off Ti3AlCN and Ti3AlC2 MAX phases, respectively. It was found that during the adsorption of mercury ions, the interaction between bimetallic and hydroxyl groups with Hg2+ was achieved, and stronger adsorption ability was achieved (Ti3C2Tx: 5473.13 mg/g; Ti3CNTx: 4606.04 mg/g). Apart from Hg2+ removal, Dong et al. used a new type of Mxene/alginate composite material, which has excellent removal ability for Cu2+ (87.6 mg/g) due to chemical coordination and ion exchange effects [107]. Elumalai et al. used amino acids to layer Ti3C2Tx thin films to study the removal of Cu2+ [108]. With the assistance of TiO2, histidine functionalized rutile TiO2@d-Ti3C2Tx The hybrid material achieves maximum adsorption capacity (95 mg/g) within 5 min. In addition, Kong et al. constructed amino functionalized MXenes (NH2-Ti3C2Tx) using in situ polymerization of (3-aminopropyl) triethoxysilane (APTES) on the surface of Ti3C2Tx [109]. Through DFT calculation, the synergistic effect between Ti and N significantly improves the binding energy of MXene to Cr(VI) and the electron density on the surface of MXene. As for radioactive elements removal, a simple example is that Liu and co-workers reported the PEI-Ti3C2 exhibited huge uranyl loading capability, high removal efficiency and remarkable sequestration selectivity [110]. Shahzad et al. reported that Ti3C2Tx mixed Prussian blue aerogel (PBMX) spheres can effectively remove Cs+ (315.91 mg/g) from wastewater [111]. In addition, there are reports on the removal of Sr2+ and Ba2+ by MXene [112,113,114,115], as shown in Figure 16.

4. Evaluation of Synergy Process

Table 3 compares the advantages and disadvantages of various membranes described in the previous chapters. In experimental design or industrial applications, it is necessary to find suitable membrane treatment methods based on the material composition and characteristics of the water to be treated. At the same time, it is also necessary to address the limitations of various membrane technologies in application to expand the scope of membrane applications.
Composite membranes should be concerned in terms of reusability. MMMs, IIMs, MOFs membranes and ENFMs all have high selectivity and adsorption capacity, but their cyclic regeneration performance is not good. Strong acids or alkalis are usually used to wash the membranes to improve their regeneration performance, but this will accelerate the aging of the membranes and reduce their service life. The nanomaterials in the composite membrane may also be leached, which can lead to a decrease in membrane treatment efficiency. On the other hand, after being discharged into the natural environment, nanoparticles may adsorb harmful substances in the environment, and may also promote the reaction among harmful substances to generate secondary pollutants. Many articles focus on the adsorption and regeneration capabilities of these composite membranes, with little attention paid to such environmental issues. For sustainability, there have been some attempts to use organic acids [119] and microorganisms [120] as leaching agents, but it will take some time for these new leaching agents to completely replace inorganic acids/bases. Also, the low dispersion rate of dissolved nanomaterials can cause the aggregation of nanoparticles, ultimately leading to the instability of membrane performance.
For PEUF and MEUF, it is necessary to study the critical point of metal complexes or micelles. If the process is not improved, it may result in a very low removal rate of metal ions. In addition, waste of ligands/micelles and the production of harmful products may also occur. For ENFMs, it is important to study in detail its adsorption mechanism, the interaction between various substances in the composite membrane, the adsorption of multi-component solutions, and real sewage experiments [77]. For example, Nauman et al. [79] elaborated on the effects of processing parameters, solution characteristics, environmental factors and collector configuration on the mechanical strength, diameter, compatibility with adsorbents and other factors of electrospun fibers; and compiled many fiber post-treatment methods (such as crosslinking, hot pressing, hot stretching and solvent welding). For IIMs, ion recognition ability needs to be given special attention, mainly because the diameter of Mg2+ is similar to that of Li+, which creates competition when adsorbing Li+. In order to enhance selectivity, modified crown ethers [121], calixarenes [122], and blended chitosan [123] are generally used as chelating agents, and there have also been studies using “crown ethers + ionic liquids” [124,125]. MOF-based membranes often suffer from insufficient stability due to the different properties of the MOF layer and support layer, which limits their binding force and the continuous growth of the MOF layer. There are some methods to solve this problem, such as using interface growth-based methods to prepare Zn-MOF membranes, and developing sacrificial strategies for preparing MOF membranes containing water unstable MOFs (MOF-5, HKUST-1, and ZIF-67) [126]. Its biocompatibility and toxicity need to be studied for industrial application for MXenes. The toxicity of MXene based materials depends on their physical and chemical properties, such as size, dosage and surface coating; and surface chemistry is also an important factor affecting toxicity. Reducing toxicity can be achieved through size reduction [127], fluorine-free synthesis, and coupling with other materials [128,129].
It is worth mentioning that, although membranes are widely used in the field of metal ion separation and recovery, there are still some common challenges that affect the industrial application of membranes. For example, fouling, swelling, reusability and sustainability remain important factors determining the commercialization ability of any membrane. Therefore, if researchers can improve the membrane process and overcome the limitations of various membranes, membrane technology will become the main way of metal ion collection.

5. Conclusions and Prospects

Over the past several years, many methods have been developed to recover metal ions from aqueous solutions. Conventional and novel methods are applied with their own advantages and limitations. Membrane technology with adsorption ability exhibits huge potential amongst these methods, as it combines the advantages of membrane separation and adsorption. These include the UF membranes (PEUF, MEUF, MMMs), NF membranes, ENFMs, MOFs membranes, IIMs and MXenes. Numerous researches have reported positive results for metal recovery by membranes. For PEUF and MEUF, selecting appropriate complexing/micelle agents is vital to achieving high removal efficacy of metal ions and ensuring feasibility without generating excessive releases of secondary pollutants. Adsorptive UF MMM is suitable for treating low concentrations of metals by enabling the complete filtration-adsorption removal of metal ions to occur at low membrane pressure. ENFM is an environmentally friendly membrane and shows abilities for metal removal, but it is still necessary to address the issues related to the adsorption mechanism. IIM needs to further improve its ion recognition ability and usability, and develop more environmentally friendly regeneration methods. Similarly, MOF-based/modified membranes and MXene also need to address their regeneration ability and stability, as well as potential secondary environmental issues caused by the detachment and agglomeration of nanoparticles. These promising technologies, if their limitations and performance are improved while reducing usage costs, can be rapidly applied in the industrial field and eventually become a universal and sustainable method for metal ion recovery in the future.

Author Contributions

Writing—original draft preparation, Y.F.; writing—review and editing, R.W.; supervision, R.W.; project administration, R.W.; funding acquisition, R.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Shandong Province (ZR2018MEE048) and Research Exchange Project of the China-Slovak Scientific & Technical Cooperation Committee (2024 No. 9-2).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Figueira, M.; Rodríguez-Jiménez, D.; López, J.; Reig, M.; Cortina, J.L.; Valderrama, C. Experimental and Economic Evaluation of Nanofiltration as a Pre-Treatment for Added-Value Elements Recovery from Seawater Desalination Brines. Desalination 2023, 549, 116321. [Google Scholar] [CrossRef]
  2. Zhu, F.; Zheng, Y.-M.; Zhang, B.-G.; Dai, Y.-R. A Critical Review on the Electrospun Nanofibrous Membranes for the Adsorption of Heavy Metals in Water Treatment. J. Hazard. Mater. 2021, 401, 123608. [Google Scholar] [CrossRef] [PubMed]
  3. Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. Simultaneous Removal of Organics and Heavy Metals from Industrial Wastewater: A Review. Chemosphere 2021, 262, 128379. [Google Scholar] [CrossRef]
  4. Abdullah, N.; Yusof, N.; Lau, W.J.; Jaafar, J.; Ismail, A.F. Recent Trends of Heavy Metal Removal from Water/Wastewater by Membrane Technologies. J. Ind. Eng. Chem. 2019, 76, 17–38. [Google Scholar] [CrossRef]
  5. Bashir, S.; Ali, U.; Shaaban, M.; Gulshan, A.B.; Iqbal, J.; Khan, S.; Husain, A.; Ahmed, N.; Mehmood, S.; Kamran, M.; et al. Role of Sepiolite for Cadmium (Cd) Polluted Soil Restoration and Spinach Growth in Wastewater Irrigated Agricultural Soil. J. Environ. Manag. 2020, 258, 110020. [Google Scholar] [CrossRef] [PubMed]
  6. Cundy, A.B.; Hopkinson, L.; Whitby, R.L.D. Use of Iron-Based Technologies in Contaminated Land and Groundwater Remediation: A Review. Sci. Total Environ. 2008, 400, 42–51. [Google Scholar] [CrossRef] [PubMed]
  7. Arana Juve, J.-M.; Christensen, F.M.S.; Wang, Y.; Wei, Z. Electrodialysis for Metal Removal and Recovery: A Review. Chem. Eng. J. 2022, 435, 134857. [Google Scholar] [CrossRef]
  8. Fu, F.; Wang, Q. Removal of Heavy Metal Ions from Wastewaters: A Review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, X.; Zhao, Y.; Zhang, C.; Zhang, D.; Yao, C.; Meng, Q.; Zhao, R.; Wei, Z. Speciation, Toxicity Mechanism and Remediation Ways of Heavy Metals during Composting: A Novel Theoretical Microbial Remediation Method Is Proposed. J. Environ. Manag. 2020, 272, 111109. [Google Scholar] [CrossRef]
  10. Kiran Marella, T.; Saxena, A.; Tiwari, A. Diatom Mediated Heavy Metal Remediation: A Review. Bioresour. Technol. 2020, 305, 123068. [Google Scholar] [CrossRef]
  11. Agarwal, A.; Upadhyay, U.; Sreedhar, I.; Singh, S.A.; Patel, C.M. A Review on Valorization of Biomass in Heavy Metal Removal from Wastewater. J. Water Process Eng. 2020, 38, 101602. [Google Scholar] [CrossRef]
  12. Hasanpour, M.; Hatami, M. Application of Three Dimensional Porous Aerogels as Adsorbent for Removal of Heavy Metal Ions from Water/Wastewater: A Review Study. Adv. Colloid Interface Sci. 2020, 284, 102247. [Google Scholar] [CrossRef]
  13. Saravanan, A.; Senthil Kumar, P.; Jeevanantham, S.; Karishma, S.; Tajsabreen, B.; Yaashikaa, P.R.; Reshma, B. Effective Water/Wastewater Treatment Methodologies for Toxic Pollutants Removal: Processes and Applications towards Sustainable Development. Chemosphere 2021, 280, 130595. [Google Scholar] [CrossRef]
  14. Zhu, S.; Yang, N.; Zhang, D. Poly(N,N-Dimethylaminoethyl Methacrylate) Modification of Activated Carbon for Copper Ions Removal. Mater. Chem. Phys. 2009, 113, 784–789. [Google Scholar] [CrossRef]
  15. Egashira, R.; Tanabe, S.; Habaki, H. Adsorption of Heavy Metals in Mine Wastewater by Mongolian Natural Zeolite. Procedia Eng. 2012, 42, 49–57. [Google Scholar] [CrossRef]
  16. Luo, L.L.; Gu, X.X.; Wu, J.; Zhong, S.X.; Chen, J.R. Advances in Graphene for Adsorption of Heavy Metals in Wastewater. Adv. Mater. Res. 2012, 550–553, 2121–2124. [Google Scholar] [CrossRef]
  17. Carolin, C.F.; Kumar, P.S.; Saravanan, A.; Joshiba, G.J.; Naushad, M. Efficient Techniques for the Removal of Toxic Heavy Metals from Aquatic Environment: A Review. J. Environ. Chem. Eng. 2017, 5, 2782–2799. [Google Scholar] [CrossRef]
  18. Hansda, A.; Kumar, V.; Anshumali. A Comparative Review towards Potential of Microbial Cells for Heavy Metal Removal with Emphasis on Biosorption and Bioaccumulation. World J. Microbiol. Biotechnol. 2016, 32, 170. [Google Scholar] [CrossRef] [PubMed]
  19. Tesh, S.J.; Scott, T.B. Nano-Composites for Water Remediation: A Review. Adv. Mater. 2014, 26, 6056–6068. [Google Scholar] [CrossRef]
  20. Abdullah, N.; Gohari, R.J.; Yusof, N.; Ismail, A.F.; Juhana, J.; Lau, W.J.; Matsuura, T. Polysulfone/Hydrous Ferric Oxide Ultrafiltration Mixed Matrix Membrane: Preparation, Characterization and Its Adsorptive Removal of Lead (II) from Aqueous Solution. Chem. Eng. J. 2016, 289, 28–37. [Google Scholar] [CrossRef]
  21. Abdullah, N.; Yusof, N.; Gohari, R.J.; Ismail, A.F.; Jaafar, J.; Lau, W.J.; Misdan, N.; Hairom, N.H.H. Characterizations of Polysulfone/Ferrihydrite Mixed Matrix Membranes for Water/Wastewater Treatment. Water Environ. Res. 2018, 90, 64–73. [Google Scholar] [CrossRef] [PubMed]
  22. Jamshidi Gohari, R.; Lau, W.J.; Matsuura, T.; Ismail, A.F. Fabrication and Characterization of Novel PES/Fe–Mn Binary Oxide UF Mixed Matrix Membrane for Adsorptive Removal of As(III) from Contaminated Water Solution. Sep. Purif. Technol. 2013, 118, 64–72. [Google Scholar] [CrossRef]
  23. Salehi, E.; Madaeni, S.S.; Rajabi, L.; Vatanpour, V.; Derakhshan, A.A.; Zinadini, S.; Ghorabi, S.; Ahmadi Monfared, H. Novel Chitosan/Poly(Vinyl) Alcohol Thin Adsorptive Membranes Modified with Amino Functionalized Multi-Walled Carbon Nanotubes for Cu(II) Removal from Water: Preparation, Characterization, Adsorption Kinetics and Thermodynamics. Sep. Purif. Technol. 2012, 89, 309–319. [Google Scholar] [CrossRef]
  24. Zhang, X.; Fang, X.; Li, J.; Pan, S.; Sun, X.; Shen, J.; Han, W.; Wang, L.; Zhao, S. Developing New Adsorptive Membrane by Modification of Support Layer with Iron Oxide Microspheres for Arsenic Removal. J. Colloid Interface Sci. 2018, 514, 760–768. [Google Scholar] [CrossRef] [PubMed]
  25. Chitpong, N.; Husson, S.M. High-Capacity, Nanofiber-Based Ion-Exchange Membranes for the Selective Recovery of Heavy Metals from Impaired Waters. Sep. Purif. Technol. 2017, 179, 94–103. [Google Scholar] [CrossRef]
  26. Kumar, M.; Shevate, R.; Hilke, R.; Peinemann, K.-V. Novel Adsorptive Ultrafiltration Membranes Derived from Polyvinyltetrazole-Co-Polyacrylonitrile for Cu(II) Ions Removal. Chem. Eng. J. 2016, 301, 306–314. [Google Scholar] [CrossRef]
  27. Mukherjee, R.; Bhunia, P.; De, S. Impact of Graphene Oxide on Removal of Heavy Metals Using Mixed Matrix Membrane. Chem. Eng. J. 2016, 292, 284–297. [Google Scholar] [CrossRef]
  28. Petrov, S.; Nenov, V. Removal and Recovery of Copper from Wastewater by a Complexation-Ultrafiltration Process. Desalination 2004, 162, 201–209. [Google Scholar] [CrossRef]
  29. Barakat, M.A.; Schmidt, E. Polymer-Enhanced Ultrafiltration Process for Heavy Metals Removal from Industrial Wastewater. Desalination 2010, 256, 90–93. [Google Scholar] [CrossRef]
  30. Huang, Y.; Feng, X. Polymer-Enhanced Ultrafiltration: Fundamentals, Applications and Recent Developments. J. Membr. Sci. 2019, 586, 53–83. [Google Scholar] [CrossRef]
  31. Qiu, Y.-R.; Mao, L.-J. Removal of Heavy Metal Ions from Aqueous Solution by Ultrafiltration Assisted with Copolymer of Maleic Acid and Acrylic Acid. Desalination 2013, 329, 78–85. [Google Scholar] [CrossRef]
  32. Chen, M.; Jafvert, C.T.; Wu, Y.; Cao, X.; Hankins, N.P. Inorganic Anion Removal Using Micellar Enhanced Ultrafiltration (MEUF), Modeling Anion Distribution and Suggested Improvements of MEUF: A Review. Chem. Eng. J. 2020, 398, 125413. [Google Scholar] [CrossRef]
  33. Salehi, E.; Daraei, P.; Arabi Shamsabadi, A. A Review on Chitosan-Based Adsorptive Membranes. Carbohydr. Polym. 2016, 152, 419–432. [Google Scholar] [CrossRef] [PubMed]
  34. Dalida, M.L.P.; Mariano, A.F.V.; Futalan, C.M.; Kan, C.-C.; Tsai, W.-C.; Wan, M.-W. Adsorptive Removal of Cu(II) from Aqueous Solutions Using Non-Crosslinked and Crosslinked Chitosan-Coated Bentonite Beads. Desalination 2011, 275, 154–159. [Google Scholar] [CrossRef]
  35. Wan, M.-W.; Kan, C.-C.; Rogel, B.D.; Dalida, M.L.P. Adsorption of Copper (II) and Lead (II) Ions from Aqueous Solution on Chitosan-Coated Sand. Carbohydr. Polym. 2010, 80, 891–899. [Google Scholar] [CrossRef]
  36. Indiasih, P.A.Y.; Syaifudin, A.D.; Himma, N.F.; Ismuyanto, B. Removal of Hexavalent Chromium from Water by Chitosan-Enhanced Ultrafiltration. Makara J. Technol. 2023, 27, 3. [Google Scholar] [CrossRef]
  37. Nikhil Chandra, P.; Nair, K.K. Functionalization of Tannic Acid and Campesterol with Polyelectrolyte Multilayer Membranes for Lithium Ions Separation and Recycling. Clean. Mater. 2021, 1, 100018. [Google Scholar] [CrossRef]
  38. Edward, K.; Dalmia, M.; Samuel, A.M.; Sivaraman, P.; Rajesh, M.P. Optimization of Process Conditions for the Removal of Zinc and Lead by Ultrafiltration Using Biopolymer. Chem. Pap. 2021, 75, 3723–3737. [Google Scholar] [CrossRef]
  39. Jahani, Z.; Mosaffa, E.; Oroujzadeh, M.; Ghafuri, H. Performance Evaluation of Polyethersulfone Membranes Modified with Poly (Acrylic Acid-Co-N-Vinyl Pyrrolidone) Grafted Mesoporous Carbon Nitride for Effective Removal of Cadmium(II) from Wastewater. Polym. Adv. Technol. 2023, 34, 3803–3817. [Google Scholar] [CrossRef]
  40. Purushothaman, M.; Harikrishnan, A.; Senthil Kumar, P.; George, J.; Rangasamy, G.; Vaidyanathan, V.K. Enhancement of Antifouling Properties, Metal Ions and Protein Separation of Poly(Ether-Ether-Sulfone) Ultrafiltration Membranes by Incorporation of Poly Ethylene Glycol and n-ZnO. Environ. Res. 2023, 216, 114696. [Google Scholar] [CrossRef]
  41. Dhume, S.; Mahajan-Tatpate, P.; Chendake, Y.; Chavan, S. Separation of Heavy Metals from Metal Industry Effluent for Acid Recovery. In Recent Advances in Recycling Engineering; Siddiqui, N.A., Baxtiyarovich, A.S., Nandan, A., Mondal, P., Eds.; Springer Nature: Singapore, 2023; pp. 207–217. [Google Scholar]
  42. Yin, N.; Wang, K.; Wang, L.; Li, Z. Amino-Functionalized MOFs Combining Ceramic Membrane Ultrafiltration for Pb (II) Removal. Chem. Eng. J. 2016, 306, 619–628. [Google Scholar] [CrossRef]
  43. Senol-Arslan, D.; Gül, A.; Uzal, N.; Yavuz, E. Ni-Zn Metal-Organic Framework Based Membranes for Rejection of Pb(II) Ions. Inorg. Chem. Commun. 2022, 146, 110084. [Google Scholar] [CrossRef]
  44. Gokcek, O.B.; Uzal, N. Arsenic Removal by the Micellar-Enhanced Ultrafiltration Using Response Surface Methodology. Water Supply 2019, 20, 574–585. [Google Scholar] [CrossRef]
  45. Köker, B.; Cebeci, M.S. Removal of Cr(VI) from Tanning Wastewater Using Chitosan-SDS Complexes in PEUF: Optimization and Analysis via Response Surface Methodology. J. Water Process Eng. 2023, 54, 103966. [Google Scholar] [CrossRef]
  46. Wołowicz, A.; Staszak, K.; Hubicki, Z. Removal of Copper(II) in the Presence of Sodium Dodecylobenzene Sulfonate from Acidic Effluents Using Adsorption on Ion Exchangers and Micellar-Enhanced Ultrafiltration Methods. Molecules 2022, 27, 2430. [Google Scholar] [CrossRef]
  47. Yu, T.; Zhou, J.; Liu, F.; Xu, B.-M.; Pan, Y. Recent Progress of Adsorptive Ultrafiltration Membranes in Water Treatment—A Mini Review. Membranes 2022, 12, 519. [Google Scholar] [CrossRef] [PubMed]
  48. Usman Farid, M.; Luan, H.-Y.; Wang, Y.; Huang, H.; An, A.K.; Jalil Khan, R. Increased Adsorption of Aqueous Zinc Species by Ar/O2 Plasma-Treated Carbon Nanotubes Immobilized in Hollow-Fiber Ultrafiltration Membrane. Chem. Eng. J. 2017, 325, 239–248. [Google Scholar] [CrossRef]
  49. Abedi, F.; Dubé, M.A.; Emadzadeh, D.; Kruczek, B. Improving Nanofiltration Performance Using Modified Cellulose Nanocrystal-Based TFN Membranes. J. Membr. Sci. 2023, 670, 121369. [Google Scholar] [CrossRef]
  50. Gao, J.; Sun, S.-P.; Zhu, W.-P.; Chung, T.-S. Chelating Polymer Modified P84 Nanofiltration (NF) Hollow Fiber Membranes for High Efficient Heavy Metal Removal. Water Res. 2014, 63, 252–261. [Google Scholar] [CrossRef]
  51. Xiang, H.; Min, X.; Tang, C.-J.; Sillanpää, M.; Zhao, F. Recent Advances in Membrane Filtration for Heavy Metal Removal from Wastewater: A Mini Review. J. Water Process Eng. 2022, 49, 103023. [Google Scholar] [CrossRef]
  52. Liu, M.; Nothling, M.D.; Tan, S.S.L.; Webley, P.A.; Qiao, G.G.; Fu, Q. Polyrotaxane-Based Thin Film Composite Membranes for Enhanced Nanofiltration Performance. Sep. Purif. Technol. 2020, 246, 116893. [Google Scholar] [CrossRef]
  53. Shao, W.; Liu, C.; Yu, T.; Xiong, Y.; Hong, Z.; Xie, Q. Constructing Positively Charged Thin-Film Nanocomposite Nanofiltration Membranes with Enhanced Performance. Polymers 2020, 12, 2526. [Google Scholar] [CrossRef] [PubMed]
  54. Abdullah, N.; Yusof, N.; Jye, L.W.; Jaafar, J.; Misdan, N.; Ismail, A.F. Removal of Lead(II) by Nanofiltration-Ranged Thin Film Nanocomposite Membrane Incorporated UiO-66-NH2: Comparative Removal Performance between Hydraulic-Driven and Osmotic-Driven Membrane Process. J. Taiwan Inst. Chem. Eng. 2021, 128, 354–369. [Google Scholar] [CrossRef]
  55. Bhowmick, K.; Roy, S.; Bhowmick, S.; Jain, R.; Majumdar, S.; Sahoo, G.C.; Mondal, P. Fabrication of Robust Ceramic Supported Polymeric Composite Nanofiltration Membrane for Heavy Metal Contaminated Wastewater Treatment. J. Water Process Eng. 2022, 50, 103287. [Google Scholar] [CrossRef]
  56. Roy, S.; Bhowmick, K.; Singh, P.; Bhowmick, S.; Mukherjee, M.; Majumdar, S.; Sahoo, G.C.; Mondal, P. Removal of Heavy Metals by Surface Tailored Copper Ion Enhanced Ceramic-Supported-Polymeric Composite Nanofiltration Membrane. J. Environ. Chem. Eng. 2021, 9, 106368. [Google Scholar] [CrossRef]
  57. Wang, Z.; Wang, X.; Zheng, T.; Mo, B.; Xu, H.; Huang, Y.; Wang, J.; Gao, C.; Gao, X. High Flux Nanofiltration Membranes with Double-Walled Carbon Nanotube (DWCNT) as the Interlayer. Membranes 2022, 12, 1011. [Google Scholar] [CrossRef] [PubMed]
  58. He, M.; Wang, L.; Lv, Y.; Wang, X.; Zhu, J.; Zhang, Y.; Liu, T. Novel Polydopamine/Metal Organic Framework Thin Film Nanocomposite Forward Osmosis Membrane for Salt Rejection and Heavy Metal Removal. Chem. Eng. J. 2020, 389, 124452. [Google Scholar] [CrossRef]
  59. Morgante, C.; Ma, X.; Chen, X.; Wang, D.; Boffa, V.; Stathopoulos, V.; Lopez, J.; Cortina, J.L.; Cipollina, A.; Tamburini, A.; et al. Metal-Organic-Framework-Based Nanofiltration Membranes for Selective Multi-Cationic Recovery from Seawater and Brines. J. Membr. Sci. 2023, 685, 121941. [Google Scholar] [CrossRef]
  60. Song, C.; Jiao, Y.; Tian, Z.; Qin, Q.; Qin, S.; Zhang, J.; Li, J.; Cui, Z. Fabrication of Hollow-Fiber Nanofiltration Membrane with Negative-Positive Dual-Charged Separation Layer to Remove Low Concentration CuSO4. Sep. Purif. Technol. 2022, 296, 121352. [Google Scholar] [CrossRef]
  61. Yang, Y.; Wang, S.; Zhang, J.; He, B.; Li, J.; Qin, S.; Yang, J.; Zhang, J.; Cui, Z. Fabrication of Hollow Fiber Nanofiltration Separation Layer with Highly Positively Charged Surface for Heavy Metal Ion Removal. J. Membr. Sci. 2022, 653, 120534. [Google Scholar] [CrossRef]
  62. Hosseini, S.M.; Moradi, F.; Farahani, S.K.; Bandehali, S.; Parvizian, F.; Ebrahimi, M.; Shen, J. Carbon Nanofibers/Chitosan Nanocomposite Thin Film for Surface Modification of Poly(Ether Sulphone) Nanofiltration Membrane. Mater. Chem. Phys. 2021, 269, 124720. [Google Scholar] [CrossRef]
  63. Xu, S.-J.; Shen, Q.; Luo, L.-H.; Tong, Y.-H.; Wu, Y.-Z.; Xu, Z.-L.; Zhang, H.-Z. Surfactants Attached Thin Film Composite (TFC) Nanofiltration (NF) Membrane via Intermolecular Interaction for Heavy Metals Removal. J. Membr. Sci. 2022, 642, 119930. [Google Scholar] [CrossRef]
  64. Thabo, B.; Okoli, B.J.; Modise, S.J.; Nelana, S. Rejection Capacity of Nanofiltration Membranes for Nickel, Copper, Silver and Palladium at Various Oxidation States. Membranes 2021, 11, 653. [Google Scholar] [CrossRef]
  65. Roy, S.; Majumdar, S.; Sahoo, G.C.; Bhowmick, S.; Kundu, A.K.; Mondal, P. Removal of As(V), Cr(VI) and Cu(II) Using Novel Amine Functionalized Composite Nanofiltration Membranes Fabricated on Ceramic Tubular Substrate. J. Hazard. Mater. 2020, 399, 122841. [Google Scholar] [CrossRef] [PubMed]
  66. Arouche, T.d.S.; Cavaleiro, R.M.d.S.; Tanoue, P.S.M.; Pereira, T.S.d.S.; Filho, T.A.; Neto, A.M.d.J.C. Heavy Metals Nanofiltration Using Nanotube and Electric Field by Molecular Dynamics. J. Nanomater. 2020, 2020, e4063201. [Google Scholar] [CrossRef]
  67. Fonseka, C.; Ryu, S.; Naidu, G.; Kandasamy, J.; Vigneswaran, S. Recovery of Water and Valuable Metals Using Low Pressure Nanofiltration and Sequential Adsorption from Acid Mine Drainage. Environ. Technol. Innov. 2022, 28, 102753. [Google Scholar] [CrossRef]
  68. Fang, W.; Shi, L.; Wang, R. Interfacially Polymerized Composite Nanofiltration Hollow Fiber Membranes for Low-Pressure Water Softening. J. Membr. Sci. 2013, 430, 129–139. [Google Scholar] [CrossRef]
  69. Feng, C.; Xu, J.; Li, M.; Tang, Y.; Gao, C. Studies on a Novel Nanofiltration Membrane Prepared by Cross-Linking of Polyethyleneimine on Polyacrylonitrile Substrate. J. Membr. Sci. 2014, 451, 103–110. [Google Scholar] [CrossRef]
  70. Li, L.; Baoguo, W.; Huimin, T.; Tianlu, C.; Jiping, X. A Novel Nanofiltration Membrane Prepared with PAMAM and TMC by in Situ Interfacial Polymerization on PEK-C Ultrafiltration Membrane. J. Membr. Sci. 2006, 269, 84–93. [Google Scholar] [CrossRef]
  71. Li, X.; Zhu, L.; Jiang, J.; Yi, Z.; Zhu, B.; Xu, Y. Hydrophilic Nanofiltration Membranes with Self-Polymerized and Strongly-Adhered Polydopamine as Separating Layer. Chin. J. Polym. Sci. 2012, 30, 152–163. [Google Scholar] [CrossRef]
  72. Liu, C.; Shi, L.; Wang, R. Crosslinked Layer-by-Layer Polyelectrolyte Nanofiltration Hollow Fiber Membrane for Low-Pressure Water Softening with the Presence of SO42− in Feed Water. J. Membr. Sci. 2015, 486, 169–176. [Google Scholar] [CrossRef]
  73. Wu, C.; Liu, S.; Wang, Z.; Zhang, J.; Wang, X.; Lu, X.; Jia, Y.; Hung, W.-S.; Lee, K.-R. Nanofiltration Membranes with Dually Charged Composite Layer Exhibiting Super-High Multivalent-Salt Rejection. J. Membr. Sci. 2016, 517, 64–72. [Google Scholar] [CrossRef]
  74. Wu, D.; Huang, Y.; Yu, S.; Lawless, D.; Feng, X. Thin Film Composite Nanofiltration Membranes Assembled Layer-by-Layer via Interfacial Polymerization from Polyethylenimine and Trimesoyl Chloride. J. Membr. Sci. 2014, 472, 141–153. [Google Scholar] [CrossRef]
  75. Xu, Y.C.; Wang, Z.X.; Cheng, X.Q.; Xiao, Y.C.; Shao, L. Positively Charged Nanofiltration Membranes via Economically Mussel-Substance-Simulated Co-Deposition for Textile Wastewater Treatment. Chem. Eng. J. 2016, 303, 555–564. [Google Scholar] [CrossRef]
  76. Zhang, H.-Z.; Xu, Z.-L.; Ding, H.; Tang, Y.-J. Positively Charged Capillary Nanofiltration Membrane with High Rejection for Mg2+ and Ca2+ and Good Separation for Mg2+ and Li+. Desalination 2017, 420, 158–166. [Google Scholar] [CrossRef]
  77. Abdulhamid, M.A.; Muzamil, K. Recent Progress on Electrospun Nanofibrous Polymer Membranes for Water and Air Purification: A Review. Chemosphere 2023, 310, 136886. [Google Scholar] [CrossRef] [PubMed]
  78. Liao, Y.; Loh, C.-H.; Tian, M.; Wang, R.; Fane, A.G. Progress in Electrospun Polymeric Nanofibrous Membranes for Water Treatment: Fabrication, Modification and Applications. Prog. Polym. Sci. 2018, 77, 69–94. [Google Scholar] [CrossRef]
  79. Nauman, S.; Lubineau, G.; Alharbi, H.F. Post Processing Strategies for the Enhancement of Mechanical Properties of ENMs (Electrospun Nanofibrous Membranes): A Review. Membranes 2021, 11, 39. [Google Scholar] [CrossRef]
  80. Uddin, Z.; Ahmad, F.; Ullan, T.; Nawab, Y.; Ahmad, S.; Azam, F.; Rasheed, A.; Zafar, M.S. Recent Trends in Water Purification Using Electrospun Nanofibrous Membranes. Int. J. Environ. Sci. Technol. 2022, 19, 9149–9176. [Google Scholar] [CrossRef]
  81. Sakib, M.N.; Mallik, A.K.; Rahman, M.M. Update on Chitosan-Based Electrospun Nanofibers for Wastewater Treatment: A Review. Carbohydr. Polym. Technol. Appl. 2021, 2, 100064. [Google Scholar] [CrossRef]
  82. Liu, X.; Jiang, B.; Yin, X.; Ma, H.; Hsiao, B.S. Highly Permeable Nanofibrous Composite Microfiltration Membranes for Removal of Nanoparticles and Heavy Metal Ions. Sep. Purif. Technol. 2020, 233, 115976. [Google Scholar] [CrossRef]
  83. 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]
  84. Mohammed, Y.A.Y.A.; Abdel-Mohsen, A.M.; Zhang, Q.-J.; Younas, M.; Zhong, L.-B.; Yang, J.-C.E.; Zheng, Y.-M. Facile Synthesis of ZIF-8 Incorporated Electrospun PAN/PEI Nanofibrous Composite Membrane for Efficient Cr(VI) Adsorption from Water. Chem. Eng. J. 2023, 461, 141972. [Google Scholar] [CrossRef]
  85. Yin, X.; Zhang, Z.; Ma, H.; Venkateswaran, S.; Hsiao, B.S. Ultra-Fine Electrospun Nanofibrous Membranes for Multicomponent Wastewater Treatment: Filtration and Adsorption. Sep. Purif. Technol. 2020, 242, 116794. [Google Scholar] [CrossRef]
  86. Li, X.-H.; Yang, H.-J.; Ma, Y.-X.; Li, T.-Z.; Meng, W.-L.; Zhong, X.-M. A Novel PAN/GO Electrospun Nanocomposite Fibrous Membranes with Rich Amino Groups for Highly Efficient Adsorption of Au(III). Diam. Relat. Mater. 2023, 137, 110175. [Google Scholar] [CrossRef]
  87. Dickey, F.H. The Preparation of Specific Adsorbents. Available online: https://www.pnas.org/doi/10.1073/pnas.35.5.227 (accessed on 25 March 2024).
  88. Nishide, H.; Tsuchida, E. Selective Adsorption of Metal Ions on Poly(4-Vinylpyridine) Resins in Which the Ligand Chain Is Immobilized by Crosslinking. Makromol. Chem. 1976, 177, 2295–2310. [Google Scholar] [CrossRef]
  89. Zhou, X.; Wang, B.; Wang, R. Insights into Ion-Imprinted Materials for the Recovery of Metal Ions: Preparation, Evaluation and Application. Sep. Purif. Technol. 2022, 298, 121469. [Google Scholar] [CrossRef]
  90. Lu, J.; Qin, Y.; Wu, Y.; Meng, M.; Yan, Y.; Li, C. Recent Advances in Ion-Imprinted Membranes: Separation and Detection via Ion-Selective Recognition. Environ. Sci. Water Res. Technol. 2019, 5, 1626–1653. [Google Scholar] [CrossRef]
  91. Yang, J.; Qu, G.; Liu, C.; Zhou, S.; Li, B.; Wei, Y. An Effective Lithium Ion-Imprinted Membrane Containing 12-Crown Ether-4 for Selective Recovery of Lithium. Chem. Eng. Res. Des. 2022, 184, 639–650. [Google Scholar] [CrossRef]
  92. He, G.; Li, Z.; Liu, Y.; Liu, M.; Zhu, C.; Zhang, L.; Zhang, H. A Novel Lithium Ion-Imprinted Membrane with Robust Adsorption Capacity and Anti-Fouling Property Based on the Uniform Multilayered Interlayer. Desalination 2022, 539, 115973. [Google Scholar] [CrossRef]
  93. Qu, G.; Yang, J.; Ran, Y.; Li, B.; Wang, H.; Wei, Y. Adsorption Performance and Mechanism of TiO2/PVDF-Based Lithium-Ion Imprinted Membrane in Leaching Solution of Spent Lithium-Ion Batteries. J. Clean. Prod. 2024, 442, 140982. [Google Scholar] [CrossRef]
  94. Jin, K.; Huang, X.; Yang, H.; Li, Y.; Zeng, J.; Zhou, H.; Liu, Y.; Zhang, R. A Acylthiourea Based Ion-Imprinted Membrane for Selective Removal of Ag+ from Aqueous Solution. Colloids Surf. Physicochem. Eng. Asp. 2024, 684, 133162. [Google Scholar] [CrossRef]
  95. Zhang, M.; Hao, H.; Shi, C.; Chen, C.; Zhou, G.; Wang, J.; Cao, Y.; Han, X. Bis-Terpyridine Imprinted Nanocage in the Confined Two-Dimensional Lamellar Membrane for Selective Adsorption of Nd(III). Sep. Purif. Technol. 2023, 310, 123095. [Google Scholar] [CrossRef]
  96. Zhang, Y.; Fan, Z.; Hua, J.; Gao, X.; Lin, L.; Li, Z. Separation and Recovery of Nd(III) via Adsorption on Three-Dimensional Wood-Based Imprinted Chitosan Composite Membranes. Sep. Purif. Technol. 2024, 333, 125924. [Google Scholar] [CrossRef]
  97. Mokhtar, M.; Dickson, S.E.; Kim, Y.; Mekky, W. Preparation and Characterization of Ion Selective Membrane and Its Application for Cu2+ Removal. J. Ind. Eng. Chem. 2018, 60, 475–484. [Google Scholar] [CrossRef]
  98. Wang, Z.; Kong, D.; Qiao, N.; Wang, N.; Wang, Q.; Liu, H.; Zhou, Z.; Ren, Z. Facile Preparation of Novel Layer-by-Layer Surface Ion-Imprinted Composite Membrane for Separation of Cu2+ from Aqueous Solution. Appl. Surf. Sci. 2018, 457, 981–990. [Google Scholar] [CrossRef]
  99. Yu, S.; Tang, H.; Zhang, D.; Wang, S.; Qiu, M.; Song, G.; Fu, D.; Hu, B.; Wang, X. MXenes as Emerging Nanomaterials in Water Purification and Environmental Remediation. Sci. Total Environ. 2022, 811, 152280. [Google Scholar] [CrossRef]
  100. Zou, J.; Wu, J.; Wang, Y.; Deng, F.; Jiang, J.; Zhang, Y.; Liu, S.; Li, N.; Zhang, H.; Yu, J.; et al. Additive-Mediated Intercalation and Surface Modification of MXenes. Chem. Soc. Rev. 2022, 51, 2972–2990. [Google Scholar] [CrossRef]
  101. Gu, P.; Liu, S.; Cheng, X.; Zhang, S.; Wu, C.; Wen, T.; Wang, X. Recent Strategies, Progress, and Prospects of Two-Dimensional Metal Carbides (MXenes) Materials in Wastewater Purification: A Review. Sci. Total Environ. 2024, 912, 169533. [Google Scholar] [CrossRef]
  102. Gu, P.; Zhang, S.; Ma, R.; Sun, M.; Wang, S.; Wen, T.; Wang, X. Layered Double Hydroxides Nanosheets In-Situ Anchored on Ultrathin MXenes for Enhanced U(VI) and Eu(III) Trapping: Excavating from Selectivity to Mechanism. Sep. Purif. Technol. 2022, 288, 120641. [Google Scholar] [CrossRef]
  103. Jun, B.-M.; Her, N.; Park, C.M.; Yoon, Y. Effective Removal of Pb(II) from Synthetic Wastewater Using Ti3C2Tx MXene. Environ. Sci. Water Res. Technol. 2019, 6, 173–180. [Google Scholar] [CrossRef]
  104. Zhang, G.; Wang, T.; Xu, Z.; Liu, M.; Shen, C.; Meng, Q. Synthesis of Amino-Functionalized Ti3C2Tx MXene by Alkalization-Grafting Modification for Efficient Lead Adsorption. Chem. Commun. 2020, 56, 11283–11286. [Google Scholar] [CrossRef]
  105. Wang, S.; Liu, Y.; Lü, Q.-F.; Zhuang, H. Facile Preparation of Biosurfactant-Functionalized Ti2CTX MXene Nanosheets with an Enhanced Adsorption Performance for Pb(II) Ions. J. Mol. Liq. 2020, 297, 111810. [Google Scholar] [CrossRef]
  106. Shahzad, A.; Rasool, K.; Iqbal, J.; Jang, J.; Lim, Y.; Kim, B.; Oh, J.-M.; Lee, D.S. MXsorption of Mercury: Exceptional Reductive Behavior of Titanium Carbide/Carbonitride MXenes. Environ. Res. 2022, 205, 112532. [Google Scholar] [CrossRef]
  107. Dong, Y.; Sang, D.; He, C.; Sheng, X.; Lei, L. Mxene/Alginate Composites for Lead and Copper Ion Removal from Aqueous Solutions. RSC Adv. 2019, 9, 29015–29022. [Google Scholar] [CrossRef]
  108. Elumalai, S.; Yoshimura, M.; Ogawa, M. Simultaneous Delamination and Rutile Formation on the Surface of Ti3C2Tx MXene for Copper Adsorption. Chem.—Asian J. 2020, 15, 1044–1051. [Google Scholar] [CrossRef]
  109. Kong, A.; Sun, Y.; Peng, M.; Gu, H.; Fu, Y.; Zhang, J.; Li, W. Amino-Functionalized MXenes for Efficient Removal of Cr(VI). Colloids Surf. Physicochem. Eng. Asp. 2021, 617, 126388. [Google Scholar] [CrossRef]
  110. Liu, C.; Li, Y.; Liu, D.; Lei, M.; Li, B.; Fu, C.; Zhang, J. Constructing of a Stable Ti3C2 MXenes Nano-Traps with Ultrahigh and Selectively Uranyl Encapsulation Capability by Polyethyleneimine Intercalation. Chem. Eng. J. 2023, 468, 143746. [Google Scholar] [CrossRef]
  111. Shahzad, A.; Moztahida, M.; Tahir, K.; Kim, B.; Jeon, H.; Ghani, A.A.; Maile, N.; Jang, J.; Lee, D.S. Highly Effective Prussian Blue-Coated MXene Aerogel Spheres for Selective Removal of Cesium Ions. J. Nucl. Mater. 2020, 539, 152277. [Google Scholar] [CrossRef]
  112. Karmaker, S.C.; Eljamal, O.; Saha, B.B. Response Surface Methodology for Strontium Removal Process Optimization from Contaminated Water Using Zeolite Nanocomposites. Environ. Sci. Pollut. Res. 2021, 28, 56535–56551. [Google Scholar] [CrossRef]
  113. Dhanapal, R.; Ravindran, R.; Seethalakshmi, N.; Selvakumar, R. Surface Functionalized Diatomaceous Earth for Effective Adsorption of Strontium from Aqueous Solution. J. Radioanal. Nucl. Chem. 2019, 319, 1301–1306. [Google Scholar] [CrossRef]
  114. Jun, B.-M.; Park, C.M.; Heo, J.; Yoon, Y. Adsorption of Ba2+ and Sr2+ on Ti3C2Tx MXene in Model Fracking Wastewater. J. Environ. Manag. 2020, 256, 109940. [Google Scholar] [CrossRef]
  115. Fard, A.K.; Mckay, G.; Chamoun, R.; Rhadfi, T.; Preud’Homme, H.; Atieh, M.A. Barium Removal from Synthetic Natural and Produced Water Using MXene as Two Dimensional (2-D) Nanosheet Adsorbent. Chem. Eng. J. 2017, 317, 331–342. [Google Scholar] [CrossRef]
  116. Buonomenna, M.G.; Mousavi, S.M.; Hashemi, S.A.; Lai, C.W. Water Cleaning Adsorptive Membranes for Efficient Removal of Heavy Metals and Metalloids. Water 2022, 14, 2718. [Google Scholar] [CrossRef]
  117. Koli, M.M.; Singh, S.P. Surface-Modified Ultrafiltration and Nanofiltration Membranes for the Selective Removal of Heavy Metals and Inorganic Groundwater Contaminants: A Review. Environ. Sci. Water Res. Technol. 2023, 9, 2803–2829. [Google Scholar] [CrossRef]
  118. Xiong, J.; Hu, Q.; Wu, J.; Jia, Z.; Ge, S.; Cao, Y.; Zhou, J.; Wang, Y.; Yan, J.; Xie, L.; et al. Structurally Stable Electrospun Nanofibrous Cellulose Acetate/Chitosan Biocomposite Membranes for the Removal of Chromium Ions from the Polluted Water. Adv. Compos. Hybrid Mater. 2023, 6, 99. [Google Scholar] [CrossRef]
  119. Zheng, R.; Zhao, L.; Wang, W.; Liu, Y.; Ma, Q.; Mu, D.; Li, R.; Dai, C. Optimized Li and Fe Recovery from Spent Lithium-Ion Batteries via a Solution-Precipitation Method. RSC Adv. 2016, 6, 43613–43625. [Google Scholar] [CrossRef]
  120. Horeh, N.B.; Mousavi, S.M.; Shojaosadati, S.A. Bioleaching of Valuable Metals from Spent Lithium-Ion Mobile Phone Batteries Using Aspergillus Niger. J. Power Sources 2016, 320, 257–266. [Google Scholar] [CrossRef]
  121. Sun, D.; Zhou, T.; Lu, Y.; Yan, Y.; Liu, C.; Che, G. Ion-Imprinted Antifouling Nanocomposite Membrane for Separation of Lithium Ion. Korean J. Chem. Eng. 2022, 39, 2482–2490. [Google Scholar] [CrossRef]
  122. Yu, C.; Lu, J.; Dai, J.; Dong, Z.; Lin, X.; Xing, W.; Wu, Y.; Ma, Z. Bio-Inspired Fabrication of Ester-Functionalized Imprinted Composite Membrane for Rapid and High-Efficient Recovery of Lithium Ion from Seawater. J. Colloid Interface Sci. 2020, 572, 340–353. [Google Scholar] [CrossRef] [PubMed]
  123. Das, D.; Abarajitha, R.; Kay, P.; Ramamurthy, V.; Goycoolea, F.M.; Das, N. Selective Recovery of Lithium from Spent Coin Cell Cathode Leachates Using Ion Imprinted Blended Chitosan Microfibers: Pilot Scale Studies Provide Insights on Scalability. J. Hazard. Mater. 2022, 431, 128535. [Google Scholar] [CrossRef]
  124. Zhang, Z.; Jia, Y.; Liu, B.; Sun, H.; Jing, Y.; Zhang, Q.; Shao, F.; Qi, M.; Yao, Y. Extraction and Separation of Lithium Isotopes by Using Organic Liquid Film Extraction System of Crown Ether-Ionic Liquid. Fusion Eng. Des. 2020, 161, 112015. [Google Scholar] [CrossRef]
  125. Shi, C.; Jia, Y.; Xiao, J.; Wang, X.; Yao, Y.; Jing, Y. Lithium Isotope Separation by Liquid-Liquid Extraction Using Ionic Liquid System Containing Dibenzo-14-Crown-4. J. Mol. Liq. 2016, 224, 662–667. [Google Scholar] [CrossRef]
  126. Shi, L.; Shi, Y.; Xu, Y.; Cao, J.; Meng, Y.; Chen, H. Metal-Organic Framework Membranes with Varying Metal Ions for Enhanced Water and Wastewater Treatment: A Critical Review. J. Environ. Chem. Eng. 2023, 11, 111468. [Google Scholar] [CrossRef]
  127. Arabi Shamsabadi, A.; Sharifian, G.M.; Anasori, B.; Soroush, M. Antimicrobial Mode-of-Action of Colloidal Ti3C2Tx MXene Nanosheets. ACS Sustain. Chem. Eng. 2018, 6, 16586–16596. [Google Scholar] [CrossRef]
  128. George, S.M.; Kandasubramanian, B. Advancements in MXene-Polymer Composites for Various Biomedical Applications. Ceram. Int. 2020, 46, 8522–8535. [Google Scholar] [CrossRef]
  129. Lim, G.P.; Soon, C.F.; Ma, N.L.; Morsin, M.; Nayan, N.; Ahmad, M.K.; Tee, K.S. Cytotoxicity of MXene-Based Nanomaterials for Biomedical Applications: A Mini Review. Environ. Res. 2021, 201, 111592. [Google Scholar] [CrossRef]
Figure 1. Number of documents found using the keywords “membrane” and “adsorption”. Sources: Scopus from Elsevier (until 20 March 2024).
Figure 1. Number of documents found using the keywords “membrane” and “adsorption”. Sources: Scopus from Elsevier (until 20 March 2024).
Materials 17 03562 g001
Figure 2. Illustrated mechanism of heavy metals adsorption on adsorbents in UF MMMs. Reprinted from [4].
Figure 2. Illustrated mechanism of heavy metals adsorption on adsorbents in UF MMMs. Reprinted from [4].
Materials 17 03562 g002
Figure 3. (a) Illustration of PEUF. Reprinted from [30]. (b) Illustration of MEUF. Reprinted from [32].
Figure 3. (a) Illustration of PEUF. Reprinted from [30]. (b) Illustration of MEUF. Reprinted from [32].
Materials 17 03562 g003
Figure 4. (a) Effect of the chitosan/Cr(VI) ratio on PEUF Performance. (b) Feed and permeate solutions obtained after PEUF using chitosan/Cr(VI) ratio of 12.5. (c) Schematic illustration of chitosan-Cr(VI) complexation and the enhanced removal in the UF membrane process. (d) Permeate flux before and after backwashing in PEUF for 200 min at 1 bar pressure. The insets show the contact angle of water droplet on membrane surface. Reprinted from [36].
Figure 4. (a) Effect of the chitosan/Cr(VI) ratio on PEUF Performance. (b) Feed and permeate solutions obtained after PEUF using chitosan/Cr(VI) ratio of 12.5. (c) Schematic illustration of chitosan-Cr(VI) complexation and the enhanced removal in the UF membrane process. (d) Permeate flux before and after backwashing in PEUF for 200 min at 1 bar pressure. The insets show the contact angle of water droplet on membrane surface. Reprinted from [36].
Materials 17 03562 g004
Figure 5. Adsorption mechanisms of Pb(II) onto the MOFs (a) FT-IR; (b) XRD patterns; (c) survey XPS; (d) N (1s); and (e) Pb (4f) spectra before and after adsorption. Reprinted from [42].
Figure 5. Adsorption mechanisms of Pb(II) onto the MOFs (a) FT-IR; (b) XRD patterns; (c) survey XPS; (d) N (1s); and (e) Pb (4f) spectra before and after adsorption. Reprinted from [42].
Materials 17 03562 g005
Figure 6. (a) AFM images of freeze-dried CNCs, ACNCs and CysCNCs. (b) ATR-FTIR spectra of TFC and TFN membranes in 0.1 wt% CNCs, ACNCs and CysCNCs. (c) Long-term performance test of TFC, CNC-TFN0.1 and CysCNC-TFN0.1 membranes within 200 ppm CuSO4 solution as a feed. Left arrow: water permeability of the three bottom lines; Right arrow: rejection rate of CuSO4 on the top three broken lines. Reprinted from [49].
Figure 6. (a) AFM images of freeze-dried CNCs, ACNCs and CysCNCs. (b) ATR-FTIR spectra of TFC and TFN membranes in 0.1 wt% CNCs, ACNCs and CysCNCs. (c) Long-term performance test of TFC, CNC-TFN0.1 and CysCNC-TFN0.1 membranes within 200 ppm CuSO4 solution as a feed. Left arrow: water permeability of the three bottom lines; Right arrow: rejection rate of CuSO4 on the top three broken lines. Reprinted from [49].
Materials 17 03562 g006
Figure 7. NH2-MIL-101(Al) powder: (a) low-magnification and high-magnification (in the insert) SEM micrographs, and (b) nitrogen adsorption/desorption isotherms (77 K). (c) FTIR analysis of chitosan, NH2-MIL101(Al), ZnO, and CMZ-35 nanocomposite. (d) EDX analysis of the CMZ-35 membrane. Reprinted from [59].
Figure 7. NH2-MIL-101(Al) powder: (a) low-magnification and high-magnification (in the insert) SEM micrographs, and (b) nitrogen adsorption/desorption isotherms (77 K). (c) FTIR analysis of chitosan, NH2-MIL101(Al), ZnO, and CMZ-35 nanocomposite. (d) EDX analysis of the CMZ-35 membrane. Reprinted from [59].
Materials 17 03562 g007
Figure 8. For article [59], the results are compared with Fang et al. (2013) [68], Feng et al. (2014) [69], Li et al. (2006) [70], Li et al. (2012) [71], Liu et al. (2015) [72], Wu et al. (2016) [73], Wu et al. (2014) [74], Xu et al. (2016) [75] and Zhang et al. (2017) [76]. And the salt couples: (a) MgCl2/NaCl, (b) CaCl2/NaCl, (c) MgCl2/Na2SO4, and (d) CaCl2/Na2SO4 at operating pressures about rejections for CMZ membranes in this work, commercial NF membranes, and positively charged NF membranes in the literature. Region of interest: high rejection values towards MgCl2 and CaCl2 with respect to NaCl and Na2SO4; Region of RO: typical rejection values for RO membranes. * Tested in their work.
Figure 8. For article [59], the results are compared with Fang et al. (2013) [68], Feng et al. (2014) [69], Li et al. (2006) [70], Li et al. (2012) [71], Liu et al. (2015) [72], Wu et al. (2016) [73], Wu et al. (2014) [74], Xu et al. (2016) [75] and Zhang et al. (2017) [76]. And the salt couples: (a) MgCl2/NaCl, (b) CaCl2/NaCl, (c) MgCl2/Na2SO4, and (d) CaCl2/Na2SO4 at operating pressures about rejections for CMZ membranes in this work, commercial NF membranes, and positively charged NF membranes in the literature. Region of interest: high rejection values towards MgCl2 and CaCl2 with respect to NaCl and Na2SO4; Region of RO: typical rejection values for RO membranes. * Tested in their work.
Materials 17 03562 g008
Figure 9. (a) Surface modification routes of PAN nanofibrous component. (b) Increase in the adsorption capacity (q) against Cr (VI) and Cd (II) ions with time using PVAm-g-PAN-PVA and hydrolyzed PAN-PVA composite membranes, respectively; and (c) corresponding pseudo-first order (logqe − qt) and pseudo-second order (t/qt) kinetics plots of the adsorption processes. Reprinted from [82].
Figure 9. (a) Surface modification routes of PAN nanofibrous component. (b) Increase in the adsorption capacity (q) against Cr (VI) and Cd (II) ions with time using PVAm-g-PAN-PVA and hydrolyzed PAN-PVA composite membranes, respectively; and (c) corresponding pseudo-first order (logqe − qt) and pseudo-second order (t/qt) kinetics plots of the adsorption processes. Reprinted from [82].
Materials 17 03562 g009
Figure 10. (a) Adsorption kinetics of Cr(VI) on Fe-MOFs particles and PAN/MOF ENFMs fitted by pseudo-first order model and pseudo-second order model; (b) Adsorption isotherms of Cr(VI) on Fe-MOFs particles and PAN/MOF ENFMs fitted by Freundlich model, Sips model and Redlich-Peterson model. Reprinted from [83].
Figure 10. (a) Adsorption kinetics of Cr(VI) on Fe-MOFs particles and PAN/MOF ENFMs fitted by pseudo-first order model and pseudo-second order model; (b) Adsorption isotherms of Cr(VI) on Fe-MOFs particles and PAN/MOF ENFMs fitted by Freundlich model, Sips model and Redlich-Peterson model. Reprinted from [83].
Materials 17 03562 g010
Figure 11. Preparation process and recognition mechanism of IIPs. Reprinted from [89].
Figure 11. Preparation process and recognition mechanism of IIPs. Reprinted from [89].
Materials 17 03562 g011
Figure 12. SEM images of (a,b) TiO2/PVDF, (d,e) LIIMs and EDS spectrum of (c) TiO2/PVDF and (f) LIIMs. (g) The nonlinear fitting curves of the Langmuir model and Freundlich model for the adsorption of LIIMs and NIIMs to Li+ (LiCl solution). (h) Schematic diagram of synthesis of LIIMs. Reprinted from [91].
Figure 12. SEM images of (a,b) TiO2/PVDF, (d,e) LIIMs and EDS spectrum of (c) TiO2/PVDF and (f) LIIMs. (g) The nonlinear fitting curves of the Langmuir model and Freundlich model for the adsorption of LIIMs and NIIMs to Li+ (LiCl solution). (h) Schematic diagram of synthesis of LIIMs. Reprinted from [91].
Materials 17 03562 g012
Figure 13. Intention of periodic table of elements related to MXene. Here, M, A, and X originate from the MAX phase, and Tx and I represent the surface termination groups and intercalants during the MAX-to-MXenes synthesis process (etching, intercalation or functionalization steps). Reprinted from [100].
Figure 13. Intention of periodic table of elements related to MXene. Here, M, A, and X originate from the MAX phase, and Tx and I represent the surface termination groups and intercalants during the MAX-to-MXenes synthesis process (etching, intercalation or functionalization steps). Reprinted from [100].
Materials 17 03562 g013
Figure 14. Number of documents on keywords “MXene” and “adsorption”. Sources: Scopus from Elsevier (until 20 March 2024).
Figure 14. Number of documents on keywords “MXene” and “adsorption”. Sources: Scopus from Elsevier (until 20 March 2024).
Materials 17 03562 g014
Figure 15. (a) SEM images of alk-MXene-NH2 and the corresponding element mapping of the alk-MXene-NH2 sample. (b) TEM and HRTEM images of alk-MXene-NH2. (c) The XRD patterns of the Ti3AlC2, DL-MXene, alk-MXene and alk-MXene-NH2 sample. (d) N2 adsorption-desorption isotherm of ML-MXene powders, alk-MXene and alk-MXene-NH2 (insets are the corresponding BJH pore size distribution). (e) Effect of competitive ions on uptake of Pb(II) onto alk-MXene-NH2 (f). (g) The possible adsorption mechanism for alkMXene-NH2 to Pb(II). Reprinted from [104].
Figure 15. (a) SEM images of alk-MXene-NH2 and the corresponding element mapping of the alk-MXene-NH2 sample. (b) TEM and HRTEM images of alk-MXene-NH2. (c) The XRD patterns of the Ti3AlC2, DL-MXene, alk-MXene and alk-MXene-NH2 sample. (d) N2 adsorption-desorption isotherm of ML-MXene powders, alk-MXene and alk-MXene-NH2 (insets are the corresponding BJH pore size distribution). (e) Effect of competitive ions on uptake of Pb(II) onto alk-MXene-NH2 (f). (g) The possible adsorption mechanism for alkMXene-NH2 to Pb(II). Reprinted from [104].
Materials 17 03562 g015
Figure 16. Adsorption mechanisms of Ba2+/Sr2+ by MXene and comparison of removal rate between Ba2+ and Sr2+ using single or mixture electrolyte solution: (A): ion exchange; (B): inner-sphere complexation (conditions: MXene dose = 100 g·L−1; Ba2+ and Sr2+ concentration of each single or mixture electrolyte solution = 2 g·L−1; temperature = 293 K; pH = 7; contact time = 1 h). Reprinted from [114].
Figure 16. Adsorption mechanisms of Ba2+/Sr2+ by MXene and comparison of removal rate between Ba2+ and Sr2+ using single or mixture electrolyte solution: (A): ion exchange; (B): inner-sphere complexation (conditions: MXene dose = 100 g·L−1; Ba2+ and Sr2+ concentration of each single or mixture electrolyte solution = 2 g·L−1; temperature = 293 K; pH = 7; contact time = 1 h). Reprinted from [114].
Materials 17 03562 g016
Table 1. Removal of heavy metals by UF membranes.
Table 1. Removal of heavy metals by UF membranes.
Membrane SystemsMetal(s) StudiedFlux (L/m2 h)Concentration of Metal (mg/L)Rejection
(%)
Ref.
TA/chitosan/PSS/PEMLi+52.08793.42[37]
CHT/chitosan/PSS/PEM81.6796.11
Chitosan/PESZn2+NA5080.00[38]
Pb2+NA90.00
Chitosan/PES+ polypropyleneCr6+9.37575.00[36]
poly(AA-co-VP)@APTES@MCN/PESCd2+322.5NA78.00[39]
PEES/n-ZnO/PEGCu2+44.310393.27[40]
Zn2+50.689.15
Ni2+57.385.81
Cd2+60.983.19
GNP/KMnO4/PSFCr6+NA10393.07[41]
Fe3+NA98.75
Amino- Zr-MOFs-CUFPb2+11003061.40[42]
Ni-Zn MOF/PSFPb2+8.858098.00[43]
BCl + MEUFAs3+NA0.1592.80[44]
DPCl + MEUF84.10
Chitosan + MEUFCr6+30.733100.00[45]
Chitosan-SDS + MEUF53.89398.33
ABSNa50 + TP220 + MEUFCu2+NA100100.00[46]
Table 2. Removal of heavy metals by NF membranes.
Table 2. Removal of heavy metals by NF membranes.
NF Membrane SystemsPressure
(MPa)
Zeta Potential (mV)Flux
(L/m2 h)
Heavy MetalConcentration of Metal (mg/L)Rejection
(%)
Ref.
PEG/α-CD/
PAN membranes
NSNS5.9Ru3+5083.10[52]
5.5Ag+88.00
5.9Co2+90.20
TFN-PEI-GO0.5NS70.3Pb2+50091.00[53]
Cu2+96.00
Cd2+96.00
TFN/UiO-66-NH20.8NS60.41Pb2+50099.00[54]
CysCNC TFN1 ± 0.1−23.5 ± 1.216Cu2+20089.20[49]
Pb2+87.00
PEI-Cu(II)-CSP1NS8.1Ni2+583.15[55]
Cd2+75.50
PEI-Cu(II)-CSP0.8NS18.1Cd2+
Zn2+
Ni2+
Pb2+
586.00
87.20
89.50
89.90
[56]
MPD/TMC/DWCNT + PES0.5NS109Na+00096.48[57]
PDA/Zr-MOF-TFN FO membranesNS−2813.2
14
13.8
Ni2+
Pb2+
Cd2+
200099.00
99.00
96.00
[58]
NH2-MIL-101 (Al)/0.35 wt% ZnO
+Cs-UF matrix
0.5+111.5
1.6
Mg2+
Ca2+
100090.10
80.49
[59]
PEI/PAA hollow fiber
membranes
NA−27.629.6 1Cu2+10094.20[60]
TA/Fe3+PEI hollow fiber
membranes
0.1+3723.18Fe3+
Cu2+
Mn2+
Zn2+
400095.70
94.10
92.90
91.10
[61]
CNFs-co-Cs PES membranesNSNS13.58Cu2+
Cr6+
Pb2+
>2097.86
98.40
96.54
[62]
1 means permeation flux, others are water flux. NS: no statistics in references; NA: not available in references.
Table 3. Comparison of membrane processes.
Table 3. Comparison of membrane processes.
MembranesAdvantagesDisadvantagesRef.
PEUFRecovery and reuse of collected retentate and polymer binding, low operational cost, and high separation efficacyRequirement of proper pre-treatment process and microligand, and not suitable for industrial wastewater[4]
MEUFHigh separation efficacy, low energy input, and rapid reactionGeneration of secondary pollutants and post-treatment, be costly and unsustainable[4,32]
Adsorptive MMMsLow cost, low-pressure requirement, high thermal and chemical strength, and high water yieldAgglomeration and low affinity between the adsorbent materials[47,116]
NFHigh heavy metals rejection at intermediate operating pressure rangeFouling issue, high energy consumption and low water permeability[4,117]
ENFMsEasy handling, high selectivity, ease of manufacturing, environmentally friendly and low costLimited reusability, fouling effect, hazardous solvents, and highly time-consuming unstable morphology[77,80,118]
MOFs membranesTunable nature of the pore, selectivity, appended functionality, fast capture kinetics and regeneration capacityParticle agglomeration, difficult to separate and recover, high cost[83,117]
IIMsHigh adsorption capacity, high extraction efficiency, easy preparation
and low cost, reusability and high species formation
Weak recognition ability, desorption difficulties, and poor reusability[89]
MxenesHigh-level hydrophilicity, chemical stability, tunable chemistry and negative surface chargesHigh cost, stability issue[99,103]
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

Feng, Y.; Wang, R. Research Progress on Metal Ion Recovery Based on Membrane Technology and Adsorption Synergy. Materials 2024, 17, 3562. https://doi.org/10.3390/ma17143562

AMA Style

Feng Y, Wang R. Research Progress on Metal Ion Recovery Based on Membrane Technology and Adsorption Synergy. Materials. 2024; 17(14):3562. https://doi.org/10.3390/ma17143562

Chicago/Turabian Style

Feng, Yiqing, and Rui Wang. 2024. "Research Progress on Metal Ion Recovery Based on Membrane Technology and Adsorption Synergy" Materials 17, no. 14: 3562. https://doi.org/10.3390/ma17143562

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

Article Metrics

Back to TopTop