Next Article in Journal
The Relationship between Structure and Performance of Different Polyimides Based on Molecular Simulations
Next Article in Special Issue
Tuning Alkaline Anion Exchange Membranes through Crosslinking: A Review of Synthetic Strategies and Property Relationships
Previous Article in Journal
Chains Stiffness Effect on the Vertical Segregation of Mixed Polymer Brushes in Selective Solvent
Previous Article in Special Issue
Stability of Properties of Layer-by-Layer Coated Membranes under Passage of Electric Current
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement of Selectivity of RALEX-CM Membranes via Modification by Ceria with a Functionalized Surface

1
Kurnakov Institute of General and Inorganic Chemistry RAS, Leninsky Prospect 31, 119991 Moscow, Russia
2
Faculty of Chemistry and High Technologies, Kuban State University, 350040 Krasnodar, Russia
3
School of Chemistry and Material Science, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(3), 647; https://doi.org/10.3390/polym15030647
Submission received: 28 December 2022 / Revised: 23 January 2023 / Accepted: 25 January 2023 / Published: 27 January 2023
(This article belongs to the Special Issue Polymers for Electrochemical Applications)

Abstract

:
Ion exchange membranes are widely used for water treatment and ion separation by electrodialysis. One of the ways to increase the efficiency of industrial membranes is their modification with various dopants. To improve the membrane permselectivity, a simple strategy of the membrane surface modification was proposed. Heterogeneous RALEX-CM membranes were surface-modified by ceria with a phosphate-functionalized surface. Despite a decrease in ionic conductivity of the prepared composite membranes, their cation transport numbers slightly increase. Moreover, the modified membranes show a threefold increase in Ca2+/Na+ permselectivity (from 2.1 to 6.1) at low current densities.

1. Introduction

Electromembrane processes, in particular, electrodialysis, are widely used in various industrial processes for the purification and separation of substances in aqueous solutions, including wastewater, feed source demineralization, seawater desalination, production of organic acids and bases, etc., providing high purity products and the efficiency of ion separation [1,2,3,4,5]. Electrodialysis involves ion transport through ion-exchange membranes under an applied electric potential gradient [6,7,8]. The efficiency of electrodialysis, like most other electromembrane processes, depends on two of the main parameters of membranes: Their ionic conductivity and selectivity [9,10,11,12]. Ionic conductivity of cation exchange membranes is determined by the transport of counterions in a thin Debye layer near membrane pore walls with negatively charged sulfonic acid groups, while the non-selective transport of co-ions is determined by their transport through an electrically neutral solution localized in membrane pore centers [13,14,15]. Relatively cheap heterogeneous cation exchange membranes generally used in electrodialysis represent laminated mixtures of ion exchange resins (generally sulfonated copolymers of styrene and divinylbenzene), an inert binder, and reinforcing fibers introduced to increase the mechanical strength of membranes. However, the selectivity of transport processes in these membranes is generally low due to the formation of an additional system of large pores between grains of membrane-forming components [16]. The membrane surface also played an important role, since its heterogeneity causes electroconvection, which can significantly increase the mass transfer rate and promote ion mixing [17].
In recent years, considerable attention has been devoted to the extraction of valuable components from natural water and wastewater [18]. For example, the problems with separation of lithium and cobalt [19,20], lithium and magnesium [21,22] are well known. There are problems of separation of sodium and calcium [23]. In this regard, the search for ways to separate singly and doubly charged ions is attracting more and more attention [24,25,26]. A number of approaches have been proposed to improve the selectivity of industrial membranes [27], e.g., membrane profiling [28], deposition of a layer of oppositely charged polyelectrolyte or several layers of alternately charged polyelectrolytes [25,29,30].
Another way to improve transport processes in ion-exchange membranes is to introduce inorganic particles (dopants) into membrane matrices [31]. One of the promising dopants is ceria. Ceria-doped membranes are highly stable during fuel cell operation [32,33], effective in water purification by reverse osmosis [34] and nanofiltration, and even have antibacterial properties [35], which can favorably affect the stability and service life of membranes. The reason for the unique antioxidant and antibacterial properties of ceria is its ability to interact with highly reactive oxygen radicals (hydroxyl radicals, superoxide radicals, etc.) by reversible transition between the Ce4+ and Ce3+ states [36,37]. This can be very important for preventing the degradation (fouling) of membranes used in electromembrane processes [38,39].
Ceria-based materials are widely used in water treatment, which was described in detail elsewhere [40]. The ability of cerium phosphate to increase the monovalent permselectivity of anion-exchange membranes is also known [41]. From this point of view, it seems interesting to modify membranes with ceria, the surface of which contains phosphate anions. However, to the best of our knowledge, there are no reports in the literature on the use of membranes modified by surface-functionalized ceria in water treatment by electrodialysis. In the case of ceria-based materials, it is especially attractive to modify only the membrane surface with them for several reasons. Modification with ceria usually leads to a decrease in the membrane conductivity [42,43,44]. However, if only the membrane surface is modified, this effect would be less pronounced. Moreover, smaller amounts of reagents are needed for modification. Finally, modification of the membrane surface, as noted above, in some cases results in an enhancement of the selective transport of singly charged ions [27].
The purpose of the present work is to understand whether it is possible to change the selectivity of ion transport (e.g., Ca2+/Na+ permselectivity) by modifying only the surface of a heterogeneous membrane with surface-modified cerium oxide. A simple procedure of manufacturing of composite ion-exchange membranes with improved permselectivity was proposed. Novel composite RALEX-CM membranes asymmetrically modified with PO4-functionalized ceria were prepared and their transport properties as well as Ca2+/Na+ permselectivity were studied.

2. Materials and Methods

Sodium chloride (NaCl), hydrochloric acid (HCl), phosphoric acid (H3PO4), sodium phosphate monobasic (NaH2PO4), calcium chloride (CaCl2) were purchased from Chimmed (Moscow, Russia). Cerium(III) nitrate hexahydrate (Ce(NO3)3.6H2O) was purchased form Sigma-Aldrich (St. Louis, MO, USA). Deionized water was used throughout all experiments.
RALEX-solutions leads to an increase in the water uptake of the R_Ce_HPM membranes based on a copolymer of sulfonated styrene and divinylbenzene (Mega a.s., Straz pod Ralskem, Czech Republic) were preliminarily conditioned according to the standard procedure [45] in aqueous NaCl solutions of various concentrations (decreasing from 2 M to 0.01 M) to let membranes swell and in 5% HCl aqueous solution to convert membranes into H+-form. For asymmetric modification (on only one side of a membrane) by in situ method, conditioned RALEX-CM membranes were placed for 10 min in a two-compartment cell, one compartment of which was filled with 0.01 M Ce(NO3)3 solution, and the other was filled with deionized water. Then, cerium nitrate solution was replaced with 8% aqueous ammonia solution and the resulting system was kept under constant stirring in both compartments for 2 h. The notation R_Ce was used to refer to these modified membranes. To prepare materials with PO4-functionalized ceria, the R_Ce membranes were treated with 1 M H3PO4 for 1 h or 1 M NaH2PO4 for 24 h also on one side. The notations R_Ce_HP and R_Ce_NaP were used to refer to these modified membranes.
To convert membranes into the H+, Na+, or Ca2+-forms, membranes were kept for 2 h in 5% HCl, 0.1 M NaCl, or 0.5 M CaCl2 solutions, respectively. Then, membranes were washed with deionized water.
X-ray diffraction (XRD) patterns were collected using a Rigaku D/MAX 2200 (Tokyo, Japan) on CuKα radiation in the 2Θ range of 10-65° at 25 °C and processed using Rigaku Application software. The attenuated total reflection FTIR spectra were recorded using a Nicolet iS5 FTIR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) with a Specac Quest attachment over the range 4000–500 cm-1. The imaging of the membrane cross sections were performed by scanning electron microscopy (SEM, Tescan Amber, Prague, Czech Republic) and energy dispersive spectroscopy (Oxford Instruments, Abingdon-on-Thames, UK). Static contact angles of water on membranes were determined using a KRUSS DSA25 instrument by the optical water drop method. The volume of a water drop was 10 µL. Before measurements, membranes were kept at 30% relative humidity for 24 h.
The water uptake (ω(H2O)) of membranes was measured on a TG209F1 thermobalance (Netzsch-Gerätebau GmbH, Selb, Germany) in the temperature range of 20–200 °C and calculated according to Equation (1):
ω(H2O) = (Wwet − Wdry)/Wdry × 100%,
where Wwet and Wdry represent the membrane weights before and after drying at 200 °C, respectively. The dopant content was determined as Wres/Wdry, where Wres represents the weight of a residue after membrane heating at 800 °C in air for 10 min. The dopant content determination is complicated by the fact that it is impossible to completely convert membranes from the Na+- to the H+-forms (remove all residual Na+ ions), and when a polymeric part of membranes is burned off, residual Na+ ions interact with the membrane sulfonic acid groups to form Na2SO4. At the same time, since the amounts of residual sodium ions in the pristine RALEX-CM and the modified membranes are different, it impossible to calculate the dopant content by subtracting Wres of the RALEX-CM membrane from Wres of the modified membranes. In addition, when the membranes are kept in HCl solution for conversion to the H+-form, ceria (dopant) partially dissolves, which also makes it difficult to correctly determine its amount in the membrane. The dopant content is estimated to be in the range of 3–7 wt %. For comparison, the dopant content of heterogeneous membranes bulk-modified with zirconia was much higher (up to 20 wt %) [46]. Ion exchange capacity (IEC) was determined by direct titration [45] of the membranes in the H+-form. IEC is given per 1 g of swollen membrane
The conductivity (σ) of the prepared membranes in Na+- and Ca2+-forms was studied using an Elins Z1500 PRO impedance meter (Elins LLC, Chernogolovka, Russia) in the frequency range of 10–1.5 × 106 Hz in a potentiostatic mode with amplitude of 80 mV. To measure the conductivity, membranes were fixed between graphite electrodes in a Plexiglas cell. The measurements were carried out in deionized water in the temperature range 25–80 °C. The resistance (1/σ) at each temperature was found by extrapolating the impedance plots to the Z’ axis (the axis of active resistances).
The current-voltage (CV) characteristics were studied using a rotating membrane disk setup that allows setting the diffusion layer thickness (δ) by changing the rotation rate of the membrane disk in accordance with the Levich theory [47]:
δ = 1.61D1/3ν1/6ω−1/2,
where D represents the electrolyte diffusion coefficient, ν represents the solution kinematic viscosity, and ω represents the angular velocity of the membrane disk rotation. The CV characteristics were recorded in the galvanostatic mode at the membrane disk rotation rates of 50–300 rpm in the mixed solution of CaCl2 and NaCl. The solution supply rate into the cathodic chamber was 7.5 ± 0.1 mL/min. The total concentration of CaCl2 and NaCl was maintained constant and equal to 0.03 mol-eq/L (cNaCl and cCaCl2 was 0.015 and 0.0075 mol-eq/L, respectively). The concentrations of Ca2+ and Na+ ions were measured by liquid chromatography (Stayer, Aquilon, Podolsk, Russia). The Hittorf transport numbers (t+) were determined by Equation (3):
t+ = (c – c0)VF/I,
where c0 and c represent the initial electrolyte concentration and the electrolyte concentration at the cathodic chamber outlet, respectively, V represents the volumetric rate of the solution supply, I represents the current, and F represents the Faraday constant. The Ca2+/Na+ permselectivity was calculated using Equation (4):
PCa/Na = tCac0Na/(tNac0Ca)
where tCa and tNa represent the effective transport numbers, c0Ca and c0Na represent the concentration of NaCl and CaCl2 solutions, respectively.

3. Results and Discussion

Figure 1 shows SEM images of the cross section of the modified membranes and the mappings of sulfur, cerium, and phosphorus over their thickness. In SEM images, ion exchanger grains in polyethylene (binder) and reinforcing fibers or holes left after their removal can be seen. Bright areas on the sulfur maps (Figure 1b,e,i) can be attributed to ion exchanger grains containing functional sulfonic acid groups. In the R_Ce membrane, cerium ions are predominantly distributed in the surface layer (the modified layer thickness does not exceed 20% of the membrane thickness (Figure 1c)). Since the cerium distribution in the modified membranes does not completely reproduce the sulfur distribution, it can be concluded that ceria is formed not only in ion exchanger grains, but also in large pores between the membrane components. A similar pattern is observed for the R_Ce_HP and R_Ce_NaP membranes (Figure 1d–k): Cerium ions are located on one side of each membrane, penetrating approximately 1/3 of the membrane thickness (about 150 μm). The layer with cerium ions is thicker in the R_Ce_HP and R_Ce_NaP membranes than in the R_Ce membrane, which may be due to a partial dissolution – precipitation of ceria in the membrane pores upon treatment with NaH2PO4 and especially H3PO4 solutions, which leads to diffusion of cerium ions and their redistribution over the membrane thickness. The distribution of phosphorus reproduces the distribution of cerium in the R_Ce_HP and R_Ce_NaP membranes. The R_Ce_NaP membrane shows the highest content of phosphorus. The Ce:P ratio in the R_Ce_HP and R_Ce_NaP samples is 9:1 and 3:1, respectively, suggesting only a partial modification of ceria by phosphate groups. It can be noticed that the phosphorus content is slightly higher in the near-surface layer of the R_Ce_HP membrane than in the total modified layer (Figure 1g).
XRD patterns of the pristine and modified RALEX-CM membranes are shown in Figure 2. On the XRD pattern of the pristine RALEX-CM membrane, there are reflections of polyethylene at 2θ ≈ 22, 24 and 25° (PDF-2 database, #53-1859) against an amorphous halo. Additional lines at 2θ ≈ 28, 47 and 56° appear on the XRD pattern of the R_Ce membrane, which correspond to the reflections of cerium dioxide (PDF-2, #34-0394). It should be noted that metal oxides synthesized in the membrane pores are generally amorphous, even cerium oxide despite high crystallinity of its nanosized particles. To the best of our knowledge, the XRD pattern of the R_Ce membrane is the first example of XRD patterns of crystalline ceria in an ion exchange membrane. The significantly lower intensity of the ceria reflections on the XRD patterns of the R_Ce_NaP and R_Ce_HP membranes indicates the transformation of cerium oxide into amorphous cerium phosphate. In this case, the size of the cerium oxide particle decreases significantly, which leads to an additional broadening of its reflections and a decrease in their intensity. The XRD pattern of the R_Ce_HP membrane is almost the same as that of the RALEX-CM membrane.
Modification of ion exchange membranes by the in situ method occurs due to the formation of dopant particles in their pores with a size of 4–5 nm. Therefore, the size of formed particles usually does not exceed 4–5 nm, and they cannot be observed by SEM. TEM study of Nafion-117 membranes modified by CeO2 particles showed that the size of latter is about 2–3 nm [32]. CeO2 particles prepared out of membrane are larger, and their treatment with phosphoric acid leads to a slight decrease in their size [48]. Metal oxide particles embedded in membranes by in situ method are generally X-ray amorphous. In this work, reflections of CeO2 introduced into the ion exchange membrane were observed on X-ray diffraction patterns for the first time. This enables us to determine their coherent scattering region (crystallite size L) using the Scherrer Equation (5):
L = 0.9 λ/(β cosΘB)
where λ is the wavelength of the X-ray beam, and β represents the full width at half maximum of the reflection at the Bragg angle ΘB. The crystallite size of ceria in the R_Ce membrane is about 3 nm. However, the treatment of ceria with NaH2PO4 or H3PO4 leads to a chemical transformation on its surface with a decrease in the size of the CeO2 nucleus, as a result of which the CeO2 reflections on the X-ray diffraction patterns of the R_Ce_HP and R_Ce_NaP membranes disappear.
Figure S1 shows FTIR spectra of the composite membranes based on RALEX-CM membranes and ceria. Frequencies of both bending and stretching vibrations of ceria are below 400 cm−1, therefore these bands are not active in the IR region. In the R_Ce_HP and R_Ce_NaP membranes containing PO4-modified ceria, the bands corresponding to phosphate groups should be in the range of 1000–1100 cm–1 [49]. However, they overlap with the bands of sulfonic acid groups of membranes, and a low content of PO4-groups leads to the fact that the FTIR spectra of the pristine RALEX-CM membrane and modified membranes coincide.
With the introduction of cerium oxide into pores and channels of RALEX-CM, ion exchange capacity the R_Ce_HP membrane decreases (Table 1). This is due to the formation of salt bridges between the dopant and sulfonic acid groups of the RALEX-CM membrane. Apparently, the binding of hydrophilic sulfonic acid groups leads to an increase in the hydrophobicity of the membrane surface; the water contact angle increases from 88 to 100° (Table 1), which is close to that of the dry RALEX-CM membrane [50]. When the R-Ce membranes are treated with phosphoric acid or NaH2PO4 solutions, the ion exchange capacity increases, but its values are somewhat lower for the R_Ce_HP and R_Ce_NaP membranes than those of the pristine RALEX-CM membrane (Table 1). At the same time, due to a partial release of the membrane functional groups and an increase in its hydrophilicity, the water contact angle decreases. Its value are 83 and 81° for the R_Ce_HP and R_Ce_NaP membranes, respectively. However, the difference between these values should not be considered significant compared to the water contact angle of the pristine membrane.
In addition to a decrease in IEC and hydrophilicity of the R_Ce membrane, the formation of salt bridges upon the cerium oxide introduction leads to the “contraction” of the membrane pore walls with the displacement of water from pores and a decrease in water uptake. A partial destruction of salt bridges upon treatment with H3PO4 or NaH2PO4 solutions leads to an increase in the water uptake of the R_Ce_HP и R_Ce_NaP membranes compared to that of the R_Ce membrane (Table 1).
The introduction of ceria has little effect on cation transport numbers in the R_Ce membrane due to two competing processes. The formation of salt bridges reduces a negative charge of pore walls and the selectivity of membranes. However, the introduction of ceria nanoparticles leads to the displacement of an electrically neutral solution from the central part of membrane pores and a decrease in the co-ion concentration in them [51]. Treatment of the R_Ce membrane with phosphoric acid and NaH2PO4 solutions leads to a partial release of sulfonic acid groups and imparts a negative charge to the ceria surface due to the dissociation of phosphate groups [41]. As a result, the cation transport numbers increase up to 94.4–94.7% in the R_Ce_HP и R_Ce_NaP membranes (Table 1). It can be noted that similar results were obtained for the RALEX-CM membrane modified with hydrogen zirconium phosphate, for which the cation transport numbers reached 93%–97% [52].
Due to reduced ion exchange capacity and water uptake, ionic conductivities of the modified membranes are somewhat lower than that of the pristine RALEX-CM membrane (Figure 3). Moreover, the conductivity of the modified membranes in the Ca2+ forms is significantly lower than that of the Na+ forms, which is associated with a lower mobility of doubly charged ions and a decrease in the water uptake of membranes in the Ca2+ form. This is also reflected in an increase in the activation energy of cation transfer with increasing cation charge (bonds between Ca2+ ions and water molecules/sulfonic acid groups are stronger) [51]. Conductivity of membranes in the Ca2+ forms decrease even more when cerium oxide is modified with PO4-groups, which have a high affinity for Ca2+ ions.
Additional information about transport processes in ion exchange membranes can be obtained from CV characteristics of membranes. The CV curves were obtained in a mixed NaCl–CaCl2 solution at different rotation rates of the membrane disk (Figure 4). With increasing rotation rate, the limiting current increases, indicating an externally diffusive nature of the limiting current. The higher the electrolyte flow rate (the higher the membrane disk rotation rate), the smaller the diffusion layer thickness, and consequently, the higher the limiting current.
Figure 4d compares the current-voltage curves for the modified membranes at a membrane disk rotation rate of 100 rpm. The ohmic components (in the low ohmic region) for all membranes are similar for all membranes (Figure 4d). However, the limiting current density is higher for the modified membranes than for the pristine RALEX-CM membrane (Figure 4d, Table 1). This is most likely due to an increase in the carrier concentration due to the dissociation of –PO4MX groups. An increase in the limiting current density will result in an improvement of the efficiency of water treatment by electrodialysis. Note also that the R_Ce_HP and R_Ce_NaP membranes pass into the over-limiting state at lower voltage (the width of the diffusion limited plateau region is narrower). Since the transport numbers t+ of the modified membranes are higher than those of the pristine membrane (Table 1), an increase in the limiting current density may be attributed to an enhancement of electro-convection on the surface of the modified membranes. Moreover, according to [45,53,54], the presence of phosphate groups in the modified membranes can catalyze water dissociation.
Differences in ionic mobilities of singly and doubly charged ions and the current-voltage characteristics of the modified membranes can be considered as their advantage for the separation of ions with different charges. Figure 5 shows the dependence of the permselectivity (PCa/Na) for the modified membranes in a mixed CaCl2–NaCl solution. Modification of the membrane surface leads to a threefold increase in the selectivity to Ca2+ ions at low current densities. The highest increase in the selectivity to doubly charged ions was achieved for the R_Ce_HP membrane. However, the permselectivity of all membranes decreases with increasing current density. The difference in the separation of sodium and calcium ions is completely lost at the limiting current density. This is due to the fact that this process no longer depends on membrane characteristics and is determined only by diffusion coefficients and charges of ions in a solution, which correlates well with the conclusions made when analyzing the dependences of current-voltage profiles on membrane disk rotation rate.
To test the stability of the R_Ce_HP and R_Ce_NaP membranes, additional measurements were performed at a current density of 6 mA/cm2 (1/2 of the limiting current density) and a membrane disk rotation speed of 100 rpm for 48 h. After these tests, the selective permeability of these membranes remains almost the same (the deviation does not exceed 5% from the initial value), which indicates their high stability and efficiency in electrodialysis applications at low current densities.
The efficiency of this approach is worth comparing to those proposed in recent works. The layer-by-layer technique of the membrane surface modification with layers of polyelectrolytes, which provide the transfer of ions of opposite charge, is most often used to increase the selectivity of membranes to single charged ions [26,27,55,56,57]. Such mem-branes exhibit very high permselectivities, which are a thousand times more than those of pristine membranes, but at the same time their resistances increase hundreds of times and the limiting current decreases [58]. This significantly reduces the efficiency of ion separation. In this regard, approaches associated with the membrane surface modifica-tion with inorganic particles or a single layer of polyelectrolyte seem to be more preferable. Malik et al. reported a decrease in calcium transfer numbers from 89.0% to 73.3% after the membrane surface modification with polyaniline [59]. The surface modification of RALEX-CM membranes with hydrogen zirconium phosphate leads to a decrease in the selectivity to calcium ions over sodium ions to 0.7 [52]. Higher limiting current densities of the R_Ce_HP membranes compared to MK-40 membranes (their composition is similar to that of RALEX-CM membranes) with a thin Nafion layer [60] allow us to expect their high efficiency in electrodialysis.

4. Conclusions

Novel composite membranes based on RALEX-CM membranes and ceria were prepared. The proposed approach allows manufacturing membrane materials asymmetrically modified with cerium oxide, including cerium oxide with a surface functionalized with PO4-groups and provides guidance for synthesis of composite ion exchange membranes with improved selectivity. The modified membranes have lower ion exchange capacities and water uptakes than a pristine RALEX-CM membrane. However, their cation transport numbers increase, indicating an increase in the selectivity to cations. The composite membranes exhibit a threefold increase in the PCa/Na permselectivity at low current densities, which indicates the prospects for their use for the separation of singly and doubly charged ions in water treatment using electrodialysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15030647/s1, Figure S1: FTIR spectra of the composite RALEX-CM membranes with ceria.

Author Contributions

Conceptualization, I.S. and P.Y.; methodology, V.Z., and A.Y; validation, V.Z., L.W. and A.Y.; investigation, A.A., P.Y. and I.S.; data curation, V.Z., L.W. and A.Y.; writing—original draft preparation, P.Y., A.A. and I.S.; writing—review and editing, I.S., P.Y. and A.Y.; supervision, A.Y. and L.W.; project administration, A.Y.; funding acquisition, A.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Russian Science Foundation (project #23-43-00138, https://rscf.ru/en/project/23-43-00138/, accessed on 27 December 2022).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

This research was performed using the equipment of the JRC PMR IGIC RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Strathmann, H.; Grabowski, A.; Eigenberger, G. Ion-Exchange Membranes in the Chemical Process Industry. Ind. Eng. Chem. Res. 2013, 52, 10364–10379. [Google Scholar] [CrossRef]
  2. Ran, J.; Wu, L.; He, Y.; Yang, Z.; Wang, Y.; Jiang, C.; Ge, L.; Bakangura, E.; Xu, T. Ion exchange membranes: New developments and applications. J. Membr. Sci. 2017, 522, 267–291. [Google Scholar] [CrossRef]
  3. Hagesteijn, K.F.L.; Jiang, S.; Ladewig, B.P. A review of the synthesis and characterization of anion exchange membranes. J. Mater. Sci. 2018, 53, 11131–11150. [Google Scholar] [CrossRef] [Green Version]
  4. Al-Amshawee, S.; Yunus, M.Y.B.M.; Azoddein, A.A.M.; Hassell, D.G.; Dakhil, I.H.; Hasan, H.A. Electrodialysis desalination for water and wastewater: A review. Chem. Eng. J. 2020, 380, 122231. [Google Scholar] [CrossRef]
  5. Gurreri, L.; Tamburini, A.; Cipollina, A.; Micale, G. Electrodialysis Applications in Wastewater Treatment for Environmental Protection and Resources Recovery: A Systematic Review. Prog. Perspect. Membr. 2020, 10, 146. [Google Scholar] [CrossRef] [PubMed]
  6. Xu, T.; Huang, C. Electrodialysis-based separation technologies: A critical review. AIChE J. 2008, 54, 3147–3159. [Google Scholar] [CrossRef]
  7. Chandra, A.; Bhuvanes, E.; Chattopadhyay, S. A critical analysis on ion transport of organic acid mixture through an anion-exchange membrane during electrodialysis. Chem. Eng. Res. Des. 2022, 178, 13–24. [Google Scholar] [CrossRef]
  8. Cao, G.; Alam, M.M.; Juthi, A.Z.; Zhang, Z.; Wang, Y.; Jiang, C.; Xu, T. Electro-desalination: State-of-the-art and prospective. Adv. Membr. 2023, 3, 100058. [Google Scholar] [CrossRef]
  9. Park, H.B.; Kamcev, J.; Robeson, L.M.; Elimelech, M.; Freeman, B.D. Maximizing the right stuff: The trade-off between membrane permeability and selectivity. Science 2017, 356, 6343. [Google Scholar] [CrossRef] [Green Version]
  10. Golubenko, D.V.; Pourcelly, G.; Yaroslavtsev, A.B. Permselectivity and ion-conductivity of grafted cation-exchange membranes based on UV-oxidized polymethylpenten and sulfonated polystyrene. Separ. Purif. Technol. 2018, 207, 329–335. [Google Scholar] [CrossRef]
  11. Campione, A.; Gurreri, L.; Ciofalo, M.; Micale, G.; Tamburini, A.; Cipollina, A. Electrodialysis for water desalination: A critical assessment of recent developments on process fundamentals, models and applications. Desalination 2018, 434, 121–160. [Google Scholar] [CrossRef]
  12. Biesheuvel, P.M.; Porada, S.; Elimelech, M.; Dykstra, J.E. Tutorial review of reverse osmosis and electrodialysis. J. Membr. Sci. 2022, 647, 120221. [Google Scholar] [CrossRef]
  13. Nikonenko, V.; Nebavsky, A.; Mareev, S.; Kovalenko, A.; Urtenov, M.; Pourcelly, G. Modelling of ion transport in electromembrane systems: Impacts of membrane bulk and surface heterogeneity. Appl. Sci. 2019, 9, 25. [Google Scholar] [CrossRef] [Green Version]
  14. Fan, H.; Yip, N.Y. Elucidating conductivity-permselectivity tradeoffs in electrodialysis and reverse electrodialysis by structure-property analysis of ion-exchange membranes. J. Membr. Sci. 2019, 573, 668–681. [Google Scholar] [CrossRef]
  15. Huang, Y.; Fan, H.; Yip, N.Y. Influence of electrolyte on concentration-induced conductivity-permselectivity tradeoff of ion-exchange membranes. J. Membr. Sci. 2023, 668, 121184. [Google Scholar] [CrossRef]
  16. Kononenko, N.; Nikonenko, V.; Grande, D.; Larchet, C.; Dammak, L.; Fomenko, M.; Volfkovich, Y. Porous structure of ion exchange membranes investigated by various techniques. Adv. Colloid Interface Sci. 2017, 246, 196–216. [Google Scholar] [CrossRef] [PubMed]
  17. Zyryanova, S.; Mareev, S.; Gil, V.; Korzhova, E.; Pismenskaya, N.; Sarapulova, V.; Rybalkina, O.; Boyko, E.; Larchet, C.; Dammak, L.; et al. How Electrical Heterogeneity Parameters of Ion-Exchange Membrane Surface Affect the Mass Transfer and Water Splitting Rate in Electrodialysis. Int. J. Mol. Sci. 2020, 21, 973. [Google Scholar] [CrossRef] [Green Version]
  18. Yaqub, M.; Lee, W. Zero-liquid discharge (ZLD) technology for resource recovery from wastewater: A review. Sci. Total Environ. 2019, 681, 551–563. [Google Scholar] [CrossRef]
  19. Verma, A.; Johnson, G.H.; Corbin, D.R.; Shiflett, M.B. Separation of Lithium and Cobalt from LiCoO2: A Unique Critical Metals Recovery Process Utilizing Oxalate Chemistry. ACS Sustain. Chem. Eng. 2020, 8, 6100–6108. [Google Scholar] [CrossRef]
  20. Zante, G.; Masmoudi, A.; Barillon, R.; Trébouet, D.; Boltoeva, M. Separation of lithium, cobalt and nickel from spent lithium-ion batteries using TBP and imidazolium-based ionic liquids. J. Ind. Eng. Chem. 2020, 82, 269–277. [Google Scholar] [CrossRef]
  21. Sun, Y.; Wang, Q.; Wang, Y.; Yun, R.; Xiang, X. Recent advances in magnesium/lithium separation and lithium extraction technologies from salt lake brine. Separ. Purif. Technol. 2021, 256, 117807. [Google Scholar] [CrossRef]
  22. Siekierka, A. Lithium and magnesium separation from brines by hybrid capacitive deionization. Desalination 2022, 527, 115569. [Google Scholar] [CrossRef]
  23. Feng, Q.; Wen, S.; Zhao, W.; Chen, Y. Effect of calcium ions on adsorption of sodium oleate onto cassiterite and quartz surfaces and implications for their flotation separation. Separ. Purif.Technol. 2018, 200, 300–306. [Google Scholar] [CrossRef]
  24. Wang, W.; Zhang, Y.; Yang, X.; Sun, H.; Wu, Y.; Shao, L. Monovalent Cation Exchange Membranes with Janus Charged Structure for Ion Separation. Engineering, 2022, in press. [CrossRef]
  25. Wang, W.; Zhang, Y.; Tan, M.; Xue, C.; Zhou, W.; Bao, H.; Lau, C.H.; Yang, X.; Ma, J.; Shao, L. Recent advances in monovalent ion selective membranes towards environmental remediation and energy harvesting. Sep. Purif. Technol. 2022, 297, 121520. [Google Scholar] [CrossRef]
  26. Ge, L.; Wu, B.; Yu, D.B.; Mondal, A.N.; Hou, L.X.; Ul Afsar, N.; Li, Q.H.; Xu, T.T.; Miao, J.B.; Xu, T.W. Monovalent cation perm-selective membranes (MCPMs): New developments and perspectives. Chin. J. Chem. Eng. 2017, 25, 1606–1615. [Google Scholar] [CrossRef]
  27. Luo, T.; Abdu, S. ⁠; Wessling, M. Selectivity of ion exchange membranes: A review. J. Membr. Sci. 2018, 555, 429–454. [Google Scholar] [CrossRef]
  28. Gurreri, L.; Filingeri, A.; Ciofalo, M.; Cipollina, A.; Tedesco, M.; Tamburini, A.; Micale, G. Electrodialysis with asymmetrically profiled membranes: Influence of profiles geometry on desalination performance and limiting current phenomena. Desalination 2021, 506, 115001. [Google Scholar] [CrossRef]
  29. Roghmans, F.; Evdochenko, E.; Martí-Calatayud, M.C.; Garthe, M.; Tiwari, R.; Walther, A.; Wessling, M. On the permselectivity of cation-exchange membranes bearing an ion selective coating. J. Membr. Sci. 2020, 600, 117854. [Google Scholar] [CrossRef]
  30. Golubenko, D.V.; Yaroslavtsev, A.B. Effect of current density, concentration of ternary electrolyte and type of cations on the monovalent ion selectivity of surface-sulfonated graft anion-exchange membranes: Modelling and experiment. J. Membr. Sci. 2021, 635, 119466. [Google Scholar] [CrossRef]
  31. Yaroslavtsev, A.B.; Stenina, I.A.; Voropaeva, E.Y.; Ilyina, A.A. Ion transfer in composite membranes based on MF-4SC incorporating nanoparticles of silica, zirconia, and polyaniline. Polym. Adv. Technol. 2009, 20, 566–570. [Google Scholar] [CrossRef]
  32. Yurova, P.A.; Malakhova, V.R.; Gerasimova, E.V.; Stenina, I.A.; Yaroslavtsev, A.B. Nafion/surface modified ceria hybrid membranes for proton exchange fuel cell application. Polymers 2021, 13, 2513. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, L.; Xing, Y.; Li, Y.; Fu, Z.; Li, Z.; Li, H. Double-Layer Expanded Polytetrafluoroethylene Reinforced Membranes with Cerium Oxide Radical Scavengers for Highly Stable Proton Exchange Membrane Fuel Cells. ACS Appl. Energy Mater. 2022, 5, 8743–8755. [Google Scholar] [CrossRef]
  34. Wang, Y.; Gao, B.; Li, S.; Jin, B.; Yue, Q.; Wang, Z. Cerium oxide doped nanocomposite membranes for reverse osmosis desalination. Chemosphere 2019, 218, 974–983. [Google Scholar] [CrossRef]
  35. Lakhotia, S.R.; Mukhopadhyay, M.; Kumari, P. Cerium oxide nanoparticles embedded thin-film nanocomposite nanofiltration membrane for water treatment. Sci. Rep. 2018, 8, 4976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Eriksson, P.; Tal, A.A.; Skallberg, A.; Brommesson, C.; Hu, Z.; Boyd, R.D.; Olovsson, W.; Fairley, N.; Abrikosov, I.A.; Zhang, X. Cerium oxide nanoparticles with antioxidant capabilities and gadolinium integration for MRI contrast enhancement. Sci. Rep. 2018, 8, 6999. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, M.; Zhang, C.; Zhai, X.; Luo, F.; Du, Y.; Yan, C. Antibacterial mechanism and activity of cerium oxide nanoparticles. Sci. China Mater. 2019, 62, 1727–1739. [Google Scholar] [CrossRef] [Green Version]
  38. Dammak, L.; Fouilloux, J.; Bdiri, M.; Larchet, C.; Renard, E.; Baklouti, L.; Sarapulova, V.; Kozmai, A.; Pismenskaya, N. A Review on Ion-Exchange Membrane Fouling during the Electrodialysis Process in the Food Industry, Part 1: Types, Effects, Characterization Methods, Fouling Mechanisms and Interactions. Membranes 2021, 11, 789. [Google Scholar] [CrossRef]
  39. Apel, P.Y.; Velizarov, S.; Volkov, A.V.; Eliseeva, T.V.; Nikonenko, V.V.; Parshina, A.V.; Pismenskaya, N.D.; Popov, K.I.; Yaroslavtsev, A.B. Fouling and Membrane Degradation in Electromembrane and Baromembrane Processes. Membr. Membr. Technol. 2022, 4, 69–92. [Google Scholar] [CrossRef]
  40. Kurian, M. Cerium oxide based materials for water treatment—A review. J. Envir. Chem. Eng. 2020, 8, 104439. [Google Scholar] [CrossRef]
  41. Golubenko, D.V.; Manin, A.D.; Wang, Y.; Xu, T.; Yaroslavtsev, A.B. The way to increase the monovalent ion selectivity of FujiFilm® anion-exchange membranes by cerium phosphate modification for electrodialysis desalination. Desalination 2022, 531, 115719. [Google Scholar] [CrossRef]
  42. Zaton, M.; Prélot, B.; Donzel, N.; Rozière, J.; Jones, D.J. Migration of Ce and Mn ions in PEMFC and its impact on PFSA membrane degradation. J. Electrochem. Soc. 2018, 165, F3281–F3289. [Google Scholar] [CrossRef]
  43. Weissbach, T.; Peckham, T.J.; Holdcroft, S. CeO2, ZrO2 and YSZ as mitigating additives against degradation of proton exchange membranes by free radicals. J. Membr. Sci. 2016, 498, 94–104. [Google Scholar] [CrossRef]
  44. Rui, Z.; Liu, J. Understanding of free radical scavengers used in highly durable proton exchange membranes. Prog. Nat. Sci. Mat. Int. 2020, 30, 732–742. [Google Scholar] [CrossRef]
  45. Novikova, S.A.; Volodina, E.I.; Pis’menskaya, N.D.; Veresov, A.G.; Stenina, I.A.; Yaroslavtsev, A.B. Ionic Transport in Cation-Exchange Membranes MK-40 Modified with Zirconium Phosphate. Russ. J. Electrochem. 2005, 41, 1070–1076. [Google Scholar] [CrossRef]
  46. Yurova, P.A.; Stenina, I.A.; Yaroslavtsev, A.B. A Comparative Study of the Transport Properties of Homogeneous and Heterogeneous Cation-Exchange Membranes Doped with Zirconia Modified with Phosphoric Acid Groups. Pet. Chem. 2018, 58, 1144–1153. [Google Scholar] [CrossRef]
  47. Gough, D.A.; Leypoldt, J.K. Membrane-Covered, Rotated Disc Electrode. Anal. Chem. 1979, 51, 439–444. [Google Scholar] [CrossRef]
  48. Yurova, P.A.; Tabachkova, N.Y.; Stenina, I.A.; Yaroslavtsev, A.B. Properties of ceria nanoparticles with surface modified by acidic groups. J. Nanopart. Res. 2020, 22, 318. [Google Scholar] [CrossRef]
  49. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds; John Wiley & Sons Marquette University: New York, NY, USA, 1978. [Google Scholar]
  50. Babilas, D.; Muszyński, J.; Milewski, A.; Leśniak-Ziółkowska, K.; Dydo, P. Electrodialysis enhanced with disodium EDTA as an innovative method for separating Cu(II) ions from zinc salts in wastewater. Chem. Eng. J. 2021, 408, 127908. [Google Scholar] [CrossRef]
  51. Stenina, I.A.; Yaroslavtsev, A.B. Ionic Mobility in Ion-Exchange Membranes. Membranes 2021, 11, 198. [Google Scholar] [CrossRef]
  52. Golubenko, D.V.; Karavanova, Y.A.; Melnikov, S.S.; Achoh, A.R.; Pourcelly, G.; Yaroslavtsev, A.B. An approach to increase the permselectivity and monovalent ion selectivity of cation-exchange membranes by introduction of amorphous zirconium phosphate nanoparticles. J. Membr. Sci. 2018, 563, 777–784. [Google Scholar] [CrossRef]
  53. Nikonenko, V.V.; Pismenskaya, N.D.; Belova, E.I.; Sistat, P.; Huguet, P.; Pourcelly, G.; Larchet, C. Intensive current transfer in membrane systems: Modelling, mechanisms and application in electrodialysis. Adv. Colloid Interface Sci. 2010, 160, 101–123. [Google Scholar] [CrossRef] [PubMed]
  54. Pismenskaya, N.D.; Pokhidnia, E.V.; Pourcelly, G.; Nikonenko, V.V. Can the electro-chemical performance of heterogeneous ion-exchange membranes be better than that of homogeneous membranes? J. Membr. Sci. 2018, 566, 54–68. [Google Scholar] [CrossRef]
  55. Afsar, N.U.; Shehzad, M.A.; Irfan, M.; Emmanuel, K.; Sheng, F.; Xu, T.; Ren, X.; Ge, L.; Xu, T. Cation exchange membrane integrated with cationic and anionic layers for selective ion separation via electrodialysis. Desalination 2019, 458, 25–33. [Google Scholar] [CrossRef]
  56. Wang, W.; Liu, R.; Tan, M.; Sun, H.; Niu, Q.J.; Xu, T.; Nikonenko, V.; Zhang, Y. Evaluation of the ideal selectivity and the performance of selectrodialysis by using TFC ion exchange membranes. J. Membr. Sci. 2019, 582, 236–245. [Google Scholar] [CrossRef]
  57. Evdochenko, E.; Kamp, J.; Femmer, R.; Xu, Y.; Nikonenko, V.; Wessling, M. Unraveling the effect of charge distribution in a polyelectrolyte multilayer nanofiltration membrane on its ion transport properties. J. Membr. Sci. 2020, 611, 118045. [Google Scholar] [CrossRef]
  58. White, N.; Misovich, M.; Alemayehu, E.; Yaroshchuk, A.; Bruening, M.L. Highly selective separations of multivalent and monovalent cations in electrodialysis through Nafion membranes coated with polyelectrolyte multilayers. Polymer 2016, 103, 478–485. [Google Scholar] [CrossRef] [Green Version]
  59. Malik, M.S.; Qaiser, A.A.; Arif, M.A. Structural and electrochemical studies of heterogeneous ion exchange membranes based on polyaniline-coated cation exchange resin particles. RSC Adv. 2016, 6, 115046–115054. [Google Scholar] [CrossRef]
  60. Belashova, E.D.; Melnik, N.A.; Pismenskaya, N.D.; Shevtsova, K.A.; Nebavsky, A.V.; Lebedev, K.A.; Nikonenko, V.V. Overlimiting mass transfer through cation-exchange membranes modified by Nafion film and carbon nanotubes. Electrochim. Acta 2012, 59, 412–423. [Google Scholar] [CrossRef]
Figure 1. SEM images (a,d,h) and mappings of sulfur (b,e,i), cerium (c,f,j) and phosphorus (g,k) across the thickness of the R_Ce (ac), R_Ce_HP (dg) and R_Ce_NaP (hk) membranes.
Figure 1. SEM images (a,d,h) and mappings of sulfur (b,e,i), cerium (c,f,j) and phosphorus (g,k) across the thickness of the R_Ce (ac), R_Ce_HP (dg) and R_Ce_NaP (hk) membranes.
Polymers 15 00647 g001aPolymers 15 00647 g001b
Figure 2. XRD patterns of the composite RALEX-CM membranes with ceria.
Figure 2. XRD patterns of the composite RALEX-CM membranes with ceria.
Polymers 15 00647 g002
Figure 3. Temperature dependences of conductivity of the modified membranes in the Na+- (a) and Ca2+- (b) forms.
Figure 3. Temperature dependences of conductivity of the modified membranes in the Na+- (a) and Ca2+- (b) forms.
Polymers 15 00647 g003
Figure 4. Current-voltage curves of the R_Ce (a), R_Ce_NaP (b) and R_Ce_HP (c) membranes in 0.03 mol-eq/L solution of NaCl and CaCl2 at different membrane disc rotation rates, and the comparison of CV curves obtained at a membrane disc rotation rate of 100 rpm (d).
Figure 4. Current-voltage curves of the R_Ce (a), R_Ce_NaP (b) and R_Ce_HP (c) membranes in 0.03 mol-eq/L solution of NaCl and CaCl2 at different membrane disc rotation rates, and the comparison of CV curves obtained at a membrane disc rotation rate of 100 rpm (d).
Polymers 15 00647 g004
Figure 5. Dependence of selective permeability (PCa/Na) of the modified membranes on the normalized current density (i/ilim) in 0.03 mol-eq/L solution of NaCl and CaCl2 at a membrane disk rotation rate of 100 rpm.
Figure 5. Dependence of selective permeability (PCa/Na) of the modified membranes on the normalized current density (i/ilim) in 0.03 mol-eq/L solution of NaCl and CaCl2 at a membrane disk rotation rate of 100 rpm.
Polymers 15 00647 g005
Table 1. Water uptake (ω(H2O)), ion exchange capacity, water contact angle (θc, from the modified side), cation transport numbers (t+), and limiting current density for the composite RALEX-CM membranes.
Table 1. Water uptake (ω(H2O)), ion exchange capacity, water contact angle (θc, from the modified side), cation transport numbers (t+), and limiting current density for the composite RALEX-CM membranes.
Membraneω(H2O), ±0.5%IEC, ±0.05 mmol/gθc, ±1°t+, ±0.1%Limiting Current Density, mA/cm2
RALEX-CM31.31.348893.811.13
R_Ce24.80.9810093.911.97
R_Ce_HP26.31.228394.413.08
R_Ce_NaP26.61.138194.711.72
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

Stenina, I.; Yurova, P.; Achoh, A.; Zabolotsky, V.; Wu, L.; Yaroslavtsev, A. Improvement of Selectivity of RALEX-CM Membranes via Modification by Ceria with a Functionalized Surface. Polymers 2023, 15, 647. https://doi.org/10.3390/polym15030647

AMA Style

Stenina I, Yurova P, Achoh A, Zabolotsky V, Wu L, Yaroslavtsev A. Improvement of Selectivity of RALEX-CM Membranes via Modification by Ceria with a Functionalized Surface. Polymers. 2023; 15(3):647. https://doi.org/10.3390/polym15030647

Chicago/Turabian Style

Stenina, Irina, Polina Yurova, Aslan Achoh, Victor Zabolotsky, Liang Wu, and Andrey Yaroslavtsev. 2023. "Improvement of Selectivity of RALEX-CM Membranes via Modification by Ceria with a Functionalized Surface" Polymers 15, no. 3: 647. https://doi.org/10.3390/polym15030647

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