Next Article in Journal
A Structural Bioinformatics-Guided Study of Adenosine Triphosphate-Binding Cassette (ABC) Transporters and Their Substrates
Previous Article in Journal
Membrane Treatment to Improve Water Recycling in an Italian Textile District
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

ZIF-8-Embedded Cation-Exchange Membranes with Improved Monovalent Ion Selectivity for Capacitive Deionization

Department of Green Chemical Engineering, College of Engineering, Sangmyung University, Cheonan 31066, Republic of Korea
*
Author to whom correspondence should be addressed.
Membranes 2025, 15(1), 19; https://doi.org/10.3390/membranes15010019
Submission received: 26 November 2024 / Revised: 27 December 2024 / Accepted: 7 January 2025 / Published: 9 January 2025
(This article belongs to the Section Membrane Applications for Water Treatment)

Abstract

:
Membrane capacitive deionization (MCDI) is an electrochemical ion separation process that combines ion-exchange membranes (IEMs) with porous carbon electrodes to enhance desalination efficiency and address the limitations of conventional capacitive deionization (CDI). In this study, a cation-exchange membrane (CEM) embedded with a metal–organic framework (MOF) was developed to effectively separate monovalent and multivalent cations in influent solutions via MCDI. To fabricate CEMs with high monovalent ion selectivity, ZIF-8 was incorporated into sulfonated poly(2,6-dimethyl-1,4-phenylene oxide) (SPPO) at various weight ratios. The resulting membranes were systematically characterized using diverse electrochemical methods. The ZIF-8-embedded CEMs demonstrated a sieving effect based on differences in ion size and hydration energy, achieving excellent permselectivity for monovalent ions. MCDI tests using the prepared CEMs showed a Na+ ion removal rate exceeding 99% in Na+/Mg2+ and Na+/Ca2+ mixed feed solutions, outperforming a commercial membrane (CSE, Astom Corp., Tokyo, Japan), which achieved a removal rate of 94.1%. These findings are expected to provide valuable insights for advancing not only MCDI but also other electro-membrane processes capable of selectively separating specific ions.

1. Introduction

Only about 0.5% of the water on Earth is freshwater, while the remaining 98% is seawater. As a result, the scarcity of freshwater has become a critical issue for humanity, especially with the rapid growth of the global population. Furthermore, the water shortage crisis is being exacerbated by worsening environmental pollution [1,2,3,4,5]. Consequently, there is growing global interest in desalination technologies, and related research is being actively pursued. Promising desalination methods beyond conventional distillation, which requires high energy consumption, include reverse osmosis (RO), electrodialysis (ED), and capacitive deionization (CDI). Among these, CDI has garnered significant attention as a cost-effective and environmentally friendly desalination technology due to its low energy requirements and lack of secondary pollution [5,6,7,8,9].
The CDI process is an electrochemical desalination technology that adsorbs cations and anions from water onto the surface of porous carbon electrodes by applying an electrical potential. Additionally, by switching the electrode polarity or turning off the power, the adsorbed ions can be released back into the water, regenerating the electrodes. However, if ionic substances are not completely removed during this process, the electrodes are only partially regenerated, leading to a reduced adsorption capacity. To address these limitations and enhance desalination efficiency, membrane capacitive deionization (MCDI)—which integrates ion-exchange membranes (IEMs) with porous carbon electrodes—has been actively researched [10,11,12,13]. In MCDI, ions are selectively adsorbed after being separated by an IEM in close contact with the carbon electrodes. This configuration effectively suppresses the undesired adsorption of oppositely charged ions onto the electrodes during the desorption process. As a result, MCDI achieves more efficient ion separation compared to CDI, improving its overall desalination performance. Figure 1 illustrates a schematic diagram of the structure and operating principle of MCDI [14].
Recently, separation processes that selectively remove specific ions from influent solutions containing various ionic species, such as seawater, have garnered significant interest [9,11,12,13,15,16]. For instance, separating monovalent from multivalent ions is often necessary to enhance the efficiency of subsequent processes that utilize treated water. Efficient separation of specific ions is also critical for purposes such as removing hazardous ionic substances from wastewater or concentrating target ions. CDI has gained attention as an efficient method for the selective desalination and concentration of specific ions [9]. In CDI, selective ion separation can be achieved by controlling the adsorptive properties of the electrode or by employing an IEM with selective permeability for specific ions. To date, most research has focused on modifying electrode properties to achieve selective ion separation [9]. Meanwhile, the development of IEMs designed to selectively separate specific ions in various electro-membrane processes has been actively pursued. As a result, the application of MCDI, which leverages the selective separation characteristics of IEMs, is also receiving growing attention.
The selective ion separation characteristics of IEMs are generally determined by electrostatic repulsion and steric hindrance effects. Electrostatic repulsion enables the separation of ions with different valences by controlling the strength and polarity of fixed charges on the membrane surface. For instance, if a positively charged layer is introduced on the surface of a cation-exchange membrane (CEM), monovalent cations can pass through, whereas divalent cations are blocked due to stronger electrostatic repulsion. Furthermore, by adjusting the degree of cross-linking in the membrane, ions with different hydration radii can be separated through the steric hindrance effect [17,18].
Metal–organic frameworks (MOFs), porous crystalline materials in which metal ions and organic molecules are linked by organic ligands to form three-dimensional skeletal structures, have recently been actively used as additives to improve the selective permeability of membranes [19,20,21]. MOFs can be synthesized into a wide variety of structures depending on the metal and organic components used, with various functionalizations possible when highly reactive organic components are included. Due to their unique features, such as diverse structures, controllable pore sizes, and modifiable frameworks, MOFs have been widely applied in separation, catalysis, electrochemistry, and biomedical fields [22,23,24,25,26,27,28,29,30,31]. Among the many types of MOFs, zeolitic imidazolate frameworks (ZIFs) offer advantages such as simple synthesis, low cost, and high specific surface area. In particular, ZIF-8, composed of dimethylimidazoles and zinc ions (Zn2+), exhibits strong hydrophobicity, and its pore window size (3.4 Å) is slightly smaller than the hydration radii of both monovalent and multivalent ions.
In this study, composite CEMs were fabricated by incorporating ZIF-8 nanoparticles as fillers into a sulfonated poly(2,6-dimethyl-1,4-phenylene oxide) (SPPO) matrix and applied to MCDI for the selective separation of monovalent cations. The well-developed pore structure and hydrophobicity of ZIF-8’s ion diffusion channels were expected to effectively separate monovalent and multivalent ions based on differences in ion size and hydration energy. To determine the optimal composition for maximizing monovalent ion selectivity and electrochemical properties, CEMs were prepared with varying amounts of ZIF-8, and their properties were systematically evaluated using various electrochemical analyses. Furthermore, the selective desalination performance was assessed by comparing commercial and fabricated membranes in MCDI applications. The findings of this study are expected to provide valuable insights for the development of IEMs and MCDI processes, not only for the separation of monovalent ions but also for the selective separation of other specific ions.

2. Experimental

2.1. ZIF-8 Synthesis and Characterizations

To synthesize ZIF-8 nanoparticles, 8 g of zinc nitrate hexahydrate and 3.6 g of 2-methylimidazole were each completely dissolved in 200 mL of methanol (MeOH) by stirring for 30 min. The two solutions were then mixed and allowed to react for 1 h. After the reaction, the mixed solution was centrifuged at 3500 rpm to precipitate the ZIF-8 nanoparticles, which were subsequently separated from the solvent. The resulting precipitate was dried in an oven (OF-12GW, JEIO TECH, Daejeon, Republic of Korea) at 120 °C for over 12 h to obtain a white powder [32,33]. The morphological characteristics of the synthesized ZIF-8 nanoparticles were examined using a field emission-scanning electron microscope (FE-SEM, SIGMA500, Carl ZEISS, Jena, Germany). Additionally, Fourier transform infrared spectroscopy (FT-IR, FT/IR-4700, Jasco, Hachioji, Japan) and X-ray diffraction (XRD, MiniFlex 600, Rigaku, Tokyo, Japan) analyses were conducted to confirm the chemical and crystal structures of the ZIF-8 nanoparticles, respectively.

2.2. Membrane Fabrication and Characterizations

To prepare SPPO, 2.75 g (23.4 mmol) of chlorosulfonic acid (CSA) was first dissolved in 8.25 mL of chloroform. Subsequently, 10 g of PPO was dissolved in 108.25 mL of chloroform to prepare a PPO solution, to which the CSA solution was slowly added dropwise while stirring vigorously. The PPO/CSA molar ratio of the mixed solution was maintained at 0.351. After the reaction was complete, the mixed solution was added dropwise to distilled water to obtain a solid polymer, which was then washed more than five times with distilled water and dried in an oven at 60 °C for over 12 h to prepare SPPO.
A casting solution was prepared by dissolving SPPO in N,N-dimethylacetamide (DMAc) at 25 wt% for fabricating the composite CEMs. The prepared ZIF-8 nanoparticles were then added to the casting solution at concentrations of 0, 1, 2.5, 5, 7.5, 10, 12.5, and 15 wt%, respectively. The nanoparticles were evenly dispersed in the solution via sonication and stirred at 60 °C for 24 h. The resulting SPPO-ZIF-8 solution was cast onto a glass plate and dried at room temperature for 4 h, followed by drying at 60 °C for 24 h, to fabricate the composite CEMs.
All reagents used in the experiments were purchased from Sigma-Aldrich Corp. (St. Louis, MO, USA) and used as received. To compare the characteristics of the prepared membranes, a commercial membrane, Neosepta CSE (Astom Corp., Tokyo, Japan), was used as a reference membrane. Neosepta CSE is a standard-grade commercial membrane known for its excellent electrochemical properties and high reliability. It is widely used in electo-membrane processes for ion separation, making it a suitable reference for comparing the basic performance of the fabricated CEMs.
To confirm the uniform dispersion of nanomaterials within the fabricated composite CEM, the membrane cross-section was observed at a magnification of 10,000× using a field-emission scanning electron microscope (FE-SEM, GeminiSEM 560, Carl Zeiss, Jena, Germany) and energy-dispersive X-ray spectroscopy (EDS, Ultim Max, Oxford Instruments, Abingdon, UK) analysis was also performed”.
Additionally, to measure the electrical resistance (ER) of the membranes, the CEM samples were immersed in a 0.5 M NaCl solution for over 6 h to reach equilibrium with the solution. The impedance of the membranes was then measured in 0.5 M NaCl solution using a lab-made clip cell connected to a potentiostat/galvanostat (SP-150, Bio-Logic Science Instruments, Grenoble, France), and the ER values were calculated using Equation (1):
E R = Z s a m p l e × cos θ s a m p l e Z b l a n k × cos θ b l a n k × A
where Z s a m p l e is the impedance of the electrolyte and membrane (Ω), Z b l a n k is the impedance of the electrolyte (Ω), θ is the phase angle (degree), and A is the membrane area (cm2). The ion conductivity of the membranes was calculated by dividing the ER value by the membrane thickness and taking the reciprocal.
The ion transport number (TN, t + ) of the membranes was determined by substituting the membrane potential measured using a pair of Ag/AgCl reference electrodes in a lab-made two-compartment cell into Equation (2) [32,33]:
E m = R T F ( 2 t + 1 ) ln a L a H
where Em is the measured membrane potential (mV), R is the gas constant, T is the absolute temperature, F is the Faraday constant, and a L and a H are the concentrations of NaCl solution (1 mM and 5 mM, respectively).
The ion-exchange capacity (IEC) of the membranes was measured using traditional acid-base titration. First, the membrane sample was immersed in a 0.5 M HCl solution for over 6 h to replace all functional groups with H+ ions. The acid solution on the membrane surface was then washed with distilled water, and the remaining water was removed using filter paper. Next, the membrane was immersed in a 0.5 M NaCl solution to re-substitute the H+ ions with Na+ ions. The amount of H+ ions in the solution was measured through acid-base titration using 0.01 M NaOH as a titrant and phenolphthalein as the indicator. After the titration was complete, the titration values and the dry weight of the membrane were substituted into Equation (3) to calculate the IEC:
I E C m e q g = C H + , s a m p l e × V s a m p l e W d r y
where Vsample is the volume (mL) of the titrant, CH+, sample is the concentration of the titrant (mol/L), and Wdry is the dry weight (g) of the membrane used in the experiment.
The water uptake (WU) of the membrane was determined by measuring the wet weight (Wwet) and dry weight (Wdry) of the membrane, respectively, and substituting them into Equation (4).
W U = W w e t W d r y W d r y × 100 %
The permselectivity (P) of the membrane was determined through electrodialysis experiments using a four-compartment cell as shown in Figure 2. For this, 200 mL of 0.1 M NaCl/MgCl2 or 0.1 M NaCl/CaCl2 was used as the feed solution, and 200 mL of 0.1 M KCl was filled and circulated in the permeate compartment. Additionally, 0.3 M Na2SO4 was circulated in the electrode compartment. The effective area of the Pt/Ti electrodes and the membrane were each 15 cm2, and the experiment was conducted for 1 h under a constant current density of 3.54 mA/cm2. The concentrations of Na+ and Mg2+ ions in the permeate compartment were measured using ion chromatography (883 Basic IC plus, Metrohm, Herisau, Switzerland), and the ion flux was calculated by substituting them into Equation (5):
J = ( C f C i ) A × t × V
where Ci and Cf are the initial and final ion concentrations (mol/L), respectively, A is the effective area (cm2), t is the experiment time (h), and V is the volume (L). In addition, the permselectivity was calculated by substituting the ion flux of the two compared ions (A+ and B2+), obtained using the above formula, into Equation (6) [17].
P B 2 + A + = J A + J B 2 +

2.3. MCDI Cell Tests for Desalination

Graphite electrodes laminated with porous activated carbon (effective area = 10 cm × 10 cm) were supplied by Siontech (Daejeon, Republic of Korea), and Neosepta ASE (Astom Corp., Tokyo, Japan) was used as the anion-exchange membrane (AEM). The effective area of the membrane was 11 cm × 11 cm, the thickness of the spacer was 2 mm, and the solution was delivered at a flow rate of 20 mL/min using a peristaltic pump. For the CDI experiment, a constant voltage of +1.5 V (for desalination) and 0 V (for regeneration) was applied, and the adsorption and desorption periods were 3 min each. A 1.7 × 10−3 M NaCl/MgCl2 or NaCl/CaCl2 solution was used as the influent. During the experiment, the influent and effluent concentrations were measured and substituted into Equation (7) to calculate the salt removal efficiency (η):
η = C i C f C i × 100 %
where C i is the ion concentration of the influent and C f is the ion concentration of the effluent. Figure 3 displays a schematic diagram explaining the operation of the MCDI process.

3. Results and Discussion

FT-IR spectra were measured to confirm the sulfonation of PPO and the synthesis of ZIF-8, with the results shown in Figure 4. The peak at 1065 cm−1 in the SPPO spectrum corresponds to the asymmetric stretching of -SO3, which is an absorption peak not present in PPO, indicating that the sulfonic acid group (-SO3H) was successfully introduced into PPO [34]. Meanwhile, the chemical structure of ZIF-8 was confirmed by observing the absorption peaks at 1380 cm−1 and 1588 cm−1, which are assigned to the methyl group of 2-methylimidazole and the C=N bond in the imidazolium ring, respectively, in the ZIF-8 spectrum. The absorption peaks corresponding to Zn-O and Zn-N at 1588 cm−1 and 692 cm−1, respectively, were also found in the spectrum. Additionally, absorption peaks corresponding to ZIF-8 were observed in the spectrum of the SPPO-ZIF-8 membrane, demonstrating the successful preparation of the composite CEM [25,35,36].
XRD analysis was performed to investigate the crystal structure of the synthesized ZIF-8 material. As shown in Figure 5a, the XRD pattern of the synthesized ZIF-8 displayed characteristic diffraction peaks at 2θ = 7.6°, 10.6°, 12.9°, 14.9°, 16.6°, 18.2°, 22.3°, 24.7°, 26.9°, and 29.8°, which were consistent with the reference (JCPDS 00-066-0091), confirming the typical sodalite structure of ZIF-8 [37,38]. Additionally, the FE-SEM image (Figure 5b) confirmed that the synthesized ZIF-8 nanoparticles had an approximate size of 30 nm and a uniform rhombic dodecahedral structure with a well-defined crystal surface, which is consistent with the XRD results [39].
To verify the even distribution of ZIF-8 throughout the thickness of the composite membrane, the cross-sectional image and EDS mapping results of the SPPO-ZIF-8 (2.5 wt%) membrane are presented in Figure 6. The results show that the S and O elements, associated with the sulfonic acid groups, were homogeneously distributed across the membrane’s cross-section [40]. Furthermore, the Zn and N elements originating from ZIF-8 were uniformly dispersed throughout the membrane thickness, confirming the even distribution of ZIF-8 nanoparticles within the composite CEM [41,42].
The results of the characterization of membranes fabricated with different ZIF-8 contents, along with the commercial membrane (Neosepta CSE), are summarized in Table 1. The thickness of the fabricated membranes was approximately 36 μm, which is much thinner than that of the CSE (ca. 143 μm). It was observed that as the ZIF-8 content in the prepared membranes increased, the ER value gradually rose, while the conductivity decreased. This indicates that as the content of ZIF-8 (which lacks ion-exchange groups) increases, ion transport through the membrane is somewhat hindered. However, due to the relatively thin membrane thickness compared to the commercial membrane, even at the maximum content of 15 wt%, the ER remained much lower than that of the CSE. The TN value of the fabricated membranes was approximately 0.97 across the range of ZIF-8 content used in the experiment, which was nearly equivalent to that of the commercial CSE membrane. The prepared membranes also exhibited an IEC value ranging from 1.99 to 1.52 meq./g, showing a decreasing trend as the ZIF-8 content increased. Meanwhile, the decrease in WU as the ZIF-8 content increased is thought to result from the increased hydrophobicity inside the membrane due to the methyl group contained in ZIF-8, along with the decrease in IEC.
The thermal properties of the fabricated CEMs were evaluated through TGA analysis, and the results are shown in Figure 7. The thermograms of the SPPO and SPPO-ZIF-8 membranes exhibited three major stages of mass loss. Up to about 200 °C, mass loss occurred due to the removal of water and residual solvent absorbed into the membrane, and decomposition of the sulfonic acid groups occurred in the range of 200–300 °C. Additionally, above 400 °C, the greatest mass loss occurred due to the decomposition of the polymer backbone [43]. In the case of ZIF-8, oxidative decomposition of 2-methylimidazole is known to occur between 100 and 300 °C, and the decomposition of the organic linker in ZIF-8 crystals appears above approximately 550 °C [44]. When the ZIF-8 content was 10 wt% or higher, it was confirmed that approximately 50% or more of the membrane mass remained at around 550 °C due to the strong bonding of the ZIF-8 crystals.
Figure 8 shows the ion flux, permselectivity, and permselectivity/ER ratio measured in Na+/Mg2+ and Na+/Ca2+ mixed solutions, highlighting the monovalent ion selective permeation characteristics of SPPO membranes with varying ZIF-8 contents. The results revealed that as the ZIF-8 content increased up to 10 wt%, the Na+ flux increased and then decreased significantly. On the other hand, for Mg2+ and Ca2+, the flux decreased as the ZIF-8 content increased up to 10 wt%, followed by a slight increase in flux. As a result, under both solution conditions, the highest permselectivity for monovalent ions was observed at 10 wt% ZIF-8 content, with permselectivity values exceeding 6 for Mg2+ and 4 for Ca2+ ions, respectively. A schematic illustrating the effect of ZIF-8 on improving the permselectivity for monovalent ions is shown in Figure 9. The basic ion separation mechanism of CEMs is based on the electrostatic attraction of counterions and the repulsion of coions by ion exchange groups (i.e., sulfonic acid groups, -SO3) within the membranes, according to traditional theory. Meanwhile, ZIF-8 does not contain ion exchange groups but can act as a selective barrier, enhancing the monovalent ion selectivity of CEMs by controlling the transport of monovalent and divalent ions. ZIF-8 is dispersed in the SPPO matrix to form a hydrophobic ion diffusion channel, through which hydrated ions must pass via a dehydration-hydration process [45,46,47,48,49,50]. In general, monovalent ions have lower hydration energies than divalent ions, making it relatively easier for monovalent ions to pass through the ion diffusion channels of ZIF-8, allowing for selective separation through the membrane. The hydration radii of Na+, Ca2+, and Mg2+ are 3.58, 4.12, and 4.28 Å, respectively, and their hydration energies are −310, −1481, and −1752 kJ/mol, respectively. Therefore, Na+ ions, which have a relatively small hydration radius and are easier to dehydrate and rehydrate, can pass through the ion diffusion channels of ZIF-8 more easily than Ca2+ and Mg2+ ions [30,50,51]. Based on this principle, the flux and permselectivity for Na+ ions are expected to increase as the amount of ZIF-8 added to SPPO increases. When comparing Ca2+ and Mg2+ ions under the same membrane conditions, it was also observed that the permselectivity for Mg2+ ions, which have a relatively larger hydration radius and hydration energy, was higher than for Ca2+ ions. However, as shown in Table 1, when an excessive amount of ZIF-8 was added (greater than 12.5 wt% in this study), the ER of the membrane increased significantly due to the aggregation of ZIF-8 particles and the decrease in IEC. It is well known that in mixed-matrix membranes (MMMs), exceeding the optimal filler content often leads to undesired particle aggregation, which can obstruct the formation of effective mass transfer channels [52]. It was believed that the ion separation through the dehydration-hydration process was hindered in this situation, leading to a decrease in permselectivity for monovalent ions. Additionally, the ratio of permselectivity to ER was calculated, revealing the same trend as permselectivity, with the optimal membrane production conditions determined at 10 wt% ZIF-8 content.
Based on the results of the previous selectivity evaluation, MCDI experiments were performed using commercial membranes and composite CEMs fabricated with 0, 5, 10, and 15 wt% ZIF-8 contents, and the results are shown in Figure 10. In this experiment, the performance of CDI without a membrane was also assessed for comparison. After checking the conductivity changes over 10 cycles (3 min for adsorption/3 min for desorption) over the course of 1 h, it was confirmed that CDI exhibited significantly lower adsorption and desorption performance compared to MCDI. In CDI, when an ionic substance is adsorbed onto the electrode surface, and a reverse potential is applied, the ions adsorbed on the electrode are desorbed and move into the bulk solution, while the counter ions move from the solution to the electrode. This prevents complete desorption of the ionic substances from the electrode, resulting in incomplete regeneration. On the other hand, the MCDI process has high salt removal efficiency by preventing desorbed ions from being re-adsorbed onto the counter electrode by the IEM [53]. In the case of CDI, the removal rates of Mg2+ and Ca2+ were found to be 98% and 95%, respectively, under Na+/Mg2+ and Na+/Ca2+ feed conditions, while the removal rate of Na+ was approximately 80%. This is because ions with higher valence have a stronger interaction with the charged electrode surface, leading to increased adsorption to the electrical double layer (EDL). Therefore, when treating influent containing a mixture of monovalent and divalent ions, divalent ions are more likely to be adsorbed onto the electrode surface than monovalent ions at equilibrium, due to their stronger electrostatic attraction [54]. In the case of MCDI, the removal rate of monovalent cations was measured to be much higher than that of divalent cations. This is because the IEM used with the electrode has a higher permselectivity for monovalent ions than for divalent ions. Moreover, the SPPO-ZIF-8 composite CEM fabricated with 10 wt% ZIF-8 content among the comparative membranes exhibited the highest selective removal efficiency for monovalent ions in the CDI test. Under both Na+/Mg2+ and Na+/Ca2+ mixed solution conditions, the Na+ removal efficiency exceeded 99%, while the Mg2+ and Ca2+ removal rates showed the lowest values among the comparative membranes. This is interpreted as a result of using a monovalent ion-selective CEM in combination with an electrode, which prevents divalent ions from passing through the membrane during adsorption and causes them to remain in the influent, resulting in a higher removal rate of Na+ compared to divalent ions. However, when the SPPO-ZIF-8(15 wt%) membrane was applied to CDI, it showed a significantly lower monovalent ion removal rate than the other membranes under both the Na+/Mg2+ and Na+/Ca2+ mixed solution conditions, which corresponds to the lower permselectivity described above.

4. Conclusions

In this study, composite CEMs with excellent monovalent ion selectivity were prepared by embedding ZIF-8 nanoparticles in an SPPO matrix at various weight ratios for MCDI applications. ZIF-8 does not have ion-exchange capabilities and has high hydrophobicity, which somewhat reduces the ion conductivity of the composite membranes. However, the membrane resistance was lower than that of the commercial membrane due to the uniform dispersion of ZIF-8 nanoparticles and the thin membrane thickness. As a result of measuring the permselectivity for monovalent ions, the highest value was observed at 10 wt% ZIF-8 content, with permselectivity for monovalent ions greater than 6 and 4 for Na+/Mg2+ and Na+/Ca2+, respectively. In addition, CDI experiments were performed using the fabricated SPPO-ZIF-8 composite CEMs, and the results revealed a high Na+ ion removal efficiency of over 99% in both Na+/Mg2+ and Na+/Ca2+ mixed solution conditions at 10 wt% ZIF-8 content. The significant improvement in monovalent ion selectivity is attributed to the pore structure and the hydrophobicity of the ion diffusion channels of ZIF-8 nanoparticles. It is believed that selective ion separation can be achieved by combining the unique characteristics of the ZIF-8 nanoparticles and the differences in the hydration radii and hydration energies of the monovalent and divalent ions. The results of this study are expected to provide useful information for the development of various electro-membrane processes, including MCDI, which can effectively separate specific ions as well as monovalent and multivalent ions.

Author Contributions

Conceptualization, M.-S.K.; methodology, M.-S.K.; data curation, E.-G.H. and J.-H.L.; validation, M.-S.K.; investigation, M.-S.K.; resources, M.-S.K.; writing—original draft preparation, E.-G.H., J.-H.L. and M.-S.K.; writing—review and editing, M.-S.K.; supervision, M.-S.K.; project administration, M.-S.K.; funding acquisition, M.-S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a 2023 Research Grant from Sangmyung University (2023-A000-0183).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liu, Y.; Nie, C.; Liu, X.; Xu, X.; Sun, Z.; Pan, L. Review on carbon-based composite materials for capacitive deionization. RSC Adv. 2015, 5, 15205–15225. [Google Scholar]
  2. Angeles, A.T.; Lee, J. Carbon-based capacitive deionization electrodes: Development techniques and its influence on electrode properties. Chem. Rec. 2021, 21, 820–840. [Google Scholar] [PubMed]
  3. Elisadiki, J.; Kibona, T.E.; Machunda, R.L.; Saleem, M.W.; Kim, W.; Jande, Y.A. Biomass-based carbon electrode materials for capacitive deionization: A review. Biomass Convers. Biorefin. 2020, 10, 1327–1356. [Google Scholar]
  4. Gaikwad, M.S.; Balomajumder, C. Capacitive deionization for desalination using nanostructured electrodes. Anal. Lett. 2016, 49, 1641–1655. [Google Scholar]
  5. Sivasubramanian, P.; Kumar, M.; Kirankumar, V.S.; Samuel, M.S.; Dong, C.; Chang, J. Capacitive deionization and electrosorption techniques with different electrodes for wastewater treatment applications. Desalination 2023, 559, 116652. [Google Scholar]
  6. Volfkovich, Y.M. Capacitive deionization of water (a review). Russ. J. Electrochem. 2020, 56, 18–51. [Google Scholar] [CrossRef]
  7. Pan, S.; Haddad, A.Z.; Kumar, A.; Wang, S. Brackish water desalination using reverse osmosis and capacitive deionization at the water-energy nexus. Water Res. 2020, 183, 116064. [Google Scholar]
  8. Sufiani, O.; Elisadiki, J.; Machunda, R.L.; Jande, Y.A. Modification strategies to enhance electrosorption performance of activated carbon electrodes for capacitive deionization applications. J. Electroanal. Chem. 2019, 848, 113328. [Google Scholar]
  9. Zhang, X.; Zuo, K.; Zhang, X.; Zhang, C.; Liang, P. Selective ion separation by capacitive deionization (CDI) based technologies: A state-of-the-art review. Environ. Sci. Water Res. Technol. 2020, 6, 243–257. [Google Scholar]
  10. Zhang, C.; He, D.; Ma, J.; Tang, W.; Waite, T.D. Faradaic reactions in capacitive deionization (CDI)-problems and possibilities: A review. Water Res. 2018, 128, 314–330. [Google Scholar]
  11. Chen, R.; Sheehan, T.; Ng, J.L.; Brucks, M.; Su, X. Capacitive deionization and electrosorption for heavy metal removal. Environ. Sci. Water Res. Technol. 2020, 6, 258–282. [Google Scholar] [CrossRef]
  12. Bao, S.; Xin, C.; Zhang, Y.; Chen, B.; Ding, W.; Luo, Y. Application of capacitive deionization in water treatment and energy recovery: A review. Energies 2023, 16, 1136. [Google Scholar] [CrossRef]
  13. Gamaethiralalage, J.G.; Singh, K.; Sahin, S.; Yoon, J.; Elimelech, M.; Suss, M.E.; Liang, P.; Biesheuvel, P.M.; Zornitta, R.L.; Smet, L.C.P.M. Recent advances in ion selectivity with capacitive deionization. Energy Environ. Sci. 2021, 14, 1095–1120. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Wang, Y.; Wang, R.; Wu, Y.; Xu, S.; Wang, J. Performance comparison and energy consumption analysis of capacitive deionization and membrane capacitive deionization processes. Desalination 2013, 324, 127–133. [Google Scholar] [CrossRef]
  15. Hou, C.; Huang, C. A comparative study of electrosorption selectivity of ions by activated carbon electrodes in capacitive deionization. Desalination 2013, 314, 124–129. [Google Scholar] [CrossRef]
  16. Kim, D.-H.; Choi, Y.-E.; Park, J.-S.; Kang, M.-S. Capacitive deionization employing pore-filled cation-exchange membranes for energy-efficient removal of multivalent cations. Electrochim. Acta 2019, 295, 164–172. [Google Scholar] [CrossRef]
  17. Pang, X.; Tao, Y.; Xu, Y.; Pan, J.; Shen, J.; Gao, C. Enhanced monovalent selectivity of cation exchange membranes via adjustable charge density on functional layers. J. Membr. Sci. 2020, 595, 117544. [Google Scholar] [CrossRef]
  18. Gao, H.; Zhang, B.; Tong, T.; Chen, Y. Monovalent-anion selective and antifouling polyelectrolytes multilayer anion exchange membrane for reverse electrodialysis. J. Membr. Sci. 2018, 567, 68–75. [Google Scholar] [CrossRef]
  19. Xu, X.; Sun, Y.; Zhang, Q.; Wang, S.; Zhang, L.; Wu, Z.; Lu, G. Synthesis of ZIF-8 hollow spheres via MOF-to-MOF conversion. ChemistrySelect 2016, 1, 1763–1767. [Google Scholar] [CrossRef]
  20. Liu, Y.; Lin, S.; Liu, Y.; Sarkar, A.K.; Bediako, J.K.; Kim, H.Y.; Yun, Y.-S. Super-stable, highly efficient, and recyclable fibrous metal–organic framework membranes for precious metal recovery from strong acidic solutions. Small 2019, 15, 1805242. [Google Scholar] [CrossRef]
  21. Ding, M.; Cai, X.; Jiang, H. Improving MOF stability: Approaches and applications. Chem. Sci. 2019, 10, 10209–10230. [Google Scholar] [CrossRef]
  22. Shekhah, O.; Liu, J.; Fischer, R.A.; Wöll, C. MOF thin films: Existing and future applications. Chem. Soc. Rev. 2011, 40, 1081–1106. [Google Scholar] [CrossRef] [PubMed]
  23. Kadhom, M.; Deng, B. Metal-organic frameworks (MOFs) in water filtration membranes for desalination and other applications. Appl. Mater. Today 2018, 11, 219–230. [Google Scholar] [CrossRef]
  24. Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O.M. High-throughput synthesis of zeolitic imidazolate frameworks and application to CO2 capture. Science 2008, 319, 939–943. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, S.; Li, H.; Zhang, X.; Du, S.; Zhang, J.; Su, B.; Gao, X.; Mandai, B. Amine-functionalized ZIF-8 nanoparticles as interlayer for the improvement of the separation performance of organic solvent nanofiltration (OSN) membrane. J. Membr. Sci. 2020, 614, 118433. [Google Scholar] [CrossRef]
  26. Yu, S.; Pang, H.; Huang, S.; Tang, H.; Wang, S.; Qiu, M.; Chen, Z.; Yang, H.; Song, G.; Fu, D.; et al. Recent advances in metal-organic framework membranes for water treatment: A review. Sci. Total Environ. 2021, 800, 149662. [Google Scholar] [CrossRef]
  27. Zhang, C.; Mu, Y.; Zhang, W.; Zhao, S.; Wang, Y. PVC-based hybrid membranes containing metal-organic frameworks for Li+/Mg2+ separation. J. Membr. Sci. 2020, 596, 117724. [Google Scholar] [CrossRef]
  28. Zhao, J.; Fan, R.; Xiang, S.; Hu, J.; Zheng, X. Preparation and lithium-ion separation property of ZIF-8 membrane with excellent flexibility. Membranes 2023, 13, 500. [Google Scholar] [CrossRef]
  29. Wang, C.; Yang, R.; Wang, H. Synthesis of ZIF-8/fly ash composite for adsorption of Cu2+, Zn2+ and Ni2+ from aqueous solutions. Materials 2020, 13, 214. [Google Scholar] [CrossRef]
  30. Xiao, T.; Liu, D. The most advanced synthesis and a wide range of applications of MOF-74 and its derivatives. Microporous Mesoporous Mater. 2019, 283, 88–103. [Google Scholar] [CrossRef]
  31. Quijia, C.R.; Lima, C.; Silva, C.; Alves, R.C.; Frem, R.; Chorilli, M. Application of MIL-100 (Fe) in drug delivery and biomedicine. J. Drug Deliv. Sci. Technol. 2021, 61, 102217. [Google Scholar] [CrossRef]
  32. Qi, C.; Li, J.; Shi, Y.; Zhang, B.; Chen, T.; Wang, C.; Liu, Q.; Yang, X. ZIF-8 penetrating composite membrane for ion sieving. J. Solid State Chem. 2022, 313, 123281. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Xian, S.; Xia, Q.; Wang, H.; Li, Z.; Li, J. Enhancement of CO2 adsorption and CO2/N2 selectivity on ZIF-8 via postsynthetic modification. AIChE J. 2013, 59, 2195–2206. [Google Scholar] [CrossRef]
  34. Sadhasivam, T.; Kim, H.-T.; Park, W.-S.; Lim, H.; Ryi, S.-K.; Roh, S.-H.; Jung, H.-Y. Low permeable composite membrane based on sulfonated poly(phenylene oxide) (sPPO) and silica for vanadium redox flow battery. Int. J. Hydrogen Energy 2017, 42, 19035–19043. [Google Scholar] [CrossRef]
  35. Abdulla, S.A.; Sabouni, R.; Ghommem, M.; Alami, A.H. Synthesis and performance analysis of zeolitic imidazolate frameworks for CO2 sensing applications. Heliyon 2023, 9, 21349. [Google Scholar] [CrossRef]
  36. Sajjadi, S.; Khataee, A.; Soltani, R.D.C.; Bagheri, N.; Karimi, A.; Azar, A.E.F. Implementation of magnetic Fe3O4@ ZIF-8 nanocomposite to activate sodium percarbonate for highly effective degradation of organic compound in aqueous solution. J. Ind. Eng. Chem. 2018, 68, 406–415. [Google Scholar] [CrossRef]
  37. Ahmad, K.; Nde, D.T.; Khan, R.A.; Raza, W. Fabrication of Zeolitic imidazolate framework-8 (ZIF-8) modified screen-printed electrode and its catalytic role for the electrochemical determination of biological molecule (dopamine). Colloids Surf. A Physicochem. Eng. Asp. 2024, 699, 134606. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Jia, Y.; Hou, L. Synthesis of zeolitic imidazolate framework-8 on polyester fiber for PM2.5 removal. RSC Adv. 2018, 8, 31471. [Google Scholar] [CrossRef]
  39. Massoudinejad, M.; Mohammadi, A.; Sadeghi, S.; Ghaderpoori, M.; Sahebi, S.; Alinejad, A. Arsenic adsorption over dodecahedra ZIF-8 from solution aqueous: Modelling, isotherms, kinetics and thermodynamics. Int. J. Environ. Anal. Chem. 2022, 102, 855–871. [Google Scholar] [CrossRef]
  40. Nagendra, B.; Salman, S.; Bloch, H.P.; Daniel, C.; Rizzo, P.; Fittipaldi, R.; Guerra, G. Poly(phenylene oxide) films with hydrophilic sulfonated amorphous phase and physically cross-linking hydrophobic crystalline phase. ACS Appl. Polym. Mater. 2023, 5, 3489–3498. [Google Scholar] [CrossRef]
  41. Si, Y.; Li, X.; Yang, G.; Mie, X.; Ge, L. Fabrication of a novel core–shell CQDs@ZIF-8 composite with enhanced photocatalytic activity. J. Mater. Sci. 2020, 55, 13049–13061. [Google Scholar]
  42. Liu, B.; Yang, Y.; Wu, H.; Wang, S.; Tian, J.; Dai, C.; Liu, T. Zeolitic imidazolate framework-8 triggers the inhibition of arginine biosynthesis to combat methicillin-resistant staphylococcus aureus. Small 2023, 19, 2205682. [Google Scholar]
  43. Petreanu, I.; Niculescu, V.-C.; Soare, A.; Iacob, C.; Teodorescu, M. Nanocomposite polyphenylene oxide with amino-functionalized silica: Structural characterization based on thermal analysis. J. Therm. Anal. Calorim. 2024, 149, 10671–10680. [Google Scholar] [CrossRef]
  44. Chirra, S.; Wang, L.-F.; Aggarwal, H.; Tsai, M.F.; Soorian, S.S.; Siliveri, S.; Goskula, S.; Gujjula, S.R.; Narayanan, V. Rapid synthesis of a novel nano-crystalline mesoporous faujasite type metal-organic framework, ZIF-8 catalyst, its detailed characterization, and NaBH4 assisted, enhanced catalytic Rhodamine B degradation. Mater. Today Commun. 2021, 26, 101993. [Google Scholar]
  45. Ji, Y.; Liu, X.; Li, H.; Jiao, X.; Yu, X.; Zhang, Y. Hydrophobic ZIF-8 covered active carbon for CO2 capture from humid gas. J. Ind. Eng. Chem. 2023, 121, 331–337. [Google Scholar] [CrossRef]
  46. Li, Y.; Lin, Z.; Wang, X.; Duan, Z.; Lu, P.; Li, S.; Ji, D.; Wang, Z.; Li, G.; Yi, D.; et al. High-hydrophobic ZIF-8@ PLA composite aerogel and application for oil-water separation. Sep. Purif. Technol. 2021, 270, 118794. [Google Scholar]
  47. Zhang, H.; Snurr, R.Q. Computational study of water adsorption in the hydrophobic metal–organic framework ZIF-8: Adsorption mechanism and acceleration of the simulations. J. Phys. Chem. C 2017, 121, 24000–24010. [Google Scholar]
  48. Sann, E.E.; Pan, Y.; Gao, Z.; Zhan, S.; Xia, F. Highly hydrophobic ZIF-8 particles and application for oil-water separation. Sep. Purif. Technol. 2018, 206, 186–191. [Google Scholar]
  49. Sun, X.; Liu, Y.; Xu, R.; Chen, Y. MOF-derived nanoporous carbon incorporated in the cation exchange membrane for gradient power generation. Membranes 2022, 12, 322. [Google Scholar] [CrossRef]
  50. Han, B.; Sun, Z.; Jiang, H.; Sun, X.; Ma, J.; He, M.; Zhang, W. Thin and defect-free ZIF-8 layer assisted enhancement of the monovalent perm-selectivity for cation exchange membrane. Desalination 2022, 529, 115637. [Google Scholar]
  51. Yu, H.; Hossain, S.M.; Wang, C.; Choo, Y.; Naidu, G.; Han, D.S.; Shon, H.K. Selective lithium extraction from diluted binary solutions using metal-organic frameworks (MOF)-based membrane capacitive deionization (MCDI). Desalination 2023, 556, 116569. [Google Scholar]
  52. Dechnik, J.; Sumby, C.J.; Janiak, C. Enhancing mixed-matrix membrane performance with metal-organic framework additives. Cryst. Growth Des. 2017, 17, 4467–4488. [Google Scholar]
  53. Kim, J.S.; Rhim, J.W. Performance study of membrane capacitive deionization (MCDI) cell constructed with Nafion and aminated polyphenylene oxide (APPO). Membr. J. 2020, 30, 350–358. [Google Scholar]
  54. Tauk, M.; Bechelany, M.; Sistat, P.; Habchi, R.; Cretin, M.; Zaviska, F. Performance analysis of zeolitic imidazolate ion-selectivity advancements in capacitive deionization: A comprehensive review. Desalination 2024, 572, 117146. [Google Scholar]
Figure 1. Schematic drawing of the MCDI operation principle.
Figure 1. Schematic drawing of the MCDI operation principle.
Membranes 15 00019 g001
Figure 2. Schematic drawing of permselectivity measurement using 4-compartment electrodialysis cell.
Figure 2. Schematic drawing of permselectivity measurement using 4-compartment electrodialysis cell.
Membranes 15 00019 g002
Figure 3. Schematic diagram showing MCDI process operation (red and blue solid lines: anode and cathode, respectively, red and blue dashed lines: effluent and influent, respectively).
Figure 3. Schematic diagram showing MCDI process operation (red and blue solid lines: anode and cathode, respectively, red and blue dashed lines: effluent and influent, respectively).
Membranes 15 00019 g003
Figure 4. FT-IR spectra of ZIF-8, PPO, SPPO, and SPPO-ZIF-8 membrane.
Figure 4. FT-IR spectra of ZIF-8, PPO, SPPO, and SPPO-ZIF-8 membrane.
Membranes 15 00019 g004
Figure 5. (a) XRD spectrum (red: ZIF-8 (measured) and blue: JCPDS 00-066-0091) and (b) FE-SEM image of ZIF-8 nanomaterial.
Figure 5. (a) XRD spectrum (red: ZIF-8 (measured) and blue: JCPDS 00-066-0091) and (b) FE-SEM image of ZIF-8 nanomaterial.
Membranes 15 00019 g005
Figure 6. (a) FE-SEM image and (b) EDS mapping of SPPO-ZIF-8(2.5 wt%) membrane (cross-sectional).
Figure 6. (a) FE-SEM image and (b) EDS mapping of SPPO-ZIF-8(2.5 wt%) membrane (cross-sectional).
Membranes 15 00019 g006
Figure 7. TGA thermograms of SPPO and SPPO-ZIF-8 membranes.
Figure 7. TGA thermograms of SPPO and SPPO-ZIF-8 membranes.
Membranes 15 00019 g007
Figure 8. Ion flux, permselectivity, and permselectivity/ER ratio through composite CEMs fabricated with various ZIF-8 contents: (a,b) in Na+/Mg2+ solution and (c,d) in Na+/Ca2+ solution.
Figure 8. Ion flux, permselectivity, and permselectivity/ER ratio through composite CEMs fabricated with various ZIF-8 contents: (a,b) in Na+/Mg2+ solution and (c,d) in Na+/Ca2+ solution.
Membranes 15 00019 g008aMembranes 15 00019 g008b
Figure 9. Schematic image showing selective transport mechanism by dehydration-hydration of ions through ion diffusion channel of ZIF-8.
Figure 9. Schematic image showing selective transport mechanism by dehydration-hydration of ions through ion diffusion channel of ZIF-8.
Membranes 15 00019 g009
Figure 10. Desalination performances of CDI and MCDI: (a) time-course change in conductivity; (b) salt removal efficiency for Na+/Mg2+ feed solution and (c) time-course change in conductivity; (d) salt removal efficiency for Na+/Ca2+ feed solution.
Figure 10. Desalination performances of CDI and MCDI: (a) time-course change in conductivity; (b) salt removal efficiency for Na+/Mg2+ feed solution and (c) time-course change in conductivity; (d) salt removal efficiency for Na+/Ca2+ feed solution.
Membranes 15 00019 g010
Table 1. Characteristics of membranes fabricated with commercial polymeric membranes and various ZIF-8 contents.
Table 1. Characteristics of membranes fabricated with commercial polymeric membranes and various ZIF-8 contents.
MembranesER
(Ω·cm2)
Conductivity
(mS/cm)
TN (−)IEC (meq./g)WU (%)
CSE (Astom, Tokyo, Japan)1.877.650.9832.2029.80
SPPO-ZIF-8 (0 wt%)0.4318.500.9661.9928.20
SPPO-ZIF-8 (1 wt%)0.5027.240.9701.9828.04
SPPO-ZIF-8 (2.5 wt%)0.5047.060.9681.9627.82
SPPO-ZIF-8 (5 wt%)0.5216.870.9711.8926.94
SPPO-ZIF-8 (7.5 wt%)0.5746.310.9621.9127.60
SPPO-ZIF-8 (10 wt%)0.5846.030.9731.8826.88
SPPO-ZIF-8 (12.5 wt%)0.6915.410.9691.6424.10
SPPO-ZIF-8 (15 wt%)0.7644.840.9671.5223.34
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

Han, E.-G.; Lee, J.-H.; Kang, M.-S. ZIF-8-Embedded Cation-Exchange Membranes with Improved Monovalent Ion Selectivity for Capacitive Deionization. Membranes 2025, 15, 19. https://doi.org/10.3390/membranes15010019

AMA Style

Han E-G, Lee J-H, Kang M-S. ZIF-8-Embedded Cation-Exchange Membranes with Improved Monovalent Ion Selectivity for Capacitive Deionization. Membranes. 2025; 15(1):19. https://doi.org/10.3390/membranes15010019

Chicago/Turabian Style

Han, Eui-Gyu, Ji-Hyeon Lee, and Moon-Sung Kang. 2025. "ZIF-8-Embedded Cation-Exchange Membranes with Improved Monovalent Ion Selectivity for Capacitive Deionization" Membranes 15, no. 1: 19. https://doi.org/10.3390/membranes15010019

APA Style

Han, E.-G., Lee, J.-H., & Kang, M.-S. (2025). ZIF-8-Embedded Cation-Exchange Membranes with Improved Monovalent Ion Selectivity for Capacitive Deionization. Membranes, 15(1), 19. https://doi.org/10.3390/membranes15010019

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