Next Article in Journal
Bis(triethoxysilyl)ethane (BTESE)–Organosilica Membranes for H2O/DMF Separation in Reverse Osmosis (RO): Evaluation and Correlation of Subnanopores via Nanopermporometry (NPP), Modified Gas Translation (mGT) and RO Performance
Previous Article in Journal
Comparison of Anion-Exchange Membranes for Diffusion Dialysis of Mixtures of Acids and Their Iron Salts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thin Film Nanocomposite Membranes Based on Zeolitic Imidazolate Framework-8/Halloysite Nanotube Composites

1
School of Chemical Engineering, Zhengzhou University, Zhengzhou 450001, China
2
School of Material Science and Engineering, North China University of Water Resources and Electric Power, Zhengzhou 450046, China
3
Pollution Prevention and Control Office, Ecological Environment Protection Commission of Zhengzhou, Zhengzhou 450007, China
4
Research Department of New Energy Technology, Zhengzhou Institute of Emerging Industrial Technology, Zhengzhou 450046, China
*
Authors to whom correspondence should be addressed.
Submission received: 28 November 2023 / Revised: 20 December 2023 / Accepted: 23 December 2023 / Published: 25 December 2023
(This article belongs to the Section Membrane Preparation and Characterization)

Abstract

:
Thin film nanocomposite (TFN) membranes have proven their unrivaled value, as they can combine the advantages of different materials and furnish membranes with improved selectivity and permeability. The development of TFN membranes has been severely limited by the poor dispersion of the nanoparticles and the weak adhesion between the nanoparticles and the polymer matrix. In this study, to address the poor dispersion of nanoparticles in TFN membranes, we proposed a new combination of m-ZIF-8 and m-HNTs, wherein the ZIF-8 and HNTs were modified with poly (sodium p-styrenesulfonate) to enhance their dispersion in water. Furthermore, the hydropathic properties of the membranes can be well controlled by adjusting the content of m-ZIF-8 and m-HNTs. A series of modified m-ZIF-8/m-HNT/PAN membranes were prepared to modulate the dye/salt separation performance of TFN membranes. The experimental results showed that our m-ZIF-8/m-HNT/PAN membranes can elevate the water flux significantly up to 42.6 L m−2 h−1 MPa−1, together with a high rejection of Reactive Red 49 (more than 80%). In particular, the optimized NFM-7.5 membrane that contained 7.5 mg of HNTs and 2.5 mg of ZIF-8 showed a 97.1% rejection of Reactive Red 49 and 21.3% retention of NaCl.

1. Introduction

A lack of clean and fresh water is one of the most pervasive problems that afflicts people throughout the world [1,2]. Membrane separation technology has undergone fast development over the past few decades, as it is a promising strategy to tackle the increasing demand for clean water [3,4,5,6]. Recently, significant advances in nanotechnologies and nanomaterials have offered us tremendous opportunities to design and prepare high-performance membranes that can overcome the limitation of the “trade-off” effect between permeability and selectivity [7,8,9]. In particular, thin film nanocomposite (TFN) membranes have attracted tremendous attention given their low resistance, ultrafast water permeation, and precise ionic sieving [10,11,12]. The addition of nanofillers can endow these membranes with improved hydrophilicity, selectivity, and antifouling properties [13].
TFN membranes with excellent performance have been exploited by utilizing various nanomaterials (such as metal–organic frameworks (MOFs), covalent organic frameworks (COFs), silica, titanium dioxide, carbon nanotubes (CNTs), graphene oxide (GO), g-C3N4 nanosheets, molybdenum disulfide, etc.) [14,15,16,17,18]. It is well known that a TFN membrane with high separation performance can be obtained through the optimization of the surface properties (hydrophilicity, charge, roughness, and interactions) or structure (including pore size, porosity, tortuosity, and thickness) of the thin film layer [19,20]. For example, CNTs are frequently applied to develop advanced TFN membranes because they can provide a one-dimensional fast water transport path [21]. After brush painting a single-walled carbon nanotube (SWCNT) network on membrane support, the resultant polyamide nanofiltration membranes can exhibit an extremely high water permeability of 40 L m−2 h−1 bar−1 and a high Na2SO4 rejection of 96.5% [22]. However, the CNTs are soft and have a high draw ratio, which may impair the water permeability of the membrane. In addition, the development of TFN membranes is currently limited by the poor dispersion of nanoparticles and the weak adhesion between nanoparticles and polymer matrix.
HNTs are natural aluminosilicate clay materials with tubular and hollow microstructures [23,24,25]. Unique among carbon nanotubes (CNTs), they are low-cost, available, and environmentally friendly materials for water treatment [26,27,28]. In addition, there are plenty of Al–OH groups on the internal lumen surface, which provides a considerable number of active sites for surface functionalization [12,29]. A previous study demonstrated that HNTs can be used to construct TFN membranes that act as nano-fillers or carriers and can endow these membranes with enhanced water permeability and selectivity [30,31,32]. Additionally, HNTs feature electropositive inner surfaces and electronegative outer surfaces, with great potential to repulse negative dyes such as Reactive Red during salt/dye separations through the Donnan effect. On the other hand, zeolitic imidazolate framework-8 (ZIF-8), which is composed of tetrahedral ions (Zn2+) and imidazole-derived organic linkers, is a commonly utilized water-stable MOF material that possesses a uniform pore size and high porosity available for the sharp molecular sieving of large molecules and fast water transport [33]. Hypothetically, taking advantage of the electronegative tube wall character of HNTs and the high porosity of ZIF-8, most negative dyes can be efficiently rejected, and water transfer can be significantly improved. Based on this assumption, a reasonable combination of HNTs and ZIF-8 may provide an alternative route to enhancing the comprehensive salt/dye separation performance of TFN membranes.
In this work, considering the poor dispersion of nanoparticles in TFN membranes, HNTs and ZIF-8 were first modified using poly (sodium p-styrenesulfonate) with a large number of hydrophilic groups to enhance their dispersion in water [34] so that the hydropathic properties of the membranes could be well controlled by adjusting the m-ZIF-8 and m-HNT contents. A series of modified ZIF-8/modified HNT/polyacrylonitrile (labeled as m-ZIF-8/m-HNT/PAN) TFN membranes were prepared, and the modulation of the separation performance of these TFN membrane sheets was achieved. The results demonstrated that HNTs can improve high water permeability by providing extra water pathways. Meanwhile, ZIF-8 can endow a membrane with enhanced separation because of its excellent molecular sieving ability. As verified by our experimental characterizations, the morphologies of the selective layers of TFN membranes can be tailored by changing the mass content of nanomaterials ranging from zero-dimensional m-ZIF-8 to one-dimensional m-HNTs. Therefore, the properties of these TFN membranes, including permeability and selectivity, can be also tuned and controlled.

2. Experimental Section

2.1. Materials and Chemicals

HNTs were provided by Henan Xiang Hu Environmental Protection Technology Co., Ltd., Zhengzhou, China. 2-Methylimidazole (Hmim), zinc nitrate hexahydrate (Zn(NO3)2·6H2O), and poly(sodium-p-styrenesulfonate) (PSS) (MW = 70 kDa) were purchased from J&K Scientific Ltd. (Shanghai, China). Disodium hydrogen phosphate sodium (Na2HPO4), triethylamine (TEA), dihydrogen phosphate sodium chloride (NaH2PO4), and glutaraldehyde (GA, 50% solution in water) were all acquired from Tianjin Kemiou Chemical Reagent Co., Ltd. (Tianjin, China). A polyacrylonitrile (PAN) ultrafiltration membrane (molecular weight cut off, MWCO = 50 kDa) was obtained from Sepro Mineral Systems Corp. (Langley, BC, Canada).

2.2. Synthesis of ZIF-8

ZIF-8 was prepared according to our previous report with a rapid room temperature synthesis method [35]. First, Zn2+ aqueous solution (4.93 mmol, 100 mL DI water) and Hmim aqueous solution (79.03 mmol, 100 mL DI water) were prepared. Then, TEA (79.06 mmol) was added to the Hmim aqueous solution to achieve the deprotonation of the Hmim. Subsequently, the deprotonated Hmim solution and the Zn2+ solution were mixed under vigorous stirring for 10 min. Following that, the same amount of TEA was added again at a molar ratio of 1:16:32 for Zn2+:Hmim:TEA in the final mixture. After an hour, the precipitate was collected and washed with DI water thrice. Finally, ZIF-8 nanomaterials were obtained after completely drying them in a vacuum oven at 60 °C.

2.3. Modification of ZIF-8 and HNTs with PSS

The modification of ZIF-8 or HNTs with PSS was performed by dispersing 0.2 g of nanomaterials (ZIF-8 or HNTs) in 50 mL of 2 wt% PSS aqueous solution with magnetic stirring for 24 h. The acquired suspension was left for 2 h, and then, the supernatant was collected via centrifugation at 3000 rpm for 10 min. The modified ZIF-8 (m-ZIF-8) or the modified HNTs (m-HNTs) could be collected via washing with DI water thrice followed by drying in a vacuum oven at 60 °C.

2.4. Preparation of m-ZIF-8/m-HNT/PAN Nanocomposite Membranes

First, 5 mg mL−1 of m-ZIF-8 and m-HNTs aqueous solution was prepared by dispersing the corresponding nanomaterials in DI water and sonicating for 10 min. Also, 0.2 wt% PVA aqueous solution was obtained by dissolving 0.2 g of PVA in 99.8 g of DI water at 90 °C for 4 h. To fabricate the m-ZIF-8/m-HNTs/PAN nanocomposite membranes, a certain volume of the m-ZIF-8 and m-HNTs aqueous solution and 1 mL of 0.2 wt% PVA aqueous solution were mixed and poured on the PAN membrane substrate, followed by evaporating the solvent at 80 °C. Finally, it was immersed in 0.2 wt% GA aqueous solution at 80 °C for about 30 min to achieve further cross-linking.
In order to analyze the effects of the used nanomaterials on membrane structures, chemical composition, separation performance, and antifouling performance, m-ZIF-8/m-HNT/PAN membranes with different mass amounts of m-ZIF and/or m-HNTs were fabricated and denominated as follows:

2.5. Characterization of HNTs, m-HNTs, ZIF-8, and m-ZIF-8

Nano-materials, including HNTs, m-HNTs, ZIF-8, and m-ZIF-8, were characterized with high-resolution transmission electron microscopy (HRTEM, JEM-2100, JEOL, Tokyo, Japan), Fourier-transform infrared spectroscopy (FTIR, Nicolet IS10, Thermo Scientific, Waltham, MA, USA), and X-ray powder diffraction (XRD, D8 Advance, Bruker, Mannheim, Germany). HRTEM images were acquired to examine the microstructures. FTIR was used to obtain the chemical composition information in a range of 4000~400 cm−1. In addition, the XRD pattern was recorded at room temperature with Cu Kα radiation at a scan rate of 10 °C min−1 in a range of 5°~80° to analyze its crystalline properties.

2.6. Characterization of the m-ZIF-8/m-HNT/PAN Membranes

The m-ZIF-8/m-HNT/PAN membranes were characterized via field emission scanning electron microscopy (FESEM, JSM-7001F, JEOL, Japan), XRD (D8 Advance, Bruker, Germany), attenuated total reflection Fourier-transform infrared spectroscopy (ATR-FTIR, Nicolet IS10, Thermo Scientific, USA), and water contact angle measurements. Each membrane sample for SEM measurement was frozen via liquid nitrogen first to preserve the integrity of the membrane structure. Then, the SEM images were obtained at an accelerating voltage of 10 kV. ATR-FTIR spectra could be obtained by utilizing an attenuated total reflectance (ATR) unit to scan ten times in a range of 4000~400 cm−1. The XRD test conditions for the membranes were the same as for the nanomaterials. Water contact angle tests were employed to evaluate membrane hydrophilicity using a contact angle goniometer (OCA20, Dataphysics Instruments, Stuttgart, Germany) via the sessile drop method.

2.7. Filtration Tests of the m-ZIF-8/m-HNT/PAN Membranes

As the two key parameters of evaluating membrane properties, cross-flow filtration was conducted to evaluate the permeability and selectivity of those as-prepared TFN membranes. Each membrane was kept at 0.8 MPa for 30 min, and then, the pressure was changed to 0.4 MPa to evaluate the filtration performance. Water flux (J, L m−2 h−1 MPa −1) for the fabricated membranes can be calculated according to Equation (1).
J = V A Δ t P
where ΔV is the permeate volume (m3), A is the effective membrane area (12.57 cm2), Δt is the filtration time (h), and ΔP is the trans-membrane pressure difference (MPa).
Rejection (R, %) for several salts or dyes, such as NaCl, Na2SO4, MgSO4, MgCl2, and Reactive Red 49, can be achieved by utilizing 0.5 g L−1 of aqueous solution as the feed. Equation (2) is employed to calculate the corresponding rejection:
R = 1 C P C f × 100 %
where Cp and Cf are the concentrations of the feed (Cf, mol L−1) and permeate (Cp, mol L−1), respectively. The salt concentration was determined via the linear relationship between concentration and conductivity, while the dye concentration was determined via the linear relationship between concentration and absorbance at an optimal absorption wavelength of 523.5 nm.

3. Results and Discussion

3.1. Characterization of the m-HNTs and m-ZIF-8 Nanomaterials

Figure 1a,b schematically show the detailed structure of HNTs and ZIF-8, respectively. Generally, HNTs can be dispersed in an aqueous solution for several hours, while ZIF-8 will agglomerate because of a lack of hydrophilic groups. Herein, PSS, with plenty of hydrophilic groups, is utilized to enhance the dispersion of HNTs and ZIF-8 in water [36]. Figure 1c,d illustrate HRTEM images of m-HNTs and m-ZIF-8 at different magnifications. It was apparent that the modification of the PSS did not affect the original structure of the HNT or ZIF-8 materials.
Figure 2a demonstrates the FTIR spectra of the HNTs and m-HNTs. The peaks between 3790~3540 cm−1 refer to the stretching vibrations of O-H bonds in the external (Si-O-H) or internal surface (Al-O-H), and those between 1160~900 cm−1 correspond to the stretching vibrations of O-Si-O or O-Al-O bonds [37]. The small peak at 1230 cm−1, which can be indexed to the stretching vibration of the –SO3Na groups, suggested the successful modification of PSS [38]. In other words, PSS can be adhered to the HNTs’ surfaces via electrostatic interactions. Figure 2b illustrates the FTIR spectra of ZIF-8 and m-ZIF-8. The peaks at 3200~2740 cm−1 can be ascribed to the stretching vibration of C-H in the Hmim [39]. Meanwhile, the stretching and in-plane or out-plane bending vibrations of the imidazole ring are located at around 1520~1350 cm−1 and 1350~900 cm−1, respectively [40]. A sharp peak at 422 cm−1 can be attributed to the stretching vibration of the Zn-N bond, indicative of the coordination of Hmin and Zn2+ [41]. Furthermore, compared with ZIF-8, m-ZIF-8 demonstrates much stronger peak intensities between 1350~900 cm−1 given the existence of –SO3Na groups. Figure 2c,d show the XRD patterns of HNTs, m-HNTs, ZIF-8, and m-ZIF-8. Both HNTs and m-HNTs demonstrate characteristic diffraction reflections at 12.2°, 19.9°, 24.7°, 54.3°, and 62.5° [42], and there are no obvious differences between these two XRD patterns. Similarly, the as-synthesized ZIF-8 and m-ZIF-8 also showcase characteristic peaks at 7.4°, 10.5°, 12.6°, 14.7°, 16.4°, 18.0°, 22.1°, 24.4°, 26.6°, 29.6°, 30.6°, and 32.4°, which correspond to the (011), (002), (112), (002), (013), (222), (114), (223), (134), (004), (244), and (235) crystal planes, respectively [43]. All the results demonstrate that the PSS was attached to the HNT or ZIF-8 surfaces via electrostatic interaction, and the modification of PSS did not affect the original crystalline structures of these two materials.

3.2. Characterization of the m-ZIF-8/m-HNT/PAN Membranes

The m-ZIF-8/m-HNT/PAN membranes could be fabricated by coating the mixtures of m-ZIF-8 and/or m-HNTs and PVA on the PAN substrate, followed by cross-linking with glutaraldehyde, as schematically shown in Figure 3. The obtained membrane is denoted as NFM-X (X = 0, 2.5, 5, 7.5, and 10, where X refers to the loading mass amount of m-HNTs), and the detailed composition of these membranes is illustrated in Table 1.
Figure 4 illustrates the surface and cross-section morphologies for the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 membranes. As shown in Figure 4a,b, all the m-ZIF-8 nanoparticles can be uniformly spread out on the PAN substrate. However, some small protrusions that were constructed by m-ZIF-8 aggregates still exist. The formation of these aggregates may be attributed to the decrease in free surface energy caused by intermolecular interactions such as Van der Waals forces, electrostatic interactions, and so on [44,45]. These interactions jointly formed a short-range driving force that helped to accelerate the m-ZIF-8’s self-assembly into clusters during solvent evaporation. As shown in Figure 4c,e,g, when 2.5 mg, 5 mg, and 7.5 mg of m-HNTs were added to replace an equal amount of m-ZIF-8, the corresponding membranes (NFM-2.5, NFM-5, and NFM-7.5) demonstrated similar surface morphologies. In the meantime, as the number of m-HNTs increased, the size of the m-ZIF-8 aggregates gradually declined, suggesting that the introduction of m-HNTs can effectively inhibit m-ZIF-8 cluster growth, possibly because of the overall increased electrostatic repulsion energy, which determined a finite aggregation. From the cross-section image shown in Figure 4f, it can be clearly observed that m-ZIF-8 is more likely to gather near the air side, while the m-HNTs tend to gather on the other side. This can be explained by the factor of gravity. Compared with m-ZIF-8 (bulk density ~0.35 g cm−3, average size < 100 nm) [46], m-HNTs with a higher density (~2.53 g cm−3, a length of 500~600 nm) [47] are prone to sinking into the PAN substrate more quickly. Figure 4i,j show the morphology of the membrane that was prepared using pure m-HNTs without the addition of m-ZIF-8. The surface morphology of the membranes showed that the m-HNTs were uniformly dispersed on the PAN substrate. Unlike the NFM-0 membranes, no obvious protrusions of m-HNTs were observed on either the surfaces or cross-sections of the NFM-10 membranes.
Furthermore, XRD measurement was implemented to analyze the crystalline properties of the NFM-0, NFM-5, and NFM-10 membranes. As presented in Figure 5a, more new diffraction reflections, attributable to the applied nanomaterials, appear in these membranes compared with the PAN substrate, demonstrating the successful introduction of the m-HNT and m-ZIF-8 nanomaterials. The ATR-FTIR spectra of the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 were then measured and are illustrated in Figure 5b. Compared with the NFM-0 membrane, all the NFM-X membranes demonstrated two obvious peaks at 3697 cm−1 and 3623 cm−1, which can be attributed to the stretching vibrations of Al-OH or Si-OH groups on m-HNTs [48]. The peaks at around 1100~850 cm−1 can be ascribed to the stretching vibrations of O-Si-O and O-Al-Si groups on m-HNTs [49]. It can be seen that, with the gradual replacement of m-ZIF-8 by m-HNTs in the corresponding membranes, the intensities of the peaks at 3760~3530 cm−1 and 1100~850 cm−1 are also enhanced. Furthermore, the in-plane bending vibrations of the imidazole ring at 1305 cm−1 and 1147 cm−1 became weak.

3.3. Hydrophilicity of the m-ZIF-8/m-HNT/PAN Membranes

The hydrophilicity of membranes is one of the critical parameters for indexing their permeability, selectivity, and antifouling performance. Herein, the hydrophilicity of the membranes is evaluated via the water contact angle, as shown in Figure 6. The NFM-0 membrane, which was fabricated by coating the m-ZIF-8 nanomaterials on the PAN substrate, illustrates an average water contact angle of 104.97°. Compared with NFM-0, the rest of the NFM-X membranes showcase lower water contact angles that can be placed in the following sequence: NFM-0 > NFM-2.5 > NFM-5 > NFM-7.5 > NFM-10. This suggests that the existence of m-HNTs can endow TFN membranes with better hydrophilicity. Intrinsically, the hydropathy properties of these TFN membranes can be controlled by altering the number of these two nanomaterials. As frequently reported [50], variations in the water contact angles of membranes are closely related to membrane surface roughness and functional groups, and a low surface roughness generally results in a high water contact angle. In our case, the NFM membrane surface gradually became smooth with the addition of HNTs, as demonstrated by Figure 4, which should theoretically enlarge the water contact angle and is the opposite of the trend observed. Therefore, it can be speculated that the HNTs primarily dominated the overall membrane hydrophilicity since a large number of hydroxyl groups were introduced.

3.4. Separation Performance of the m-ZIF-8/m-HNT/PAN Membranes

Water flux and the rejection of salts and dyes were measured to estimate the separation performance of the m-ZIF-8/m-HNT/PAN membranes, as presented in Figure 7. Water fluxes for the NFM-0, NFM-2.5, NFM-5.0, NFM-7.5, and NFM-10 membranes were 16.7 L m−2 h−1 MPa−1, 21.4 L m−2 h−1 MPa−1, 26.9 L m−2 h−1 MPa−1, 38.3 L m−2 h−1 MPa−1, and 42.6 L m−2 h−1 MPa−1, respectively. Obviously, the water fluxes show the following sequence: NFM-0 < NFM-2.5 < NFM-5 < NFM-7.5 < NFM-10. This is clearly related to the hydrophilicity and structure of the TFN membranes. Specifically, the water flux of the series of TFN membranes was negatively related to the water contact angles. The layer structure of the TFN membranes was also vital. As can be seen in Figure 4 and Figure 8, the water path generated by small m-ZIF-8 nanoparticles was zigzag, long, and narrow for the NFM-0 membrane. Therefore, water molecules were subject to great resistance, hindering the water penetration through the NFM-0 membrane and resulting in low water flux. By contrast, for the NFM-10 membrane, the water path generated by aligned m-HNTs was regular, short, and interconnected. Thus, the water flux was higher than that of the NFM-0 membrane. The water flux of the NFM-2.5, NFM-5, and NFM-7.5 membranes, all composed of two nanomaterials, was intermediate. This scenario has also been reported for other kinds of membranes. For instance, porous graphene oxide can endow the corresponding composite membrane with short transport channels and help water molecules to pass by, and thus, its water flux is higher than a membrane prepared with graphene oxide [51].
The rejection rates of the m-ZIF-8/m-HNT/PAN membranes for Reactive Red 49, NaCl, Na2SO4, MgCl2, and MgSO4 were also measured. All these membranes showed a high rejection rate (exceeding 80%) for Reactive Red 49. In addition to the NFM-2.5 membrane, which showed a slight decrease in dye rejection, other membranes showed an increase in the rejection of Reactive Red 49 with the addition of m-HNTs, and the highest rejection rate achieved was 97.1%. The relatively high retention rate for reactive dyes indicates that the pore size range of our prepared TFN membranes was approximate to the size of the reactive dyes, and the introduction of m-HNTs and m-ZIF-8 particles could not degrade the separation performance of the reactive dyes. Interestingly, rejection rates for the salts conformed with the following order: Na2SO4 > MgSO4 > NaCl (≈MgCl2). This mainly depended on the interactions of the sieve effect and the Donnan effect in the thin film layer of the TFN membrane. Table 2 illustrates the stock radii and hydrated radii of several ions [52]. If the surface charge in the membrane is neutral, considering r(Na+) < r(Mg2+), the rejection rate for Na2SO4 is lower than for MgSO4 (Na2SO4 < MgSO4). The rejection demonstrates a reverse sequence (Na2SO4 > MgSO4). According to the Donnan effect, negatively charged membranes tend to adsorb high-valent cations and repel high-valent anions. Therefore, the TFN membrane can attract stronger positive charges and possess a negatively charged surface. In addition, since r(SO42−) > r(Cl) and the membrane possesses a negatively charged surface, the rejection rate for SO42− is higher than for Cl, resulting in Na2SO4 > NaCl and MgSO4 > MgCl2. The Na2SO4 retention of the NFM-2.5, NFM-5, and NFM-7.5 membranes, all composed of these two nanomaterials, was significantly enhanced by up to 62%, and the retention of other salts varied slightly. Overall, the introduction of hydrophilic m-HNTs and m-ZIF-8 greatly improved the water flux of TFN membranes by up to 38.3 L m−2 h−1 MPa−1 while also maintaining good dye retention (~97.1%) and stable salt retention. Table 3 summarizes the performance of TFN nanofiltration membranes containing other nanoparticles reported in other literature. Obviously, our m-ZIF-8/m-HNT/PAN membranes have both high water flux and good dye/salt retention.

4. Conclusions

In summary, we proposed a new combination of m-ZIF-8 and m-HNTs, attempting to address the poor dispersion of nanoparticles in TFN membranes, and prepared a series of m-ZIF-8/m-HNT/PAN membranes by using the drop-coating method. It was demonstrated that the morphologies of the thin film layer of TFN membranes can be readily changed, varying from zero-dimensional m-ZIF-8 to one-dimensional m-HNTs. Consequently, the performance of these TFN membranes, including water flux and dye rejection, can be also adjusted and controlled. The experimental results showed that the modified m-ZIF-8/m-HNT/PAN membranes increased water flux by up to 42.6 L m−2 h−1 MPa−1. All these membranes showed a high inhibition of Reactive Red 49 (more than 80%) and maintained good salt retention. In particular, the optimized NFM-7.5 membrane showed an optimized separation performance with a 97.1% rejection of Reactive Red 49 and a 21.3% retention of NaCl. We emphasize that the properties of nanomaterials, including dimensionality, hydrophilicity, etc., should be taken into serious consideration when preparing advanced TFN membranes. This work is expected to promote the application of HNT and MOF nanomaterials in membrane separation technology and provide a new way to solve the dispersion of nanofillers in TFN membranes.

Author Contributions

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

Funding

This research was funded by [National Natural Science Foundation of China] [No. 22178327] and [Distinguished Youth Foundation of Henan Scientific Committee] [No. 222300420018]. Furthermore, the authors thank the Center of Advanced Analysis and Computational Science, Zhengzhou University, for characterizations.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, J.; Xie, Y.; Cheng, L.; Li, X.; Liu, F.; Wang, Z. Photo-Fenton reaction derived self-cleaning nanofiltration membrane with MOFs coordinated biopolymers for efficient dye/salt separation. Desalination 2023, 553, 116459. [Google Scholar] [CrossRef]
  2. Yang, L.; Liu, X.; Zhang, X.; Chen, T.; Ye, Z.; Rahaman, M.S. High performance nanocomposite nanofiltration membranes with polydopamine-modified cellulose nanocrystals for efficient dye/salt separation. Desalination 2022, 521, 115385. [Google Scholar] [CrossRef]
  3. Liu, B.; Zhu, T.; Liu, W.; Zhou, R.; Zhou, S.; Wu, R.; Deng, L.; Wang, J.; Van der Bruggen, B. Ultrafiltration pre-oxidation by boron-doped diamond anode for algae-laden water treatment: Membrane fouling mitigation, interface characteristics and cake layer organic release. Water Res. 2020, 187, 116435. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, P.; Zhou, Y.; Cheng, H. Large-scale orientated self-assembled halloysite nanotubes membrane with nanofluidic ion transport properties. Appl. Clay Sci. 2019, 180, 105184. [Google Scholar] [CrossRef]
  5. Zhang, R.; Liu, Y.; Li, Y.; Han, Q.; Zhang, T.; Zeng, K.; Zhao, C. Polyvinylidene fluoride membrane modified by tea polyphenol for dye removal. J. Mater. Sci. 2020, 55, 389–403. [Google Scholar] [CrossRef]
  6. Ji, D.; Xiao, C.; An, S.; Zhao, J.; Hao, J.; Chen, K. Preparation of high-flux PSF/GO loose nanofiltration hollow fiber membranes with dense-loose structure for treating textile wastewater. Chem. Eng. J. 2019, 363, 33–42. [Google Scholar] [CrossRef]
  7. Chen, L.; Shi, G.; Shen, J.; Peng, B.; Zhang, B.; Wang, Y.; Bian, F.; Wang, J.; Li, D.; Qian, Z.; et al. Ion sieving in graphene oxide membranes via cationic control of interlayer spacing. Nature 2017, 550, 380–383. [Google Scholar] [CrossRef]
  8. Chen, X.; Mohammed, S.; Yang, G.; Qian, T.; Chen, Y.; Ma, H.; Xie, Z.; Zhang, X.; Simon, G.P.; Wang, H. Selective permeation of water through angstrom-channel graphene membranes for bioethanol concentration. Adv. Mater. 2020, 32, 2002320. [Google Scholar] [CrossRef]
  9. Xue, S.; Ji, C.; Kowal, M.D.; Molas, J.C.; Lin, C.-W.; McVerry, B.T.; Turner, C.L.; Mak, W.H.; Anderson, M.; Muni, M.; et al. Nanostructured graphene oxide composite membranes with ultrapermeability and mechanical robustness. Nano Lett. 2020, 20, 2209–2218. [Google Scholar] [CrossRef]
  10. Wang, Q.; Wang, Y.; Chen, B.-Z.; Lu, T.-D.; Wu, H.-L.; Fan, Y.-Q.; Xing, W.; Sun, S.-P. Designing high-performance nanofiltration membranes for high-salinity separation of sulfate and chloride in the chlor-alkali process. Ind. Eng. Chem. Res. 2019, 58, 12280–12290. [Google Scholar] [CrossRef]
  11. Rajakumaran, R.; Kumar, M.; Chetty, R. Morphological effect of ZnO nanostructures on desalination performance and antibacterial activity of thin-film nanocomposite (TFN) membrane. Desalination 2020, 495, 114673. [Google Scholar] [CrossRef]
  12. Li, Q.; Liao, Z.; Fang, X.; Wang, D.; Xie, J.; Sun, X.; Wang, L.; Li, J. Tannic acid-polyethyleneimine crosslinked loose nanofiltration membrane for dye/salt mixture separation. J. Membr. Sci. 2019, 584, 324–332. [Google Scholar] [CrossRef]
  13. Van Goethem, C.; Verbeke, R.; Pfanmöller, M.; Koschine, T.; Dickmann, M.; Timpel-Lindner, T.; Egger, W.; Bals, S.; Vankelecom, I. The role of MOFs in Thin-Film Nanocomposite (TFN) membranes. J. Membr. Sci. 2018, 563, 938–948. [Google Scholar] [CrossRef]
  14. Li, C.; Li, S.; Tian, L.; Zhang, J.; Su, B.; Hu, M.Z. Covalent organic frameworks (COFs)-incorporated thin film nanocomposite (TFN) membranes for high-flux organic solvent nanofiltration (OSN). J. Membr. Sci. 2019, 572, 520–531. [Google Scholar] [CrossRef]
  15. Wang, Y.; Wang, X.; Guan, J.; Yang, L.; Ren, Y.; Nasir, N.; Wu, H.; Chen, Z.; Jiang, Z. 110th Anniversary: Mixed matrix membranes with fillers of intrinsic nanopores for gas separation. Ind. Eng. Chem. Res. 2019, 58, 7706–7724. [Google Scholar] [CrossRef]
  16. Yuan, Y.; Tong, C.; Lü, C. Mussel-inspired functionalized LDH as covalent crosslinkers for constructing micro-crosslinking fluorenyl-containing polysulfone-based composite anion exchange membranes with enhanced properties. Appl. Clay Sci. 2020, 199, 105878. [Google Scholar] [CrossRef]
  17. Zhijiang, C.; Cong, Z.; Ping, X.; Jie, G.; Kongyin, Z. Calcium alginate-coated electrospun polyhydroxybutyrate/carbon nanotubes composite nanofiners as nanofiltration membrane for dye removal. J. Mater. Sci. 2018, 53, 14801–14820. [Google Scholar] [CrossRef]
  18. Zhao, D.; Japip, S.; Zhang, Y.; Weber, M.; Maletzko, C.; Chung, T. Emerging thin-film nanocomposite (TFN) membranes for reverse osmosis: A review. Water Res. 2020, 173, 115557. [Google Scholar] [CrossRef]
  19. Eykens, L.; De Sitter, K.; Dotremont, C.; Pinoy, L.; Van der Bruggen, B. How to optimize the membrane properties for membrane distillation: A review. Ind. Eng. Chem. Res. 2016, 55, 9333–9343. [Google Scholar] [CrossRef]
  20. Hao, S.; Wen, J.; Li, S.; Wang, J.; Jia, Z. Preparation of COF-LZU1/PAN membranes by an evaporation/casting method for separation of dyes. J. Mater. Sci. 2020, 5, 14817–14828. [Google Scholar] [CrossRef]
  21. Wang, J.; Lang, W.Z.; Xu, H.P.; Zhang, X.; Guo, Y.J. Improved poly(vinyl butyral) hollow fiber membranes by embedding multi-walled carbon nanotube for the ultrafiltrations of bovine serum albumin and humic acid. Chem. Eng. J. 2015, 260, 90–98. [Google Scholar] [CrossRef]
  22. Zhao, F.Y.; Ji, Y.L.; Weng, X.D.; Mi, Y.F.; Ye, C.C.; An, Q.F.; Gao, C.J. High-flux positively charged nanocomposite nanofiltration membranes filled with poly(dopamine) modified multiwall carbon nanotubes. ACS Appl. Mater. Interfaces 2016, 8, 6693–6700. [Google Scholar] [CrossRef] [PubMed]
  23. Kamal, N.; Ahzi, S.; Kochkodan, V. Polysulfone/halloysite composite membranes with low fouling properties and enhanced compaction resistance. Appl. Clay Sci. 2020, 199, 105873. [Google Scholar] [CrossRef]
  24. Wang, Q.; Cui, J.; Liu, S.; Gao, J.; Lang, J.; Li, C.; Yan, Y. Facile preparation of halloysite nanotube-modified polyvinylidene fluoride composite membranes for highly efficient oil/water emulsion separation. J. Mater. Sci. 2019, 54, 8332–8345. [Google Scholar] [CrossRef]
  25. Xu, Y.; Zhu, Y.; Qiu, Q.; Qi, Z.; Liu, S.; Weng, J.; Shen, J. Development of Mixed-Dimensional Membranes Comprising Halloysite Nanotubes and Kevlar Aramid Nanofiber for Enhanced Small-Molecule Dye/Salt Separation. Ind. Eng. Chem. Res. 2023, 62, 1558–1570. [Google Scholar] [CrossRef]
  26. Kausar, A. Review on polymer/halloysite nanotube nanocomposite. Polym. Plast. Technol. Eng. 2018, 57, 548–564. [Google Scholar] [CrossRef]
  27. Zeng, G.; He, Y.; Zhan, Y.; Zhang, L.; Shi, H.; Yu, Z. Preparation of a novel poly(vinylidene fluoride) ultrafiltration membrane by incorporation of 3-aminopropyltriethoxysilane-grafted halloysite nanotubes for oil/water separation. Ind. Eng. Chem. Res. 2016, 55, 1760–1767. [Google Scholar] [CrossRef]
  28. Zeng, G.; He, Y.; Ye, Z.; Yang, X.; Chen, X.; Ma, J.; Li, F. Novel halloysite nanotubes intercalated graphene oxide based composite membranes for multifunctional applications: Oil/water separation and dyes removal. Ind. Eng. Chem. Res. 2017, 56, 10472–10481. [Google Scholar] [CrossRef]
  29. Mishra, G.; Mukhopadhyay, M. Enhanced antifouling performance of halloysite nanotubes (HNTs) blended poly(vinyl chloride) (PVC/HNTs) ultrafiltration membranes: For water treatment. J. Ind. Eng. Chem. 2018, 63, 366–379. [Google Scholar] [CrossRef]
  30. Ghanbari, M.; Emadzadeh, D.; Lau, W.J.; Lai, S.O.; Matsuura, T.; Ismail, A.F. Synthesis and characterization of novel thin film nanocomposite (TFN) membranes embedded with halloysite nanotubes (HNTs) for water desalination. Desalination 2015, 358, 33–41. [Google Scholar] [CrossRef]
  31. Asempour, F.; Akbari, S.; Kanani-Jazi, M.H.; Atashgar, A.; Matsuura, T.; Kruczek, B. Chlorine-resistant TFN RO membranes containing modified poly(amidoamine) dendrimer-functionalized halloysite nanotubes. J. Membr. Sci. 2021, 623, 119039. [Google Scholar] [CrossRef]
  32. Zhu, L.; Wang, H.; Bai, J.; Liu, J.; Zhang, Y. A porous graphene composite membrane intercalated by halloysite nanotubes for efficient dye desalination. Desalination 2017, 420, 145–157. [Google Scholar] [CrossRef]
  33. Qiu, S.; Xue, M.; Zhu, G. Metal-organic framework membranes: From synthesis to separation application. Chem. Soc. Rev. 2014, 43, 6116–6140. [Google Scholar] [CrossRef] [PubMed]
  34. Beh, J.J.; Ooi, B.S.; Lim, J.K.; Ng, E.P.; Mustapa, H. Development of high water permeability and chemically stable thin film nanocomposite (TFN) forward osmosis (FO) membrane with poly(sodium 4-styrenesulfonate) (PSS)-coated zeolitic imidazolate framework-8 (ZIF-8) for produced water treatment. J. Water Process. Eng. 2020, 33, 101031. [Google Scholar] [CrossRef]
  35. Wang, J.; Wang, Y.; Zhang, Y.; Uliana, A.; Zhu, J.; Liu, J.; Van der Bruggen, B. Zeolitic imidazolate framework/graphene oxide hybrid nanosheets functionalized thin film nanocomposite membrane for enhanced antimicrobial performance. ACS Appl. Mater. Interfaces 2016, 8, 25508–25519. [Google Scholar] [CrossRef] [PubMed]
  36. Gambinossi, F.; Mylon, S.E.; Ferri, J.K. Aggregation kinetics and colloidal stability of functionalized nanoparticles. Adv. Colloid Interface Sci. 2015, 222, 332–349. [Google Scholar] [CrossRef] [PubMed]
  37. Zhao, X.; Zhou, C.; Lvov, Y.; Liu, M. Clay nanotubes aligned with shear forces for mesenchymal stem cell patterning. Small 2019, 15, 1900357. [Google Scholar] [CrossRef] [PubMed]
  38. Zhao, B.; Guo, Z.; Wang, H.; Wang, L.; Qian, Y.; Long, X.; Ma, C.; Zhang, Z.; Li, J.; Zhang, H. Enhanced water permeance of a polyamide thin-film composite nanofiltration membrane with a metal-organic framework interlayer. J. Membr. Sci. 2021, 625, 119154. [Google Scholar] [CrossRef]
  39. Wu, X.; Yang, L.; Meng, F.; Shao, W.; Liu, X.; Li, M. ZIF-8-incorporated thin-film nanocomposite (TFN) nanofiltration membranes: Importance of particle deposition methods on structure and performance. J. Membr. Sci. 2021, 632, 119356. [Google Scholar] [CrossRef]
  40. Wee, L.H.; Martens, J.A.; Vankelecom, I.F. Interfacial synthesis of ZIF-8 membranes with improved nanofiltration performance. J. Membr. Sci. 2017, 523, 561–566. [Google Scholar]
  41. Liu, G.; Jiang, Z.; Cao, K.; Nair, S.; Cheng, X.; Zhao, J.; Gomaa, H.; Wu, H.; Pan, F. Pervaporation performance comparison of hybrid membranes filled with two-dimensional ZIF-L nanosheets and zero-dimensional ZIF-8 nanoparticles. J. Membr. Sci. 2017, 523, 185–196. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Shen, Y.; Hou, J.; Zhang, Y.; Fam, W.; Liu, J.; Bennett, T.D.; Chen, V. Ultraselective pebax membranes enabled by templated microphase separation. ACS Appl. Mater. Interfaces 2018, 10, 20006–20013. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, T.H.; Oh, J.Y.; Hong, S.P.; Lee, J.M.; Roh, S.M.; Kim, S.H.; Park, H.B. ZIF-8 particle size effects on reverse osmosis performance of polyamide thin-film nanocomposite membranes: Importance of particle deposition. J. Membr. Sci. 2019, 570, 23–33. [Google Scholar] [CrossRef]
  44. Liu, M.L.; Li, L.; Sun, Y.X.; Fu, Z.J.; Cao, X.L.; Sun, S.P. Scalable conductive polymer membranes for ultrafast organic pollutants removal. J. Membr. Sci. 2021, 617, 118644. [Google Scholar] [CrossRef]
  45. Luo, D.; Yan, C.; Wang, T. Interparticle forces underlying nanoparticle self-assemblies. Small 2015, 11, 5984–6008. [Google Scholar] [CrossRef]
  46. Basu, S.; Maes, M.; Cano-Odena, A.; Alaerts, L.; De Vos, D.E.; Vankelecom, I.F. Solvent resistant nanofiltration (SRNF) membranes based on metal-organic frameworks. J. Membr. Sci. 2009, 344, 190–198. [Google Scholar] [CrossRef]
  47. Lvov, Y.; Wang, W.; Zhang, L.; Fakhrullin, R. Halloysite clay nanotubes for loading and sustained release of functional compounds. Adv. Mater. 2016, 28, 1227–1250. [Google Scholar] [CrossRef]
  48. He, H.; Xu, P.; Wang, D.; Zhou, H.; Chen, C. Polyoxometalate-modified halloysite nanotubes-based thin-film nanocomposite membrane for efficient organic solvent nanofiltration. Sep. Purif. Technol. 2022, 295, 121348. [Google Scholar] [CrossRef]
  49. Shah, A.A.; Cho, Y.H.; Choi, H.G.; Nam, S.E.; Kim, J.F.; Kim, Y.; Park, Y.-I.; Park, H. Facile integration of halloysite nanotubes with bioadhesive as highly permeable interlayer in forward osmosis membranes. J. Ind. Eng. Chem. 2019, 73, 276–285. [Google Scholar] [CrossRef]
  50. Ang MB, M.Y.; Tang, C.L.; De Guzman, M.R.; Maganto HL, C.; Caparanga, A.R.; Huang, S.H.; Tsai, H.; Hu, C.; Lee, K.; Lai, J. Improved performance of thin-film nanofiltration membranes fabricated with the intervention of surfactants having different structures for water treatment. Desalination 2020, 481, 114352. [Google Scholar] [CrossRef]
  51. Thebo, K.H.; Qian, X.; Zhang, Q.; Chen, L.; Cheng, H.M.; Ren, W. Highly stable graphene-oxide-based membranes with superior permeability. Nat. Commun. 2018, 9, 1486. [Google Scholar] [CrossRef] [PubMed]
  52. Nightingale, E.R. Phenomenological theory of ion solvation. Effective radii of hydrated ions. J. Phys. Chem. 1959, 63, 1381–1387. [Google Scholar] [CrossRef]
  53. Ji, M.; Wang, Z.; Zhu, Y.; Shan, L.; Lu, Y.; Zhang, Y.; Zhang, Y.; Jin, J. Thin-film composite nanofiltration membrane with unprecedented stability in strong acid for highly selective dye/NaCl separation. J. Membr. Sci. 2022, 645, 120189. [Google Scholar] [CrossRef]
  54. Huang, L.; Li, Z.; Luo, Y.; Zhang, N.; Qi, W.; Jiang, E.; Bao, J.; Zhang, X.; Zheng, W.; An, B.; et al. Low-pressure loose GO composite membrane intercalated by CNT for effective dye/salt separation. Sep. Purif. Technol. 2021, 256, 117839. [Google Scholar] [CrossRef]
  55. Li, P.; Wang, Z.; Yang, L.; Zhao, S.; Song, P.; Khan, B. A novel loose-NF membrane based on the phosphorylation and cross-linking of polyethyleneimine layer on porous PAN UF membranes. J. Membr. Sci. 2018, 555, 56–68. [Google Scholar] [CrossRef]
  56. Jin, J.; Du, X.; Yu, J.; Qin, S.; He, M.; Zhang, K.; Chen, G. High performance nanofiltration membrane based on SMA-PEI cross-linked coating for dye/salt separation. J. Membr. Sci. 2020, 611, 118307. [Google Scholar] [CrossRef]
  57. Cao, X.L.; Yan, Y.N.; Zhou, F.Y.; Sun, S.P. Tailoring nanofiltration membranes for effective removing dye intermediates in complex dye-wastewater. J. Membr. Sci. 2020, 595, 117476. [Google Scholar] [CrossRef]
  58. Fan, L.; Ma, Y.; Su, Y.; Zhang, R.; Liu, Y.; Zhang, Q.; Jiang, Z. Green coating by coordination of tannic acid and iron ions for antioxidant nanofiltration membranes. RSC Adv. 2015, 5, 107777–107784. [Google Scholar] [CrossRef]
  59. Liu, S.; Wang, Z.; Song, P. Free Radical Graft Copolymerization Strategy To Prepare Catechin-Modified Chitosan Loose Nanofiltration (NF) Membrane for Dye Desalination, ACS Sustain. Chem. Eng. 2018, 6, 4253–4263. [Google Scholar]
  60. Zhao, S.; Wang, Z. A loose nano-filtration membrane prepared by coating HPAN UF membrane with modified PEI for dye reuse and desalination. J. Membr. Sci. 2017, 524, 214–224. [Google Scholar] [CrossRef]
  61. Kong, G.; Fan, L.; Zhao, L.; Feng, Y.; Cui, X.; Pang, J.; Guo, H.; Sun, H.; Kang, Z.; Sun, D.; et al. Spray-dispersion of ultra-small EMT zeolite crystals in thin-film composite membrane for high-permeability nanofiltration process. J. Membr. Sci. 2021, 622, 119045. [Google Scholar] [CrossRef]
Figure 1. Structures of HNTs (a) and ZIF-8 (b) and HRTEM images of m-HNTs (c) and m-ZIF-8 (d).
Figure 1. Structures of HNTs (a) and ZIF-8 (b) and HRTEM images of m-HNTs (c) and m-ZIF-8 (d).
Membranes 14 00007 g001
Figure 2. (a,b) FITR spectra and (c,d) XRD patterns of the HNTs and ZIF-8 before or after modification via PSS.
Figure 2. (a,b) FITR spectra and (c,d) XRD patterns of the HNTs and ZIF-8 before or after modification via PSS.
Membranes 14 00007 g002
Figure 3. Schematic diagrams for preparing the NFM-X series (a) NFM-0, (b) NFM-2.5, NFM-5 or NFM-7.5, and (c) NFM-10.
Figure 3. Schematic diagrams for preparing the NFM-X series (a) NFM-0, (b) NFM-2.5, NFM-5 or NFM-7.5, and (c) NFM-10.
Membranes 14 00007 g003
Figure 4. SEM images of membrane surfaces and cross-sections of NFM-0 (a,b), NFM-2.5 (c,d), NFM-5 (e,f), NFM-7.5 (g,h), and NFM-10 (i,j) membranes.
Figure 4. SEM images of membrane surfaces and cross-sections of NFM-0 (a,b), NFM-2.5 (c,d), NFM-5 (e,f), NFM-7.5 (g,h), and NFM-10 (i,j) membranes.
Membranes 14 00007 g004
Figure 5. XRD pattern (a) and ATR-FTIR spectra (b) of the PAN and NFM-X membranes.
Figure 5. XRD pattern (a) and ATR-FTIR spectra (b) of the PAN and NFM-X membranes.
Membranes 14 00007 g005
Figure 6. Water contact angle results for the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 membranes.
Figure 6. Water contact angle results for the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 membranes.
Membranes 14 00007 g006
Figure 7. The pure water flux (black dotted line) and rejection of Reactive Red 49 (light gray dotted line) and salts for the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 membranes.
Figure 7. The pure water flux (black dotted line) and rejection of Reactive Red 49 (light gray dotted line) and salts for the NFM-0, NFM-2.5, NFM-5, NFM-7.5, and NFM-10 membranes.
Membranes 14 00007 g007
Figure 8. Schematic diagrams of the water transport paths for the NFM-0, NFM-5, and NFM-10 membranes.
Figure 8. Schematic diagrams of the water transport paths for the NFM-0, NFM-5, and NFM-10 membranes.
Membranes 14 00007 g008
Table 1. The fabrication of m-ZIF-8/m-HNT/PAN membranes.
Table 1. The fabrication of m-ZIF-8/m-HNT/PAN membranes.
SamplesTotal Number of Nanomaterials (mg)m-HNTs (mg)m-ZIF-8 (mg)
NFM-010010
NFM-2.5102.57.5
NFM-51055
NFM-7.5107.52.5
NFM-1010100
Table 2. The stock radii and hydrated radii of several ions.
Table 2. The stock radii and hydrated radii of several ions.
Ionsrs (Å)rH (Å)
Na+1.843.58
Mg2+3.474.28
Cl1.213.32
SO42−2.33.79
Table 3. Comparison of dye/salt separation performance of m-ZIF-8/m-HNT/PAN membranes with data from the literature.
Table 3. Comparison of dye/salt separation performance of m-ZIF-8/m-HNT/PAN membranes with data from the literature.
MembranesDye MoleculeDye Rejection (%)NaCl Rejection (%)Water Flux (L m−2 h−1)Ref.
bisAPAF-TMC/PESDirect Red 2398.719.517.0[53]
c-CNT@GOMethyl Blue94.11.326.3[54]
Fe(III)-phos-(PEI)/HPANMethyl Blue99.97.58.5[55]
SMA-PEI/PESCongo Red99.72.423[56]
M-PIPReactive black 599.011.026.4[57]
TA-Fe3+/PESCongo red99.0527.2[58]
Catechin-chitosan/HPANAcid fuchsin98.712.514.4[59]
GA/PEI-MMethyl Blue90.05.025.5[60]
S-EMT/PA-2Methyl blue98.928.424.4[61]
m-ZIF-8/m-HNTs/PANReactive Red 4997.121.317.0This work
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

Wang, Y.; Duan, S.; Wang, H.; Wei, C.; Qin, L.; Dong, G.; Zhang, Y. Thin Film Nanocomposite Membranes Based on Zeolitic Imidazolate Framework-8/Halloysite Nanotube Composites. Membranes 2024, 14, 7. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes14010007

AMA Style

Wang Y, Duan S, Wang H, Wei C, Qin L, Dong G, Zhang Y. Thin Film Nanocomposite Membranes Based on Zeolitic Imidazolate Framework-8/Halloysite Nanotube Composites. Membranes. 2024; 14(1):7. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes14010007

Chicago/Turabian Style

Wang, Yan, Shaofan Duan, Huixian Wang, Can Wei, Lijuan Qin, Guanying Dong, and Yatao Zhang. 2024. "Thin Film Nanocomposite Membranes Based on Zeolitic Imidazolate Framework-8/Halloysite Nanotube Composites" Membranes 14, no. 1: 7. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes14010007

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