Next Article in Journal
Toward Suppressing Charge Trapping Based on a Combined Driving Waveform with an AC Reset Signal for Electro-Fluidic Displays
Next Article in Special Issue
New Insights into the Fouling of a Membrane during the Ultrafiltration of Complex Organic–Inorganic Feed Water
Previous Article in Journal
A Summary of Practical Considerations for the Application of the Steric Exclusion Chromatography for the Purification of the Orf Viral Vector
Previous Article in Special Issue
Development of Ultrafiltration Kaolin Membranes over Sand and Zeolite Supports for the Treatment of Electroplating Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tailoring Morphology and Properties of Tight Utrafiltration Membranes by Two-Dimensional Molybdenum Disulfide for Performance Improvement

1
Key Laboratory of Ecology of Rare and Endangered Species and Environmental Protection, College of Life Sciences, Guangxi Normal University, Ministry of Education, Guilin 541000, China
2
Key Laboratory of Urban Pollutant Conversion, Institute of Urban Environment, Chinese Academy of Sciences, Xiamen 361021, China
3
CSIRO Manufacturing, Clayton South, Victoria 3169, Australia
4
Key Laboratory of Marine Environment and Ecology, Ministry of Education, Ocean University of China, Qingdao 266100, China
*
Authors to whom correspondence should be addressed.
Membranes 2022, 12(11), 1071; https://doi.org/10.3390/membranes12111071
Submission received: 13 October 2022 / Revised: 24 October 2022 / Accepted: 25 October 2022 / Published: 29 October 2022
(This article belongs to the Special Issue UF/NF/RO Membranes for Wastewater Treatment and Reuse)

Abstract

:
To enhance the permeation and separation performance of the polyethersulfone (PES) tight ultrafiltration (TUF) membrane, two-dimensional molybdenum disulfide (MoS2) was applied as a modifier in low concentrations. The influence of different concentrations of MoS2 (0, 0.25, 0.50, 1.00, and 1.50 wt%) on TUF membranes was investigated in terms of morphology, mechanical strength properties, permeation, and separation. The results indicate that the blending of MoS2 tailored the microstructure of the membrane and enhanced the mechanical strength property. Moreover, by embedding an appropriate amount of MoS2 into the membrane, the PES/MoS2 membranes showed improvement in permeation and without the sacrifice of the rejection of bovine serum protein (BSA) and humic acid (HA). Compared with the pristine membrane, the modified membrane embedded with 0.5 wt% MoS2 showed a 36.08% increase in the pure water flux, and >99.6% rejections of BSA and HA. This study reveals that two-dimensional MoS2 can be used as an effective additive to improve the performance and properties of TUF membranes for water treatment.

1. Introduction

To meet the growing demand for clean water, water recovery from wastewater by membrane technologies has been an emerging strategy [1,2]. The ultrafiltration membrane (UF) has become an interesting strategy because of its effectiveness in removing particles, microorganisms, and other organic contaminants [3]. Among them, the tight ultrafiltration (TUF) membrane (a molecular weight cut-off (MWCO) of ~1000–10000 Da [4]) has a higher retention performance on organic material. Therefore, using TUF for the removal of natural organic matter (NOM) in water has attracted more and more attention [5,6,7]. However, current TUF membranes face a critical challenge: the trade-off between water permeability and selectivity. In general, conventional polymer-based UF membranes could not break the “upper bound” between the separation factor and membrane permeability [8]. Since the steric effect is the main separation mechanism of the UF membrane [9], and the selectivity of the UF membrane is determined by the pore radius (R) and pore radius distribution (σ), while permeability is mainly determined by pore radius (R), porosity (ε) and the thickness of selectivity layer (δm). The selectivity and permeability are inversely related [10]. Therefore, conventional modification approaches make it difficult to break through the trade-off limitation.
Recently, to enhance the water flux and separation performance of the membranes, polymer matrix membranes are advanced membranes incorporating inorganic materials into an organic or a polymer matrix [11,12,13,14,15,16,17,18]. With a high surface area and adjustable interlayer spacing, two-dimensional (2D) materials such as graphene and its derivatives can create a powerful platform to accommodate convenient transport carriers and facilitate transportation [19]. For membrane modification, previous research demonstrated that graphene oxide (GO) enhances the water permeation and fouling resistance in the membrane [20,21]. Molybdenum disulfide (MoS2), another typical 2D material, is one of the transition-metal dichalcogenides (TMDCs) that has been studied extensively [22]. MoS2 occurs naturally on the earth’s crust as a molybdenum mineral, making it easier to produce on a large scale [23], and multilayered MoS2 can be obtained by simply exfoliating [24,25,26]. Moreover, the distance between two-layer MoS2 is about 0.62 nm, and the laminar channel spacing is about 0.29 nm, which is slightly larger than the size of water molecules [24,27]. In addition, MoS2 does not have additional functional groups, like the GO surface. Therefore, the water channels in MoS2 are smooth, resulting in a 2–10 times faster water passage rate compared to GO [28,29,30]. In addition, MoS2 exhibits good stability in a wide range of pH aqueous solutions and mechanical stability under pressure [31,32]. Therefore, MoS2 can be used as a suitable 2D building block for the fabrication of separation membranes with relatively fixed-size nanochannels [29]. These advantages of MoS2 may bring more water molecular channels to the mixed-matrix membrane and improve permeability without sacrificing membrane interception. Recently, there are only a few studies which use MoS2 to modify membranes. For example, in the preparation of thin film nanocomposite (TFN) film, 2D MoS2 was used to introduce polyimide (PA), a selective layer, to improve the salt rejection, water permeability, hydrophilicity, electronegativity, and anti-pollution properties of the membrane [24,33,34]. In addition, in the UF membrane, functionalized MoS2 was added to the membrane matrix to improve the membrane permeability and anti-pollution properties [35,36]. However, most studies ignored the detailed study of the regulation of membrane pores and membrane morphology by 2D MoS2. In addition, there are few detailed studies on the NOM removal of MoS2 mixed-matrix membranes [23]. In our previous study, MoS2 was introduced into the PES matrix, and we found that a high concentration (3.0wt%) of MoS2 could reduce the pore size of the membrane and improve the permeation flux of the membrane [37]. However, the high concentration of MoS2 was bound to burden the cost of membrane preparation.
In this work, to reduce the adverse effect caused by the high concentration of MoS2 and investigate the influence of MoS2 on membrane performance, we fabricated a series of PES/MoS2 TUF membranes by the phase inversion method and used a low content of 2D MoS2 as additives. The morphology and filtration performance of as-prepared membranes were studied and the optimal loading of MoS2 in the casting solution of TUF membranes was investigated. We obtained a higher permeability membrane without reducing the rejection of humic acid (HA) and bovine serum (BSA) protein, which indicates that commercial MoS2 has great potential for membrane modification.

2. Experimental

2.1. Materials

Polyethersulfone (Ultrason E6020P) was purchased from BASF (Ludwigshafen, Germany). Molybdenum disulfide (MoS2) (99.5% metals basis, <2 μm) and bovine serum albumin (BSA, 67 KDa) were obtained from Aladdin Chemistry Co. Ltd., Shanghai, China. N,N-Dimethylacetamide (DMAc) and polyvinylpyrrolidone (PVP) were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. Polyethylene glycol PEG (2000 Da, 4000 Da, 6000 Da, 8000 Da, 10,000 Da) and HA was acquired from Sigma Aldrich (Sigma–Aldrich, Inc., St. Louis, MO, USA).

2.2. Membrane Preparation

Different amounts of dry MoS2 were added to DMAc solution and ultrasonicated for 4 h for well dispersion (QSONICA SONICATORS, Newtown, CT, USA, 500W, 75%). Then, polymer (PES) and membrane pore-forming agents (PVP) were dissolved in the solvent, and the homogeneous casting solution was obtained by heating and stirring. The viscosity of the casting solution containing different concentrations of MoS2 additives is shown in Table 1. To prepare the membrane, the solution was cast on non-woven fabrics, and then immersed into the deionized water (DI) water bath. Finally, the prepared membrane was immersed in another water bath for 24 h to remove residues and then stored in DI water prior to use. The synthesis and filtration process diagram of PES/MoS2 membrane is shown in Figure 1.

2.3. Characterization of MoS2 and Membranes

The morphologies of MoS2 were characterized by file emission scanning electron microscopy (FESEM, HITACHIS-4800, Hitachi, Tokyo, Japan) and transmission electron microscopy (TEM, H-7650, Hitachi, Japan). The zeta potential (Zeta PALS, Malvern Instruments Ltd., Malvern, UK) was used to characterize the electronegativity of MoS2.
To verify that the 2D MoS2 sheets were incorporated into the UF membranes successfully, X-ray diffraction (XRD X’Pert, Pro, PANalytical, Almelo, Netherlands) was used to investigate the MoS2 and the prepared membranes in a range of 10–70°. Morphological structures of the membranes were examined by using a FESEM (HITACHIS-4800, Hitachi, Japan) and TEM (H-7650, Hitachi, Japan). The hydrophilicity of the membranes was determined using a contact angle goniometer (Dataphysics OCA20, Dataphysics, Filderstadt, Germany). A water drop with a volume of 5 μL was dropped onto the top surface of each membrane. At least three contact angles at different locations on each sample were recorded to obtain a reliable value. The mechanical properties of prepared membranes were analyzed by measuring tensile stress using tensile testing equipment (LRK-500N, NTS, Tokyo, Japan). The membrane samples were stretched at an elongation rate of 100 mmmin−1. The membrane was initially fixed by grips at a distance of 55 mm, after which the movable crosshead containing the load cell of 500N pulled the membrane until the membrane broke.

2.4. Molecular Weight Cut-off, Pore Size, Porosity, and Filtration Performance of Membranes

The MWCO of membranes was defined by polyethylene glycol (PEG) solutions with a concentration of 1.0 gL−1, and the retention rate was 90% [38]. Rejection measurements were performed using the same dead-end filtration cell (Model 8010, Millipore Corp, Burlington, MA, USA) at a pressure of 0.1 MPa after a 30 min pre-compaction at 0.15 MPa. The concentrations of PEG in the feed solution and permeate were measured by a total organic carbon analyzer (TOC,TOC-LCSH, Shimadzu, Tokyo, Japan) [39]. The PEG rejection of as-prepared membranes was calculated by Equation (1):
R (%) = (1 − Cp/Cf) × 100
where Cp and Cf were the PEG concentrations of permeate and feed solutions, respectively (gL−1).
The pore diameter of the membrane was equal to the Stokes radius (ds) of PEG at a 50% rejection, which could be calculated by Equation (2) [40]:
d s = 2 × 16.73 × 10 12 × M P E G 0.557
The porosity ε (%) was determined by a gravimetric method, as defined by Equation (3) [41]:
ε   ( % ) = ( w w w d ) D W ( w w w d ) D W   + W d D P × 100 %  
where ε is the porosity of membranes (%); Ww and Wd are related to the wet weight and the dry weight of the membrane (g), respectively; Dw (0.998 g cm−3) is the density of the water; and Dp (0.37 g cm−3) is the density of polymer.
The water flux was tested by a dead-end filtration cell with a volume capacity of 10 mL and the effective area of the membrane was 4.1 cm2 at room temperature (25 ± 1 °C). The samples were prepressed at 0.15 MPa for 30 min with DI water as the feed solution, ensuring that the flux reaches a steady state, and then the pure water flux of each membrane sample was recorded at 0.1 MPa by monitoring the weight change in permeate with an electronic balance (Sartorius BS224S, Sartorius AG, Goettingen, Germany). The permeate flux (J0) was calculated by Equation (4):
J 0 = Δ V A m Δ t
where J0 (Lm−2h−1) was the membrane flux, ΔV (L) was the volume of permeated water during the period of permeation time Δt (h), and A (m2) was the testing membrane area.
The filtration performances of the membranes were investigated by dead-end filtration cell and using BSA and HA as model solutes. The BSA (1.0 gL−1) solution in PBS (pH = 7.4), and HA (20 mgL−1) in DI water. The concentration of these two chemical substances was measured by ultraviolet–visible (UV–vis) spectrophotometer (Spectra Max M2, Molecular) at 280 nm (BSA) and 254 nm (HA), respectively. The permeation flux and BSA/HA rejections were calculated using Equations (1) and (4), respectively.

3. Results and Discussion

3.1. Characterization of MoS2

The morphology of original MoS2 and MoS2 after 4 h of sonication was observed by SEM and TEM images. As shown in Figure 2, it can be observed that the MoS2 flakes with ultrasonic treatment (Figure 2b) have enhanced dispersion compared to the untreated MoS2 flakes (Figure 2a). The original MoS2 comprised a multilayered structure, as shown in Figure 2c. Compared to that, the lamellar number of MoS2 was significantly reduced (Figure 2d). The reduced layer number was further confirmed by TEM images in Figure 2e,f.
The crystallite structure of MoS2 was measured by XRD analysis, and these patterns are represented in Figure 3a. The peaks at 2θ = 14.4°, 29°, 39.6°, 44.2°, 49.8°, and 60.2° were attributed to (002), (004), (103), (104), (105), and (110) planes of MoS2, respectively [42,43]. The zeta potential of MoS2 is shown in Figure 3b. The zeta potential of MoS2 was −31.5 ± 5.3 mV at pH 3 and gradually reduced to −41.35 ± 2.90 mV at pH 10. The decrease zeta potential of MoS2 was because of the oxidation of Mo. Mo can exist in the form of HMoO4 or MoO42− in the aqueous solution, making MoS2 negatively charged [44,45].

3.2. Characterization of MoS2/PES Membrane °

The microstructure of the fabricated membranes was observed by SEM (Figure 4). Compared to the smooth surface of the pure PES membrane M0, there was some MoS2 exposed on the surface of PES/MoS2 membranes. With an increase in the concentration of MoS2 in casting solutions, more MoS2 was observed on the membrane surface. This result was associated with the influence of MoS2 on the membrane fabrication process. In the phase inversion process, the MoS2 migrated from the PES matrix to the water bath, which reduced the interfacial energy between the casting solution and the water bath. As a result, the content of MoS2 locally increased on the skin layer of PES membranes and more MoS2 can be observed on the surface of PES membranes [46]. It was observed that the M0 membrane was full of sponge-like structures. The introduction of MoS2 led to more finger-like pore structures in the membrane, as confirmed by the comparison between SEM images of the M0 membrane and the other PES/MoS2 membranes in Figure 4. Moreover, with an increase in the MoS2 concentration from 0 (M0) to 0.50 (M2) wt%, the finger-like structures became longer and wider. Additionally, the microvoids were closer to the skin layer. The major reason for this result was due to the addition of MoS2 which accelerated the instantaneous exchange of solvent and non-solvent in the casting solution, leading to the formation of microvoids in the membranes [47,48]. In addition, MoS2 sheets migrated to the skin layer of the membrane when the solvent and non-solvent exchange happened, which resulted in the formation of finger-like structures in the membrane interior.
The cross-sectional structures were further investigated by TEM, as shown in Figure 5. This indicates that the MoS2 dispersed evenly in the PES matrix in the PES/MoS2 membrane. With an increase in the MoS2 concentration, the MoS2 brought various porous structures to the PES matrix. This was due to the formation of cavities between 2D sheets and polymers. Interestingly, some cavities appeared in the dense skin layer of the PES/MoS2 membranes. By using Image J software, the thickness of the dense skin layer in each membrane was measured. This indicates that the skin layer thickness increased with the increased concentration of MoS2 in membranes, as M0 (3.46 μm) < M2 (3.83 μm) < M4 (4.80 μm). This may be related to the increased viscosity of the casting solution by adding MoS2, as shown in Table 1.
Figure 6 shows the XRD results of the PES membrane and the PES/MoS2 membranes. All the membranes showed typical peaks representing the amorphous region of PES in 2θ at 17.6°, 22.6°, and 25.9° [49]. Compared to the XRD result of the M0 membrane, there were new peaks in the spectra of PES/MoS2 membranes. Peaks at 2θ=14.4°, 29°, 39.6°, 44°, 49.8°, and 60.1° were assigned to (002), (004), (103), (104), (105) and (110) planes of MoS2, which were matched well with JCPDS card no. 37–1492 [50,51,52]. Moreover, the M4 membrane showed more obvious typical MoS2 peaks due to the high concentration of MoS2 in the membrane. These results indicate that MoS2 was successfully blended in PES/MoS2 membranes.
To investigate the effect of the MoS2 concentration on the hydrophilicity of membrane surface, the water contact angles (CAs) of the membranes were measured. It was found that the contact angle of the M0 membrane was approximately 47.9° (Figure 7a). With the concentration of the MoS2 increasing, the CA values of the TUF membranes increased gradually, which can be ascribed to the hydrophobic nature of the MoS2 [53].
The mechanical strength properties of the PES and PES/MoS2 membranes were expressed in terms of tensile strength and elongation at break. The result in Figure 7 indicates that the addition of MoS2 improved the tensile strength of the modified membranes which was in line with other studies on MoS2 [25]. For example, the M0 membrane was found to have a tensile strength of 36.09 MPa at elongation at a break of 7.98%. Compared to that, the M4 membrane had a tensile strength of 48.59 MPa at a break of 10.97%. In the mixed-matrix membranes, MoS2 could cross-link with the polymeric chains and increase the rigidity of polymeric chains. Due to this, the mechanical strength of mixed-matrix membranes was enhanced. However, the tensile strength did not show a gradual increase with increasing MoS2 content for membranes M1-M4. When the concentration of MoS2 was 0.50 wt%, the membrane M2 showed lower tensile strength than that of the membrane M1, which was due to the longer and wider finger-like structures in the membrane M2. The porous structures in the M2 membranes reduced the tensile strength.
The MWCOs of as-prepared membranes were measured, and the results are shown in Figure 8. The fabricated membranes exhibited an MWCO range between 7.80 and 9.78 kDa, which confirmed that these membranes were TUF membranes. Compared to the pristine PES membrane with 7.83 kDa of MWCO, the MWCO increased to 9.78 kDa for the membrane M1 and 9.67 kDa for the membrane M2, but then dropped to 7.83 kDa for the membrane M3 and 7.80 kDa for the membrane M4. The corresponding pore radius of the TUF membranes is shown in Table 2. The results indicate that the pore sizes of the membranes increased as the concentration of MoS2 in the casting solutions was 0 to 0.50 wt%, while higher contents of nanofillers (1.00–1.50 wt%) may lead to the MoS2 having a strong interaction with PES and DMAc, which reflected an increase in the viscosity of the casting solution. These effects led to a delay in PES precipitation, a decrease in the diffusion rate of water in the PES matrix, and a delay in liquid–liquid stratification, which inhibited the synthesis of macropores [54,55] and consequently reduced the membrane pore size.
Table 2 also shows the pure water flux of the prepared TUF membranes. The result indicates that the pure water flux of membranes initially improved with the increased content of MoS2 when the concentration was 0 to 0.50 wt%. The pure water flux of the M0 membrane without MoS2 was found to be 72.37 Lm−2h−1, and it increased to 91.86 Lm−2h−1 for the M1 membrane. The water flux permeability of UF membranes was determined by properties including the porosity, pore size, and thickness of the membranes’ skin layer. Compared to the M0 membrane, the M1 membrane has similar porosity but a larger MWCO and mean pore radius. Therefore, the improved permeability in the M1 membrane was potentially dominated by the influence of the large pore size. Similar results could also be found when comparing the performance of the M2 and M3 membranes. Compared to the M0 membrane, the M2 membrane exhibited both improved porosity and pore size, which resulted in 1.36 times higher water flux (98.48 Lm−2h−1) than that of the M0 membrane. However, a further increase in the MoS2 concentration increased the membrane surface hydrophobicity, increased the thickness of the skin layer, and the reduced the membrane pore size, which consequently decreased the water flux of the M4 membrane to 80.18 Lm−2h−1.
Figure 9 shows the filtration performance of as-prepared TUF membranes by using BSA solution as the feed solution. The result indicates that all as-prepared PES/MoS2 membranes have higher BSA rejections than the pristine membrane (M0), and all of them were over 99.50%. This was related to the fact that the negative charge brought by MoS2 increased the negative charge on the membrane surface. The permeation flux of the BSA solution for the M0 membrane was 62.37 Lm−2h−1. The water flux increased to 68.29 Lm−2h−1 for the M1 membrane, and the rejection increased to 99.85%. As a function of the MoS2 concentration, the permeation flux decreased gradually. An initial increase in the permeation flux was attributed to the pore size change, while the following decrease was due to the increased hydrophobicity of the membranes [56].
The filtration performance of prepared TUF membranes was further investigated by using HA aqueous solution as the feed solution. This result shows in Figure 10, indicates that all the membranes have excellent separation performance for HA with the rejection rate between 99.21 and 99.60%. Moreover, the permeation flux of the HA solution for the M1 membrane (0.25 wt%) was 77.24 Lm−2h−1, which was 33.4% higher than that of the M0 membrane (57.90 Lm−2h−1). This increase was primarily dependent on the change in pore size. However, it was also found that the effective pore sizes of the M1 and M3 membranes were almost the same and the permeation was changed. This phenomenon may be caused by an increase in the surface hydrophobicity of the membrane due to an increase in the content of MoS2. HA adsorption onto a PES membrane has been confirmed in other research and mainly correlated to the hydrophobic groups of molecules [57]. The effect of adsorption reduced the permeability of the membrane. The results show that further research on enhanced hydrophilicity of 2D MoS2 would improve the antifouling performance of the mixed-matrix membrane. Table 3 compares the HA separation performance of the PES/MoS2 membrane with some other membranes in the previous literature [23,58,59,60,61,62], and indicates that the PES/MoS2 membrane has better HA selectivity compared to UF membranes and improved water permeability compared to NF membranes with similar selectivity properties.

4. Conclusions

In this work, 2D MoS2 was blended into the PES matrix to enhance the performance and properties of membranes. A series of PES/MoS2 membranes were fabricated and their morphologies and separation performance were investigated. The results indicate that the mechanical strength property of membranes was improved by the addition of MoS2. SEM images proved the presence of MoS2 in the mixed-matrix membrane, and the finger-like macrovoid appeared. Further, the dispersed morphology was given by TEM. These changes enhanced the permeability of both pure water and the BSA/HA solution. By optimizing the concentration of MoS2, the PES/MoS2 membrane with 0.50 wt% MoS2 exhibited a 36.08% increase in the water flux compared to the pristine TUF membrane, and excellent rejection properties to BSA (99.85%) and HA (99.60%). This study exhibited the feasibility of applying low concentrations of MoS2 to tune the structure and properties of TUF membranes, which could be an effective strategy to fabricate high-performance TUF membranes.

Author Contributions

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

Funding

This research was funded by grants from the Bureau of Frontier Sciences and Education (QYZDB-SSW-DQC044), the Bureau of International Cooperation (132C35KYSB20160018), the Chinese Academy of Sciences, and the Joint Project between CAS-CSIRO (132C35KYSB20170051).

Data Availability Statement

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

Acknowledgments

This work was supported through the grants from the Bureau Frontier Sciences and Education (QYZDB-SSW-DQC044), the Bureau of International Cooperation (132C35KYSB20160018), Chinese Academy of Sciences, and the Joint Project between CAS-CSIRO. The authors acknowledge Manying Zhang, Li Yi, Qingliang Jiang, Olusegun K. Abass, and Fang Fang for their assistance in the membrane preparation. The authors also thank the technical help from Hongyun Ren for SEM and TEM analysis.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. McCloskey, B.D.; Park, H.B.; Ju, H.; Rowe, B.W.; Miller, D.J.; Freeman, B.D. A bioinspired fouling-resistant surface modification for water purification membranes. J. Membr. Sci. 2012, 413, 82–90. [Google Scholar] [CrossRef]
  2. Yu, W.; Liu, T.; Crawshaw, J.; Liu, T.; Graham, N. Ultrafiltration and nanofiltration membrane fouling by natural organic matter: Mechanisms and mitigation by pre-ozonation and pH. Water Res. 2018, 139, 353–362. [Google Scholar] [CrossRef] [PubMed]
  3. Lee, J.; Jeong, S.; Ye, Y.; Chen, V.; Vigneswaran, S.; Leiknes, T.; Liu, Z. Protein fouling in carbon nanotubes enhanced ultrafiltration membrane: Fouling mechanism as a function of pH and ionic strength. Sep. Purif. Technol. 2017, 176, 323–334. [Google Scholar] [CrossRef] [Green Version]
  4. Winter, J.; Barbeau, B.; Berube, P. Nanofiltration and Tight Ultrafiltration Membranes for Natural Organic Matter Removal-Contribution of Fouling and Concentration Polarization to Filtration Resistance. Membranes 2017, 7, 34. [Google Scholar] [CrossRef] [Green Version]
  5. Lowe, J.; Hossain, M.M. Application of ultrafiltration membranes for removal of humic acid from drinking water. Desalination 2008, 218, 343–354. [Google Scholar] [CrossRef]
  6. Yu, Y.; Wu, Y.; Xie, C.; Sun, X.; Wang, Y.; Liu, P.; Wang, Y.; Liu, C.; Wan, Y.; Pan, W.; et al. High-flux, antifouling and highly hydrophilic tight ultrafiltration membranes based on crosslinked PEEKWC/PEI containing positively charged water channel for dyes removal. Chem. Eng. Res. Des. 2022, 188, 1–14. [Google Scholar] [CrossRef]
  7. Shang, R.; Verliefde, A.R.D.; Hu, J.; Heijman, S.G.J.; Rietveld, L.C. The impact of EfOM, NOM and cations on phosphate rejection by tight ceramic ultrafiltration. Sep. Purif. Technol. 2014, 132, 289–294. [Google Scholar] [CrossRef]
  8. Mehta, A.; Zydney, A.L. Permeability and selectivity analysis for ultrafiltration membranes. J. Membr. Sci. 2005, 249, 245–249. [Google Scholar] [CrossRef]
  9. Lin, C.-E.; Wang, J.; Zhou, M.-Y.; Zhu, B.-K.; Zhu, L.-P.; Gao, C.-J. Poly(m-phenylene isophthalamide) (PMIA): A potential polymer for breaking through the selectivity-permeability trade-off for ultrafiltration membranes. J. Membr. Sci. 2016, 518, 72–78. [Google Scholar] [CrossRef]
  10. Mochizuki, S.; Zydney, A.L. Theoretical analysis of pore size distribution effects on membrane transport. J. Membr. Sci 1993, 82, 211–227. [Google Scholar] [CrossRef]
  11. Choi, H.G.; Son, M.; Choi, H. Integrating seawater desalination and wastewater reclamation forward osmosis process using thin-film composite mixed matrix membrane with functionalized carbon nanotube blended polyethersulfone support layer. Chemosphere 2017, 185, 1181–1188. [Google Scholar] [CrossRef] [PubMed]
  12. Zhao, X.; Ma, J.; Wang, Z.; Wen, G.; Jiang, J.; Shi, F.; Sheng, L. Hyperbranched-polymer functionalized multi-walled carbon nanotubes for poly (vinylidene fluoride) membranes: From dispersion to blended fouling-control membrane. Desalination 2012, 303, 29–38. [Google Scholar] [CrossRef]
  13. Moghadam, F.; Park, H.B. Two-dimensional materials: An emerging platform for gas separation membranes. Curr. Opin. Chem. Eng. 2018, 20, 28–38. [Google Scholar] [CrossRef]
  14. Mohan, V.B.; Lau, K.-T.; Hui, D.; Bhattacharyya, D. Graphene-based materials and their composites: A review on production, applications and product limitations. Compos. Part B Eng. 2018, 142, 200–220. [Google Scholar] [CrossRef]
  15. Zhang, M.; Field, R.W.; Zhang, K. Biogenic silver nanocomposite polyethersulfone UF membranes with antifouling properties. J. Membr. Sci. 2014, 471, 274–284. [Google Scholar] [CrossRef]
  16. Zodrow, K.; Brunet, L.; Mahendra, S.; Li, D.; Zhang, A.; Li, Q.; Alvarez, P.J. Polysulfone ultrafiltration membranes impregnated with silver nanoparticles show improved biofouling resistance and virus removal. Water Res. 2009, 43, 715–723. [Google Scholar] [CrossRef] [Green Version]
  17. Nguyen, T.P.; Yang, S.H. Hybrid materials based on polymer nanocomposites for environmental applications. In Polymer-based Nanocomposites for Energy and Environmental Applications; Woodhead Publishing: Sawston, UK, 2018; pp. 507–551. [Google Scholar]
  18. Goh, P.S.; Ng, B.C.; Lau, W.J.; Ismail, A.F. Inorganic Nanomaterials in Polymeric Ultrafiltration Membranes for Water Treatment. Sep. Purif. Rev. 2014, 44, 216–249. [Google Scholar] [CrossRef]
  19. Li, H.; Lu, G.; Yin, Z.; He, Q.; Li, H.; Zhang, Q.; Zhang, H. Optical identification of single- and few-layer MoS(2) sheets. Small 2012, 8, 682–686. [Google Scholar] [CrossRef]
  20. Wang, Z.; Yu, H.; Xia, J.; Zhang, F.; Li, F.; Xia, Y.; Li, Y. Novel GO-blended PVDF ultrafiltration membranes. Desalination 2012, 299, 50–54. [Google Scholar] [CrossRef]
  21. Ganesh, B.M.; Isloor, A.M.; Ismail, A.F. Enhanced hydrophilicity and salt rejection study of graphene oxide-polysulfone mixed matrix membrane. Desalination 2013, 313, 199–207. [Google Scholar] [CrossRef]
  22. Ganatra, R.; Zhang, Q. Few-Layer MoS2 A Promising Layered Semiconductor. ACS Nano 2014, 8, 4074–4099. [Google Scholar] [CrossRef] [PubMed]
  23. Mallya, D.S.; Yang, G.; Lei, W.; Muthukumaran, S.; Baskaran, K. Functionalized MoS2 nanosheets enabled nanofiltration membrane with enhanced permeance and fouling resistance. Environ. Technol. Innov. 2022, 27, 102719. [Google Scholar] [CrossRef]
  24. Li, Y.; Yang, S.; Zhang, K.; Van der Bruggen, B. Thin film nanocomposite reverse osmosis membrane modified by two dimensional laminar MoS2 with improved desalination performance and fouling-resistant characteristics. Desalination 2019, 454, 48–58. [Google Scholar] [CrossRef]
  25. Feng, X.; Wang, X.; Xing, W.; Zhou, K.; Song, L.; Hu, Y. Liquid-exfoliated MoS2 by chitosan and enhanced mechanical and thermal properties of chitosan/MoS2 composites. Compos. Sci. Technol. 2014, 93, 76–82. [Google Scholar] [CrossRef]
  26. Feng, X.; Xing, W.; Yang, H.; Yuan, B.; Song, L.; Hu, Y.; Liew, K.M. High-Performance Poly(ethylene oxide)/Molybdenum Disulfide Nanocomposite Films: Reinforcement of Properties Based on the Gradient Interface Effect. ACS Appl. Mater. Interfaces 2015, 7, 13164–13173. [Google Scholar] [CrossRef]
  27. Dai, R.; Han, H.; Wang, T.; Li, X.; Wang, Z. Enhanced removal of hydrophobic endocrine disrupting compounds from wastewater by nanofiltration membranes intercalated with hydrophilic MoS2 nanosheets: Role of surface properties and internal nanochannels. J. Membr. Sci. 2021, 628, 119267. [Google Scholar] [CrossRef]
  28. Wang, Z.; Tu, Q.; Zheng, S.; Urban, J.J.; Li, S.; Mi, B. Understanding the Aqueous Stability and Filtration Capability of MoS2 Membranes. Nano Lett. 2017, 17, 7289–7298. [Google Scholar] [CrossRef]
  29. Wang, Z.; Mi, B. Environmental Applications of 2D Molybdenum Disulfide (MoS2) Nanosheets. Environ. Sci. Technol. 2017, 51, 8229–8244. [Google Scholar] [CrossRef]
  30. Guo, B.Y.; Jiang, S.D.; Tang, M.J.; Li, K.; Sun, S.; Chen, P.Y.; Zhang, S. MoS2 Membranes for Organic Solvent Nanofiltration: Stability and Structural Control. J. Phys. Chem. Lett. 2019, 10, 4609–4617. [Google Scholar] [CrossRef]
  31. Jiang, J.W.; Qi, Z.; Park, H.S.; Rabczuk, T. Elastic bending modulus of single-layer molybdenum disulfide (MoS2): Finite thickness effect. Nanotechnology 2013, 24, 435705. [Google Scholar] [CrossRef]
  32. Deng, M.; Kwac, K.; Li, M.; Jung, Y.; Park, H.G. Stability, Molecular Sieving, and Ion Diffusion Selectivity of a Lamellar Membrane from Two-Dimensional Molybdenum Disulfide. Nano Lett 2017, 17, 2342–2348. [Google Scholar] [CrossRef] [PubMed]
  33. Ma, M.-Q.; Zhang, C.; Zhu, C.-Y.; Huang, S.; Yang, J.; Xu, Z.-K. Nanocomposite membranes embedded with functionalized MoS2 nanosheets for enhanced interfacial compatibility and nanofiltration performance. J. Membr. Sci. 2019, 591, 117316. [Google Scholar] [CrossRef]
  34. Zhang, H.; Gong, X.-Y.; Li, W.-X.; Ma, X.-H.; Tang, C.Y.; Xu, Z.-L. Thin-film nanocomposite membranes containing tannic acid-Fe3+ modified MoS2 nanosheets with enhanced nanofiltration performance. J. Membr. Sci. 2020, 616, 118605. [Google Scholar] [CrossRef]
  35. Liang, X.; Wang, P.; Wang, J.; Zhang, Y.; Wu, W.; Liu, J.; Van der Bruggen, B. Zwitterionic functionalized MoS2 nanosheets for a novel composite membrane with effective salt/dye separation performance. J. Membr. Sci. 2019, 573, 270–279. [Google Scholar] [CrossRef]
  36. Saraswathi, M.S.S.A.; Rana, D.; Vijayakumar, P.; Alwarappan, S.; Nagendran, A. Tailored PVDF nanocomposite membranes using exfoliated MoS2 nanosheets for improved permeation and antifouling performance. New J. Chem. 2017, 41, 14315–14324. [Google Scholar] [CrossRef]
  37. Tian, H.; Wu, X.; Zhang, K. Mixed matrix polyethersulfone tight ultrafiltration (TUF) membrane with improved dye removal by physical blending of 2D MoS2. Desalination Water Treat. 2020, 204, 93–106. [Google Scholar] [CrossRef]
  38. Dolar, D.; Kosutic, K.; Strmecky, T. Hybrid processes for treatment of landfill leachate: Coagulation/UF/NF-RO and adsorption/UF/NF-RO. Sep. Purif. Technol. 2016, 168, 39–46. [Google Scholar] [CrossRef]
  39. Wang, K.; Lin, X.; Jiang, G.; Liu, J.Z.; Jiang, L.; Doherty, C.M.; Hill, A.J.; Xu, T.; Wang, H. Slow hydrophobic hydration induced polymer ultrafiltration membranes with high water flux. J. Membr. Sci. 2014, 471, 27–34. [Google Scholar] [CrossRef]
  40. Wu, X.; Xie, Z.; Wang, H.; Zhao, C.; Ng, D.; Zhang, K. Improved filtration performance and antifouling properties of polyethersulfone ultrafiltration membranes by blending with carboxylic acid functionalized polysulfone. RSC Adv. 2018, 8, 7774–7784. [Google Scholar] [CrossRef] [Green Version]
  41. Salimi, E.; Ghaee, A.; Ismail, A.F. Improving Blood Compatibility of Polyethersulfone Hollow Fiber Membranes via Blending with Sulfonated Polyether Ether Ketone. Macromol. Mater. Eng. 2016, 301, 1084–1095. [Google Scholar] [CrossRef]
  42. Arefi-Oskoui, S.; Khataee, A.; Ucun, O.K.; Kobya, M.; Hanci, T.O.; Arslan-Alaton, I. Toxicity evaluation of bulk and nanosheet MoS2 catalysts using battery bioassays. Chemosphere 2021, 268, 128822. [Google Scholar] [CrossRef] [PubMed]
  43. Rajendhran, N.; Palanisamy, S.; Periyasamy, P.; Venkatachalam, R. Enhancing of the tribological characteristics of the lubricant oils using Ni-promoted MoS2 nanosheets as nano-additives. Tribol. Int. 2018, 118, 314–328. [Google Scholar] [CrossRef]
  44. Jia, F.; Sun, K.; Yang, B.; Zhang, X.; Wang, Q.; Song, S. Defect-rich molybdenum disulfide as electrode for enhanced capacitive deionization from water. Desalination 2018, 446, 21–30. [Google Scholar] [CrossRef]
  45. Yein, W.T.; Wang, Q.; Liu, Y.; Li, Y.; Jian, J.; Wu, X. Piezo-potential induced molecular oxygen activation of defect-rich MoS2 ultrathin nanosheets for organic dye degradation in dark. J. Environ. Chem. Eng. 2020, 8, 103626. [Google Scholar] [CrossRef]
  46. Huang, J.; Zhang, K.; Wang, K.; Xie, Z.; Ladewig, B.; Wang, H. Fabrication of polyethersulfone-mesoporous silica nanocomposite ultrafiltration membranes with antifouling properties. J. Membr. Sci. 2012, 423, 362–370. [Google Scholar] [CrossRef]
  47. Zhu, L.-J.; Zhu, L.-P.; Jiang, J.-H.; Yi, Z.; Zhao, Y.-F.; Zhu, B.-K.; Xu, Y.-Y. Hydrophilic and anti-fouling polyethersulfone ultrafiltration membranes with poly(2-hydroxyethyl methacrylate) grafted silica nanoparticles as additive. J. Membr. Sci. 2014, 451, 157–168. [Google Scholar] [CrossRef]
  48. Huang, H.; Yu, J.; Guo, H.; Shen, Y.; Yang, F.; Wang, H.; Liu, R.; Liu, Y. Improved antifouling performance of ultrafiltration membrane via preparing novel zwitterionic polyimide. Appl. Surf. Sci. 2018, 427, 38–47. [Google Scholar] [CrossRef]
  49. Jamshidi Gohari, R.; Halakoo, E.; Nazri, N.A.M.; Lau, W.J.; Matsuura, T.; Ismail, A.F. Improving performance and antifouling capability of PES UF membranes via blending with highly hydrophilic hydrous manganese dioxide nanoparticles. Desalination 2014, 335, 87–95. [Google Scholar] [CrossRef]
  50. Zheng, J.; Zhang, H.; Dong, S.; Liu, Y.; Nai, C.T.; Shin, H.S.; Jeong, H.Y.; Liu, B.; Loh, K.P. High yield exfoliation of two-dimensional chalcogenides using sodium naphthalenide. Nat. Commun. 2014, 5, 2995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Saraswathi, M.; Rana, D.; Nagendran, A.; Alwarappan, S. Custom-made PEI/exfoliated-MoS2 nanocomposite ultrafiltration membranes for separation of bovine serum albumin and humic acid. Mater. Sci. Eng. C 2018, 83, 108–114. [Google Scholar] [CrossRef]
  52. Cheng, Z.; Xiao, Y.; Wu, W.; Zhang, X.; Fu, Q.; Zhao, Y.; Qu, L. All-pH-Tolerant In-Plane Heterostructures for Efficient Hydrogen Evolution Reaction. ACS Nano 2021, 15, 11417–11427. [Google Scholar] [CrossRef] [PubMed]
  53. Song, Y.; Jiang, Z.; Gao, B.; Wang, H.; Wang, M.; He, Z.; Cao, X.; Pan, F. Embedding hydrophobic MoS 2 nanosheets within hydrophilic sodium alginate membrane for enhanced ethanol dehydration. Chem. Eng. Sci. 2018, 185, 231–242. [Google Scholar] [CrossRef]
  54. Vatanpour, V.; Madaeni, S.S.; Rajabi, L.; Zinadini, S.; Derakhshan, A.A. Boehmite nanoparticles as a new nanofiller for preparation of antifouling mixed matrix membranes. J. Membr. Sci. 2012, 401, 132–143. [Google Scholar] [CrossRef]
  55. Garcia-Ivars, J.; Iborra-Clar, M.-I.; Alcaina-Miranda, M.-I.; Van der Bruggen, B. Comparison between hydrophilic and hydrophobic metal nanoparticles on the phase separation phenomena during formation of asymmetric polyethersulphone membranes. J. Membr. Sci. 2015, 493, 709–722. [Google Scholar] [CrossRef] [Green Version]
  56. Maruyama, T.; Katoh, S.; Nakajima, M.; Nabetani, H.; Abbott, T.P.; Shono, A.; Satoh, K. FT-IR analysis of BSA fouled on ultrafiltration and microfiltration membranes. J. Membr. Sci. 2001, 192, 201–207. [Google Scholar] [CrossRef]
  57. Jermann, D.; Pronk, W.; Meylan, S.; Boller, M. Interplay of different NOM fouling mechanisms during ultrafiltration for drinking water production. Water Res. 2007, 41, 1713–1722. [Google Scholar] [CrossRef] [PubMed]
  58. Chai, P.V.; Mahmoudi, E.; Teow, Y.H.; Mohammad, A.W. Preparation of novel polysulfone-Fe3O4/GO mixed-matrix membrane for humic acid rejection. J. Water Process Eng. 2017, 15, 83–88. [Google Scholar] [CrossRef]
  59. Liu, X.; Yuan, H.; Wang, C.; Zhang, S.; Zhang, L.; Liu, X.; Liu, F.; Zhu, X.; Rohani, S.; Ching, C.; et al. A novel PVDF/PFSA-g-GO ultrafiltration membrane with enhanced permeation and antifouling performances. Sep. Purif. Technol. 2020, 233, 116038. [Google Scholar] [CrossRef]
  60. Chu, K.H.; Huang, Y.; Yu, M.; Heo, J.; Flora, J.R.V.; Jang, A.; Jang, M.; Jung, C.; Park, C.M.; Kim, D.-H.; et al. Evaluation of graphene oxide-coated ultrafiltration membranes for humic acid removal at different pH and conductivity conditions. Sep. Purif. Technol. 2017, 181, 139–147. [Google Scholar] [CrossRef]
  61. Abdikheibari, S.; Lei, W.; Dumée, L.F.; Milne, N.; Baskaran, K. Thin film nanocomposite nanofiltration membranes from amine functionalized-boron nitride/polypiperazine amide with enhanced flux and fouling resistance. J. Mater. Chem. A 2018, 6, 12066–12081. [Google Scholar] [CrossRef]
  62. Song, H.; Shao, J.; He, Y.; Hou, J.; Chao, W. Natural organic matter removal and flux decline with charged ultrafiltration and nanofiltration membranes. J. Membr. Sci. 2011, 376, 179–187. [Google Scholar] [CrossRef]
Figure 1. Synthesis and filtration process diagram of PES/MoS2 membrane.
Figure 1. Synthesis and filtration process diagram of PES/MoS2 membrane.
Membranes 12 01071 g001
Figure 2. SEM images of original commercial MoS2 (a,c) and MoS2 after 4 h of sonication (b,d), TEM images of original commercial MoS2 (e), and MoS2 after 4 h of sonication (f).
Figure 2. SEM images of original commercial MoS2 (a,c) and MoS2 after 4 h of sonication (b,d), TEM images of original commercial MoS2 (e), and MoS2 after 4 h of sonication (f).
Membranes 12 01071 g002
Figure 3. X-ray diffraction (XRD) characteristic (a) and zeta potential (b) of MoS2.
Figure 3. X-ray diffraction (XRD) characteristic (a) and zeta potential (b) of MoS2.
Membranes 12 01071 g003
Figure 4. SEM surface and cross-sectional images of PES and PES/MoS2 membranes.
Figure 4. SEM surface and cross-sectional images of PES and PES/MoS2 membranes.
Membranes 12 01071 g004
Figure 5. Cross-sectional TEM images of the M0, M2 and M4 membranes.
Figure 5. Cross-sectional TEM images of the M0, M2 and M4 membranes.
Membranes 12 01071 g005
Figure 6. XRD patterns of PES membrane and PES/MoS2 membranes.
Figure 6. XRD patterns of PES membrane and PES/MoS2 membranes.
Membranes 12 01071 g006
Figure 7. Water contact angle values (a) and the stress–strain behaviors (b) of PES and PES/MoS2 membranes.
Figure 7. Water contact angle values (a) and the stress–strain behaviors (b) of PES and PES/MoS2 membranes.
Membranes 12 01071 g007
Figure 8. Molecular weight cut-off of PES/MoS2 membranes.
Figure 8. Molecular weight cut-off of PES/MoS2 membranes.
Membranes 12 01071 g008
Figure 9. BSA rejection of TUF membranes with different concentrations of MoS2.
Figure 9. BSA rejection of TUF membranes with different concentrations of MoS2.
Membranes 12 01071 g009
Figure 10. HA rejection of prepared membranes with different concentrations of MoS2.
Figure 10. HA rejection of prepared membranes with different concentrations of MoS2.
Membranes 12 01071 g010
Table 1. The composition of PES and PES/MoS2 casting solutions.
Table 1. The composition of PES and PES/MoS2 casting solutions.
MembraneDMAc (wt%)MoS2 (wt%)Viscosity (Pa.s)
M061.000.0053.5 ± 0.1
M160.750.2564.1 ± 1.0
M260.500.5081.5 ± 0.6
M360.001.0090.4 ± 0.6
M459.501.50115.5 ± 2.6
Table 2. Effect of different concentrations of 2D MoS2 on membrane porosity, mean pore radius, and water flux.
Table 2. Effect of different concentrations of 2D MoS2 on membrane porosity, mean pore radius, and water flux.
MembranePorosity(%)Mean Pore Radius (nm)Pure Water Flux (Lm−2h−1)
M052.44 ± 0.151.5172.37 ± 0.14
M151.98 ± 0.861.6691.86 ± 5.79
M253.38 ± 0.311.7198.48 ± 3.82
M353.50 ± 0.621.6491.00 ± 6.75
M452.23 ± 0.981.6080.18 ± 6.01
Table 3. Comparison of the HA separation performance of the PES/MoS2 membrane with some other membranes.
Table 3. Comparison of the HA separation performance of the PES/MoS2 membrane with some other membranes.
MembraneRejection (%) UV254Permeation Flux (Lm−2h−1 bar−1)Pure Water Flux (Lm−2h−1 bar−1)Membrane TypeRef
PSf/GO-Fe3O484 ± 2156.99-UF[58]
PSf89 ± 251.78-UF[58]
PVDF/PFSA-g-GO79.6 587.4UF[59]
PES/GO85.3–93.9-~36–108UF[60]
PPA-BN-497.91-14.24NF[61]
PES-PPA-OH-MoS299.20-14.023NF[23]
NF27099.4-16NF[62]
PES/MoS2 (M2 membrane)99.6061.2498.48UFThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tian, H.; Wu, X.; Zhang, K. Tailoring Morphology and Properties of Tight Utrafiltration Membranes by Two-Dimensional Molybdenum Disulfide for Performance Improvement. Membranes 2022, 12, 1071. https://doi.org/10.3390/membranes12111071

AMA Style

Tian H, Wu X, Zhang K. Tailoring Morphology and Properties of Tight Utrafiltration Membranes by Two-Dimensional Molybdenum Disulfide for Performance Improvement. Membranes. 2022; 12(11):1071. https://doi.org/10.3390/membranes12111071

Chicago/Turabian Style

Tian, Huali, Xing Wu, and Kaisong Zhang. 2022. "Tailoring Morphology and Properties of Tight Utrafiltration Membranes by Two-Dimensional Molybdenum Disulfide for Performance Improvement" Membranes 12, no. 11: 1071. https://doi.org/10.3390/membranes12111071

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