Next Article in Journal
cRGD-Conjugated GdIO Nanoclusters for the Theranostics of Pancreatic Cancer through the Combination of T1–T2 Dual-Modal MRI and DTX Delivery
Previous Article in Journal
Tolyporphins–Exotic Tetrapyrrole Pigments in a Cyanobacterium—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Water Treatment Performance of Porous Polysulfone Hollow Fiber Membranes through Atomic Layer Deposition

1
Institut Européen des Membranes—IEM, UMR-5635, University of Montpellier, ENSCM, CNRS, Place Eugene Bataillon, 34095 Montpellier, France
2
Applied Mathematics and Bioinformatics (CAMB), Gulf University for Science and Technology—GUST, Kuwait City 32093, Kuwait
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(16), 6133; https://doi.org/10.3390/molecules28166133
Submission received: 12 July 2023 / Revised: 11 August 2023 / Accepted: 15 August 2023 / Published: 18 August 2023
(This article belongs to the Section Materials Chemistry)

Abstract

:
Polysulfone (PSF) is one of the most used polymers for water treatment membranes, but its intrinsic hydrophobicity can be detrimental to the membranes’ performances. By modifying a membrane’s surface, it is possible to adapt its physicochemical properties and thus tune the membrane’s hydrophilicity or porosity, which can achieve improved permeability and antifouling efficiency. Atomic layer deposition (ALD) stands as a distinctive technology offering exceedingly even and uniform layers of coatings, like oxides that cover the surfaces of objects with three-dimensional (3D) shapes, porous structures, and particles. In the context of this study, the focus was on titanium dioxide (TiO2), zinc oxide (ZnO), and alumina (Al2O3), which were deposited on polysulfone hollow fiber (HF) membranes via ALD using TiCl4, diethyl zinc (DEZ), and trimethylamine (TMA), respectively, and H2O as precursors. The morphology and mechanical properties of membranes were changed without damaging their performances. The deposition was confirmed mainly by energy-dispersive X-ray spectroscopy (EDX). All depositions offered great performances with a maintained permeability and BSA retention and a 20 to 40° lower water contact angle (WCA) than the raw PSF HF membrane. The deposition of TiO2 offered the best results, showing an enhancement of 50% for the water permeability and 20% for the fouling resistance of the PSF HF membranes.

1. Introduction

Among all polymers, polysulfone (PSF) is widely chosen for membrane applications and especially hollow fiber (HF) membrane fabrication due to its remarkable chemical and thermal resistance, which makes it suitable for membrane washing and ensuring long-term usage with good mechanical properties [1,2]. However, the inherent hydrophobicity of this polymer affects a membrane’s performances especially in water treatment applications because of contaminant fouling on its surface. That is why surface modification of these hydrophobic polymeric membranes appears to be an interesting way to overcome this issue and thus enhance both the membranes’ water permeability and resistance to fouling [3,4]. Coating methods offer a viable solution for surface modification involving the deposition of a thin layer of material onto the membrane surface. Various techniques [5], such as sol-gel, physical vapor deposition (PVD), chemical vapor deposition (CVD), or atomic layer deposition (ALD), can be employed for this purpose. For instance, Zhang et al. prepared and characterized a PSF-TiO2 HF ultrafiltration membrane using the sol–gel method. With only 2 wt. % TiO2, they observed a water contact angle (WCA) that was significantly reduced from 85° to 55° and a maximum of 96.3% bovine serum albumin (BSA) rejection [6]. While all of these methods show promise and deliver interesting results, ALD stands out due to its exceptional surface conformality. Unlike PVD, which only affects the surface; sol-gel, which can block pores; and CVD, which may not provide uniform coating throughout, ALD enables the deposition of highly uniform coatings over three-dimensional (3D) surfaces, deep trenches, porous materials, and particles thanks to its surface-controlled film growth [7]. ALD is a relatively recent technique developed over the past four decades that allows for the deposition of precise sub-nanometer-thickness coatings of various inorganic materials. This method has already attracted interest for water treatment applications. Yang et al. laid out a pathway to establishing an ALD-based universal functionalization platform for water treatment and detailed a numerous of water applications for ALD-modified materials [8].
Recent studies have demonstrated the applicability of ALD to flexible and temperature-sensitive polymeric membranes [9,10]. Deposition of an oxide thin layer through ALD has been shown to enhance membrane hydrophilicity, modify the pore size [11], and reduce membrane fouling, albeit at the expense of water permeability. Wang et al. deposited polyimide on the surface of nanoporous anodized alumina using ALD. They found that 50 cycles resulted in a remarkable improvement in retention, increasing it from nearly none for the neat alumina substrate to 82% while maintaining a high water permeability of 800 L/hm2bar, albeit about three times lower compared to the uncoated membrane [12]. Feng et al. ALD-deposited ZnO on fabrics of multi-walled carbon nanotubes. ZnO was grown on the CNT surface as nanoparticles initially; it then formed conformal layers wrapping the CNTs. The originally hydrophobic surface of the CNTs was progressively tuned to be highly hydrophilic with rising ALD cycles, and the WCA was reduced from 110° to almost 20° after 100 ALD cycles. They also observed a six-fold increase in BSA rejection with the highest permeability value of 430 L/hm2bar after 40 ALD cycles [13]. Jia et al. proved that ALD was an effective strategy to conveniently upgrade the filtration performances of polypropylene HF (PPHF) membranes by studying the deposition of Al2O3. In fact, they showed simultaneously enhanced water permeability and retention after deposition with moderate ALD cycle numbers. After 50 ALD cycles, they obtained a 17% increase in water permeability and a one-fold increase in BSA rejection accompanied with an improved fouling resistance for the modified PPHF [14]. Nikkola et al. presented the surface modification of a thin-film composite (TFC) polyamide (PA) reverse osmosis (RO) membrane via ALD using trimethyl aluminum (AlMe3) in order to improve its antifouling performances. The most hydrophilic surface was obtained with 10 and 50 ALD cycles at a temperature of 70 °C and 10 ALD cycles at 100 °C, respectively leading to a 16° and 27° WCA compared to the initial 53°. The antifouling properties of the membranes were also improved [15]. Li et al. focused on the hydrophilic modification of polyvinylidene fluoride (PVDF) membranes via ZnO ALD using nitrogen dioxide and diethylzinc functionalization. In this way, they attempted to counter the granular deposits that can sometimes sacrifice membrane flux even if they enhance hydrophilicity. They observed that after 200 ALD cycles, the WCA decreased from 140° of the original membrane to 67° and 20° for the direct modified membrane and the NO2/DEZ activated membrane, respectively. In addition, compared with direct modified membranes, a maximum water flux of 5261.4 L/hm2bar and a superior separation performance with 97% BSA retention were achieved for activated membranes at only 100 ALD cycles [16]. Wang et al. studied TiO2 deposition on PVDF membranes. They explained that the deposition of TiO2 enhanced the hydrophilicity and fouling resistance of the modified membranes, which was more evident at higher ALD cycle numbers. They simultaneously achieved an almost three-fold improved water flux and retention at 120 ALD cycles. This was as a result of the competing effect of the increased hydrophilicity from 67° to 30° and reduced pore sizes. They added that the water flux could be sensitively tuned by changing the exposure time of the precursors [17].
For the first time, this paper focused on the deposition of three different oxides on PSF HF membranes and their interest for water treatment applications. In a previous work, the impact of 10 to 100 cycles of TiO2 ALD on PSF HF membranes was investigated. The results were encouraging, as we obtained increases of 50% in water permeability and 20% in fouling resistance for a thickness deposition of only 2 nm [18]. In this study, the focus was on comparing the deposition of Al2O3, ZnO, and TiO2 using ALD on PSF HF membranes. The effect of ALD treatment on membrane morphology was investigated using scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy analysis (EDX), and water contact angle (WCA) measurements to assess hydrophilicity. The performance of the HF membranes for water-purification applications was characterized in terms of water permeability, antifouling properties using bovine serum albumin (BSA) protein, and mechanical properties. Prior to ALD treatment, the polymeric membranes underwent a conditioning step to ensure their ability to withstand vacuum and heat treatment. Based on our previous study [18] and the best results obtained for a 2 nm TiO2 deposition, all membranes were placed in the deposition chamber with glycerin conditioning while aiming for a targeted thickness of 2 nm.

2. Results and Discussion

2.1. Structural, Physical, and Chemical Characterizations

The formation of oxide layers on the PSF HF membranes was investigated via ATR-FTIR spectroscopy. In Figure S1, the characteristic peaks of PSF are visible in the raw and modified membranes. Signals observed around 1100, 1490, 1510, and 1590 cm−1 corresponded to the vibrations of PSF aromatic rings (C=C stretching). Bands at 1150 cm−1 and 1240 cm−1 were attributed to the –O–S–O– symmetric and aromatic ether bond (–C–O–C–), respectively. The peak at 1295 cm−1 was related to SO2 group stretching vibration [19]. Additionally, the absorbance peak around 1680 cm−1 could be attributed either to the C=O stretching vibration of NMP or PVP that might remain on the membrane surface. For modified membranes, any new peaks could be observed in the investigated frequency range. Thus, we can assume that the amount of oxides of around 2 nm thickness was too low to be detected via this technique as a consequence of a low number of ALD cycles. We expected to observe the metal oxide interaction peak with the PSF HF membranes. For example, for Al, we could have seen the Al-O peak that would appeared around 700 cm−1 [14]. In the case of Zn, two peaks could be observed between 500 and 630 cm−1 corresponding to the Zn-O bond [20]. Jia et al. studied the deposition of Al2O3 on porous polypropylene HF via the ALD technique, and they only noticed a new absorption peak after more than 100 deposition cycles [14].
SEM analyses performed on the top surface of a raw PSF HF membrane and modified PSF HF membranes are presented in Figure 1. No noticeable change was observed on the surface of oxide-deposited membranes. In the case of the ZnO- and Al2O3-modified membranes, some aggregates were observable on the surface. However, the number of cycles used in the deposition process was relatively low, making it challenging to clearly determine the oxide thickness via SEM. For comprehensive visualization, primary cross-section and top-surface SEM images of all membranes are provided in Supplementary Figure S2.
In order to confirm the oxide deposition and attempt to quantify the amount deposited on the membranes, EDX of all raw and modified PSF HF membranes was performed. The measured atomic percentages of C, O, and Ti are presented in Table 1. The detected atomic percentages of Ti were very low due to the small layer thicknesses of the oxides. However, for the ZnO- and Al2O3-modified membranes, the atomic percentages of Zn and Al respectively attained 1.7 and 3.2% and displayed a much higher O percentage (34.3 and 47.8%, respectively) compared to the raw PSF HF membrane with 9.9%. This confirmed the deposition of ZnO and Al2O3 on the entire membrane surface. There are two plausible explanations for the significant difference in titanium content compared to zinc and aluminum in the sample. Firstly, the EDX method used for analysis may exhibit varying detection efficiency for different elements. Moreover, the detection may be affected when a peak appears alone or overlaps with other peaks. In the case of titanium, it seemed to appear at the same energy (keV) as oxygen, which could contribute to its reduced detection. Secondly, the variation in precursor materials and their interactions with the substrate plays a crucial role. Factors like steric hindrance and diffusion can differ depending on the specific precursor used, leading to variable diffusion and deposition within the pores of the membranes as well as within the polymeric chain of the membranes. Consequently, certain oxides, such as titanium in our case, may not be detected as effectively as the others. In addition, EDX surface and cross-sectioned mapping were also investigated; these are presented in Figure 2. All modified membranes demonstrated a uniform deposition of the oxide on the surface, but only TiO2 and ZnO also had a great diffusion of the oxide onto the membrane network as suggested by the cross-sectioned images. These results could be optimized by modifying the deposition conditions of Al2O3.
Due to the considerably low Ti atomic percentages observed in the EDX analysis, XPS analyses were carried out in a previous study [18]. For the 20-cycle samples, the results were once again below the detection limit, but when using 50 cycles, the presence of Ti was clearly evident in the XPS analysis. The XPS survey spectra of both the raw and TiO2 50-cycle modified PSF HF membranes are presented in Figure S3. In both cases, the survey spectra showed two peaks at 284.85 and 532.05 eV, which were assigned to C and O atoms, respectively. Additionally, a new peak appeared at 458.36 eV in the ALD-modified PSF HF membrane, corresponding to Ti atoms. The C 1s spectra for both membranes were similar, indicating that the PSF HF membrane remained undamaged during deposition. The deconvoluted O 1s spectra showed signals for O-C and O=S bonds, and a new peak at 529.86 eV was observed for the modified membrane, indicating the presence of a Ti-O bond. This confirmed the successful deposition of TiO2, which reacted with the oxygen atoms of the PSF matrix. Moreover, in the deconvoluted Ti 2p spectra (Figure S3f) for the ALD-modified PSF HF membrane, two peaks were identified at 458.36 and 464.16 eV, corresponding to Ti 2p3/2 and Ti 2p1/2 states, respectively, indicating that Ti existed in the 4+ valence state. In this work, XPS analyses were not conducted on the ZnO- and Al2O3-modifed PSF HF membranes, but we trusted a previous study of similar deposition investigated via XPS. In fact, Viter et al. prepared ZnO/Al2O3 nanolaminate coated electrospun nanofibers and were able to confirm the presence of Zn and Al from the deposition via XPS. In addition, it was confirmed that no hydrated oxides were formed on the nanofibers’ surfaces [21].

2.2. Membrane Performances

Mechanical tests were performed on raw and modified membranes; the Young’s modulus, strength at break, and elongation at break values are presented in Table 2. The results showed that the modified membranes displayed a higher Young’s modulus of around 150 MPa on average compared to the raw membrane (132 MPa), which corresponded to a lower resistance to stress deformation. Because the ALD-modified membranes were more rigid, the strength applied to break them was lower than for the raw PSF HF membrane, decreasing from 2.10 ± 0.02 to 1.51 N on average. Also, the elongation at break was increased for the TiO2-modified membranes, whereas it remained stable or decreased for the other ones. Tensile strength curves for all membranes are presented in Figure S5. Nevertheless, TiO2 might induce a nanocomposite impact within the polymer chains, thereby enhancing the overall robustness of the network structure and consequently bolstering its mechanical resilience [22]. This observation aligns with the findings of Diez-Pascual et al., who investigated the impact of introducing TiO2 nanoparticles into polyphenylsulfone. They documented a progressive increase in the Young’s modulus as the nanoparticle loading was raised, reaching up to a 62% enhancement at the highest concentration evaluated. This enhancement was attributed to the influence of the inflexible TiO2 nanoparticles, which established interactions (such as H-bonding) with the membrane matrix. These interactions led to a reduction in polymer chain mobility and facilitated stronger adhesion between the two phases [23]. Figure 3 provides the permeability data, which indicated a notable enhancement in the water permeability for the TiO2-modified membranes, while the permeability values remained consistent for the membranes modified with ZnO and Al2O3. According to the mechanical properties and permeability and compared to other oxides, TiO2 seemed to be the best candidate for improving the PSF HF membranes’ properties. Interestingly, Al2O3 deposition showed differences in terms of mechanical properties compared to the other modified membranes. One hypothesis could be that Al2O3 spread between polymer chains and not the pores, causing polymer chain infiltration and thus alteration of the mechanical properties; this would be in good agreement with EDX analysis above.
To complete the experiment, the WCAs were measured for the raw and modified membranes, and the results are presented along with the water permeability values in Figure 3. Initially, the raw membrane had a WCA of around 90°, and this value was gradually reduced down to 70° with TiO2 deposition. These findings aligned well with the existing literature. For instance, Zhang et al. fabricated a hybrid ultrafiltration flat membrane by incorporating TiO2-g-HEMA and examined its resistance to fouling. They achieved the most substantial enhancement in hydrophilicity marked by a water contact angle (WCA) of 72° at a concentration of 2%, as the study reported [24]. The WCA was also reduced to 70° and even to 45° with ZnO and Al2O3 oxides, respectively, accompanied by a stable water permeability compared to the raw membrane. Li et al. studied the hydrophilic modification of polyvinylidene fluoride membranes via ZnO atomic layer deposition and observed a stable water flux and a WCA reduction only above 50 cycles [16]. Lin et al. prepared Al2O3/PES composite hollow fiber UF membranes via facile in situ vapor-induced hydrolyzation and observed a WCA reduced from 80° to 40° but with almost twice the reduced water permeability [25]. Once again, TiO2 seemed to be the best oxide candidate for improving the PSF HF membranes properties by combining a lower WCA and higher water permeability. A previous study thoroughly investigated and ruled out the possibility that the superior performances of TiO2-modified PSF HF membranes were solely attributed to the thermal conditions of the ALD process [18]. In Figure S4, SEM images and permeability values are presented for a comparison between the TiO2-modified PSF HF membrane and a membrane that underwent the same vacuum and temperature process for the same duration but without the deposition. The membrane without deposition displayed a permeability of only 27 L/hm2bar, whereas the ALD-modified PSF HF membrane exhibited a significantly higher permeability of 200 L/hm2bar. These results provided strong evidence that the outstanding performances of TiO2-modified membranes cannot be solely attributed to the thermal parameters of the ALD process.
The filtration performances of the raw and modified PSF HF membranes are depicted in Figure 3 and Figure 4a. The raw membrane exhibited a pure water permeability of 135 L·h−1m−2bar−1 and a retention of BSA of around 80%. For all modified membranes, the BSA retention rate was quite higher or stable except for that with the Al2O3 deposition, which decreased to 60%. Antifouling performances of the modified membranes were investigated using BSA as a fouling agent; for this, the F R R , R t , R r ,   and R i r of the membranes were calculated and are presented in Figure 4b. The F R R was increased from 88% for the raw PSF HF membrane to 100% only for TiO2, whereas it decreased to 60% on average for all other oxides. Following this trend, once again only the TiO2-modified membrane exhibited a lower R t , R r ,   and R i r . A decreasing R r   and R i r means that BSA adsorbed reversibly was removed after water backwashing and none of it remained, which was a good result because membrane lifetimes are threatened by fouling and most of all irreversible fouling, and chemical backwashing is needed to clean them. This result means that due to TiO2 addition, the ALD-modified PSF HF membranes were recovering all the flux after BSA filtration with no irreversible fouling. These results are very interesting since the values are even better than those reported in other studies. Yang et al. [26] and Zhang et al. [24] inserted, for instance, TiO2 nanoparticles into flat PSF membranes and obtained 97 and 95% BSA rejection, respectively. In fact, BSA is a complex protein of 67kDa, the behavior of which is widely influenced by its environment. Several studies focused on this understanding and can provide some information to comprehend the adsorption mechanism of the protein onto the substrate [27,28,29]. BSA can be adsorbed either by electrostatic or hydrophobic interactions, meaning that both the WCA and zeta potential of the membranes are key parameters. The isoelectric point of a protein plays an important role in transport because it determines the surface charge of the protein, depending on the pH of the solution. When the pH value of a buffer solution is equal to the isoelectric point of the protein, the surface charge of the protein is neutral; at a pH above that point, the protein is negatively charged (and vice versa at a lower pH) [29]. So, BSA has an isoelectric point of 4.7, which means that at the experimental pH of 7.4, its surface is negatively charged. In the case of TiO2, multiple works have shown that adding it to a polymeric membrane would reinforce the negative charge of the surface, enhancing the repulsive interactions between the membrane and BSA [30,31,32]. TiO2 both improved the hydrophilicity and negatively charged surface of the PSF HF membranes with an initial zeta potential known to be around −10 mV [33], which explains the great BSA retention as well as the high fouling parameters because it was widely spread onto the matrix as proved by EDX observations. TiO2, ZnO, and Al2O3 should have the save behavior in terms of zeta potential at around pH 7.4 [34]. Huang et al. incorporated oleic acid-modified Ag@ZnO core–shell nanoparticles into thin-film composite membranes and observed enhanced antifouling properties against BSA. They showed that ZnO could increase the membrane’s negative surface charge or slowly reduce it, depending on the proportion of nanoparticles [35]. In our case, the ZnO-modified membrane had the same WCA properties as the TiO2 one, but the negative surface charge possibly was not enough to repulse the BSA and prevent it from entering the polymer matrix, causing lower fouling parameters. Concerning Al2O3, Jia et al. deposited Al2O3 on PP HF membranes and obtained a maximum of 75% BSA rejection [14]. We can suppose than despite the highly reduced WCA of the Al2O3-modified membrane, the surface charge and roughness, which could not be measured in the study, were too high to allow better antifouling properties against BSA compared to the raw PSF HF membrane. Also, we utilized fouling data to confirm the stability of the membrane. The data demonstrated that the modified membrane retained the same performance even after several hours of filtration, which validated the presence and maintenance of the oxide layer on the surface of our support.

3. Materials and Methods

3.1. Materials

This study utilized PSF hollow fiber (PSF HF) membranes supplied by Polymem (Fort Worth, TX, USA) as the substrates for atomic layer deposition (ALD). The membranes underwent chlorine washing and glycerin conditioning steps before the deposition process. The following materials were employed for the ALD process: titanium (IV) chloride (TiCl4, 99.9%, CAS: 7550-45-0), diethylzinc (DEZ, (C2H5)2Zn, >95%, CAS: 93-3030), trimethylaluminum (TMA, Al(CH3)3, >98%, CAS: 93-1360), and deionized water. Argon gas, which was purchased from Linde (Dublin, Ireland), was utilized as received. Phosphate-buffered saline (PBS) from Roth (Karlsruhe, Germany) and bovine serum albumin (BSA) solution with a molecular weight of 67 kDa and a purity of ≥96% from Sigma-Aldrich (St. Louis, MO, USA) were used to evaluate the fouling resistance of both the untreated and modified PSF HF membranes. Deionized water from Millipore Milli-Q (Burlington, MA, USA) was used in all aqueous solutions throughout the experiment.

3.2. Oxide Deposition on PSF HF Membranes

The commercial PSF membranes were received preconditioned in a mixture of water and glycerin. These membranes were directly placed in the deposition chamber for the atomic layer deposition (ALD) process. The efficacy of glycerin conditioning for ALD was previously demonstrated in another study [18]. After the deposition process, the membranes were subjected to successive rinses in water baths to remove any remaining glycerin and to determine their functional properties. To initiate the deposition process, the preconditioned PSF membranes were introduced into the home-built ALD reactor and allowed to acclimate for 15 min under vacuum conditions (~10−2 mbar). The reactor was preheated to 100 °C. For all deposition runs, we aimed for a 2 nm thickness while guided by the successful outcomes of a prior study involving TiO2 deposition on PSF HF membranes [18]. The number of cycles for each oxide was determined theoretically while taking into account the growth rates observed in a previous study [36] to achieve the targeted 2 nm deposition thickness. Specifically, TiO2 exhibited a growth rate of 0.1 nm per cycle [18], while both ZnO and Al2O3 demonstrated a growth rate of 0.2 nm per cycle [36], resulting in the application of 20 and 10 cycles, respectively. The deposition processes utilized specific precursors, and all parameters (including pulse, exposure, and purge times) are detailed in Table 3. The precursors were contained in stainless steel cylinders, and the lines leading to the reactor were heated to 80 °C to prevent any condensation.

3.3. Characterizations

The morphology of the membranes, both in cross-section and on the surface, was examined using a Hitachi S4800 scanning electron microscopy system (SEM). Prior to imaging, the membranes underwent nitrogen cold-cutting. Additionally, energy-dispersive X-ray spectroscopy (EDX) analysis was conducted using a Zeiss EVO HD15 microscope equipped with an Oxford X-MaxN EDX detector. To investigate the changes in chemical structure and confirm the successful deposition of oxides on the PSF HF membranes, Fourier transform infrared (FTIR) spectra were acquired. An FTIR spectrometer (NEXUS instrument) equipped with an attenuated total reflection (ATR) accessory was used to obtain spectra in the frequency range of 600–4000 cm−1. The ATR-FTIR spectra were recorded at a resolution of 4 cm−1, and an average of 64 scans were taken. Moreover, X-ray photoelectron spectroscopy (XPS) was performed using a Kratos AX-IS NOVA spectrometer (Kratos Analytical Ltd., Manchester, UK) equipped with a mono-chromatized Al Kα X-ray source (1486.6 eV) operating at a power of 150 W (10 mA, 15 kV). Survey and high-resolution spectra were obtained using 160 and 20 eV pass energies, respectively. The binding energy (BE) measurements were corrected based on the C1s energy at 284.4 eV. The experiments were conducted on PSF HF segments securely mounted on a support with a primary focus on analyzing the external surface of the membrane. The mechanical properties of both the raw and modified PSF HF membranes were evaluated using a dynamic mechanical analysis device (Z005, 5 kN Proline, Zwick Roell, Ulm, Germany). A 10N sensor and a tensile testing speed of 0.4 mm·min−1 were applied to determine the Young’s modulus of the membranes. Water contact angles (WCAs) were measured to assess the hydrophilicity of the membranes. An ILMS GBX tensiometer/goniometer equipped with an optic telecentric F55 double focal monochrome and GBX software (County Dublin D24, Ireland, https://gbxonline.com/ accessed on 10 August 2023) was used for the measurements. Approximately 0.5 μL of ultrapure water was deposited on the membranes using a needle for each sample.

3.4. Filtration Experiments

In the experimental setup, a U-tube-shaped hollow fiber bundle was created using a homemade polyvinyl chloride (PVC) housing module. The bundle consisted of 10 wet fibers, each measuring 40 cm in length, resulting in an effective membrane area of 90 ± 5 cm2. The module was sealed using epoxy resin. To measure the pure water flux ( Jw ) for each membrane, pure water was circulated through the membrane system under applied pressures ranging from 0.5 to 1.5 bar. The flow passing through the membrane was measured for 60 s. Jw was calculated using the following formula:
Jw = Q A   L · h 1 · m 2
where Q (L·h−1) represents the amount of water that passed through the membrane and A (m2) is the membrane geometric area. The permeability was determined by analyzing the linear variation of Jw versus the applied pressure.
For the rejection experiments, BSA was used as a solute. An I-shaped hollow fiber bundle consisting of 10 wet fibers measuring 27 cm in length was prepared using a PVC module and epoxy resin for sealing. The effective calculated membrane geometric area was 63.28 cm2. The BSA contents in the feed (Cf) and permeate (Cp) solutions were determined using a spectrophotometer. The solute rejection (R) was calculated using the following formula:
R = 1 C p C f × 100   %
The membrane was first conditioned with distilled water at a pressure of 1.5 bar for 30 min. External–internal solute filtration was then carried out for 20 min at 0.5 bar. To estimate the antifouling properties of the PSF HF membranes, static protein adsorption was performed. A BSA solution was prepared by dissolving 2 g of BSA in 2 L of PBS solution at pH 7.4. The membrane was conditioned by applying a pressure of 1.5 bar with distilled water for 30 min. Initial external–internal water filtration was conducted for 20 min at 0.5 bar, and the steady pure water flux (   J w 0 ) was recorded.
The tank was then filled with the BSA solution, and filtration was carried out at 0.5 bar for 2 h, during which the foulant flux (   J w f ) was recorded. The BSA was subsequently removed via mechanical agitation and backwashing with water at a pressure of 1.5 bar internally–externally for 5 min. Finally, a last external–internal filtration cycle with water for 20 min at 0.5 bar was performed, and the steady flux ( J w 1 ) was recorded.
Several ratios were calculated to assess the fouling properties, including the total fouling ratio ( R t ), flux recovery ratio ( FRR ), reversible fouling ratio ( R r ), and irreversible fouling ratio ( R ir ), using the following equations [37]:
R t = J w 0 J w f J w 0 × 100   %
FRR = J w 1 J w 0 × 100   %
R r = J w 1 J w f J w 0 × 100   %
R ir = J w 0 J w 1 J w 0 × 100   %
The feed foulant solution, foulant retentate, and permeate concentrations were calculated using the Beer–Lambert law with a constant value of k = 0.6439 L/g and an ultraviolet (UV) spectrometer for the absorbance at 280 nm (Uviline Connect Series 940). The rejection of the foulant (R) could be determined using Equation (2), where Cp represents the concentration of the foulant in the permeate and Cf represents the concentration of the foulant in the feed. The rejection was calculated as a percentage and represented the percentage of the foulant that was retained or rejected by the membrane during the filtration process.

4. Conclusions

Atomic layer deposition of TiO2, ZnO, and Al2O3 on PSF HF membranes was studied. Based on the first study of TiO2 deposition on PSF HF membranes, different oxides were tested following a glycerin conditioning to protect the pores during the process time, and a 2 nm deposition thickness was targeted. The presence of each oxide was proved via EDX analysis. SEM and analysis of the mechanical properties allowed us to ensure that the morphology and integrity of the membranes were not affected by the deposition. Hydrophilicity improvement was confirmed by a WCA reduction from 90° for the raw PSF HF membrane to 70° for the TiO2 and ZnO oxides and even 46.5° for the Al2O3-modified PSF HF membranes. All modified membranes exhibited a great water permeability and stable BSA retention rate after deposition, but only the TiO2-modified PSF HF membranes showed enhanced fouling parameters with an optimum 100% F R R and a 0% R i r . Expanding the study to include the filtration of other molecules such as humic acid or bacteria would provide a more comprehensive understanding of the performance of the modified PSF HF membranes. By considering a broader range of foulants, such a study can provide valuable insights into the versatility and potential applications of the modified PSF HF membranes in diverse water treatment methods and other relevant fields.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28166133/s1, Figure S1. Raw P and modified P-1 (TiO2), P-2 (ZnO), P-3 (Al2O3), P-4 (alucone) PSF membranes ATR-FTIR spectra. Figure S2. Cross section and top surface primary SEM images of (a) raw, (b) TiO2, (c) ZnO and (d) Al2O3 modified PSF HF membranes. Figure S3. Survey XPS spectra (a) and deconvoluted XPS spectra of raw membrane C 1s (b) and O 1s (c) and 50 TiO2 cycles PSF HF membrane C 1s (d), O 1s (e) and Ti 2p (f). Figure S4. SEM Images of (a) external surface and (b) cross section of PSF HF membrane that underwent same vacuum and temperature conditions but without TiO2 deposition and (c) external surface and (d) cross section of TiO2 ALD modified PSF HF membrane. Table (e) is presenting permeability values of both these membranes. Figure S5. Tensile strength curves of (a) raw, (b) TiO2, (c) ZnO and (d) Al2O3 modified PSF HF membranes

Author Contributions

Methodology, J.C., C.P.-B., D.C. and M.B.; Validation, J.C. and M.B.; Formal analysis, J.C.; Resources, D.C. and M.B.; Data curation, J.C.; Writing—original draft, J.C.; Writing—review & editing, C.P.-B., D.C., M.B. and P.M.; Supervision, C.P.-B., D.C. and M.B.; Funding acquisition, C.P.-B., D.C., M.B. and P.M. All authors have read and agreed to the published version of the manuscript.

Funding

This project was part of INNOMEM and received funding from the European Union’s Horizon 2020 Research and Innovation Program under Grant Agreement No. 862330.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Acknowledgments

The authors would like to thank the company Polymem for providing the PSF HF membranes. The authors acknowledge D. Cot and B. Rebiere for SEM and EDX observations and the ALD team of IEM for its technical support in the deposition processes.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Nicolaisen, B. Developments in membrane technology for water treatment. Desalination 2003, 153, 355–360. [Google Scholar] [CrossRef]
  2. Kheirieh, S.; Asghari, M.; Afsari, M. Application and modification of polysulfone membranes. Rev. Chem. Eng. 2018, 34, 657–693. [Google Scholar] [CrossRef]
  3. Ganj, M.; Asadollahi, M.; Mousavi, S.A.; Bastani, D.; Aghaeifard, F. Surface modification of polysulfone ultrafiltration membranes by free radical graft polymerization of acrylic acid using response surface methodology. J. Polym. Res. 2019, 26, 231. [Google Scholar] [CrossRef]
  4. Mu, Y.; Feng, H.; Wang, S.; Zhang, S.; Luan, J.; Zhang, M.; Yu, Z.; Wang, G. Combined strategy of blending and surface modification as an effective route to prepare antifouling ultrafiltration membranes. J. Colloid Interface Sci. 2021, 589, 1–12. [Google Scholar] [CrossRef] [PubMed]
  5. Fotovvati, B.; Namdari, N.; Dehghanghadikolaei, A. On coating techniques for surface protection: A review. J. Manuf. Mater. Process. 2019, 3, 28. [Google Scholar] [CrossRef]
  6. Zhang, H.; Guo, Y.; Zhang, X.; Hu, X.; Wang, C.; Yang, Y. Preparation and characterization of PSF-TiO2 hybrid hollow fiber UF membrane by sol–gel method. J. Polym. Res. 2020, 27, 376. [Google Scholar] [CrossRef]
  7. Pakkala, A.; Putkonen, M. Atomic Layer Deposition. In Handbook of Deposition Technologies for Films and Coatings, 3rd ed.; Elsevier Ltd.: Amsterdam, The Netherlands, 2010; ISBN 9780815520313. [Google Scholar]
  8. Yang, X.; Martinson, A.B.F.; Elam, J.W.; Shao, L.; Darling, S.B. Water treatment based on atomically engineered materials: Atomic layer deposition and beyond. Matter 2021, 4, 3515–3548. [Google Scholar] [CrossRef]
  9. Lee, J.; Kim, I.S. Environmental Science membrane surface engineering methods for water treatment: A short review. Environ. Sci. Water Res. Technol. 2020, 6, 1765–1785. [Google Scholar] [CrossRef]
  10. Weber, M.; Julbe, A.; Ayral, A.; Miele, P.; Bechelany, M. Atomic Layer Deposition for Membranes: Basics, Challenges, and Opportunities. Chem. Mater. 2018, 30, 7368–7390. [Google Scholar] [CrossRef]
  11. Li, F.; Li, L.; Liao, X.; Wang, Y. Precise pore size tuning and surface modifications of polymeric membranes using the atomic layer deposition technique. J. Memb. Sci. 2011, 385–386, 1–9. [Google Scholar] [CrossRef]
  12. Wang, H.; Wei, M.; Zhong, Z.; Wang, Y. Atomic-layer-deposition-enabled thin-film composite membranes of polyimide supported on nanoporous anodized alumina. J. Memb. Sci. 2017, 535, 56–62. [Google Scholar] [CrossRef]
  13. Feng, J.; Xiong, S.; Wang, Z.; Cui, Z.; Sun, S.P.; Wang, Y. Atomic layer deposition of metal oxides on carbon nanotube fabrics for robust, hydrophilic ultrafiltration membranes. J. Memb. Sci. 2018, 550, 246–253. [Google Scholar] [CrossRef]
  14. Jia, X.; Low, Z.; Chen, H.; Xiong, S.; Wang, Y. Atomic layer deposition of Al2O3 on porous polypropylene hollow fibers for enhanced membrane performances. Chin. J. Chem. Eng. 2018, 26, 695–700. [Google Scholar] [CrossRef]
  15. Nikkola, J.; Sievänen, J.; Raulio, M.; Wei, J.; Vuorinen, J.; Tang, C.Y. Surface modi fi cation of thin fi lm composite polyamide membrane using atomic layer deposition method. J. Memb. Sci. 2014, 450, 174–180. [Google Scholar] [CrossRef]
  16. Li, N.; Zhang, J.; Tian, Y.; Zhang, J.; Zhan, W.; Zhao, J.; Ding, Y.; Zuo, W. Hydrophilic modification of polyvinylidene fluoride membranes by ZnO atomic layer deposition using nitrogen dioxide/diethylzinc functionalization. J. Memb. Sci. 2016, 514, 241–249. [Google Scholar] [CrossRef]
  17. Wang, Q.; Wang, X.; Wang, Z.; Huang, J.; Wang, Y. PVDF membranes with simultaneously enhanced permeability and selectivity by breaking the tradeoff effect via atomic layer deposition of TiO2. J. Memb. Sci. 2013, 442, 57–64. [Google Scholar] [CrossRef]
  18. Casetta, J.; Ortiz, D.G.; Pochat-Bohatier, C.; Bechelany, M.; Miele, P. Atomic layer deposition of TiO2 on porous polysulfone hollow fibers membranes for water treatment. Sep. Purif. Technol. 2023, 312, 123377. [Google Scholar] [CrossRef]
  19. Cabasso, I.; Klein, E.; Smith, J.K.; South, G. Polysulfone Hollow Fibers. I. Spinning and Properties. 1976, 20, 2377–2394. [Google Scholar]
  20. Janaki, A.C.; Sailatha, E.; Gunasekaran, S. Synthesis, Characteristics and Antimicrobial activity of ZnO nanoparticles. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 144, 17–22. [Google Scholar] [CrossRef]
  21. Viter, R.; Iatsunskyi, I.; Fedorenko, V.; Tumenas, S.; Balevicius, Z.; Ramanavicius, A.; Balme, S.; Nowaczyk, G.; Jurga, S.; Bechelany, M. Enhancement of Electronic and Optical Properties of Enhancement of Electronic and Optical Properties of ZnO/Al2O3 Nanolaminate Coated Electrospun Nanofibers. J. Phys. Chem. 2016, 120, 5124–5132. [Google Scholar] [CrossRef]
  22. Galiano, F.; Song, X.; Marino, T.; Boerrigter, M.; Saoncella, O.; Simone, S.; Faccini, M.; Chaumette, C.; Drioli, E.; Figoli, A. Novel photocatalytic PVDF/Nano-TiO2 hollow fibers for Environmental remediation. Polymers 2018, 10, 1134. [Google Scholar] [CrossRef] [PubMed]
  23. Díez-Pascual, A.M.; Díez-Vicente, A.L. Effect of TiO2 nanoparticles on the performance of polyphenylsulfone biomaterial for orthopaedic implants. J. Mater. Chem. B 2014, 2, 7502–7514. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, G.; Lu, S.; Zhang, L.; Meng, Q.; Shen, C.; Zhang, J. Novel polysulfone hybrid ultrafiltration membrane prepared with TiO2-g-HEMA and its antifouling characteristics. J. Memb. Sci. 2013, 436, 163–173. [Google Scholar] [CrossRef]
  25. Lin, Y.; Loh, C.H.; Shi, L.; Fan, Y.; Wang, R. Preparation of high-performance Al2O3/PES composite hollow fiber UF membranes via facile in-situ vapor induced hydrolyzation. J. Memb. Sci. 2017, 539, 65–75. [Google Scholar] [CrossRef]
  26. Yang, Y.; Wang, P.; Zheng, Q. Preparation and properties of polysulfone/TiO2 composite ultrafiltration membranes. J. Polym. Sci. Part B Polym. Phys. 2006, 44, 879–887. [Google Scholar] [CrossRef]
  27. Wu, X.; Hao, P.; He, F.; Yao, Z.; Zhang, X. Applied Surface Science Molecular dynamics simulations of BSA absorptions on pure and formate-contaminated rutile (1 1 0) surface. Appl. Surf. Sci. 2020, 533, 147574. [Google Scholar] [CrossRef]
  28. Membranes, B.C. Selective Separation of Similarly Sized Proteins with Tunable Nanoporous. ACS Nano 2013, 7, 768–776. [Google Scholar]
  29. Kubiak-ossowska, K.; Tokarczyk, K.; Jachimska, B.; Mulheran, P.A. Bovine Serum Albumin Adsorption at a Silica Surface Explored by Simulation and Experiment. J. Phys. Chem. B 2017, 121, 3975–3986. [Google Scholar] [CrossRef]
  30. Abba, M.U.; Man, H.C.; Azis, S.; Idris, A.I.; Hamzah, M.H.; Yunos, K.F.; Katibi, K.K. Novel PVDF-PVP Hollow Fiber Membrane Augmented with TiO2 Nanoparticles: Preparation, Characterization and Application for Copper Removal from Leachate. Nanomaterials 2021, 11, 399. [Google Scholar] [CrossRef]
  31. Mahdi, N.; Kumar, P.; Goswami, A.; Perdicakis, B.; Shankar, K. Robust Polymer Nanocomposite Membranes Incorporating Discrete TiO2 Nanotubes for Water Treatment. Nanomaterials 2019, 9, 1186. [Google Scholar] [CrossRef]
  32. Li, T.; Xiao, Y.; Guo, D.; Shen, L.; Li, R.; Jiao, Y.; Xu, Y.; Lin, H. Journal of Colloid and Interface Science In-situ coating TiO2 surface by plant-inspired tannic acid for fabrication of thin film nanocomposite nanofiltration membranes toward enhanced separation and antibacterial performance. J. Colloid Interface Sci. 2020, 572, 114–121. [Google Scholar] [CrossRef] [PubMed]
  33. Modi, A.; Bellare, J. Efficiently improved oil/water separation using high flux and superior antifouling polysulfone hollow fiber membranes modified with functionalized carbon nanotubes/graphene oxide nanohybrid. J. Environ. Chem. Eng. 2019, 7, 102944. [Google Scholar] [CrossRef]
  34. Xia, Z.; Rozyyev, V.; Mane, A.U.; Elam, W.; Darling, S.B. Surface Zeta Potential of ALD-Grown Metal-Oxide Films. Langmuir 2021, 37, 11618–11624. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, X.; Chen, Y.; Feng, X.; Hu, X.; Zhang, Y.; Liu, L. Incorporation of oleic acid-modified Ag @ ZnO core-shell nanoparticles into thin film composite membranes for enhanced antifouling and antibacterial properties. J. Memb. Sci. 2020, 602, 117956. [Google Scholar] [CrossRef]
  36. Chaaya, A.A.; Viter, R.; Baleviciute, I.; Bechelany, M.; Ramanavicius, A.; Europeen, I.; Montpellier, U.; Bataillon, P.E. Tuning Optical Properties of Al2O3/ZnO Nanolaminates Synthesized by Atomic Layer Deposition. J. Phys. Chem. C 2014, 118, 3811–3819. [Google Scholar] [CrossRef]
  37. Alkhouzaam, A.; Qiblawey, H. Novel polysulfone ultrafiltration membranes incorporating polydopamine functionalized graphene oxide with enhanced flux and fouling resistance. J. Memb. Sci. 2021, 620, 118900. [Google Scholar] [CrossRef]
Figure 1. SEM surface images of raw and TiO2-, ZnO-, and Al2O3-modified PSF HF membranes.
Figure 1. SEM surface images of raw and TiO2-, ZnO-, and Al2O3-modified PSF HF membranes.
Molecules 28 06133 g001
Figure 2. EDX mapping of TiO2, ZnO (cross section), and Al2O3 (surface) modified PSF HF membranes.
Figure 2. EDX mapping of TiO2, ZnO (cross section), and Al2O3 (surface) modified PSF HF membranes.
Molecules 28 06133 g002
Figure 3. WCA and water permeability of raw and TiO2-, ZnO-, and Al2O3-modified membranes. For water permeability, the error refers to the standard deviation of 3 samples. In the case of the WCA, the contact angle was measured 3 times on each droplet image, and this was repeated on 5 to 10 samples.
Figure 3. WCA and water permeability of raw and TiO2-, ZnO-, and Al2O3-modified membranes. For water permeability, the error refers to the standard deviation of 3 samples. In the case of the WCA, the contact angle was measured 3 times on each droplet image, and this was repeated on 5 to 10 samples.
Molecules 28 06133 g003
Figure 4. (a) BSA retention and (b) fouling resistance parameters for raw and TiO2-, ZnO-, and Al2O3-modified membranes (the error refers to the standard deviation of 3 samples).
Figure 4. (a) BSA retention and (b) fouling resistance parameters for raw and TiO2-, ZnO-, and Al2O3-modified membranes (the error refers to the standard deviation of 3 samples).
Molecules 28 06133 g004
Table 1. EDX data showing the composition of the raw and modified membranes (the error was calculated based on a 10% EDX measurement uncertainty).
Table 1. EDX data showing the composition of the raw and modified membranes (the error was calculated based on a 10% EDX measurement uncertainty).
Atomic Percentages (%)
MembranesCOTiZnAl
Raw86.8 ± 8.69.9 ± 0.9///
TiO287.2 ± 8.79.9 ± 0.9≤0.1//
ZnO62.4 ± 6.234.3 ± 3.4/1.7 ± 0.2/
Al2O348.3 ± 4.847.8 ± 4.8//3.2 ± 0.3
Table 2. Tensile properties of raw and modified HF membranes (the error refers to the standard deviation of 3 samples).
Table 2. Tensile properties of raw and modified HF membranes (the error refers to the standard deviation of 3 samples).
OxidesYoung’s Modulus (MPa)Strength at Break (N)Elongation at Break (mm)
Raw132 ± 52.10 ± 0.0255 ± 5
TiO2143 ± 61.84 ± 0.0562 ± 7
ZnO157 ± 51.31 ± 0.1749 ± 8
Al2O3132 ± 31.12 ± 0.0538 ± 5
Table 3. Deposition protocols of modified PSF HF membranes.
Table 3. Deposition protocols of modified PSF HF membranes.
OxideTiO2ZnOAl2O3
Precursor 1TiCl4DEZTMA
Pulse time 1 (s)0.50.40.4
Precursor 2H2OH2OH2O
Pulse time 2 (s)222
Temperature (°C)100100100
Exposure time for both precursors (s)103040
Purge time for both precursors (s)604060
Argon mass flow during purge (sccm)100100100
Number of cycles201010
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

Casetta, J.; Pochat-Bohatier, C.; Cornu, D.; Bechelany, M.; Miele, P. Enhancing Water Treatment Performance of Porous Polysulfone Hollow Fiber Membranes through Atomic Layer Deposition. Molecules 2023, 28, 6133. https://doi.org/10.3390/molecules28166133

AMA Style

Casetta J, Pochat-Bohatier C, Cornu D, Bechelany M, Miele P. Enhancing Water Treatment Performance of Porous Polysulfone Hollow Fiber Membranes through Atomic Layer Deposition. Molecules. 2023; 28(16):6133. https://doi.org/10.3390/molecules28166133

Chicago/Turabian Style

Casetta, Jeanne, Céline Pochat-Bohatier, David Cornu, Mikhael Bechelany, and Philippe Miele. 2023. "Enhancing Water Treatment Performance of Porous Polysulfone Hollow Fiber Membranes through Atomic Layer Deposition" Molecules 28, no. 16: 6133. https://doi.org/10.3390/molecules28166133

Article Metrics

Back to TopTop