Next Article in Journal
Special Aspects of Nitrocellulose Molar Mass Determination by Dynamic Light Scattering
Next Article in Special Issue
Catalytic Pyrolysis of Polystyrene Waste in Hydrocarbon Medium
Previous Article in Journal
Biopolymeric Fibrous Aerogels: The Sustainable Alternative for Water Remediation
Previous Article in Special Issue
Studying the Physical and Chemical Properties of Polydimethylsiloxane Matrix Reinforced by Nanostructured TiO2 Supported on Mesoporous Silica
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement in Acid Resistance of Polyimide Membranes: A Sustainable Cross-Linking Approach via Green-Solvent-Based Fenton Reaction

1
School of Urban and Environmental Engineering, Ulsan National Institute of Science and Technology (UNIST), Ulsan 44919, Republic of Korea
2
Membrane Research Center, Korea Research Institute of Chemical Technology, Daejeon 34114, Republic of Korea
*
Author to whom correspondence should be addressed.
Both authors have contributed equally.
Polymers 2023, 15(2), 264; https://doi.org/10.3390/polym15020264
Submission received: 12 October 2022 / Revised: 22 November 2022 / Accepted: 1 January 2023 / Published: 4 January 2023
(This article belongs to the Special Issue Durability and Degradation of Polymeric Materials II)

Abstract

:
In this study, we present a facile surface modification method using green solvents for a commercial polyimide (PI) nanofiltration membrane to exhibit good acid stability. To enhance acid stability, the PI organic solvent nanofiltration membrane was modified using Fenton’s reaction, an oxidative cross-linking process, using environmentally friendly solvents: water and ethanol. The surface properties of the pristine and modified PI membranes were investigated and compared using various analytical tools. We studied the surface morphology using scanning electron microscopy, performed elemental analysis using X-ray photoelectron spectroscopy, investigated chemical bonds using attenuated total reflectance-Fourier transform infrared spectroscopy, and studied thermal stability using thermogravimetric analysis. The acid resistances of the pristine and modified membranes were confirmed through performance tests. The pristine PI nanofiltration membrane exposed to a 50 w/v% sulfuric acid for 4 h showed an increase in the normalized water flux to 205% and a decrease in the MgSO4 normalized rejection to 44%, revealing damage to the membrane. The membrane modified by the Fenton reaction exhibited a decline in flux and improved rejection, which are typical performance changes after surface modification. However, the Fenton-modified membrane exposed to 50 w/v% sulfuric acid for 4 h showed a flux increase of 7% and a rejection increase of 4%, indicating improved acid resistance. Furthermore, the Fenton post-treatment enhanced the thermal stability and organic solvent resistance of the PI membrane. This study shows that the acid resistance of PI membranes can be successfully improved by a novel and facile Fenton reaction using green solvents.

1. Introduction

Water is vital for living organisms and is a crucial resource for the sustainable development of the earth’s environment. Today, approximately 88% of developing countries face water shortages and scarcity [1]. Therefore, state-of-the-art water and wastewater treatment technologies are required to overcome this problem. Membrane separation processes are widely used in water and wastewater treatment. Among the various pressure-driven membranes, nanofiltration (NF) membranes, which have a pore diameter of approximately 0.5–2.0 nm, have recently attracted significant attention [2]. This is mainly attributed to their lower required pressure, high permeate flux, cost-effectiveness, high retention of multivalent salts, and ability to separate low molecular weight organic compounds (i.e., those with molecular weight in the range 200–1000 g mol−1)) from solutions containing monovalent ions. Recently, there has been an increasing demand for acid-resistant NF membranes, particularly for the recovery of target chemical compounds from highly acidic aqueous and organic solutions. NF membranes are used particularly widely in the pulp and paper industries for the treatment of acidic effluents [3], the mining and metal industries for the removal of heavy metals [4] and sulfate ions [5], dairy cleaning-in-place processes for the regeneration of acid streams [6], gold ore processing for the removal of gold from solutions with high sulfuric acid concentrations [4], and sewage sludge treatment for the recovery of phosphorous [7,8]. Other uses of acid-resistant NF membranes include the separation of rare-earth elements, such as lanthanum and yttrium, from strong acids, such as nitric acid (HNO3), hydrochloric acid (HCl), and sulfuric acid (H2SO4) [9]. NF membranes are also used in the semiconductor industry for the treatment of the etching wastewater containing HCl, hydrobromic acid (HBr), and hydroiodic acid (HI) [10,11]. In all the aforementioned applications, membranes must operate in low pH solutions. Although high-performance commercial NF full-aromatic polyamide membranes, which are fabricated from piperazine (PIP) or m-phenylenediamine (MPD) with trimesoyl chloride (TMC), are available, their application is limited to the pH range of 2–11 [11,12,13].
Among a wide range of polymer materials, polyimide (PI) is known for its good chemical, thermal, and mechanical stability, as well as exceptional engineering processability [14,15]. Modified PI show enhanced chemical stability, that is, excellent resistance to many polar aprotic organic solvents, because of the rigid structure of the polymer backbone [16,17]. Diamines are the most commonly used crosslinkers for modifying PI membranes. Diamines react with a carboxylic group in the form of poly(amic acid) to form a bridge between two polymer chains. Modification with diamines is performed by adding them to the polymer casting solution. Zhao et al. [18] added a cross-linking agent to a polymer solution with a concentration of less than 5 wt.% to avoid gel formation. The modified PI membrane was more flexible than the unmodified PI membrane and demonstrated higher gas permeability and selectivity. This method of adding a cross-linking agent to the casting solution can only be used with polymer solutions of very low concentration to avoid gelation of the casting solution before film synthesis. Hence, its use is limited to the production of thin and dense PI films suitable for gas separation. At the same time, there are cases in which cross-linking is performed via post-treatment of the fabricated membrane. Toh et al. [19] modified Lenzing P84 PI membranes with aliphatic diamines (ethylenediamine (EDA), propylenediamine, 1,6-hexanediamine, and 1,8-octanediamine). The modified membranes showed excellent resistance to N,N-dimethylacetamide (DMA), dimethylsulfoxide (DMSO), N-methylpyrrolidone (NMP), and N,N’-dimethylformamide (DMF); furthermore, using the filtration test with toluene, they elucidated that cross-linking reduces the flux by making the active layer denser. The membrane modified using EDA showed a high flux in DMF; moreover, its stable performance was confirmed even after 120 h. Vanherck et al. [20] modified Matrimid PI membranes with aromatic diamine p-xylenediamine to prepare stable membranes for the filtration of DMA, DMSO, NMP, tetrahydrofuran, and DMF. With increasing cross-linking time, the flux decreased, indicating that the active layer became denser. The membrane showed excellent performance in DMF, a flux of up to 5 LMH, and the rejection of methyl orange of up to 98%. In the development of membranes for the treatment of solutions containing both organic solvents and acids, it may be a promising strategy to impart acid resistance to organic-solvent-resistant PI membranes via post-treatment.
Recently, sustainability and green chemistry have attracted considerable interest for reducing the worldwide usage of toxic chemicals and pollution. However, in most membrane fabrication and modification processes, conventional toxic solvents are used [21,22]; therefore, studies have been currently focused on replacing these solvents with non-toxic or green alternatives. Figure 1 presents the major merits of using green solvents as replacements for toxic solvents. The expression “green chemistry” was introduced to attract the attention of scientists worldwide to limit the utilization of hazardous and toxic reagents or solvents in the production of chemicals [23].
Zhao et al. cross-linked the polybenzimidazole (PBI) backbone using potassium persulfate as an oxidant; this method is categorized as “green” because of the use of an aqueous solution instead of toxic solvents [24]. In another study, Shin et al. modified a PBI organic-solvent-resistant NF membrane using oxidative cross-linking with an aqueous solution of potassium permanganate (KMnO4) [25]. Membrane science and technology is one of the crucial components of green chemistry because it plays a major role in the environmental friendliness of industrial processes. Although recent advancements in the field of membrane fabrication and performance enhancement are commendable, the materials used must be revised to reduce the pollution and pernicious effects of toxic chemicals [26,27,28].
Herein, a novel facile method for cross-linking the commercially available PI NF membranes via post-treatment is introduced to improve their acid stability. The Fenton reaction, an oxidative cross-linking process involving catalytic decomposition of hydrogen peroxide (H2O2) to hydroxyl radicals in presence of ferrous iron, was applied to crosslink and improve the acid resistance of the membrane. In this method, green solvents (water and ethanol) are used, making it a sustainable approach to cross-linking/modification of PI membranes. The surface properties of the pristine and modified PI membranes were studied using various characterization methods; the acid stability of the membranes was investigated using 50 w/v% H2SO4, and the water flux and MgSO4 rejection were determined through a filtration test. Changes in the surface properties of the PI membranes after exposure to acid were evaluated using various analytical tools.

2. Materials and Methods

2.1. Materials

DuraMem®900 (Flat sheet P84 PI membrane, Evonik (Essen, Germany)), a commercially available PI membrane, was used as a PI NF membrane, the acid stability of which was improved. Milli-Q water, ethyl alcohol (Daejung (Siheung, Republic of Korea), 95%), iron (II) sulfate heptahydrate (Sigma-Aldrich (St. Louis, MO, USA), 99%), and H2O2 (Daejung (Siheung, Republic of Korea), 30%) were used for the Fenton reaction. MPD (Sigma-Aldrich (St. Louis, MO, USA), 99%) was used as an additive in the Fenton reaction. H2SO4 (Daejung (Siheung, Republic of Korea), 95%) was used in the acid stability test, and MgSO4 (Daejung (Siheung, Republic of Korea), 99%) was used to prepare the conductive feed solution. All chemicals were used as received.

2.2. Modification of PI Membrane via Fenton Reaction

Prior to modification, the PI membrane was immersed in running deionized (DI) water for 24 h to remove any residual chemicals present on its surface. Iron sulfate (2.8 g), ethanol (100 g), and Milli-Q water (100 g) were mixed in a 250 mL glass screw-cap reagent bottle. The mixture was continuously magnetically stirred at 50 °C. The pristine PI membrane was immersed in the reaction mixture and purged with nitrogen. Under a nitrogen atmosphere, 12 g of H2O2 was added to the reaction mixture and allowed to react for 1 h. After the reaction completion, the membrane was thoroughly washed with running DI water and stored in Milli-Q water for further characterization and performance tests.
The effect of the MPD additive was evaluated during the Fenton reaction. In a 250 mL glass screw-cap reagent bottle, aqueous solutions (100 g) of MPD with various wt.% were prepared by dissolving 0, 0.2, 1.5, and 2.5 g of MPD in Milli-Q water. The mixture was stirred thoroughly to obtain a homogeneous solution. To each of these 100 g aqueous MPD solutions, 2.8 g of iron sulfate and 100 g of ethanol were added and stirred. Thereafter, the Fenton reaction was conducted as described above.

2.3. Filtration Performance of the Membrane

The performance of the pristine and modified PI membranes was evaluated using filtration test equipment similar to that used in other studies [29]. The filtration test was performed at 25 °C at a flow rate of 2.5 L/min using 2000 ppm MgSO4 aqueous solution. Four membrane samples with an active layer area of 19.6 cm2 were installed and processed simultaneously in the filtration system. Before sampling, the filtration test procedure was followed by compaction at 150 psi for 1 h and stabilization at 75 psi for 30 min. Water flux and salt rejection were measured three times, each for 10 min at 75 psi. Water flux was calculated as the change in volume of the permeate solution (converted from weight change and density of the solution) per unit membrane area and time, as in Equation (1). The concentrations of the feed and permeate solutions were calculated using a conductivity meter, and the salt rejection of MgSO4 was determined according to Equation (2).
J = w t × A × ρ
  • J = water flux (L/m2h)
  • w = weight of permeate water (g)
  • t = time (h)
  • A = effective membrane area (m2)
  • ρ = density of water (g/L)
R % = 1 C p e r m e a t e C f e e d × 100

2.4. Acid Resistance and Organic Solvent Stability Test

The acid resistance of the pristine and modified membranes was evaluated by exposing them to 50 w/v% H2SO4 at room temperature for 4 h. The acid-exposed membrane was thoroughly washed with DI water, and a performance test was conducted as described in Section 2.3. Unless indicated otherwise, acid-resistance tests were performed under these conditions. However, for the long-term exposure test, the membrane was exposed to 15 w/v% H2SO4 for 240 h. J
The organic solvent stability of the PI membranes was determined by comparing the weights of the membranes before and after exposure to the organic solvent of interest for a specified time. Membrane samples were washed with DI water and completely dried in an oven at 100 °C. The active layer was separated from the support layer to exclude the effect of the support layer during the solvent exposure test. It was weighed and exposed to dimethylacetamide (DMAc) in a glass vial at 80 °C for 20 h. After 20 h of exposure, the residual active layer was washed and then dried for at least 24 h in an oven at 100 °C. The ratio of the weight before exposure to that after exposure was calculated using Equation (3). The solvent resistance of the membrane samples was then compared.
W % = W e i g h t   a f t e r   s o l v e n t   e x p o s u r e   W e i g h t   b e f o r e   s o l v e n t   e x p o s u r e × 100

2.5. Characterization

2.5.1. Scanning Electron Microscopy (SEM) and Attenuated Total Reflectance-Fourier Transform Infrared Spectroscopy (ATR-FTIR)

SEM (S-4800, Hitachi High-Technology (Tokyo, Japan)) was used to investigate the surface morphology of the membrane before and after exposure to acid and before and after the modification of the membrane. The analyzed membrane was immersed in DI water and freeze-dried for at least 72 h prior to analysis. The freeze-dried membrane was coated with Pt at 20 mA and 2 × 10−3 mbar for 60 s in a turbo-pumped high-resolution chromium sputter coater (K575X, EMITECH, Lohmar, Germany) to reduce image artifacts caused by static electricity.
ATR-FTIR (Nicolet 6700, Thermo Fisher, Waltham, MA, USA) was used to investigate the chemical properties of the membranes. The freeze-dried membrane samples were analyzed using Ge crystals. ATR-FTIR was conducted at a resolution of 4 cm−1 in the range 700–3700 cm−1, and the obtained spectra were analyzed using Omnic 8.1 software.

2.5.2. X-ray Photoelectron Spectroscopy (XPS) and Thermo-Gravimetric Analysis (TGA)

XPS was used to evaluate the effect of the Fenton reaction on the elemental composition of PI membranes. XPS spectra were obtained in the range of the electron binding energy of 0–1000 eV with a resolution of 1 eV. In this study, the binding energies of C1s, O1s, and N1s were detected at 285, 531, and 399 eV, respectively.
TGA (Q500, TA Instruments (New Castle, DE, USA) was used to investigate the thermal stability of the PI membranes. The membrane sample immersed in DI water was cut into 2 cm × 2 cm pieces and freeze-dried for at least 72 h. The test was performed at 30–800 °C and a ramp rate of 10 °C/min. The effect of the Fenton reaction on thermal stability was investigated based on the difference in weight between the pristine and modified membranes.

2.5.3. Molecular Weight Cut-Off (MWCO) and Pore Size of Membranes

MWCO of the membranes was determined by the molecular weight of the neutral solute, in which 90% of the solute was retained by the membrane. The MWCO values of the pristine and Fenton-modified membranes were obtained using the previously reported method [30,31,32]. Filtration tests were conducted using the feed solutions of 50 ppm polyethylene glycol (PEG) with molecular weights of 400, 600, 1000, and 2000 Da as a model solute at 25 °C and a flow rate of 2.5 L/min. The concentrations of PEGs in the feed and permeate were measured using a total organic carbon analyzer (TOC-V, Shimadzu, Japan). Because the molecular weight (Mw, Da) of the solute, PEG, is related to the Stokes–Einstein radius (R, cm), the average pore size of the membrane can be estimated [33] using
R = 16.73 × 10 10 M w 0.557

3. Result and Discussion

3.1. Performance Degradation of the PI Membrane Owing to Acid Exposure

The acid resistance of PI membranes was evaluated by exposure to 15 w/v% H2SO4 for 240 h. The pH values of the H2SO4 solutions with pKa of −3 and 1.9 [34] are approximately −0.2. The flux and rejection of the PI membrane after exposure to 15 w/v% H2SO4 (Figure 2a and 2b, respectively) can be categorized into three phases. In phase I, the flux slightly increases, and the rejection remains almost constant. In phase II, the flux gradually increases, but rejection monotonically decreases. In phase III, the flux rapidly increases, and the membrane loses its capability to reject ions from the feed solution. From these results, it can be inferred that the PI membrane has a certain tolerance for acid, but beyond this range, the membrane degrades under acidic conditions.
To ensure that the experiment is systematic and time-efficient, the concentration of H2SO4 was increased to 50 w/v%. The pH of 50 w/v% H2SO4 was −0.7~−0.8. The performance of the acid-exposed membrane was expressed as the normalized flux and normalized rejection and compared to the performance of the pristine membrane. The normalized flux after exposure to 50 w/v% H2SO4 for various time intervals is shown in Figure 2c. The flux of the PI membrane after exposure to the acidic solution for 1, 2, 4, or 8 h increased to 139, 149, 205, and 278%, respectively, and the normalized rejection, as shown in Figure 2d, decreased to 60, 60, 44, and 28%, respectively. Thus, the damage to the PI membrane is strongly dependent on the exposure time. Exposure to 50 w/v% H2SO4 for 4 h was used as the standard exposure condition in further experiments.

3.2. Fenton Reaction to Improve acid Resistance of PI Membranes

The performance of pristine, modified, and acid-exposed membranes is shown in Figure 3. After exposure to 50 w/v% H2SO4, the normalized flux increased to 205%, and the normalized rejection decreased to 44% (dashed line in Figure 3), indicating that the PI membrane was damaged by acid exposure. When ferrous ions (Fe2+) are added to H2O2 solution, ferrous ions are oxidized to ferric ions (Fe3+), and H2O2 is converted into hydroxyl free radicals and hydroxyl anions. Then, the ferric ions are again reduced to ferrous ions, generating protons and hydroperoxyl free radicals from other H2O2 molecules (Figure 4) [35,36]. Iron ions serve as catalysts for the generation of free radicals during the reaction. The generated free radicals reacted with the PI to lower the flux and slightly increase the rejection. This is likely due to the cross-linking of the membrane. More details on the reaction mechanism are provided in the analysis of ATR-FTIR spectra.
The change in the pore size of the membrane by the Fenton reaction was evaluated through the rejection of a neutral solute passing through the membrane. The molecular weight of the solute at which the rejection reaches 90% is called MWCO. Figure 5 shows the rejection of PEG solutes with various molecular weights for both pristine and Fenton-modified PI membranes. The rejection of both membranes increases with increasing molecular weight of the PEG solute from 400 to 1200 Da. The MWCO of the pristine PI membrane was ~1060 Da, whereas that of the Fenton post-treated PI membrane was ~920 Da. These values correspond to the mean pore size of ~0.81 and ~0.74 nm, respectively. The reduction of pore size by the Fenton reaction explains the decrease in flux and a slight increase in the rejection, as shown in Figure 3.
At the same time, exposure of the Fenton-modified membrane to 50 w/v% H2SO4 resulted in 7% and 4% increases in normalized flux and rejection, respectively. Compared with the pristine membrane, the modified membrane showed little change in performance after exposure to H2SO4. Thus, Fenton modification enhanced the acid resistance of the membrane.
The structure of PI membranes was characterized using SEM. Cross-sectional SEM images of pristine and Fenton-modified PI membranes after exposure to 50 w/v% H2SO4 for 0, 1, 2, 4, and 8 h are shown in Figure 6. As the exposure time to the H2SO4 solution increases, the pore size of the pristine PI membrane also increases. This indicates that harsh acidic conditions led to polymer degradation. This result is consistent with the deterioration of the membrane performance shown by the increase in flux and decrease in the rejection upon exposure to acid (Figure 2). However, the pore size of the Fenton-modified membrane did not increase with the exposure time to the H2SO4 solution. Therefore, the Fenton-modified membrane has superior acid resistance, which is attributed to the cross-linking of the PI membrane by the Fenton modification.
Figure 7 shows the FTIR spectra of the DuraMem®900 PI membranes. Symmetric C=O stretching and C–N–C stretching of the imide peak are observed at 1720 cm−1 and at 1360 cm−1, respectively [19,20]. The peak of O–H stretching is distributed at 2500–3300 cm−1 [37], and the peaks of the CONH structure of the amide group, known as amide I and amide II bands, are observed at 1650 and 1540 cm−1, respectively [21]. The amide I band was mainly attributed to C=O stretching, and the amide II band was ascribed to the combination of N–H in-plane bending and N–C stretching of the amide bond [38]. Both imide and imide functional groups were also detected in the pristine PI membranes. In the Fenton-modified membrane, the peak of O–H stretching located between 3300 and 2500 cm−1 is more pronounced, and the CO peak of the secondary alcohol at 1100 cm−1 [39,40] is significantly more intense than in the pristine membrane. The PI membrane used in this study was a DuraMem®900 PI membrane (Evonik), which is a commercially available organic-solvent-resistant membrane. In the post-treatment for imparting organic solvent resistance to PI membranes, imide functional groups are commonly converted into amide and carboxylic groups [19]. When PI is immersed in an alkaline solution, the imide ring can be broken to form polyamic acid, which can react with other amines to form polyamide functional group [41]. According to ATR-FTIR, the organic-solvent-resistant PI membrane contains both amide and imide groups. It has been reported that amide bonds are vulnerable to acids [12,29]. The hydrolysis of amides begins when (i) the proton attacks the oxygen or nitrogen of the amide bond, followed by (ii) nucleophilic addition of water molecules to form a dihydroxy tetrahedral intermediate structure, (iii) transfer of the hydroxyl group from the oxygen to the nitrogen of the amide bond, and (iv) breakage of the amide bond [13,30]. During Fenton post-treatment of the membrane, the amine of the amide functional group can be attacked by hydroxyl radicals produced by Fenton reagents, leading to the radicalization of the amine. The radicalized amines react with each other to form a cross-linking bond (Figure 8). The cross-linking of the membrane might enhance the acid resistance of the membrane. The degree of stability of the amide bond is affected by the resonance between the electrons of the C=O π-bond and lone pair electrons of nitrogen [14]. When the nitrogen atoms of the amide bonds are bonded to each other (–CO– N ¨ N ¨ –O–) by Fenton post-treatment, the resonance structures of the two amide bonds overlap and expand, thereby strengthening the resonance structure and enhancing the acid tolerance of the membrane.
The surface chemistry changes in the PI membrane owing to the Fenton post-treatment were evaluated using XPS, and the results are shown in Table 1. The elemental composition of the pristine membrane was 73.0% carbon, 10.7% nitrogen, and 16.3% oxygen. However, in the Fenton-modified membrane, the atomic percentage of oxygen is significantly higher (58.3%). This result can be explained by the increase in the contents of hydroxyl and carboxylic functional groups generated in the PI membrane during the oxidation by H2O2 [42,43], as shown in the ATR-FTIR spectra (Figure 7). The decrease in the atomic percentages of carbon and nitrogen is attributed to the decrease in the total carbon and nitrogen content owing to the significantly increased oxygen content.
Figure 9 shows the TGA results for the membrane samples used in this study. Both samples show a rapid mass loss at ~400 °C, which was attributed to the degradation of the imide group [44,45]. At 800 °C, the residual masses of the pristine and Fenton-modified membranes were 20.3 and 26.7%, respectively. The Fenton-modified membrane shows better thermal durability, suggesting that cross-linking was realized through the Fenton reaction. The effect of the cross-linking of the PI membrane can also be observed by its solubility in polar aprotic solvents such as DMAc. Figure 9b shows the residual weight of the pristine and Fenton-modified membranes after the solvent exposure test. The pristine PI membrane retained 69.4% of its original weight after soaking at 80 °C for 20 h, whereas the Fenton-modified membrane retained 81.6%, demonstrating successful cross-linking and enhanced solvent resistance compared to that of the pristine membrane. Thus, Fenton post-treatment not only increased the acid resistance but also further improved the thermal stability and organic solvent resistance.
Table 2 presents a comparison of the solvents used for cross-linking the PI membrane and the performance of the membrane. Compared to other studies, the Fenton method requires less time for modification (1 h) and uses green solvents and non-toxic chemicals. The Fenton reaction of DuraMem® PI is a facile and sustainable process to enhance the acid resistance of PI membranes for the treatment of solutions containing organic solvents and strong acids.

3.3. Effect of MPD Additive in the Fenton Reaction

PI membranes have been applied in various separation processes owing to their excellent chemical resistance. In certain cases, depending on the material to be separated from the feed solution, the rejection of commercially available membranes needs to be tailored. In particular, in pharmaceutical processes using both strong acids and organic solvents, the recovery of specific substances is important. A simple method for obtaining the desired performance by controlling the rejection (or recovery) via the addition of MPD during Fenton modification is proposed.
The performance of the PI membrane modified by the Fenton reaction in the presence of MPD is shown in Figure 10. The concentrations of MPD added were 0, 0.25, 0.75, and 1.25%. When the concentration was increased to 2.25 and 3.5%, the flux significantly decreased to 0.8 and 0.5%, respectively, and the amount of permeate was insufficient for the rejection calculations. After the Fenton reaction, the J/J0 value was calculated to easily determine the change in the filtration characteristics of the membrane. J0 is the initial flux and J is the flux after modification. The J/J0 values of 0.67, 0.55, 0.30, and 0.18 were obtained at 0, 0.25, 0.75, and 1.25% MPD concentrations, respectively. An increase in the concentration of MPD resulted in a decrease in the flux after modification. Rejection was assessed by R/R0, which was calculated in the same way as the flux. The initial rejection is indicated by R0, and rejection after modification is indicated by R. R/R0 values of 1.14, 2.21, 2.37, and 3.76 are observed at MPD concentrations of 0, 0.25, 0.75, and 1.25%, respectively. An R/R0 value greater than 1 indicates that the rejection was increased by modification. The addition of MPD during the Fenton reaction increased the cross-linking of the polymeric membranes, thereby reducing the permeability and increasing the removal rate. However, the addition of MPD did not further increase acid resistance. According to the permeation test results, the performance of the acid-resistant and organic solvent-resistant DuraMem® PI membranes can be controlled by the addition of MPD, leading to the potential extension of the application scope.

4. Conclusions

In this study, the DuraMem® 900 PI membrane was modified using the strong oxidizing agent generated by the green solvent-based Fenton reaction. Pristine PI membrane exhibited a flux increase to 205% and a decrease in rejection to 44% after exposure to 50 w/v% H2SO4 at 25 °C for 4 h, which was used as the evaluation standard for acid resistance. These changes resulted from the degradation of the PI membrane in low pH environment. Fenton post-treatment induced cross-linking of the PI membrane, thereby reducing the flux and increasing the rejection. However, exposure of the Fenton-modified membrane to H2SO4 led to only a slight increase in flux and almost no changes in rejection. The modification through the Fenton reaction significantly improved the acid resistance of the PI membrane. The Fenton post-treatment of the PI membrane not only increased the acid resistance but also improved its thermal stability and organic solvent resistance. Furthermore, the performance of the acid-resistant and organic-solvent-resistant PI membranes was controlled using the MPD additive, thereby extending the application scope of the PI membrane modified via Fenton post-treatment.

Author Contributions

Conceptualization, Y.-N.K.; methodology, W.-S.K. and H.-K.L.; formal analysis, Y.-I.P., H.P. and I.-C.K.; investigation, S.R., W.-S.K. and H.-K.L.; resources, Y.-I.P., H.P. and I.-C.K.; writing—original draft preparation, S.R. and W.-S.K.; writing—review and editing, S.R. and Y.-N.K.; supervision, Y.-N.K.; funding acquisition, Y.-N.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant financed by the Korea Government (MOTIE) (20202020800330), and by the National Research Foundation (NRF) of Korea as funded by the Ministry of Education, Science and Technology (NRF-2018R1D1A1B07043609), and by Korea Ministry of Environment (MOE) as ‘Graduate School Specialized in Integrated Water Resources Management’.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gil, A.; Galeano, L.A.; Vicente, M.Á. Applications of Advanced Oxidation Processes (AOPs) in Drinking Water Treatment; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  2. Jye, L.W.; Ismail, A.F. Nanofiltration Membranes: Synthesis, Characterization, and Applications; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar]
  3. Zeng, Y.; Wang, L.; Zhang, L.; Yu, J.Q. An acid resistant nanofiltration membrane prepared from a precursor of poly (s-triazine-amine) by interfacial polymerization. J. Membr. Sci. 2018, 546, 225–233. [Google Scholar] [CrossRef]
  4. Platt, S.; Nyström, M.; Bottino, A.; Capannelli, G. Stability of NF membranes under extreme acidic conditions. J. Membr. Sci. 2004, 239, 91–103. [Google Scholar] [CrossRef]
  5. Visser, T.; Modise, S.; Krieg, H.; Keizer, K. The removal of acid sulphate pollution by nanofiltration. Desalination 2001, 140, 79–86. [Google Scholar] [CrossRef]
  6. Räsänen, E.; Nyström, M.; Sahlstein, J.; Tossavainen, O. Purification and regeneration of diluted caustic and acidic washing solutions by membrane filtration. Desalination 2002, 149, 185–190. [Google Scholar] [CrossRef]
  7. Niewersch, C.; Koh, C.; Wintgens, T.; Melin, T.; Schaum, C.; Cornel, P. Potentials of using nanofiltration to recover phosphorus from sewage sludge. Water Sci. Technol. 2008, 57, 707–714. [Google Scholar] [CrossRef]
  8. Blöcher, C.; Niewersch, C.; Melin, T. Phosphorus recovery from sewage sludge with a hybrid process of low pressure wet oxidation and nanofiltration. Water Res. 2012, 46, 2009–2019. [Google Scholar] [CrossRef]
  9. López, J.; Reig, M.; Gibert, O.; Torres, E.; Ayora, C.; Cortina, J.L. Application of nanofiltration for acidic waters containing rare earth elements: Influence of transition elements, acidity and membrane stability. Desalination 2018, 430, 33–44. [Google Scholar] [CrossRef] [Green Version]
  10. Vitale, S.A.; Chae, H.; Sawin, H.H. Silicon etching yields in F2, Cl2, Br2, and HBr high density plasmas. J. Vac. Sci. Technol. A 2001, 19, 2197–2206. [Google Scholar] [CrossRef]
  11. Brion, C.E.; Dyck, M.; Cooper, G. Absolute photoabsorption cross-sections (oscillator strengths) for valence and inner shell excitations in hydrogen chloride, hydrogen bromide and hydrogen iodide. J. Electron Spectrosc. Relat. Phenom. 2005, 144, 127–130. [Google Scholar] [CrossRef]
  12. Jun, B.M.; Lee, H.K.; Park, Y.I.; Kwon, Y.N. Degradation of full aromatic polyamide NF membrane by sulfuric acid and hydrogen halides: Change of the surface/permeability properties. Polym. Degrad. Stab. 2019, 162, 1–11. [Google Scholar] [CrossRef]
  13. Jun, B.M.; Kim, S.H.; Kwak, S.K.; Kwon, Y.N. Effect of acidic aqueous solution on chemical and physical properties of polyamide NF membranes. Appl. Surf. Sci. 2018, 444, 387–398. [Google Scholar] [CrossRef]
  14. Vanherck, K.; Koeckelberghs, G.; Vankelecom, I.F. Crosslinking polyimides for membrane applications: A review. Prog. Polym. Sci. 2013, 38, 874–896. [Google Scholar] [CrossRef]
  15. Priske, M.; Lazar, M.; Schnitzer, C.; Baumgarten, G. Recent applications of organic solvent nanofiltration. Chem. Ing. Tech. 2016, 88, 39–49. [Google Scholar] [CrossRef]
  16. Vanherck, K.; Cano-Odena, A.; Koeckelberghs, G.; Dedroog, T.; Vankelecom, I. A simplified diamine crosslinking method for PI nanofiltration membranes. J. Membr. Sci. 2010, 353, 135–143. [Google Scholar] [CrossRef]
  17. Li, C.; Li, S.; Lv, L.; Su, B.; Hu, M.Z. High solvent-resistant and integrally crosslinked polyimide-based composite membranes for organic solvent nanofiltration. J. Membr. Sci. 2018, 564, 10–21. [Google Scholar] [CrossRef]
  18. Zhao, H.-Y.; Cao, Y.-M.; Ding, X.-L.; Zhou, M.-Q.; Liu, J.-H.; Yuan, Q. Poly (ethylene oxide) induced cross-linking modification of Matrimid membranes for selective separation of CO2. J. Membr. Sci. 2008, 320, 179–184. [Google Scholar] [CrossRef]
  19. Toh, Y.H.S.; Lim, F.W.; Livingston, A.G. Polymeric membranes for nanofiltration in polar aprotic solvents. J. Membr. Sci. 2007, 301, 3–10. [Google Scholar] [CrossRef]
  20. Vandezande, P.; Gevers, L.E.; Weyens, N.; Vankelecom, I.F. Compositional optimization of polyimide-based SEPPI membranes using a genetic algorithm and high-throughput techniques. J. Comb. Chem. 2009, 11, 243–251. [Google Scholar] [CrossRef] [PubMed]
  21. Chen, D.; Yan, C.; Li, X.; Liu, L.; Wu, D.; Li, X. A highly stable PBI solvent resistant nanofiltration membrane prepared via versatile and simple crosslinking process. Sep. Purif. Technol. 2019, 224, 15–22. [Google Scholar] [CrossRef]
  22. Nguyen Thi, H.Y.; Nguyen, B.T.D.; Kim, J.F. Sustainable fabrication of organic solvent nanofiltration membranes. Membranes 2020, 11, 19. [Google Scholar] [CrossRef]
  23. Mehrabani, S.A.N.; Vatanpour, V.; Koyuncu, I. Green solvents in polymeric membrane fabrication: A review. Sep. Purif. Technol. 2022, 298, 121691. [Google Scholar] [CrossRef]
  24. Zhao, B.W.; Shi, G.M.; Wang, K.Y.; Lai, J.Y.; Chung, T.S. Employing a green cross-linking method to fabricate polybenzimidazole (PBI) hollow fiber membranes for organic solvent nanofiltration (OSN). Sep. Purif. Technol. 2021, 255, 117702. [Google Scholar] [CrossRef]
  25. Shin, S.J.; Park, Y.-I.; Park, H.; Cho, Y.H.; Won, G.Y.; Yoo, Y. A facile crosslinking method for polybenzimidazole membranes toward enhanced organic solvent nanofiltration performance. Sep. Purif. Technol. 2022, 299, 121783. [Google Scholar] [CrossRef]
  26. Figoli, A.; Marino, T.; Simone, S.; Di Nicolò, E.; Li, X.-M.; He, T.; Tornaghi, S.; Drioli, E. Towards non-toxic solvents for membrane preparation: A review. Green Chem. 2014, 16, 4034–4059. [Google Scholar] [CrossRef]
  27. Xie, W.; Li, T.; Tiraferri, A.; Drioli, E.; Figoli, A.; Crittenden, J.C.; Liu, B. Toward the next generation of sustainable membranes from green chemistry principles. ACS Sustain. Chem. Eng. 2020, 9, 50–75. [Google Scholar] [CrossRef]
  28. Kim, D.; Nunes, S.P. Green solvents for membrane manufacture: Recent trends and perspectives. Curr. Opin. Green Sustain. Chem. 2021, 28, 100427. [Google Scholar] [CrossRef]
  29. Jun, B.M.; Kwon, Y.N.; Kim, S.H.; Lim, H.Y.; Kwak, S.K. Acid stability of polyamide membranes. Polymer 2022, 241, 124516. [Google Scholar] [CrossRef]
  30. Jun, B.M.; Lee, H.K.; Kwon, Y.N. Acid-catalyzed hydrolysis of semi-aromatic polyamide NF membrane and its application to water softening and antibiotics enrichment. Chem. Eng. J. 2018, 332, 419–430. [Google Scholar] [CrossRef]
  31. Weng, X.D.; Ji, Y.L.; Ma, R.; Zhao, F.Y.; An, Q.F.; Gao, C.J. Superhydrophilic and antibacterial zwitterionic polyamide nanofiltration membranes for antibiotics separation. J. Membr. Sci. 2016, 510, 122–130. [Google Scholar] [CrossRef]
  32. Wei, X.Z.; Ong, X.; Yang, J.; Zhang, G.L.; Chen, J.Y.; Wang, J.D. Structure influence of hyperbranched polyester on structure and properties of synthesized nanofiltration membranes. J. Membr. Sci. 2013, 440, 67–76. [Google Scholar] [CrossRef]
  33. Jayalakshmi, A.; Kim, I.C.; Kwon, Y.N. Cellulose acetate graft-(glycidylmethacrylate-g-PEG) for modification of AMC ultrafiltration membranes to mitigate organic fouling. RSC Adv. 2015, 5, 48290–48300. [Google Scholar] [CrossRef] [Green Version]
  34. Snoeyink, V.L.; Jenkins, D. Water Chemistry; Wiley: Hoboken, NJ, USA, 1980. [Google Scholar]
  35. Reyhani, A.; McKenzie, T.G.; Fu, Q.; Qiao, G.G. Fenton-Chemistry-Mediated Radical Polymerization. Macromol. Rapid Commun. 2019, 40, e1900220. [Google Scholar] [CrossRef]
  36. Salgado, P.; Melin, V.; Contreras, D.; Moreno, Y.; Mansilla, H.D. Fenton reaction driven by iron ligands. J. Chil. Chem. Soc. 2013, 58, 2096–2101. [Google Scholar] [CrossRef] [Green Version]
  37. Chaffraix, T.; Voda, A.S.; Dumee, L.F.; Magniez, K. Surface ionic charge dependence on the molecular mobility and self-assembly behavior of ionomers produced from carboxylic acid-terminated dendrimers. Polym. J. 2017, 49, 245–254. [Google Scholar] [CrossRef]
  38. Kwon, Y.N.; Leckie, J.O. Hypochlorite degradation of crosslinked polyamide membranes—II. Changes in hydrogen bonding behavior and performance. J. Membr. Sci. 2006, 282, 456–464. [Google Scholar] [CrossRef]
  39. Xu, L.F.; Li, C.; Ng, K.Y.S. In-situ monitoring of urethane formation by FTIR and Raman spectroscopy. J. Phys. Chem. A 2000, 104, 3952–3957. [Google Scholar] [CrossRef]
  40. Sudhamani, S.R.; Prasad, M.S.; Sankar, K.U. DSC and FTIR studies on Gellan and Polyvinyl alcohol (PVA) blend films. Food Hydrocolloid 2003, 17, 245–250. [Google Scholar] [CrossRef]
  41. Kim, H.J.; Park, Y.J.; Choi, J.H.; Han, H.S.; Hong, Y.T. Surface modification of polyimide film by coupling reaction for copper metallization. J. Ind. Eng. Chem. 2009, 15, 23–30. [Google Scholar] [CrossRef]
  42. Radwan, N.R.E.; Selim, M.M. Catalytic hydroxylation of benzene by H2O2 in the presence of surfactants. Afinidad 1998, 55, 139–142. [Google Scholar]
  43. Liu, G.Y.; Huang, H.B.; Xie, R.J.; Feng, Q.Y.; Fang, R.M.; Shu, Y.J.; Zhan, Y.J.; Ye, X.G.; Zhong, C. Enhanced degradation of gaseous benzene by a Fenton reaction. RSC Adv. 2017, 7, 71–76. [Google Scholar] [CrossRef] [Green Version]
  44. Wang, L.; Zhang, L.C.; Fischer, A.; Zhong, Y.H.; Drummer, D.; Wu, W. Enhanced thermal conductivity and flame retardancy of polyamide 6/flame retardant composites with hexagonal boron nitride. J. Polym. Eng. 2018, 38, 767–774. [Google Scholar] [CrossRef]
  45. Deligoz, H.; Ozgumus, S.; Yalcinyuva, T.; Yildirim, S.; Deger, D.; Ulutas, K. A novel cross-linked polyimide film: Synthesis and dielectric properties. Polymer 2005, 46, 3720–3729. [Google Scholar] [CrossRef]
  46. Luo, X.; Wang, Z.; Wu, S.; Fang, W.; Jin, J. Metal ion cross-linked nanoporous polymeric membranes with improved organic solvent resistance for molecular separation. J. Membr. Sci. 2021, 621, 119002. [Google Scholar] [CrossRef]
  47. Liu, Q.; Xu, S.; Xiong, S.; Yi, M.; Wang, Y. Coordination-crosslinked polyimide supported membrane for ultrafast molecular separation in multi-solvent systems. Chem. Eng. J. 2022, 427, 130941. [Google Scholar] [CrossRef]
  48. Xu, Y.C.; Cheng, X.Q.; Long, J.; Shao, L. A novel monoamine modification strategy toward high-performance organic solvent nanofiltration (OSN) membrane for sustainable molecular separations. J. Membr. Sci. 2016, 497, 77–89. [Google Scholar] [CrossRef]
  49. Gao, J.; Japip, S.; Chung, T.-S. Organic solvent resistant membranes made from a cross-linked functionalized polymer with intrinsic microporosity (PIM) containing thioamide groups. Chem. Eng. J. 2018, 353, 689–698. [Google Scholar] [CrossRef]
  50. Feng, W.; Li, J.; Fang, C.; Zhang, L.; Zhu, L. Controllable thermal annealing of polyimide membranes for highly-precise organic solvent nanofiltration. J. Membr. Sci. 2022, 643, 120013. [Google Scholar] [CrossRef]
  51. Yang, C.; Xu, W.; Nan, Y.; Wang, Y. Novel solvent-resistant nanofiltration membranes using MPD co-crosslinked polyimide for efficient desalination. J. Membr. Sci. 2020, 616, 118603. [Google Scholar] [CrossRef]
  52. Li, S.; Li, C.; Song, X.; Su, B.; Mandal, B.; Prasad, B.; Gao, X.; Gao, C. Graphene quantum dots-doped thin film nanocomposite polyimide membranes with enhanced solvent resistance for solvent-resistant nanofiltration. ACS Appl. Mater. Interfaces 2019, 11, 6527–6540. [Google Scholar] [CrossRef]
  53. Li, C.; Li, S.; Tian, L.; Zhang, J.; Su, B.; Hu, M.Z. Covalent organic frameworks (COFs)-incorporated thin film nanocomposite (TFN) membranes for high-flux organic solvent nanofiltration (OSN). J. Membr. Sci. 2019, 572, 520–531. [Google Scholar] [CrossRef]
Figure 1. Merits of green and sustainable solvents over conventional solvents.
Figure 1. Merits of green and sustainable solvents over conventional solvents.
Polymers 15 00264 g001
Figure 2. (a) Flux and (b) MgSO4 rejection of the PI membrane after long-term exposure to 15 w/v% H2SO4. (c) Normalized flux and (d) normalized rejection of the PI membrane after exposure to 50 w/v% H2SO4.
Figure 2. (a) Flux and (b) MgSO4 rejection of the PI membrane after long-term exposure to 15 w/v% H2SO4. (c) Normalized flux and (d) normalized rejection of the PI membrane after exposure to 50 w/v% H2SO4.
Polymers 15 00264 g002
Figure 3. Normalized flux and normalized rejection of pristine and Fenton-modified PI membranes. The membrane was modified by the Fenton reaction for 1 h.
Figure 3. Normalized flux and normalized rejection of pristine and Fenton-modified PI membranes. The membrane was modified by the Fenton reaction for 1 h.
Polymers 15 00264 g003
Figure 4. Schematic of Fenton modification of the PI membrane.
Figure 4. Schematic of Fenton modification of the PI membrane.
Polymers 15 00264 g004
Figure 5. PEGs rejection curves of pristine and Fenton-modified PI membranes for MWCO measurement.
Figure 5. PEGs rejection curves of pristine and Fenton-modified PI membranes for MWCO measurement.
Polymers 15 00264 g005
Figure 6. SEM images of pristine and Fenton-modified PI membranes after exposure to 50 w/v% H2SO4 for 0, 1, 2, 4, and 8 h.
Figure 6. SEM images of pristine and Fenton-modified PI membranes after exposure to 50 w/v% H2SO4 for 0, 1, 2, 4, and 8 h.
Polymers 15 00264 g006
Figure 7. ATR-FTIR spectra of pristine and Fenton-modified PI membranes.
Figure 7. ATR-FTIR spectra of pristine and Fenton-modified PI membranes.
Polymers 15 00264 g007
Figure 8. Cross-linking of Duramem900 PI membrane using the Fenton reaction.
Figure 8. Cross-linking of Duramem900 PI membrane using the Fenton reaction.
Polymers 15 00264 g008
Figure 9. (a) TGA spectrum of pristine and modified membranes and (b) residual weight of pristine and Fenton-modified membranes soaked in DMAC at 80 °C for 20 h.
Figure 9. (a) TGA spectrum of pristine and modified membranes and (b) residual weight of pristine and Fenton-modified membranes soaked in DMAC at 80 °C for 20 h.
Polymers 15 00264 g009
Figure 10. J/J0 and R/R0 values of Fenton-MPD-modified membranes.
Figure 10. J/J0 and R/R0 values of Fenton-MPD-modified membranes.
Polymers 15 00264 g010
Table 1. Elemental compositions of pristine and Fenton cross-linked membranes according to XPS.
Table 1. Elemental compositions of pristine and Fenton cross-linked membranes according to XPS.
PristineFenton
Carbon (%)73.036.2
Nitrogen (%)10.75.5
Oxygen (%)16.358.3
Table 2. Comparison of crosslinking methods of PI membranes.
Table 2. Comparison of crosslinking methods of PI membranes.
MembranesCrosslinking MethodSolvents Used for Crosslinking Reaction Time Used for CrosslinkingSoluteRejection (%)ApplicationReference
DuraMem®900 PIFenton reaction
(oxidative crosslinking—Iron sulfate + Hydrogens peroxide)
Ethanol and Water1 hPEG 1000~93Acid Resistant Membrane (ARM)This work
Synthesized PI (6FDA + Durene + DABA)Cu2+ crosslinking between carboxylic acid of synthesized PIMethanol24 hCBB~99%OSN [46]
Lenzing P84 PICrosslinking of PI with diamine (HDA)Isopropyl alcohol24 hRDB (479)
RB (1017)
9899.9OSN[17]
Matrimid® 5218 PIImide ring opening with NaOH followed by Ca2+ coordination crosslinking Water~48 hRose Bengal>98Support layer[47]
P84 PIActivation of PI with Tris and then crosslinking with Diamine (HDA)Isopropyl alcohol24 hPS (200)∼93OSN[48]
P84 PICrosslinking with diamine (HDA)Ethanol 24 h. RB (974)97.74Substrate[49]
P84 PI Thermal annealing at 330 °C followed by solvent activationDMF (Solvent activation)~ 4 h + 12 h (Solvent activation)Food yellow 3 (452.36Da)99.0OSN[50]
Synthesized PI (TMPDA + DDM + PMDA)Crosslinking with MPDWater 24 hSalt rejections ~97.45Substrate[51]
Lenzing P84 PICrosslinking with diamine (HDA) Isopropyl alcohol, DMF (Solvent activation)1 h + (30 min (Solvent activation)Rhodamine B98.6%Substrate [52]
Lenzing P84 PICrosslinking with diamine (HDA)Isopropyl alcohol, DMF (Solvent activation) alcohol, Hexane 1 h + (30 min (Solvent activation)Rhodamine B >99% Substrate[53]
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

Ravi, S.; Kang, W.-S.; Lee, H.-K.; Park, Y.-I.; Park, H.; Kim, I.-C.; Kwon, Y.-N. Improvement in Acid Resistance of Polyimide Membranes: A Sustainable Cross-Linking Approach via Green-Solvent-Based Fenton Reaction. Polymers 2023, 15, 264. https://doi.org/10.3390/polym15020264

AMA Style

Ravi S, Kang W-S, Lee H-K, Park Y-I, Park H, Kim I-C, Kwon Y-N. Improvement in Acid Resistance of Polyimide Membranes: A Sustainable Cross-Linking Approach via Green-Solvent-Based Fenton Reaction. Polymers. 2023; 15(2):264. https://doi.org/10.3390/polym15020264

Chicago/Turabian Style

Ravi, Srinath, Woo-Seok Kang, Hyung-Kae Lee, You-In Park, Hosik Park, In-Chul Kim, and Young-Nam Kwon. 2023. "Improvement in Acid Resistance of Polyimide Membranes: A Sustainable Cross-Linking Approach via Green-Solvent-Based Fenton Reaction" Polymers 15, no. 2: 264. https://doi.org/10.3390/polym15020264

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