Next Article in Journal
Ring-Opening Reaction of 1-Phospha-2-Azanorbornenes via P-N Bond Cleavage and Reversibility Studies
Previous Article in Journal
Chemical Composition, Functional and Anticancer Properties of Carrot
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deep Eutectic Solvent Coated Cerium Oxide Nanoparticles Based Polysulfone Membrane to Mitigate Environmental Toxicology

1
Department of Chemical Engineering, COMSATS University Islamabad, Lahore Campus, Defence Road, Off Raiwind Road, Lahore 54000, Punjab, Pakistan
2
Interdisciplinary Research Center in Biomedical Materials, COMSATS University Islamabad, Lahore Campus, Defence Road, Off Raiwind Road, Lahore 54000, Punjab, Pakistan
3
Department of Chemical Engineering, Ghulam Ishaq Khan Institute of Engineering Sciences and Technology, Topi 23460, Khyber Pakhtunkhwa, Pakistan
4
Department of Physics, GC University Faisalabad, Faisalabad 38000, Punjab, Pakistan
5
State Key Laboratory of Metastable Materials Science and Technology, School of Materials Science and Engineering, Yanshan University, Qinhuangdao 066004, China
6
Department of Chemistry, Kulliyyah of Science, International Islamic University, Malaysia, Jalan Sultan Ahmad Shah, Kuantan 25200, Pahang, Malaysia
7
MEU Research Unit, Middle East University, Amman 541350, Jordan
8
Chemistry Department, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
9
Department of Chemical Engineering, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
10
Department of Chemical, Polymer and Composite Materials Engineering, University of Engineering and Technology Lahore, New Campus, Lahore 39161, Punjab, Pakistan
11
Department of Dental Materials, Institute of Basic Medical Sciences, Khyber Medical University, Peshawar 25100, Khyber Pakhtunkhwa, Pakistan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(20), 7162; https://doi.org/10.3390/molecules28207162
Submission received: 12 September 2023 / Revised: 4 October 2023 / Accepted: 11 October 2023 / Published: 19 October 2023
(This article belongs to the Section Materials Chemistry)

Abstract

:
In this study, ceria nanoparticles (NPs) and deep eutectic solvent (DES) were synthesized, and the ceria-NP’s surfaces were modified by DES to form DES-ceria NP filler to develop mixed matrix membranes (MMMs). For the sake of interface engineering, MMMs of 2%, 4%, 6% and 8% filler loadings were fabricated using solution casting technique. The characterizations of SEM, FTIR and TGA of synthesized membranes were performed. SEM represented the surface and cross-sectional morphology of membranes, which indicated that the filler is uniformly dispersed in the polysulfone. FTIR was used to analyze the interaction between the filler and support, which showed there was no reaction between the polymer and DES-ceria NPs as all the peaks were consistent, and TGA provided the variation in the membrane materials with respect to temperature, which categorized all of the membranes as very stable and showed that the trend of stability increases with respect to DES-ceria NPs filler loading. For the evaluation of efficiency of the MMMs, the gas permeation was tested. The permeability of CO2 was improved in comparison with the pristine Polysulfone (PSF) membrane and enhanced selectivities of 35.43 ( α CO2/CH4) and 39.3 ( α CO2/N2) were found. Hence, the DES-ceria NP-based MMMs proved useful in mitigating CO2 from a gaseous mixture.

Graphical Abstract

1. Introduction

CO2 is the major impurity found in natural gas, which is a major contributor to methane and is an important source of renewable energy [1]. The CO2 in natural gas causes the corrosion of pipelines and reduces its calorific value [2]. Also, as it is a greenhouse gas, it results in a thermal effect on the environment by absorbing radiation which gives rise to global warming. Concerning the sources of CO2 emissions, power generation is the major CO2 discharge source, which is due to the combustion of natural gas and coal or other fossil fuels for the production of electricity [3]. The burning of fuels in vehicles on a domestic level, as well as in other industries, are further sources of CO2 emissions. Hence, to reduce greenhouse gas emissions and global warming, to improve economic efficiency, and to provide a sustainable source of energy, it has become necessary to separate CO2 from natural gas [4]. As well as CO2/CH4 separation, CO2/N2 separation is also necessary as it improves the cleanliness of the atmosphere and reduces the percentage of CO2 in the environment [5].
Membrane technology has attracted researchers due to its attributes of having a small footprint, being easy to operate and having a low energy consumption in separating CO2 [6]. Membrane-based materials have great significance in this field [7] and organic polymeric membranes are leading the market, as compared to other two-dimensional materials [8] and organics [9,10,11]. The reasons for this are the low cost, easy processing, and high mechanical strength of the polymers [12,13]. The major problem faced in this field is overcoming the permeability and selectivity trade-off as explained by Rosebeson [14]. The incorporation of novel advanced fillers on polymeric support to synthesize membranes may provide an effective solution to optimize the perm-selectivity trade-off [15,16].
The synthesis of mixed matrix membranes (MMMs) has some drawbacks, like fewer interactions between filler particles and polymers, the non-uniform dispersion of the filler on polymer, and chances of a reaction between the filler and polymers [17,18]. The interface morphology of MMMs is also a significant factor in separation via membrane and is controlled by novel supervenient materials as filler. Covalent organic frameworks (COFs), metal organic frameworks (MOFs), metal oxides, alumina, silica, zeolites, nanoparticles, nanoghraphenes, and ceramics are the materials that are being used as filler for CO2 separation in different separation techniques, as well as in the fabrication of MMMs [19,20,21]. Sometimes inimical problems occur due to an acidic or humid environment, which can become the cause of interface incompatibility between the filler and polymer [22]. The specification of the filler plays a key role in overcoming this type of problem, and desirable features can be imparted to the membranes by introducing specific fillers in the membrane matrix [23,24].
Recent studies have employed nanoparticles as the filler in mixed matrix membranes. Hasebe et al. [25] has applied silica nanoparticles in polymeric membranes to separate CO2 as a way to overcome the Robeson upper bound limit, which resulted in a cost-effective solution for high performance CO2 separation in the form of permeability and selectivity (CO2/N2) on large scale. Raouf et al. [26] have used graphene hydroxyl nanoparticles and polysulfone/polyethylene glycol to synthesize MMMs and obtained a permeability in the range of 15.9–28.2 Barrer and a selectivity (CO2/CH4) range of 12.23–12.81 at 2 bar pressure with a maximum of 22.39 at 8 bar pressure. Sainath et al. [27] grew ZIF-67 NPs on PSF/GO hollow fiber membranes (HFMs) to increase the CO2 removal from natural gas and reported selectivities of 44.94 ± 3.00 and 22.38 ± 0.30 for mixed and pure, respectively. Ruhaimi et al. [28] synthesized spherical CeO2 nanoparticles to apply in egg-shell membranes as a bio-template for the high efficiency of CO2 adsorption. Farashi et al. [29] has improved CO2 separation from CH4 using a Pebax-1657 membrane with the addition of alumina (Al2O3) NPs in the membrane matrix and reported CO2 permeability as 159.27 Barrer and selectivity as 24.73 at 8% Al2O3 NP loadings by weight with different characterizations of membrane. And Xu et al. [30] reported the high NPs loadings in MMMs with the help of chemical bridging-crosslinking for the high performance separation of CO2 and found the permeation of CO2 and selectivity (CO2/N2) to be 1295 GPU and 91 at 0.3 MPa, respectively.
In this study, the purpose was to mitigate the amount of CO2 from the gaseous mixture to reduce the toxicity of the gases, like natural gas, flue gases etc. A lot of materials have been used for this purpose, so there was the need to make a novel combination of the compounds that could have the potential to give better outcomes for this purpose. Following this purpose, DES immobilized cerium oxide (CeO2) is used in MMMs for CO2 separation. DESs are popular candidates and green solvents in the membrane technology because of their nontoxicity, low viscosity, low vapor pressure, prominent tunabilty, easy preparation, high biocompatibility and biodegradability [31]. It can be synthesized by mixing and heating at least two chemicals that are hydrogen bond donors (HBD) and hydrogen bond acceptors (HBA) [32]. Therefore, DES was synthesized by cetrimonium bromide (CTMB) and acetic acid as HBA and HBD, respectively. The nanoparticles (NPs) of ceria were synthesized by using cerium (IV) ammonium nitrate, ethylene glycol and isopropanol. DES was immobilized on ceria NPs by using ethanol via a solvent evaporation technique and filler for the MMMs was obtained. Polysulfone was used as support material to maintain the strength of the membrane matrix as it is easy to process, mechanically strong, and thermally and chemically stable. Finally, mixed matrix membranes of 2%, 4%, 6% and 8% DES immobilized ceria NP loadings were synthesized and analyzed by FTIR, SEM and TGA. Also, gas permeation was evaluated on the permeation of CO2 and the selectivities of CO2/N2 and CO2/CH4 for the detection of CO2 mitigation from the gaseous mixture. The results obtained proved that this combination tries to overcome and to optimize the trade-off between permeability and selectivity in the Robeson plot and lie near the line that might be a potential region of the plot. No doubt, it was not able to completely overcome the Robeson upper bound limit, but it still has potential improvement to be reported in comparison with pure polysulfone membranes.

2. Results and Discussion

2.1. Scanning Electron Microscopy (SEM)

SEM (Tescan Vega (LMU)) was utilized to observe the cross-sectional and surface morphology of the membranes of different NP loadings (Figure 1 and Figure S1). Membranes were broken in liquid nitrogen to fix the cross-sectional ends of the membranes and a voltage of 15–20 Kv at different magnifications was applied with gold coating of the samples in the SEM machine. A SEM micrograph is used to observe the aggregation, dispersion and the existence of the filler NPs in the membrane matrix. SEM images show the increasing NP loadings as 2%, 4%, 6% and 8% in the membrane and the filler is highly dispersed across the whole membrane in each case. All of the membranes are accurately synthesized, and the structural density of the composition is visible without any imperfections. Some of the ripples can be seen with the increase of filler loadings that are due to inert NP aggregation that leads to agglomeration in greater than 8% NP loadings. Obviously, the nanoparticles form agglomerates due to the strong interaction between the modified small particles which can be a consequence of generating channels or voids that indicate the molecular chains’ flexibility in the membrane matrix [33]. The agglomeration is also due to the fact that the shape of the DES-modified ceria NPs cannot be assumed [34]. The main contribution of the selective gas permeation and membrane strength depends upon the pore structure of MMMs [35], which is obvious from the cross-section morphology. The springy asymmetric structures shown are due to the strong contact between the filler and polysulfone, and poor attraction of tetrahydrofuran (THF), which is the organic solvent that caused the skin-like structures while drying [36]. Hence the uniform and homogeneous dispersion of the filler exhibited by the images point towards the enhanced permeation of CO2 and the relative selectivities. The pore formation increases with the increase of DES-ceria which supports the hypothesis.

2.2. Fourier Transform Infrared (FTIR) Spectroscopy

Figure 2 demonstrates the FTIR spectra for ceria, DES-immobilized ceria and all of the four compositions of the membranes in a 650–1800 cm−1 range. The major peak of ceria, that is the O-Ce-O stretching vibration, is at 1071 cm−1 [37,38,39]. It can be observed in all of the membrane compositions that indicated the presence of ceria filler in the synthesized MMMs. The peak at 1582 cm−1 indicates C=C stretching vibrations and benzene ring stretching [40]. The peaks at 1291 cm−1, 1321 cm−1 and 1483 cm−1 are the peaks of symmetric O=S=O stretching vibrations and C-H bending vibrations. The peaks at 1168 cm−1 shows C-O bending vibrations. The vibrational elongating symmetric behavior and asymmetric stretching vibrations of O=S=O bonds is represented by peaks at 1101 cm−1 and at 1146 cm−1, respectively, which is a clear indication that sulfone groups exist. The peak at 1233 cm−1 is a clear exhibition of the elongating vibration of Benz-O-Benz bonds, where Benz represents the benzene rings/aromatic functional group of PSF. The C=C stretching mode in aromatic compounds lies in the range 1485–1590 cm−1, which covers the peak at 1503 cm−1. The peak near 1710 cm−1, that is 1740 cm−1, indicates the symmetric stretching of a C=O bond, and at 829 cm−1 indicates the C6H6 ring bending. The peak at 1010 cm−1 represents C-H stretching of the C6H6 ring in PSF. The small shift in the peaks of the membrane in comparison with ceria NPs indicates the strong interface of the filler and polymer at the molecular level. A summary of all the peaks found can be seen in Table 1.

2.3. Thermogravimetric Analysis

The thermogravimetric analysis was performed using a TGA analyzer (Perkin-Elmer STA 6000, Waltham, MA, USA), which is shown in Figure 3a, and the differential thermogravimetry is shown in Figure 3b to assess the thermal stability of MMMs. The MMMs were cut with a sample weight of 10 mg and placed in an oven at 100 °C to get rid of the enduring solvents and moisture, and then were cooled to room temperature. The cooled samples were placed in a sample holder and the temperature was adapted from 30 °C to 800 °C under a N2 environment with a 10 °C/min heating rate. The TGA is better than other isothermal conventional methods for the evaluation of thermal decomposition; this is because very small amounts of the sample are enough for the investigation. All of the MMMs were found to be stable up to approximately 500 °C, which is an indication of quite high stability and the complex behavior of the TGA curves indicate the presence of multiple aromatic rings. The initial flatness of the curves is an indication of the removal of residual solvents, like THF, used in the synthesis process. The first 6% loss in the range of 110–230 °C leads to the amputation of the molecules. The further 4% decomposition between 230 °C to 500 °C represents decomposition of DES-ceria NPs, as the variations in the curves started with the change of composition which represents the increase of thermal stability with the increase of DES-ceria NPs filler loading. This improvement of thermal stability proves the strong interaction between DES-ceria NPs and the polymer matrix. The sudden weight loss after 500 °C demonstrates breakage of the structure of major aromatic portions of PSF [33]. The normal temperature of flue gases is not more than 100 °C [30,41,42], hence it is right to say that MMMs are suitable for the capturing of CO2 from flue gases under operating temperatures. All of these weight losses can also be observed in the DTG curves. These results are similar to other already reported studies [29,43,44,45].

2.4. Gas Permeation Analysis

The permeation of gases was evaluated using a custom-built system for analyzing the permeabilities and selectivities of mixed and pure gases. Permeate and retentate compositions were estimated with the help of a gas chromatograph (Agilent 7890, Santa Clara, CA, USA) that has twin thermal conductivity detectors. The overall scheme of the gas infiltration process has been reported elsewhere, as well as the construction and working of the setup [46]. Metallic support was utilized to place the membranes and seal them. The inlet gas mixture flow rate was 1 L/min. For membrane testing, the feed temperature and pressure of the system was kept at 25 °C and 10 bar. All of the measurements were taken a minimum of three times and average values were calculated. The solution-diffusion model was utilized for the transport mechanism via MMMs.
The permeabilities of CO2 and the selectivities (CO2/CH4 and CO2/N2) as a function of DES-ceria nanoparticle filler loadings are depicted in the Figure 4a,b. The variation in the permeabilities of different membrane compositions in comparison with pristine PSF membrane is termed as the facilitation ratio (FR). The FR is enhanced with the rise of filler from 2% to 4% DES-ceria NPs. The gas with smaller kinetic data (CO2 = 3.3Å) and a greater affinity with the amine and carboxyl groups present in DES exhibited an increased FR in comparison with gases of higher kinetic diameter (N2 = 3.64Å and CH4 = 3.8Å), which have a negligible affinity with the constituent functional groups of DES. This suggests that these results, within the same filler loading range, are due to the increment in pore density and the loss in pore size of membranes. The comparative study of the permeability of all three gases revealed the enhanced molecular sieving properties of DES-ceria NP-based MMMs that is the consequence of an overall increase in FR of CO2, N2 and CH4; this has a minor effect on the non-selective gases and gave better separation of the selective gas CO2. Hence, we conclude that the existence of DES-supported ceria NPs enhanced the molecular sieving properties of the membranes and improved the whole process of CO2 separation.
DES-ceria NP-based MMMs increased the CO2 permeability from 6.06 Barrer to 16.3 Barrer and the selectivity from 23.31 to 35.43 in the case of CO2/CH4 mixed gas. CO2 permeability increased from 6.36 Barrer to 16.9 Barrer and the selectivity from 25.44 to 39.3 in the case of CO2/N2 mixed gas. This is the FR of the gas permeation results. The difference between the results of the pure and mixed gases can be seen in Tables S1 and S2. This difference occurs because of the formation of the non-selective channels in MMMs. The incorporation of unambiguous functional groups was attempted to enhance CO2 separation from the mixed gases with MMMs, due to the affinity of the amine and carboxyl groups towards CO2. Hence, this combination was highly efficient at separation without any high selectivity loss.
The results obtained from the permeation of gas analysis were calibrated with the famous Robeson plots as shown in the Figure 5a,b. The Robeson plot is an explanation of the association between permeability and selectivity, and it demonstrates the trade-off between these; if the permeability increases, the selectivity decreased; and if permeability decreases, the selectivity increases. The Robeson lines are considered the ideal lines for researchers to approach [14,29,47]. It is apparent that the points are moving towards the Robeson lines with respect to an increase in the filler loadings and the trade-off for the M-4 membrane is less than M-0. Therefore, in comparison with the pure PSF membrane, the 8% DES-ceria NP-based membrane has a smaller trade-off and lies in the technically strong region of the plot. Hence, the synthesized DES-ceria NP-based MMMs can be considered as potential candidates to enhance the efficiency of CO2 separation from gaseous mixtures [48].

3. Materials and Methods

3.1. Materials

PSF of A.M.W ~22,000, cetrimonium bromide (CTMB) (Purity ≥ 98%), and acetic acid (Purity ≥ 99%) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Tetrahydrofuran (THF) (Purity = 99.6%), isopropyl alcohol (Purity ≥ 99.5%), ethylene glycol (Purity = 99%) and acetone (Purity ≥ 95%) for washing purposes were purchased from Fisher Scientific. Ammonium cerium (IV) nitrate (Purity = 99.5%) was obtained from Scharlau. A Water Purification Unit (Adrona SIA-B30, Riga, Latvia) is used to obtain Ultrapure water for washing purposes.

3.2. Deep Eutectic Solvent (DES) Preparation

The synthesis of DES was reported in our previous publication [33]. Briefly, DES was formed by combining equal masses of cetrimonium bromide (CTMB) and acetic acid at 70 °C for 3 h to obtain a homogeneous mixture. The synthesis was confirmed physically by analyzing the lowering of melting point as compared to the separate constituents and chemically by FTIR analysis. Figure 6 shows that CTMB was the hydrogen bond acceptor (HBA) and acetic acid was the hydrogen bond donor (HBD) [37].

3.3. Ceria Nanoparticles formation

The method of formation of the ceria nanoparticles has already been reported. Briefly, the uniform red-colored solution was formed by adding 0.3 g ammonium cerium (IV) nitrate into a mixture of 10 mL ethylene glycol and 10 mL isopropyl alcohol. The small amount of acetic acid (1 mL) was added to the above mixture dropwise with continuous stirring. The red solution was converted to a uniform colorless solution after 2 h of stirring. A Teflon-lined hydrothermal autoclave was used as a batch reactor and the resultant mixture was placed in an oven at 130 °C for 7 h. The yield was in the form of a greenish-brown mixture when cooled to room temperature. Ceria nanoparticles were obtained by centrifugation of the resultant mixture for 5 min at 8000 rpm. The solvents that were unreacted were recycled to avoid losses. The product was cleaned 3 times with ethanol and 1 time using acetone to achieve dried cerium oxide nanoparticles. After drying, 0.67 g of NPs was formed. In the recycling process, 0.3 g ammonium cerium (IV) nitrate was mixed with the unreacted solvents, then 0.7 mL acetic acid was added to the solution and the same procedure was repeated. The yield was 0.59 g in the first recycling process. In the second recycling process, the same procedure was repeated with 0.5 mL acetic acid and the yield was 0.50 g.

3.4. Surface Modification of Cerium Oxide Nanoparticles by DES

The DES-modified cerium oxide NPs were obtained by mixing the NPs with DES with a 8:1 ratio by weight into the small amount of evaporated organic solvent that is ethanol; the agglomeration was reduced using a mortar and pestle before using ethanol. The mixture was stirred to homogenize and placed at room temperature to evaporate the complete solvent for 24 h. The yield obtained after drying, that is the filler for MMMs, was named DES-ceria [33]. The work-flow of the surface tuning of ceria NPs is shown in Figure 7.

3.5. Mixed Matrix Membranes Fabrication

MMMs of various filler loadings were synthesized using DES-ceria nanoparticles as a filler with the help of a solution casting technique for dense membrane synthesis as shown in Table 2. PSF was mixed in THF for 2 h in a viol at room temperature with stirring of 500 rpm. After complete dissolution of the PSF, filler was added to the solution and mixed for 24 h under the same conditions to form a uniform composition. The air bubbles were removed using the gravity method and the solution was placed into a flat-bottom petri dish and left for 24 h to dry. After complete solvent evaporation, the membrane was peeled off from the bottom. The whole process of MMM synthesis is shown in Figure 8.

3.6. Characterization of Membrane Samples

The Tescan Vega LMU variable pressure Scanning Electron Microscope (Brno-Kohoutovice Czech Republic) was utilized to observe the cross-sectional and surface morphologies of the membrane samples. The Thermo Scientific Nicolet iS5 instrument from the United States was used to conduct the FTIR (Fourier Transform Infrared) spectroscopic analysis of membrane samples. A TGA analyzer (Perkin-Elmer STA 6000, Waltham, MA, USA), was utilized to examine the thermal stability of membrane samples. The composition of gases was evaluated with the help of a gas chromatograph (Agilent 7890, Santa Clara, CA, USA).

4. Conclusions

Ceria NPs were synthesized and combined with DES to fabricate a novel filler. MMMs of DES-ceria NPs were developed using a solvent evaporation technique. The SEM, FTIR and TGA of the membranes were performed. The SEM micrograph revealed the cross-sectional and surface morphologies of the synthesized MMMs, which described that the DES-ceria NPs were uniformly distributed over the polymer matrix and also making agglomerates because of the magnetic effects among the particles. The FTIR spectra indicated characteristic peaks of all of the constituents that suggests that no chemical reaction took place in the preparation of the membranes as all of the characteristic peaks can be easily indicated with respect to their functional groups. The TGA provided the comparison of thermal stabilities of the membrane and that the membrane with the highest filler loading was the most stable membrane which highlights the thermal stability of DES-ceria NPs, and that the overall trend for thermal stability was ascending from 0% to 8% DES-ceria NPs. The gas permeation analysis was carried out using pure and mixed gases and highlighted the enhanced permeability of CO2 and the selectivity as compared to a pristine PSF membrane. The improvement in the permeability of CO2 in pure and mixed gases (CO2/CH4 and CO2/N2) was from 6.7 Barrer to 17.2 Barrer, from 6.06 Barrer to 16.3 Barrer and from 6.36 Barrer to 16.9 Barrer, respectively. The improvement in the selectivities of the pure gases was from 24.81 to 41.95 (CO2/CH4) and from 26.8 to 40.95 (CO2/N2), and for the mixed gases was from 23.31 to 35.43 (CO2/CH4) and from 25.44 to 39.3 (CO2/N2). Thus, the DES-ceria NP-based MMMs enhanced the permeability as well as the selectivity and proved to be a promising filler for the mitigation of CO2 from a gaseous mixture. In future, mathematical modeling can also be applied by obtaining statistical data for different parameters relevant to the presented work, and also further modification of the nanoparticles by tunable DESs can open up a way for CO2 separation via MMMs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28207162/s1, Figure S1: Surface images of synthesized membranes; Table S1: The pure gas permeability and ideal gas selectivity of DES-ceria NPs based MMMs; Table S2: The mixed gas permeabilities and selectivities of DES-ceria NPs based MMMs.

Author Contributions

S.-u.-R.: Methodology, Writing—Original draft preparation. M.S.M.: Data curation, Supervision. M.J.: Review and Editing. S.U.Z.: Review and Editing. S.R.: Visualization, Investigation. N.M.: Conceptualization. J.b.F.: Validation, Software, M.A.: paper format preparation. M.F.-e.-A.: Review and Editing. J.R.: Review and Editing. M.U.: Review and Editing. R.A.A.: Funding acquisition. S.A.S.: Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Researchers Supporting Project Number (RSP2023R265), King Saud University, Riyadh, Saudi Arabia, and also supported by the International Islamic University Malaysia Research Management Centre Grant (RMCG20-038-0038).

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

The data will be available on request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not Applicable.

References

  1. Chuah, C.Y.; Goh, K.; Yang, Y.; Gong, H.; Li, W.; Karahan, H.E.; Guiver, M.D.; Wang, R.; Bae, T.-H. Harnessing filler materials for enhancing biogas separation membranes. Chem. Rev. 2018, 118, 8655–8769. [Google Scholar] [CrossRef]
  2. Hua, Y.; Wang, H.; Li, Q.; Chen, G.; Liu, G.; Duan, J.; Jin, W. Highly efficient CH4 purification by LaBTB PCP-based mixed matrix membranes. J. Mater. Chem. A 2018, 6, 599–606. [Google Scholar] [CrossRef]
  3. Norahim, N.; Yaisanga, P.; Faungnawakij, K.; Charinpanitkul, T.; Klaysom, C. Recent membrane developments for CO2 separation and capture. Chem. Eng. Technol. 2018, 41, 211–223. [Google Scholar] [CrossRef]
  4. Xie, K.; Fu, Q.; Qiao, G.G.; Webley, P.A. Recent progress on fabrication methods of polymeric thin film gas separation membranes for CO2 capture. J. Membr. Sci. 2019, 572, 38–60. [Google Scholar] [CrossRef]
  5. Olajire, A.A. CO2 capture and separation technologies for end-of-pipe applications—A review. Energy 2010, 35, 2610–2628. [Google Scholar] [CrossRef]
  6. Bernardo, P.; Drioli, E.; Golemme, G. Membrane gas separation: A review/state of the art. Ind. Eng. Chem. Res. 2009, 48, 4638–4663. [Google Scholar] [CrossRef]
  7. Koros, W.J.; Zhang, C. Materials for next-generation molecularly selective synthetic membranes. Nat. Mater. 2017, 16, 289–297. [Google Scholar] [CrossRef]
  8. Liu, G.; Jin, W.; Xu, N. Two-dimensional-material membranes: A new family of high-performance separation membranes. Angew. Chem. Int. Ed. 2016, 55, 13384–13397. [Google Scholar] [CrossRef]
  9. Hou, J.; Zhang, H.; Simon, G.P.; Wang, H. Polycrystalline advanced microporous framework membranes for efficient separation of small molecules and ions. Adv. Mater. 2020, 32, e1902009. [Google Scholar] [CrossRef]
  10. Kosinov, N.; Gascon, J.; Kapteijn, F.; Hensen, E.J. Recent developments in zeolite membranes for gas separation. J. Membr. Sci. 2016, 499, 65–79. [Google Scholar] [CrossRef]
  11. Anderson, M.; Wang, H.; Lin, Y. Inorganic membranes for carbon dioxide and nitrogen separation. Rev. Chem. Eng. 2012, 28, 101–121. [Google Scholar] [CrossRef]
  12. Wang, S.; Li, X.; Wu, H.; Tian, Z.; Xin, Q.; He, G.; Peng, D.; Chen, S.; Yin, Y.; Jiang, Z.; et al. Advances in high permeability polymer-based membrane materials for CO2 separations. Energy Environ. Sci. 2016, 9, 1863–1890. [Google Scholar] [CrossRef]
  13. Galizia, M.; Chi, W.S.; Smith, Z.P.; Merkel, T.C.; Baker, R.W.; Freeman, B.D. 50th anniversary perspective: Polymers and mixed matrix membranes for gas and vapor separation: A review and prospective opportunities. Macromolecules 2017, 50, 7809–7843. [Google Scholar] [CrossRef]
  14. Robeson, L.M. The upper bound revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
  15. Dechnik, J.; Gascon, J.; Doonan, C.J.; Janiak, C.; Sumby, C. Mixed-matrix membranes. Angew. Int. Ed. 2017, 56, 9292–9310. [Google Scholar] [CrossRef] [PubMed]
  16. Cheng, Y.; Wang, Z.; Zhao, D. Mixed matrix membranes for natural gas upgrading: Current status and opportunities. Ind. Eng. Chem. Res. 2018, 57, 4139–4169. [Google Scholar] [CrossRef]
  17. Dorosti, F.; Omidkhah, M.; Abedini, R. Fabrication and characterization of Matrimid/MIL-53 mixed matrix membrane for CO2/CH4 separation. Chem. Eng. Res. Des. 2014, 92, 2439–2448. [Google Scholar] [CrossRef]
  18. Sadeghi, Z.; Omidkhah, M.; Masoumi, M.E.; Abedini, R. Modification of existing permeation models of mixed matrix membranes filled with porous particles for gas separation. Can. J. Chem. Eng. 2016, 94, 547–555. [Google Scholar] [CrossRef]
  19. Jamshidi, M.; Pirouzfar, V.; Abedini, R.; Pedram, M.Z. The influence of nanoparticles on gas transport properties of mixed matrix membranes: An experimental investigation and modeling. Korean J. Chem. Eng. 2017, 34, 829–843. [Google Scholar] [CrossRef]
  20. Zhang, S.; Zhang, J.; Zhang, Y.; Deng, Y. Nanoconfined ionic liquids. Chem. Rev. 2017, 117, 6755–6833. [Google Scholar] [CrossRef]
  21. Liu, Y.; Wu, H.; Wu, S.; Song, S.; Guo, Z.; Ren, Y.; Zhao, R.; Yang, L.; Wu, Y.; Jiang, Z. Multifunctional covalent organic framework (COF)-Based mixed matrix membranes for enhanced CO2 separation. J. Membr. Sci. 2021, 618, 118693. [Google Scholar] [CrossRef]
  22. Wang, M.; Wang, Z.; Zhao, S.; Wang, J.; Wang, S. Recent advances on mixed matrix membranes for CO2 separation. Chin. J. Chem. Eng. 2017, 25, 1581–1597. [Google Scholar] [CrossRef]
  23. Kim, S.; Lee, Y.M. High performance polymer membranes for CO2 separation. Curr. Opin. Chem. Eng. 2013, 2, 238–244. [Google Scholar] [CrossRef]
  24. Aroon, M.; Ismail, A.; Matsuura, T.; Montazer-Rahmati, M. Performance studies of mixed matrix membranes for gas separation: A review. Sep. Purif. Technol. 2010, 75, 229–242. [Google Scholar] [CrossRef]
  25. Hasebe, S.; Aoyama, S.; Tanaka, M.; Kawakami, H. CO2 separation of polymer membranes containing silica nanoparticles with gas permeable nano-space. J. Membr. Sci. 2017, 536, 148–155. [Google Scholar] [CrossRef]
  26. Raouf, M.; Abedini, R.; Omidkhah, M.; Nezhadmoghadam, E. A favored CO2 separation over light gases using mixed matrix membrane comprising polysulfone/polyethylene glycol and graphene hydroxyl nanoparticles. Process. Saf. Environ. Prot. 2020, 133, 394–407. [Google Scholar] [CrossRef]
  27. Sainath, K.; Modi, A.; Bellare, J. In-situ growth of zeolitic imidazolate framework-67 nanoparticles on polysulfone/graphene oxide hollow fiber membranes enhance CO2/CH4 separation. J. Membr. Sci. 2020, 614, 118506. [Google Scholar] [CrossRef]
  28. Ruhaimi, A.H.; Ab Aziz, M.A. Spherical CeO2 nanoparticles prepared using an egg-shell membrane as a bio-template for high CO2 adsorption. Chem. Phys. Lett. 2021, 779, 138842. [Google Scholar] [CrossRef]
  29. Farashi, Z.; Azizi, S.; Arzhandi, M.R.D.; Noroozi, Z.; Azizi, N. Improving CO2/CH4 separation efficiency of Pebax-1657 membrane by adding Al2O3 nanoparticles in its matrix. J. Nat. Gas Sci. Eng. 2019, 72, 103019. [Google Scholar] [CrossRef]
  30. Xu, R.; Wang, Z.; Wang, M.; Qiao, Z.; Wang, J. High nanoparticles loadings mixed matrix membranes via chemical bridging-crosslinking for CO2 separation. J. Membr. Sci. 2019, 573, 455–464. [Google Scholar] [CrossRef]
  31. Taghizadeh, M.; Taghizadeh, A.; Vatanpour, V.; Ganjali, M.R.; Saeb, M.R. Deep eutectic solvents in membrane science and technology: Fundamental, preparation, application, and future perspective. Sep. Purif. Technol. 2020, 258, 118015. [Google Scholar] [CrossRef]
  32. Jablonský, M.; Škulcová, A.; Šima, J. Use of deep eutectic solvents in polymer chemistry–A review. Molecules 2019, 24, 3978. [Google Scholar] [CrossRef]
  33. Reed, K.; Cormack, A.; Kulkarni, A.; Mayton, M.; Sayle, D.; Klaessig, F.; Stadler, B. Exploring the properties and applications of nanoceria: Is there still plenty of room at the bottom? Environ. Sci. Nano 2014, 1, 390–405. [Google Scholar] [CrossRef]
  34. Feng, Y.; Han, G.; Chung, T.-S.; Weber, M.; Widjojo, N.; Maletzko, C. Effects of polyethylene glycol on membrane formation and properties of hydrophilic sulfonated polyphenylenesulfone (sPPSU) membranes. J. Membr. Sci. 2017, 531, 27–35. [Google Scholar] [CrossRef]
  35. Elahi, B.; Mirzaee, M.; Darroudi, M.; Oskuee, R.K.; Sadri, K.; Amiri, M.S. Preparation of cerium oxide nanoparticles in Salvia MacrosiphonBoiss seeds extract and investigation of their photo-catalytic activities. Ceram. Int. 2019, 45, 4790–4797. [Google Scholar] [CrossRef]
  36. Sreekanth, T.; Dillip, G.; Lee, Y.R. Picrasmaquassioides mediated cerium oxide nanostructures and their post-annealing treatment on the microstructural, morphological and enhanced catalytic performance. Ceram. Int. 2016, 42, 6610–6618. [Google Scholar] [CrossRef]
  37. Zhang, Q.; De Oliveira Vigier, K.; Royer, S.; Jérôme, F. Deep eutectic solvents: Syntheses, properties and applications. Chem. Soc. Rev. 2012, 41, 7108–7146. [Google Scholar] [CrossRef]
  38. Shahbaz, K.; Mjalli, F.S.; Hashim, M.A.; AlNashef, I.M. Using deep eutectic solvents based on methyl triphenyl phosphunium bromide for the removal of glycerol from palm-oil-based biodiesel. Energy Fuels 2011, 25, 2671–2678. [Google Scholar] [CrossRef]
  39. Durand, E.; Lecomte, J.; Villeneuve, P. Deep eutectic solvents: Synthesis, application, and focus on lipase-catalyzed reactions. Eur. J. Lipid Sci. Technol. 2013, 115, 379–385. [Google Scholar] [CrossRef]
  40. Lian, S.; Li, R.; Zhang, Z.; Liu, Q.; Song, C.; Lu, S. Improved CO2 separation performance and interfacial affinity of composite membranes by incorporating amino acid-based deep eutectic solvents. Sep. Purif. Technol. 2021, 272, 118953. [Google Scholar] [CrossRef]
  41. Muduli, S.K.; Wang, S.; Chen, S.; Ng, C.F.; Huan, C.H.A.; Sum, T.C.; Soo, H.S. Mesoporous cerium oxide nanospheres for the visible-light driven photocatalytic degradation of dyes. Beilstein J. Nanotechnol. 2014, 5, 517–523. [Google Scholar] [CrossRef]
  42. Yang, L.; Zhang, S.; Wu, H.; Ye, C.; Liang, X.; Wang, S.; Wu, X.; Wu, Y.; Ren, Y.; Liu, Y.; et al. Porous organosilicon nanotubes in pebax-based mixed-matrix membranes for biogas purification. J. Membr. Sci. 2019, 573, 301–308. [Google Scholar] [CrossRef]
  43. Singh, K.; Devi, S.; Bajaj, H.C.; Ingole, P.; Choudhari, J.; Bhrambhatt, H. Optical resolution of racemic mixtures of amino acids through nanofiltration membrane process. Sep. Sci. Technol. 2014, 49, 2630–2641. [Google Scholar] [CrossRef]
  44. Shan, M.; Seoane, B.; Andres-Garcia, E.; Kapteijn, F.; Gascon, J. Mixed-matrix membranes containing an azine-linked covalent organic framework: Influence of the polymeric matrix on post-combustion CO2-capture. J. Membr. Sci. 2018, 549, 377–384. [Google Scholar] [CrossRef]
  45. Chen, Y.; Ho, W.W. High-molecular-weight polyvinylamine/piperazine glycinate membranes for CO2 capture from flue gas. J. Membr. Sci. 2016, 514, 376–384. [Google Scholar] [CrossRef]
  46. Khan, A.L.; Basu, S.; Cano-Odena, A.; Vankelecom, I.F. Novel high throughput equipment for membrane-based gas separations. J. Membr. Sci. 2010, 354, 32–39. [Google Scholar] [CrossRef]
  47. Robeson, L.M. Correlation of separation factor versus permeability for polymeric membranes. J. Membr. Sci. 1991, 62, 165–185. [Google Scholar] [CrossRef]
  48. Vakharia, V.; Salim, W.; Wu, D.; Han, Y.; Chen, Y.; Zhao, L.; Ho, W.W. Scale-up of amine-containing thin-film composite membranes for CO2 capture from flue gas. J. Membr. Sci. 2018, 555, 379–387. [Google Scholar] [CrossRef]
Figure 1. SEM images of cross-sections of MMMs of different filler loadings (A) 2% (B) 4% (C) 6% (D) 8%.
Figure 1. SEM images of cross-sections of MMMs of different filler loadings (A) 2% (B) 4% (C) 6% (D) 8%.
Molecules 28 07162 g001
Figure 2. FTIR spectra for the ceria, DES-ceria and membrane compositions.
Figure 2. FTIR spectra for the ceria, DES-ceria and membrane compositions.
Molecules 28 07162 g002
Figure 3. (a) Thermogravimetric analysis of membrane compositions. (b) Differential thermogravimetric analysis of membranes.
Figure 3. (a) Thermogravimetric analysis of membrane compositions. (b) Differential thermogravimetric analysis of membranes.
Molecules 28 07162 g003aMolecules 28 07162 g003b
Figure 4. (a) Permeability and selectivity of MMMs for mixed gases CO2/CH4. (b) Permeability and selectivity of MMMs for mixed gases CO2/N2.
Figure 4. (a) Permeability and selectivity of MMMs for mixed gases CO2/CH4. (b) Permeability and selectivity of MMMs for mixed gases CO2/N2.
Molecules 28 07162 g004
Figure 5. (a) Robeson plot comparison of MMMs for mixed gas CO2/CH4. (b) Robeson plot comparison of MMMs for mixed gas CO2/N2.
Figure 5. (a) Robeson plot comparison of MMMs for mixed gas CO2/CH4. (b) Robeson plot comparison of MMMs for mixed gas CO2/N2.
Molecules 28 07162 g005aMolecules 28 07162 g005b
Figure 6. Synthesis of DES and H-bonding between CTMB and acetic acid [33].
Figure 6. Synthesis of DES and H-bonding between CTMB and acetic acid [33].
Molecules 28 07162 g006
Figure 7. Surface tuning of ceria nanoparticles using CTMB-based deep eutectic solvent [33].
Figure 7. Surface tuning of ceria nanoparticles using CTMB-based deep eutectic solvent [33].
Molecules 28 07162 g007
Figure 8. Schematic diagram for synthesis of mixed matrix dense membrane [33].
Figure 8. Schematic diagram for synthesis of mixed matrix dense membrane [33].
Molecules 28 07162 g008
Table 1. Summary of FTIR spectra.
Table 1. Summary of FTIR spectra.
Sr No.Wavenumber (cm−1)Functional Groups
1829Benzene ring bending
21010C-H stretching in benzene ring
31071O-Ce-O stretching vibration
41101Symmetric elongating vibration of O=S=O
51146Asymmetric stretching vibration of O=S=O
61168Bending vibration of C-O
71233Elongating vibration of -C6H4-O-C6H4-
81291, 1321Symmetric stretching vibration of O=S=O
91483Bending vibration of C-H
101503, 1582Stretching mode of C=C in aromatics
111740Symmetric stretching of C=O
Table 2. MMMs composition by weight percentage.
Table 2. MMMs composition by weight percentage.
Membrane IdMembrane TypePSFCeriaTHF
M-0PSF4.8-95.2
M-12% ceria membrane4.80.195.1
M-24% ceria membrane4.80.295.0
M-36% ceria membrane4.80.394.9
M-48% ceria membrane4.80.494.8
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

Saif-ur-Rehman; Shozab Mehdi, M.; Fakhar-e-Alam, M.; Asif, M.; Rehman, J.; A. Alshgari, R.; Jamal, M.; Uz Zaman, S.; Umar, M.; Rafiq, S.; et al. Deep Eutectic Solvent Coated Cerium Oxide Nanoparticles Based Polysulfone Membrane to Mitigate Environmental Toxicology. Molecules 2023, 28, 7162. https://doi.org/10.3390/molecules28207162

AMA Style

Saif-ur-Rehman, Shozab Mehdi M, Fakhar-e-Alam M, Asif M, Rehman J, A. Alshgari R, Jamal M, Uz Zaman S, Umar M, Rafiq S, et al. Deep Eutectic Solvent Coated Cerium Oxide Nanoparticles Based Polysulfone Membrane to Mitigate Environmental Toxicology. Molecules. 2023; 28(20):7162. https://doi.org/10.3390/molecules28207162

Chicago/Turabian Style

Saif-ur-Rehman, Muhammad Shozab Mehdi, Muhammad Fakhar-e-Alam, Muhammad Asif, Javed Rehman, Razan A. Alshgari, Muddasar Jamal, Shafiq Uz Zaman, Muhammad Umar, Sikander Rafiq, and et al. 2023. "Deep Eutectic Solvent Coated Cerium Oxide Nanoparticles Based Polysulfone Membrane to Mitigate Environmental Toxicology" Molecules 28, no. 20: 7162. https://doi.org/10.3390/molecules28207162

Article Metrics

Back to TopTop