Next Article in Journal
Developing Enzyme Immobilization with Fibrous Membranes: Longevity and Characterization Considerations
Next Article in Special Issue
Using the Log Mean Temperature Difference (LMTD) and ε-NTU Methods to Analyze Heat and Mass Transfer in Direct Contact Membrane Distillation
Previous Article in Journal
Enhanced Anti-Biofouling Properties of BWRO Membranes via the Deposition of Poly (Catechol/Polyamine) and Ag Nanoparticles
Previous Article in Special Issue
Versatile Silver-Nanoparticle-Impregnated Membranes for Water Treatment: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption Performance of Heavy Metal Ions under Multifactorial Conditions by Synthesized Organic-Inorganic Hybrid Membranes

1
Shanghai Civil Aviation College, 1 Longhua West Road, Shanghai 200232, China
2
State Key Laboratory of Molecular Engineering of Polymers, Department of Macromolecular Science, Fudan University, 2005 Songhu Road, Shanghai 200438, China
3
School of Pharmacy, Fudan University, Shanghai 201203, China
*
Author to whom correspondence should be addressed.
Membranes 2023, 13(5), 531; https://doi.org/10.3390/membranes13050531
Submission received: 30 March 2023 / Revised: 1 May 2023 / Accepted: 17 May 2023 / Published: 19 May 2023
(This article belongs to the Special Issue Membrane-Based Technologies for Water/Wastewater Treatment)

Abstract

:
A series of hybridized charged membrane materials containing carboxyl and silyl groups were prepared via the epoxy ring-opening reaction and sol–gel methods using 3-glycidoxypropyltrimethoxysilane (WD-60) and polyethylene glycol 6000 (PEG-6000) as raw materials and DMF as a solvent. Scanning electron microscopy (SEM), fourier transform infrared spectroscopy (FTIR), and thermal gravimetric analyzer/differential scanning calorimetry (TGA/DSC) analysis showed that the heat resistance of the polymerized materials could reach over 300 °C after hybridization. A comparison of the results of heavy metal lead and copper ions’ adsorption tests on the materials at different times, temperatures, pHs, and concentrations showed that the hybridized membrane materials have good adsorption effects on heavy metals and better adsorption effects on lead ions. The maximum capacity obtained from optimized conditions for Cu2+ and Pb2+ ions were 0.331 and 5.012 mmol/g. The experiments proved that this material is indeed a new environmentally friendly, energy-saving, high-efficiency material. Moreover, their adsorptions for Cu2+ and Pb2+ ions will be evaluated as a model for the separation and recovery of heavy metal ions from wastewater.

1. Introduction

With the rapid development of the global economy, mankind’s best efforts to exploit the Earth’s environmental resources have also made our living environment increasingly harsh [1,2,3]. Heavy metal cations and certain toxic anions can cause serious pollution and hazards in the water environment, affecting fisheries and agriculture directly, and these ions are either directly or indirectly harmful to human health. For example, a single dose of 60–l00 mg/kg body weight of lead will stimulate the digestive system, causing abdominal pain, nausea, vomiting, and gastroenteritis, and a long-term overdose can lead to liver cirrhosis. Membrane separation is a new separation technology that emerged in the early 20th century and became increasingly popular after the 1960s. Membrane separation technology has been widely used in food, medicine, biology, environmental protection, the chemical industry, metallurgy, energy, petroleum, water treatment [4], electronics, bionics, and other fields, has produced huge economic and social benefits, and has become one of the most important methods in separation science today. Nowadays, membrane separation technology is widely used in water treatment [5,6] and the chemical, biological, and pharmaceutical fields at an industrial scale, showing its broad application prospects. Polyethylene glycol (PEG) is an excellent biodegradable and water-soluble biopolymer. Due to its environmentally friendly properties, PEG and its derivatives have been used extensively in the manufacture of PEG-based polymeric membranes or materials for various industrial applications. Although these membranes show some favorable properties, such as structural flexibility and excellent film-forming properties, they also show disadvantages, such as low mechanical strength and thermal stability. Therefore, improving their mechanical strength and thermal stability has turned out to be a key consideration in membrane preparation [7,8,9]. First of all, membrane separation is an efficient process that can achieve high purity. In addition, the membrane does not undergo phase change during the separation process, so the energy consumption is low. Secondly, membrane separation is usually carried out at room temperature, which is particularly suitable for handling heat-sensitive materials, and the separation equipment itself has no moving parts, making it highly reliable and easy to operate and maintain, which has very important theoretical and practical significance for the study of membrane separation technology to treat and recover heavy metal ions in order to reduce hazards. The performance of the membrane material is one of the most critical factors affecting the separation effect, so improving or enhancing the performance of the membrane material has become one of the core issues of membrane separation technology research.
Organic-inorganic hybrid materials are a new field of composite materials that combine many excellent properties of organic and inorganic materials, such as the low density, good viscoelasticity, high toughness, and good processability of organic materials, and the high hardness, high modulus of elasticity, high strength, better light transmission, and high refractive index of inorganic materials [10,11,12,13]. Organic-inorganic hybrid membrane materials combine the excellent properties of both organic and inorganic components, and their rapid development provides new ideas for the design and development of membrane materials [14,15,16,17,18]. In 1985, Schmidt H successfully prepared a class of amorphous solids using organically modified alkoxysilanes R′nSi(OR)4−n (n = 1–3, R is alkyl and R′ is an organic short chain) as precursors via a sol-gel reaction [19]. Wilkes G. L. et al. condensed low-molecular-weight (500–1700) polydimethylsiloxanes with hydroxyl groups at the end groups (PDMS) and tetraethoxysilane (TEOS) in a sol-gel reaction under acidic conditions. During the sol-gel reaction, the Si-OH groups at the chain end of PDMS and TEOS undergo condensation to form Si-O-Si bonds so that the PDMS and inorganic SiO2 components are bonded together through covalent bonds, thus successfully preparing a PDMS-SiO2 system of organic-inorganic hybrid materials. The inorganic SiO2 network in these hybrids is bonded to polymer molecular chains, which Wilkes et al. named “ceramic polymers (creamers)” [20]. Following Schmidt and Wilkes, a number of researchers have explored organic-inorganic hybrids and have prepared them by various methods (e.g., sol-gel method [21], partial pyrolysis of organic polymers [22], nanotechnology copolymerization [23], chemical vapor deposition (CVD), etc.) [24,25,26]. To produce materials containing a variety of organic components, such as alkyl, phenyl, polystyrene, epoxy, polyether, and poly(ethylene glycol), epoxy resins, polyethers, polyimides, aminopropyl, aminophenyl, polymethyl methacrylate, and inorganic components, such as SiO2, Al2O3, and TiO2 have been combined with each other, and reviews of these materials have been published [27,28].
In recent years, sol-gel technology has developed rapidly in the preparation of organic-inorganic hybrid materials, and many papers have summarized it. [29]. The principle of the sol-gel method for the preparation of hybrid materials involves using inorganic compounds of metals as precursors, which are hydrolytically depolymerized, dehydrated and condensed to form inorganic networks [30], and adding organic substances or organic monomers with a common solvent to the inorganic substances to the inorganic sol to form inorganic-organic hybrid materials via condensation gelation. In recent years, many researchers have prepared a series of inorganic-organic hybrid membranes using inorganic and organic hybrid materials [31], which can not only significantly improve the thermal stability and mechanical strength of the materials and membranes, but can also use electrostatic effects to separate and adsorb heavy metal ions in industrial waste; thus, they have broad application prospects in wastewater treatment and environmental protection.
In this paper, a novel negatively charged siloxane hybrid membrane material is prepared by epoxy ring opening and sol-gel reactions to achieve the efficient and optimized separation of heavy metals at the molecular level. For this study, we chose Pb2+ and Cu2+ ions as typical adsorbed species because of their water polluting properties. The aim of this study is to optimize the molar ratio of organic and inorganic components, which is the most important factor affecting the degree of thermal stability and the separation effect. On this basis, the membrane material’s heavy metal ion adsorption properties were then investigated. The best adsorption pH, adsorption temperature, Cu2+/Pb2+ concentration, and adsorption time were identified, respectively, via the single variable method.

2. Materials and Methods

2.1. Materials

Silane coupler WD-60 (3-Glycidoxypropyltrimethoxysilane, 99%) was purchased from WD Silicone New Material Corporation (Hubei, China). Polyethylene glycol 6000 (PEG-6000), N,N-Dimethylformamide (DMF, 99.8%), maleic anhydride (99%), Cu(NO3)2, KI solution (10%, w/w), K2Cr2O7 (reference substance, dried at 140 °C for 2 h), Na2S2O3·5H2O solid, Na2CO3 solid, sulfuric acid solution (0.1 mol/L), hydrochloric acid solution (0.1 mol/L), nitric acid (0.1 mol/L), glacial acetic acid, sodium acetate, starch solution (0.5%, v/v), Pb(NO3)2, disodium ethylenediaminetetra acetate (EDTA), dimethoate orange indicator (0.2%, w/w), and methyltetramine buffer solution (20%, w/w) were purchased from Sinopharm Group Chemical Reagent Corporation (Shanghai, China). All chemicals were used as received unless otherwise stated.

2.2. Preparation of Hybrid Membranes

This class of hybrid membrane materials was prepared using N,N-dimethylformamide (DMF) as the solvent, and charged hybrid membrane materials containing acidic groups were prepared via the epoxy ring opening reaction and sol-gel methods. Four different ratios of materials were designed as presented in Table 1.
We started the DF-101S magnetic stirrer with a stirrer (900 rpm), thermometer, and heating device, increased its temperature to 80 °C, added 6.00 g of polyethylene glycol 6000 (PEG-6000) to the beaker, and then added 20 mL of DMF. We placed the beaker in the magnetic stirrer with heat collection in a water bath to dissolve it and then waited for it to cool for a while before using it. After adding 1.00 g of maleic acid and dissolving it in a water bath, we added an equal amount of WD-60 and heated the mixture in a DF-101S magnetic stirrer for 8 h. After the reaction was complete, the mixture was extracted and cooled to room temperature (27 °C). Then, we used a glass rod to coat it evenly on a Teflon board. We made sure to remove any air bubbles from the solution before coating, and we pumped it with a vacuum pump if required. We left it for 2–3 days at room temperature. After the material had been produced, we baked it for one day in a constant-temperature oven at 35 °C initially, then raised the temperature by 5 °C every other day. At 50 °C, the material was fully dry (usually 5–6 days). The samples were put in self-sealing bags and stored for testing after they had dried completely.

2.3. Membrane Characterization

To examine the microstructures, surface and cross-sectional morphology of the membranes, SEM (400 FEG, FEI) was performed. FTIR spectra were recorded on a Shimadzu FTIR-8400S spectrophotometer to determine the chemical compositions of the membranes. Solid samples were subjected to the KBr tablet method. All infrared spectra were accumulated 32 times with a 4 cm−1 resolution. The samples were analyzed thermally using a Netzsch STA 409 PC/PG integrated TGA/DSC (differential scanning calorimetry) thermal analyzer protected by nitrogen with a ramp-up rate of 10 °C/min; the temperature was increased from room temperature to 500 °C.

2.4. Adsorption Experiment

The adsorption experiment of organic-inorganic hybrid membranes for Cu2+/Pb2+ ions was conducted in a similar way as described in the previous articles [32,33,34], in which Cu2+/Pb2+ ions were used as the adsorption medium [35]. The detailed test procedure can be described in the Supplementary Materials (File S1). The adsorption capacity (qCu2+) of Cu2+ ions (mmol/g) is calculated as follows (Equation (1)):
q Cu 2 + = ( C 0 C R ) V W
where V is the volume of aqueous Cu(NO3)2 solution (mL), C0 and CR are the concentrations of initial and remaining Cu(NO3)2, respectively (mol/L), and W is the weight of hybrid membrane (g). The adsorption capacity (qPb2+) of Pb2+ ions (mmol/g) can also be calculated by Equation (1), where V should be changed to the volume of aqueous Pb(NO3)2 solution (mL), C0 and CR are the concentrations of initial and remaining Pb(NO3)2, respectively (mol/L), and W is the weight of hybrid membrane (g) (File S2).
The effect of adsorption pH, adsorption temperature, Cu2+/Pb2+ concentration, and adsorption time were performed to determine optimal conditions. The macroscopic morphology of the prepared hybrid membrane materials before and after adsorption can be visualized in Figure 1.

3. Results and Discussion

3.1. Characterization of Hybrid Membranes

The microscopic surface structure and cross-sectional morphology of the organic-inorganic hybrid membrane were clearly characterized in scanning electron microscopy (SEM) micrographs. As shown in Figure 2, it is obvious to find that the surface of the hybrid membrane (sample A) exhibits a larger pore-size structure [36].
PEG/Si-O-Si organic-inorganic hybrid materials were synthesized via an epoxy ring-opening reaction, as shown in Figure 3a. Figure 3b illustrates the FTIR spectra of the hybrid membrane materials, which all exhibited Si-O-Si stretching vibration absorption peaks at 1107 cm1; in the range of 2700–2900 cm1, there were larger broad peaks, primarily methylene and methylene absorption peaks, demonstrating that the carbon chain was not broken in the reaction; 2900 cm1 was the asymmetric stretching vibration peak of –CH2; 3437 cm1 was the –OH stretching and bending vibration peak; and the characteristic absorption peak of –COOH appeared at 1718 cm1, and the peak increased as the ratio increased, indicating that the content of –COOH also increased [19,20,28]. In addition, the spectra of the other three ratios showed an increase with the amount of WD-60. The characteristic absorption peak for epoxy was near 962 cm1, showing that there were epoxy group residues in the system, but the peak area was tiny, indicating that there were fewer epoxy residues in the system.
Figure 4a shows TGA curves with different ratios of silica-based hybrid membranes. As can be seen from the figure, there were a total of three weight loss processes with an onset temperature of 44 °C. The first weight loss process occurred between 44 °C and 300 °C, with a mass reduction of approximately 10% of the total initial mass of the specimen. Significant weight loss occurred between 300 and 440 °C, with a mass reduction of approximately 50% of the initial total mass of the specimen, indicating that the thermal stability of the hybrid material could reach above 300 °C. With an increase in inorganic WD-60, the thermal stability of the hybrid material increased. The third, which occurred between 440 °C and 500 °C, was a plateau process with essentially constant mass, where an increase in the inorganic WD-60 meant that the amount of inorganic material remaining in the hybrid material also increased. Meanwhile, the weight loss beyond 330 °C can be ascribed to the further degradation of organic groups and the production of crystallized silica [35]. It can be seen from Figure 2a that the thermal stability of the hybridized membrane material is best when the ratio is 6:1; the weight loss around 330 °C was 7.4%. This sample, D, demonstrates a relatively higher compatibility between the inorganic and organic ingredients [35].
The DSC curves for the various ratios of the hybridized membrane materials are shown in Figure 4b. The glass transition temperature of the hybridized material, where the absorbing group exerted its heat, was around 63 °C, where a downward peak can be seen. The peak shifted to the right as the ratio rose at about 380 °C, showing that the stability of the material rises in step with the ratio.

3.2. Absorption Properties of Cu2+ by Hybrid Membrane Materials

Figure 5a demonstrates that there was no significant change of Cu2+ adsorption effect. As the amount of WD-60 increased, the adsorption effect slightly increased, and the qCu2+ was about 0.33 mmol/g, which was three times higher than in previous studies (0.075 mmol/g) in Table 2. The Cu2+ adsorption schemes of the hybrid membranes (A, B, C, and D) under multifactorial conditions are shown in Figure 5b.
The constant temperature vacuum drying oven was set at 30, 35, 40, 45, and 50 °C. The adsorption capacity of the hybrid membrane material for Cu2+ ions fluctuated with temperature, as can be seen in Figure 6a, with the best adsorption effect occurring at 45 °C. The optimal temperature for Cu2+ adsorption by the hybrid membrane material was therefore determined to be 45 °C. The same conclusion can be found in the previous articles for the zwitterionic hybrid adsorbents [32].
Different concentrations of 0.005 mol/L, 0.01 mol/L, 0.03 mol/L, 0.05 mol/L, 0.1 mol/L, 0.2 mol/L, 0.3 mol/L, and 0.5 mol/L of Cu(NO3)2 were prepared, and their pH was adjusted to pH = 4. The volume of Na2S2O3 consumed can be used to compute the amount of sample adsorbed. The adsorption of the membrane materials all increased with the concentration of Cu2+ ions, as can be seen in Figure 6b,c. The reason for this is that as the Cu(NO3)2 solution’s concentration increased, meaning that more Cu2+ ions engaged in ion exchange, which increased the amount of Cu2+ ions that the hybrid polymer was able to adsorb.
Figure 6d shows that the hybrid membrane material’s adsorption amount gradually increased over time and eventually tended to be constant, indicating saturation adsorption. This is because the ion exchange capacity of the hybrid material, which was correlated with adsorption, progressively rose over time and eventually approached saturation at the optimum adsorption duration. The equilibrium adsorption time was 9 h, at which time the membrane’s adsorption capacity was essentially saturated.
The material adsorbed Cu2+ ions very quickly, and Table 3 demonstrates that the amount of adsorption progressively increased with time and that the rate of adsorption was speedy, with the first 9 h accounting for 99.3% of the maximum adsorption of the entire process.

3.3. Absorption Properties of Pb2+ by Hybrid Membrane Materials

The 6:1 material has the highest adsorption capacity of 5.012 mmol/g, as can be seen in Figure 7a. In the following experiments, we looked at the degree of Pb2+ ion adsorption for materials A, B, C, and D at various times, concentrations, and pH values (Figure 5b). As can be seen in Figure 5a, the adsorption capacity of the material for Pb2+ ions became stronger as the organic composition increased.
The theoretical explanation for this phenomenon is that the amount of adsorption of the hybrid material may be related both to the charging characteristics of the hybrid material and to the network structure formed by the hybrid material. Different structures also show different adsorption for different metal ions. For the adsorption of Pb2+ ions, which was mainly based on ion exchange, a large number of hydrogen ions were exchanged with Pb2+ ions in solution [14,19,28], so the larger the WD-60 ratio, the greater the amount of adsorption (Figure 7b).
At a pH of 5, the four materials were best at adsorbing Pb2+. The optimal pH for Pb2+ adsorption by materials A, B, C, and D hybrids was therefore determined to be 5 (Figure 8a). The maximum adsorption capacity (qPb2+) for the four materials was reached at 21 h. (Figure 8b). Because ion exchange adsorption was the primary cause of Pb2+ ion adsorption, the best adsorption effect of the hybrid material was obtained at a Pb2+ ion concentration 0.2 mol/L, as can be seen in Figure 8c. The adsorption performance of the hybrid material was subsequently improved by increasing the amount of WD-60 at various concentrations. The adsorption capability of the hybrid membrane material for Pb2+ ions fluctuated with temperature, as can be seen in Figure 8d, with the best adsorption occurring at 40 °C. Therefore, 40 °C is the ideal temperature for the hybrid membrane material to adsorb Pb2+ ions. The theoretical explanation to this phenomenon can be ascribed to the electrostatic attraction between Pb2+ ions and carboxylic groups on the polymer chains [36].
Commonly, when metal ions are adsorbed by an adsorbent, the metal ions transport from the solution through the interface between the solution and adsorbent into the pores of the network structure. To have an insight into the effect of intraparticle diffusion on adsorption rate, sample A–D can be calculated based on the relationship of the adsorption capacity and time, which usually is expressed as (Equation (2)) [32,36]:
q t = x i + k p t 0.5
where qt is the adsorbed amount (mmol/g) at time t, kp is the intraparticle diffusion rate constant, and xi is the intercept of the straight line, which is related to the boundary layer thickness. It is now well accepted that if the plot of qt vs. t0.5 gives a straight line, the adsorption process is solely controlled by intraparticle diffusion. If the data exhibit multilinear plots, however, two or more steps will influence the adsorption process [32,36].
Figure 9 illustrated the intraparticle diffusion curves for Cu2+ and Pb2+ ions adsorption in this case. As shown in Figure 9, it can be found that the adsorption of Cu2+ and Pb2+ ions on samples A–D is non-linear over the range of adsorption time and three steps are involved: the rapid interface diffusion from 0 to 3 h; subsequently, the intraparticle diffusion (increased slowly and approached equilibrium), suggesting that Cu2+ and Pb2+ adsorption on samples A–D is not governed by intraparticle diffusion; diffusion-controlled adsorption mechanisms might be the major process, as reported in other articles [32,36].
Table 4 below shows the adsorption capacity (qPb2+) percentage of the first 9 h to 21 h (maximum adsorption capacity). Table 4 shows that the rate of adsorption was quick and that the amount of adsorption gradually increased over time, with the first 9 h accounting for a maximum of 93.0% of the entire process.

4. Conclusions

Using epoxy ring opening and sol-gel reactions, a new siloxane hybrid membrane material has been devised to enhance the heavy metal ion separation capabilities of organic-inorganic membranes under multifactorial reaction conditions. FTIR results proved that sol-gel technology was suitable for the preparation of organic-inorganic hybrid materials and the PEG and inorganic Si-O-Si components were bonded together through covalent bonds, thus successfully preparing a PEG/Si-O-Si system of ion-exchange membrane materials. TGA/DSC measurements revealed significantly improved thermal stability following the hybridization of these membranes, with the increased amount of inorganic material (WD-60) contributing to the molecular structural stability of the hybrid materials. The hybrid membrane material surfaces displayed high and efficient separation activity against Cu2+ and Pb2+ with varying degrees of heavy metal wastewater pollution. This study describes a low-cost but effective approach to preparing membrane separation surfaces and solves the problem of multifactorial reaction conditions and complicated operations from the perspective of industrial controllability. The best adsorption pH (Figure 5b), adsorption temperature, adsorption concentration, and adsorption time (Figure 7b) were found in turn via the single variable method, suggesting that these materials are likely to be widely used in water treatment and the chemical, biological, and pharmaceutical fields at an industrial scale with broad application prospects. In the practical condition, the membrane has to face the complicated feed solution. Future research will continue to investigate the microscopic properties of membrane materials to adapt to complex environments.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/membranes13050531/s1, File S1: Determination of the adsorption capacity for Cu2+; File S2: Determination of the adsorption capacity for Pb2+.

Author Contributions

C.W.: conceptualization, theoretical analysis, writing—review and editing, writing—original draft, supervision, funding acquisition, visualization, and project administration. J.Z.: methodology, investigation, validation, formal analysis, data curation, and visualization; L.H.: resources, investigation, and theoretical analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This research was sponsored by“Chenguang Program” supported by Shanghai Education Development Foundation and Shanghai Municipal Education Commission. The funding number is 21CGB13.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liang, L.W.; Wang, Z.B.; Li, J.X. The Effect of Urbanization on Environmental Pollution in Rapidly Developing Urban Agglomerations. J. Clean Prod. 2019, 237, 117649. [Google Scholar] [CrossRef]
  2. Li, K.; Fang, L.; He, L. How Population and Energy Price Affect China s Environmental Pollution? Energ. Policy 2019, 129, 386–396. [Google Scholar] [CrossRef]
  3. Liu, P.; Wang, J.; Sangaiah, A.K.; Xie, Y.; Yin, X.C. Analysis and Prediction of Water Quality Using LSTM Deep Neural Networks in IoT Environment. Sustainability 2019, 11, 2058. [Google Scholar] [CrossRef]
  4. Muhammad, Z.; Anum, R.; Saba, A. A Comprehensive Review On Polymeric Nano-composite Membranes for Water Treatment. J. Membr. Sci. Technol. 2018, 8, 1–20. [Google Scholar] [CrossRef]
  5. Teodor, S.; Andrei, S.; Simona, C.; Elena-Bianca, S.; Tanța-Verona, I.; Anita-Laura, C. Polymer Membranes as Innovative Means of Quality Restoring for Wastewater Bearing Heavy Metals. Membranes 2022, 12, 1179. [Google Scholar] [CrossRef]
  6. Caprarescu, S.; Modrogan, C.; Purcar, V.; Dancila, A.M.; Orbulet, O.D. Study of Polyvinyl Alcohol-SiO2 Nanoparticles Polymeric Membrane in Wastewater Treatment Containing Zinc Ions. Polymers 2021, 13, 1875. [Google Scholar] [CrossRef] [PubMed]
  7. Cheng, Z.J.; Li, C.; Lai, H.; Du, Y.; Liu, H.W.; Liu, M.; Jin, L.G.; Zhang, C.G.; Zhang, N.Q.; Sun, K.N. A PH-responsive Superwetting Nanostructured Copper Mesh Membrane for Separating Both Water-in-oil and Oil-in-water Emulsions. RSC Adv. 2016, 6, 72317–72325. [Google Scholar] [CrossRef]
  8. Yuan, S.S.; Zhu, J.Y.; Li, J.; Volodine, A.; Yang, J.; Puyvelde, P.V.; Bruggen, B.V. Nano/microstructure Decorated Thin Membrane Composite Poly (arylene Sulfide Sulfone) Membrane Constructed By Induced Fouling in Organic Solvent Ultrafiltration. Chem. Eng. J. 2018, 348, 180–190. [Google Scholar] [CrossRef]
  9. Kim, D.; Park, M.S.; Choi, Y.; Lee, K.B.; Kim, J.H. Synthesis of PVA-g-POEM Graft Copolymers and Their Use in Highly Permeable Thin Membrane Composite Membranes. Chem. Eng. J. 2018, 346, 739–747. [Google Scholar] [CrossRef]
  10. Nichol, T.; Callaghan, J.; Townsend, R.; Stockley, I.; Hatton, P.V.; Le Maitre, C.; Smith, T.J.; Akid, R. The Antimicrobial Activity and Biocompatibility of a Controlled Gentamicin-Releasing Single-Layer Sol-Gel Coating on Hydroxyapatite-Coated Titanium. Bone Jt. J. 2021, 103-B, 522–529. [Google Scholar] [CrossRef]
  11. Zhang, L.K.; Gong, Y.J.; Wang, T.; Xiao, J.H.; Pang, Y.P.; Hu, Q.Z.; Yu, L. Morphological Transformation of Ultrasonically Obtained Nanofibers During Living Self-assembly. New J. Chem. 2020, 44, 10813–10818. [Google Scholar] [CrossRef]
  12. El-Sayed, A.; Mahdy, I.A.; Ibraheem, F.; Mahmoud, E.A.; Ortega, J.E.; Rogero, C. Harvesting Multiple Optical Energies Using ZnPc/CdS-QDs Hybrid Organic/inorganic Semiconductors. J. Mater. Sci-Mater. Electron. 2020, 31, 12735–12742. [Google Scholar] [CrossRef]
  13. Park, W.; Shin, H.; Choi, B.; Rhim, W.; Na, K.; Han, D.K. Advanced Hybrid Nanomaterials for Biomedical Applications. Prog. Mater. 2020, 114, 100686. [Google Scholar] [CrossRef]
  14. Goodman, E.D.; Zhou, C.S.; Cargnello, M. Design of Organic/Inorganic Hybrid Catalysts for Energy And Environmental Applications. ACS Cent. Sci. 2020, 6, 1916–1937. [Google Scholar] [CrossRef]
  15. Tsai, Y.S.; Tsai, S.C.; Kuo, C.C.; Chan, W.L.; Lin, W.H.; Wu, Y.C.S.; Lin, Y.S.; Li, M.H.; Kuo, M.Y.; Chen, H. Organic/inorganic Hybrid Nanostructures of Polycrystalline Perylene Diimide Decorated ZnO Nanorods Highly Enhanced Dual Sensing Performance of UV Light/CO Gas Sensors. Results Phys. 2021, 24, 104173. [Google Scholar] [CrossRef]
  16. Wei, X.F.; Li, W.Z.; Chen, H.F.; Lv, M.G.; Wang, H.W.; Qi, X.W. An Easy Preparation of BaTiO3/PAM Organic/Inorganic Hybrid Material Using Surface-Initiated Process. J. Phys. Conf. Ser. 2021, 2194, 012045. [Google Scholar] [CrossRef]
  17. Bachevillier, S.; Yuan, H.-K.; Tetzner, K.; Bradley, D.D.C.; Anthopoulos, T.D.; Stavrinou, P.N.; Stingelin, N. Planar Refractive Index Patterning through Microcontact Photo-Thermal Annealing of a Printable Organic/Inorganic Hybrid Material. Mater. Horiz. 2021, 9, 411–416. [Google Scholar] [CrossRef]
  18. Guo, H.L.; Li, L.; Wang, F.; Kim, S.W.; Sun, H.J. Mitigating The Negative Piezoelectricity in Organic/Inorganic Hybrid Materials for High-performance Piezoelectric Nanogenerators. ACS Appl. Mater. Interfaces 2022, 14, 34733–34741. [Google Scholar] [CrossRef]
  19. Schmidt, H. New Type of Non-Crystalline Solids Between Inorganic and Organic Materials. J. Non-Cryst. Solids 1985, 73, 681–691. [Google Scholar] [CrossRef]
  20. Huang, H.H.; Orler, B.; Wilkes, G.L. Structure-Property Behavior of New Hybrid Materials Incorporating Oligomeric Species into Sol-Gel Glasses. 3. Effect of Acid Content, Tetraethoxysilane Content, and Molecular Weight of Poly(dimethylsiloxane). Macromolecules 1987, 20, 1322–1330. [Google Scholar] [CrossRef]
  21. Baxamusa, S.H.; Im, S.G.; Gleason, K.K. Initiated And Oxidative Chemical Vapor Deposition: A Scalable Method for Conformal and Functional Polymer Membranes on Real Substrates. Phys. Chem. Chem. Phys. 2009, 11, 5227–5240. [Google Scholar] [CrossRef] [PubMed]
  22. Moein, M.M.; Abdel-Rehim, A.; Abdel-Rehim, M. Recent Applications of Molecularly Imprinted Sol-Gel Methodology in Sample Preparation. Molecules 2019, 24, 2889. [Google Scholar] [CrossRef] [PubMed]
  23. Fan, Q.Q.; Li, D.N.; Li, J.H.; Wang, C.H. Structure and Piezoelectricity Properties of V-Doped ZnO Thin Membranes Fabricated By Sol-gel Method. J. Alloys Compd. 2020, 829, 154483. [Google Scholar] [CrossRef]
  24. Sarjono; Riyanto, S.; Mutiara, E.; Yusnitha, E.; Yulianto, T.; Langenati, R.; Sumaryanto, A. R&D on Surrogate Kernel Fabrication in Support of Reaktor Daya Eksperimental (RDE) Project. J. Phys. Conf. Ser. 2021, 2048, 012011. [Google Scholar] [CrossRef]
  25. Khiterer, M.; Loy, D.A.; Cornelius, C.J.; Fujimoto, C.H.; Small, J.H.; McIntire, T.M.; Shea, K.J. Hybrid Polyelectrolyte Materials for Fuel Cell Applications: Design, Synthesis, and Evaluation of Proton-Conducting Bridged Polysilsesquioxanes. Chem. Mater. 2006, 18, 3665–3673. [Google Scholar] [CrossRef]
  26. Liu, J.S.; Xu, T.W.; Gong, M.; Yu, F.; Fu, Y.X. Fundamental Studies of Novel Inorganic-organic Charged Zwitterionic Hybrids 4. New Hybrid Zwitterionic Membranes Prepared From Polyethylene Glycol (PEG) And Silane Coupling Agent. J. Membr. Sci. 2006, 283, 190–200. [Google Scholar] [CrossRef]
  27. Liu, X.Y.; Zheng, Z.; Wang, J.Q.; Wang, Y.F.; Xu, B.W.; Zhang, S.Q.; Hou, J.H. Fluidic Manipulating of Printable Zinc Oxide for Flexible Organic Solar Cells. Adv. Mater. 2022, 34, 2106453. [Google Scholar] [CrossRef]
  28. Nagarale, R.K.; Gohil, G.S.; Shahi, V.K.; Rangarajan, R. Organic−Inorganic Hybrid Membrane: Thermally Stable Cation-Exchange Membrane Prepared by the Sol−Gel Method. Macromolecules 2004, 37, 10023–10030. [Google Scholar] [CrossRef]
  29. Pogorelov, S.N.; Semenyak, G.S.; Kolmogorova, A. Sol-Gel Technology for the Production of High-Strength Refractory Materials Based on Binders. IOP Conf. Ser. Mater. Sci. Eng. 2020, 962, 022024. [Google Scholar] [CrossRef]
  30. Periyasamy, A.P.; Venkataraman, M.; Kremenakova, D.; Militky, J.; Zhou, Y. Progress in Sol-Gel Technology for the Coatings of Fabrics. Materials 2020, 13, 1838. [Google Scholar] [CrossRef]
  31. Nair, P.A.K.; Vasconcelos, W.L.; Paine, K.; Calabria-Holley, J. A review on applications of sol-gel science in cement. Constr. Build. Mater. 2021, 291, 123065. [Google Scholar] [CrossRef]
  32. Dong, Q.; Liu, J.S.; Song, L.; Shao, G.Q. Novel Zwitterionic Inorganic-Organic Hybrids: Synthesis of Hybrid Adsorbents and Their Applications for Cu2+ Removal. J. Haz. Mat. 2011, 186, 1335–1342. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, X.; Zhang, W.X.; Liu, J.S.; Wu, L.L. Removal of Lead(II) from Aqueous Solution by Adsorption Using Zwitterionic Hy-brid Copolymers. Asian J. Chem. 2013, 25, 6575–6578. [Google Scholar] [CrossRef]
  34. Liu, J.S.; Wang, X. Novel Silica-Based Hybrid Adsorbents: Lead(II) Adsorption Isotherms. Sci. World J. 2013, 2013, 897159. [Google Scholar] [CrossRef]
  35. Liu, J.S.; Li, T.; Hu, K.Y.; Shao, G.Q. Preparation and Adsorption Performances of Novel Negatively Charged Hybrid Materials. J. Appl. Polym. Sci. 2009, 112, 2179–2184. [Google Scholar] [CrossRef]
  36. Liu, J.S.; Si, J.Y.; Zhang, Q.; Zheng, J.H.; Han, C.L.; Shao, G.Q. Preparation of Negatively Charged Hybrid Adsorbents and Their Applications for Pb2+ Removal. Ind. Eng. Chem. Res. 2011, 50, 8645–8657. [Google Scholar] [CrossRef]
Figure 1. Macrographs of membrane material before (left) and after (right) adsorption.
Figure 1. Macrographs of membrane material before (left) and after (right) adsorption.
Membranes 13 00531 g001
Figure 2. (a) Surface structure of PEG/Si-O-Si organic-inorganic hybrid materials; (b) Cross-sectional morphology of PEG/Si-O-Si organic-inorganic hybrid materials.
Figure 2. (a) Surface structure of PEG/Si-O-Si organic-inorganic hybrid materials; (b) Cross-sectional morphology of PEG/Si-O-Si organic-inorganic hybrid materials.
Membranes 13 00531 g002
Figure 3. (a) Synthesis of PEG/Si-O-Si organic-inorganic hybrid materials; (b) FTIR spectra of the hybrid membrane materials (A, B, C, and D).
Figure 3. (a) Synthesis of PEG/Si-O-Si organic-inorganic hybrid materials; (b) FTIR spectra of the hybrid membrane materials (A, B, C, and D).
Membranes 13 00531 g003
Figure 4. (a) TGA curves for different ratios of silica-based hybrid membranes; (b) DSC curves for different ratios of silica-based hybrid membranes.
Figure 4. (a) TGA curves for different ratios of silica-based hybrid membranes; (b) DSC curves for different ratios of silica-based hybrid membranes.
Membranes 13 00531 g004
Figure 5. (a) Adsorption capacity of different samples (A, B, C, and D) for Cu2+ ions; (b) Cu2+ adsorption experiment with the hybrid membrane under multifactorial conditions.
Figure 5. (a) Adsorption capacity of different samples (A, B, C, and D) for Cu2+ ions; (b) Cu2+ adsorption experiment with the hybrid membrane under multifactorial conditions.
Membranes 13 00531 g005
Figure 6. (a) Cu2+ ion adsorption curves of A, B, C, and D at different temperatures; (b) Adsorption curves of A, B, C, and D at different Cu2+ ion concentrations of 0.005 mol/L, 0.01 mol/L, 0.03 mol/L and 0.05 mol/L; (c) Adsorption curves of A, B, C, and D at different Cu2+ ion concentrations of 0.1 mol/L, 0.2 mol/L, 0.3 mol/L, and 0.5 mol/L; (d) Cu2+ ion adsorption curves of A, B, C, and D at different time.
Figure 6. (a) Cu2+ ion adsorption curves of A, B, C, and D at different temperatures; (b) Adsorption curves of A, B, C, and D at different Cu2+ ion concentrations of 0.005 mol/L, 0.01 mol/L, 0.03 mol/L and 0.05 mol/L; (c) Adsorption curves of A, B, C, and D at different Cu2+ ion concentrations of 0.1 mol/L, 0.2 mol/L, 0.3 mol/L, and 0.5 mol/L; (d) Cu2+ ion adsorption curves of A, B, C, and D at different time.
Membranes 13 00531 g006
Figure 7. (a) Adsorption capacity of different samples (A, B, C, and D) for Pb2+ ions; (b) Pb2+ adsorption experiment by the hybrid membrane under multifactorial conditions.
Figure 7. (a) Adsorption capacity of different samples (A, B, C, and D) for Pb2+ ions; (b) Pb2+ adsorption experiment by the hybrid membrane under multifactorial conditions.
Membranes 13 00531 g007
Figure 8. (a) Pb2+ ion adsorption curves of A, B, C, and D at different pHs; (b) Pb2+ ion adsorption curves of A, B, C, and D at different times; (c) Adsorption curves of A, B, C, and D at different Pb2+ ion concentrations of 0.03 mol/L, 0.1 mol/L, 0.2 mol/L and 0.5 mol/L; (d) Pb2+ ion adsorption curves of A, B, C, and D at different temperatures.
Figure 8. (a) Pb2+ ion adsorption curves of A, B, C, and D at different pHs; (b) Pb2+ ion adsorption curves of A, B, C, and D at different times; (c) Adsorption curves of A, B, C, and D at different Pb2+ ion concentrations of 0.03 mol/L, 0.1 mol/L, 0.2 mol/L and 0.5 mol/L; (d) Pb2+ ion adsorption curves of A, B, C, and D at different temperatures.
Membranes 13 00531 g008
Figure 9. (a) Intraparticle diffusion curves for Cu2+ adsorption on samples A–D; (b) Intraparticle diffusion curves for Pb2+ adsorption on samples A–D.
Figure 9. (a) Intraparticle diffusion curves for Cu2+ adsorption on samples A–D; (b) Intraparticle diffusion curves for Pb2+ adsorption on samples A–D.
Membranes 13 00531 g009
Table 1. Proportion of raw materials for the preparation of hybrid membranes.
Table 1. Proportion of raw materials for the preparation of hybrid membranes.
SampleRatio (x:y)Amount of WD-60 (g)Amount of PEG-6000 (g)DMF Volume (mL)Amount of Maleic Anhydride (g)
A1:12.366.00201.00
B2:14.726.00201.00
C4:19.446.00201.00
D6:114.146.00201.00
x:y indicates the molar ratio of WD-60 to PEG-6000.
Table 2. Comparison of Qm values obtained from samples A–D with those of different types of sorbents reported in the literature.
Table 2. Comparison of Qm values obtained from samples A–D with those of different types of sorbents reported in the literature.
Sorbent TypeqCu2+
(mmol/g)
qPb2+
(mmol/g)
Literature
zwitterionic hybrid copolymers0.0750.244[32,33]
silica-based hybrid adsorbents/0.506[34]
pyromellitic acid dianhydride (PMDA)
hybrid adsorbents
/6.379[36]
negatively charged siloxane hybrid membrane0.3315.012this work
Table 3. The proportion of Cu2+ adsorption in the first 9 h of the overall process.
Table 3. The proportion of Cu2+ adsorption in the first 9 h of the overall process.
SamplesAdsorption Rate (%)Standard Deviation (%)
A98.50.08
B98.50.05
C98.60.06
D99.30.05
Table 4. Proportion of Pb2+ adsorption in the first 9 h of the overall process.
Table 4. Proportion of Pb2+ adsorption in the first 9 h of the overall process.
SamplesAdsorption Rate (%)Standard Deviation (%)
A88.50.3
B91.10.2
C91.30.3
D93.00.4
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

Wu, C.; Zheng, J.; Han, L. Adsorption Performance of Heavy Metal Ions under Multifactorial Conditions by Synthesized Organic-Inorganic Hybrid Membranes. Membranes 2023, 13, 531. https://doi.org/10.3390/membranes13050531

AMA Style

Wu C, Zheng J, Han L. Adsorption Performance of Heavy Metal Ions under Multifactorial Conditions by Synthesized Organic-Inorganic Hybrid Membranes. Membranes. 2023; 13(5):531. https://doi.org/10.3390/membranes13050531

Chicago/Turabian Style

Wu, Chaoqun, Jiuhan Zheng, and Limei Han. 2023. "Adsorption Performance of Heavy Metal Ions under Multifactorial Conditions by Synthesized Organic-Inorganic Hybrid Membranes" Membranes 13, no. 5: 531. https://doi.org/10.3390/membranes13050531

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