Next Article in Journal
Fault Detection and Efficiency Assessment for HVAC Systems Using Non-Intrusive Load Monitoring: A Review
Next Article in Special Issue
Influence of Calcination Temperature on Electrochemical Properties of Perovskite Oxide Nanofiber Catalysts
Previous Article in Journal
Evaluation of Changes in Psychophysical Performance during the Afternoon Drop off in Work Capacity after the Exposure to Specific Color of Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparison of Nature and Synthetic Zeolite for Waste Battery Electrolyte Treatment in Fixed-Bed Adsorption Column

1
Department of Chemical Engineering and Analytical Science, Faculty of Science and Engineering, The University of Manchester, Manchester M1 3BB, UK
2
School of Mechanical Engineering and Automation, Harbin Institute of Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(1), 347; https://doi.org/10.3390/en15010347
Submission received: 7 November 2021 / Revised: 8 December 2021 / Accepted: 30 December 2021 / Published: 4 January 2022
(This article belongs to the Special Issue Advances in Fuel Cells and Metal-Air Batteries)

Abstract

:
To support a sustainable energy development, CO2 reduction for carbon neutralization and water-splitting for hydrogen economy are two feasible technical routes, both of which require a significant input of renewable energies. To efficiently store renewable energies, secondary batteries will be applied in great quantity, so that a considerable amount of energy needs to be invested to eliminate the waste battery electrolyte pollution caused by heavy metals including Cu2+, Zn2+ and Pb2+. To reduce this energy consumption, the removal behaviors of these ions by using clinoptilolite and zeolite A under 5, 7 and 10 BV h−1 in a fixed-bed reactor were investigated. The used zeolites were then regenerated by a novel NH4Cl solution soaking, coupled with the ultrasonication method. Further characterizations were carried out using scanning electron microscopy, N2 adsorption and desorption test, and wide-angle X-ray diffraction. The adsorption breakthrough curves revealed that the leaching preference of clinoptilolite was Pb2+ > Cu2+ > Zn2+, while the removal sequence for zeolite A was Zn2+ > Cu2+ > Pb2+. The maximum removal percentage of Zn2+ ions for clinoptilolite under 5 BV h−1 was 21.55%, while it was 83.45% for zeolite A. The leaching ability difference was also discussed combining with the characterization results. The fact that unit cell stayed the same before and after the regeneration treatment approved the efficacy of the regeneration method, which detached most of the ions while doing little change to both morphology and crystallinity of the zeolites. By evaluating the pH and conductivity changes, the leaching mechanisms by adsorption and ion exchange were further studied.

1. Introduction

Sustainable energy development is of key importance to the human society, which is especially the case when facing the fossil fuel depletion and the associated global warming issues. To tackle this, great efforts have been made to achieve the carbon neutralization via CO2 reduction to hydrocarbon fuels [1,2]. Alternatively, the hydrogen fuel can be utilized instead in order to totally eliminate the carbon-related issue, which can be produced simply via the water-splitting reaction [3,4]. In both ways, a significant amount of energy is required to facilitate the reduction reaction, which has to be coming from various renewable sources such as solar, wind, geothermal and nuclear energy in order to ensure the sustainability. On the other hand, the urge to preserve renewable energies or smoothly couple them with electrical grids requires an effective large-scale energy storage/conversion system, such as batteries and electrolyzers [5]. Batteries are well-known for their high energy efficiency and fast response. Among them, the commercial lithium-ion battery is already very mature in technology, but it still needs further improvement in cost and safety before being applied into the large-scale electrical grid [6]. On the other hand, the rechargeable metal-air battery such as aqueous Zn-air battery is also promising for this mission considering its great cost-efficiency and high safety level [7]. As for electrolyzers, they can convert the renewable energy into stable chemicals such as hydrogen gas, which have unique advantages such as negligible self-discharge for long-term storage, low-cost expansion of storage capacity and high energy density. However, the energy efficiency would be lower due to the overpotential loss.
For both batteries and electrolyzers, the electrode is one of the most important components where the electrochemical reaction takes place. Its active surface area, catalytic capability as well as stability, strongly determine the device performance, including power density, energy efficiency and long-term durability. Therefore, significant research has been conducted in exploring new catalysts, catalyst structures and synergistic catalyst combinations, among which the nanostructured electrode is more advanced than conventional monolithic electrodes. On the other hand, the electrolyte is also a prerequisite component which determines the ohmic resistance of the energy device. In general, it contains a large extent of heavy metal ions, which can be a very serious environmental problem considering the ever-increasing demand for energy storage. For instance, as the mostly used heavy metal in daily applications, the high concentration and the high toxicity of the Cu2+ made it urgent to be removed from the water system [8]. Hence, the heavy metal ions in the waste electrolyte need to be removed efficiently, especially for preventing the water pollution resulted from copper, lead, zinc, mercury and manganese ions that have been reported to be increasingly serious in the estuarine floodplain areas [9,10,11]. Heavy metals including Pb2+ in the sediment can accumulate through the food chain and eventually enter the human body [12], causing damage to the kidney, red blood cells and even DNA damage, eventually leading to conformational changes [13]. These are likely to result in cell-cycle modulation, carcinogenesis, or apoptosis, even at very low concentrations [14,15].
Conventional methods have been used to remove the heavy metal ions from contaminated water, including chemical precipitation [16], electrochemical methods [17], membrane filtration [18] and ion exchange and adsorption. These either require extra electricity input or create additional slurry that demands further treatment. Among these methods, ion exchange and adsorption stand out for its considerable treatment capacity, high removal efficiency, fast response and low energy consumption [14,19]. Moreover, adsorption was considered to be a cost-efficient, easy, flexible and durable process [20]. Due to their abundant resources, renewable property and relatively low cost, zeolites were massively applied in the fields of adsorption, catalysis and ion exchange [21,22]. Zeolites are microporous crystalline aluminosilicates minerals with cage-like structures [23]. The primary building unit of zeolite is based on TO4 tetrahedra, where T is aluminum or silicon atom, and this structure makes zeolite a suitable material for ion exchange [24]. Some silicon cations are substituted by aluminum cations, yielding a net negative charge in the tectosilicate framework, which gives zeolites a cation exchangeability [25]. Several pieces of literature proved that natural zeolite such as clinoptilolite has been widely applied into the fields of catalytic activities [26] and electrodes [27]. Meanwhile, it was also implemented globally over the years as the most widespread adsorbent to remove heavy metals from wastewater [28,29,30,31]. Zeolite A is usually synthesized and used as the carrier for basis of catalyst [32] or selective electrode [33]. Zeolites received the spotlight because of the high catalytic adsorption capacity, controlled acidity, confinement effects and high stability when applied as the catalyst [34].
Many studies have been conducted for leaching heavy metals from contaminated water [35,36] in batch mode. On the other hand, using zeolite for the removal of heavy metals in a fixed-bed column has been attracting attention due to the fact that it can better mimic the real contaminated water treatment process [37,38,39]. In fixed-bed studies, most of them focus on the selectivity of a certain type of zeolite towards different ions, while little attention is paid to the difference in the leaching performances among different zeolites. Therefore, in this work, we examined clinoptilolite and also zeolite A in the fixed-bed column to compare their leaching performances as well as their corresponding selectivity towards copper, lead and zinc ions.
The reactivation of used adsorbents is also important in industrial applications. Usually, acid, including nitric acid [40], base or salts were utilized in the regeneration process. However, zeolite is material with aluminol groups at the edge; when zeolite is put under a low pH value (pH < 4) environment, the gibbsite edge begins to dissolve [41]. Thus, harsh treatment should be avoided. Herein, the combination of soaking and ultrasonic bathing method was proposed as a novel regeneration method. It has been proved that most natural zeolites in their exchanged form are selective for NH4+ experimentally [42] and theoretically [43]. Therefore, the combination of NH4Cl solution soaking and ultrasonic bathing is likely to enhance regeneration efficiency, which has never been reported before to the best of our knowledge. In general, this work will provide a sustainable treatment method with low energy consumption for waste battery electrolyte by reusing the zeolites, which will contribute to the achievement of clean water and sanitation (goal 6), responsible consumption and reproduction (goal 12), protecting the life on land (goal 15) as well as the life below water (goal 14) of the 17 Sustainable Development Goals (SDGs).

2. Materials and Methods

2.1. Adsorption Experiment Raw Materials

Lead (II) nitrate (Pb(NO3)2, 99%), copper (II) sulphate pentahydrate (CuSO4⋅5H2O), zinc chloride (ZnCl2) and ammonium chloride (NH4Cl ≥ 99.5%) were purchased from Sigma-Aldrich with analytical grade. Heavy metal bulk solutions at a concentration of 0.5 g L−1 were prepared by dissolving the corresponding salts in the deionized water to mimic the waste battery electrolyte. A 0.01 M nitric acid (HNO3) was used to adjust the initial solution pH to 4 in order to avoid precipitation during all ion exchange experiments. The average size of clinoptilolite was in the range of 9–16 mm, while that of zeolite A is approximately 5 mm. All zeolites were washed with distilled water five times to remove residue dust on the surface and then dried at 60 °C for 3 days.

2.2. Fixed-Bed Adsorption

The fixed-bed adsorption column was designed as shown in Figure 1. The solution with a specific concentration (C) was injected at a constant volumetric rate (Q) through a peristaltic pump from the bottom of the column, and then the liquid samples were collected from the top of the column. To gain more precise breakthrough curves, samples were taken out every 10 min at the beginning of the experiment for 1 h. After that, the sampling time was set to be 20 min. The diameter and the length of the column were 4.2 cm and 21 cm, respectively, and the empty bed volume (BV) was calculated to be 291 mL.
Fixed bed experiments were carried out in order to compare the leaching performances of zeolite A and clinoptilolite. Three different heavy metal ions (Cu2+, Pb2+ and Zn2+) and three different volumetric flowrates (5 BV h−1 = 24.25 mL min−1, 7 BV h−1 = 33.95 mL min−1 and 10 BV h−1 = 48.5 mL min−1) were investigated in this project, and each experiment lasted for 3 h to ensure that the leaching process reached a stable state.

2.3. Characterization of the Zeolites Associated with Heavy Metal Adsorption Ions

The samples were dried under 60 °C for 3 days and then reserved for the following characterizations: Scanning electron microscopy (FEI Quanta 200 Environmental SEM) was used to study the surface morphology of zeolite samples before and after the adsorption and regeneration. Energy Dispersive X-ray Spectroscopy (EDX) was used to analyze the elemental composition of the adsorbent. The dried samples were further ground into powders and sent for characterizations including N2 adsorption and desorption test, also called Brunauer, Emmett and Teller (BET) test, for the measurement of the pore size and the surface area of the samples. An XRD diffractometer (PANalytical XRD5 (Phillips)) was also used to investigate the crystal phases and the mineral identity of these samples. The X-ray source was generated by copper Kα radiation, the scanning range started from 15° to 100° at a rate of 5°/min, 0.02°/step.
The heavy metal concentrations in the liquid samples were measured by inductively coupled plasma optical emission spectrometry (ICP-OES). The exit relative metal concentration (Ct/C0) was defined by the following equation:
C t / C 0 = ( C 0 C R ) C 0 × 100 %
where C0 is the initial solution concentration, Ct is the solution concentration at time t, CR represents the removed solution concentration. Therefore, the smaller the Ct/C0 value, the more the corresponding contaminants are removed. By plotting Ct/C0 versus bed volume, the breakthrough curves were then obtained. Meanwhile, the pH and the conductivity of the samples were measured simultaneously by the pH meter (Eutech PC 2700, Thermo Scientific, Waltham, MA, USA).

2.4. Regeneration Experiments

The used zeolites were washed 5 times with distilled water and placed in an oven at 60 °C for 3 days to be fully dried. Then, they were weighed and poured into a 500 mL beaker before adding 300 mL regeneration solution (NH4Cl at the concentration of 1 g L−1), and the mixture was ultrasonicated at room temperature for 2 h. The regenerated zeolite was also washed 5 times with distilled water and placed in an oven at 60 °C for 3 days to be fully dried.

3. Results

3.1. Heavy Metal Ions Leaching Experiments

The adsorption behavior of both zeolites toward 0.5 g L−1 Cu2+, Zn2+ and Pb2+ was studied under 5 BV h−1, 7 BV h−1 and 10 BV h−1 in a fixed-bed adsorption column. As shown in Figure 2, the leaching speed of Cu2+, Zn2+ and Pb2+ ions via both zeolites gradually decreased over time. The lower flow rate of the feed solution in the column provided a longer contact time for the adsorption process between the heavy metal ions and the adsorbents. Therefore, 5 BV h−1 flow rate provided the longest contact time, hence enhancing the leaching performance.
Overall, Zn2+ has more affinity to Zeolite A, while Pb2+ was more prone to be adsorbed by clinoptilolite. The corresponding pH and conductivity evaluations were also illustrated in Figure 3 and Figure 4. After the Zn2+ solutions passed through the fixed-bed adsorption column with clinoptilolite, the pH increased from 4 to 6.5 (on average) and there were negligible changes during the whole experiment time. This indicates that the amount of H+ ion in the solution decreased due to the H+/Na+ exchange according to Kurtoğlu and Atun [44]. In the case of zeolite A, the pH increased to 8 (on average) and then decreased to 6.5 (on average), which could be attributed to the increased concentration of heavy metal ions in the effluent solution. The ions bounded with the H2O molecular, leading to the hydrolysis effect and resulting in a decrease of pH value [45].
As shown in Figure 4, the conductivity values decreased for all the experiments as a function of time, except for clinoptilolite leaching Pb2+. This indicates that the number of movable electrolyte ions in the solution decreased, which may be resulted from the adsorption mechanism. The existing ions in the bulk were adsorbed into the zeolite matrix, leading to a decrease in the conductivity of the effluent solution. It is worth noting that in the experiment of clinoptilolite leaching Pb2+, another removal behavior appeared. The conductivities in the effluent solutions nearly stayed at the same level around 400 μS (the conductivity of raw solution was 404 μS), revealing that there might be only ion exchange in the removal process [38].
As shown in Figure 5, at the end of experiments when using clinoptilolite, the removal percentage of Pb2+ was the most (62.4%), followed by Cu2+ (32.7%), and the removed Zn2+ was the least (21.55%). Under the same experimental conditions, zeolite A adsorbed the most amount of Zn2+ and the least Pb2+. This shows that the selectivity sequence among these three ions for clinoptilolite is: Pb2+ > Cu2+ > Zn2+, while it was Zn2+ > Cu2+ >Pb2+ for zeolite A. In addition, the removal percentage for zeolite A towards Pb2+ is the least among all three ions, which can be explained by the Hard-Soft interaction theory: the harder the cations, the higher the charge density they have, and their d-orbitals do not participate in π-bond formation. This led to the higher possibility to conduct the electrostatic interaction [46,47]. However, as a soft cation, Pb2+ was least adsorbed by the zeolite A.

3.2. Regeneration Experiments

As can be seen from Figure 6a, the regenerated clinoptilolite followed a similar adsorption behavior with the clean clinoptilolite at the beginning of the experiment. However, as the adsorption proceeded, the regenerated clinoptilolites showed a weaker leaching ability than the pristine clinoptilolites. At the end of the adsorption, the removal percentage (1 − Ct/C0) × 100% dropped from 23.91% to 16.63% after the regeneration. This phenomenon also occurred frequently in other studies [12,48], illustrating two facts: 1. The regeneration method was effective, which detached most of the Zn2+ ions that had been bound to the zeolite adsorption sites; 2. Some of the residue Zn2+ ions were not completely removed from the zeolite matrix and still occupied part of the adsorption sites, resulting in a decrease in the removal percentage after regeneration.
The removal performance was represented by the removal percentage of Zn2+ ions in the solution at the end of the experiment. The corresponding removal percentages are listed in Table 1. Our findings suggest that the adsorption ability of each zeolite was similar after the regeneration, and the regeneration method to reactivate the used zeolite was effective. This work is a novel empirical example of the application of ultrasonication to treat contaminated zeolites. Our results encourage further research to investigate the processes and mechanisms using different classes of adsorbents and contaminants.

3.3. Supportive Characterization of the Zeolites Associated with Heavy Metal Adsorption

3.3.1. SEM and EDX

SEM-EDX analysis was conducted on clean and post-treatment samples of each zeolite. Clinoptilolite is a kind of material with a rough surface as shown in Figure 7a, and the ultrasonic bath did not change much on its morphology. As shown in Figure 7b, there are some plate-like coagulations on the surface. Additionally, smaller cubic-structured particles were formed on zeolite A after the adsorption process in Figure 7d, which indicates that there existed some substances on the surface after the adsorption of zinc ions.
In terms of the EDX analysis, the elements present in each zeolite sample and their atomic percentages were displayed in Table 2. Zinc was absent from the clean zeolites, and after the adsorption reaction, the zinc content increased to 1.33% for clinoptilolite, and it increased to 2.03% for zeolite A. These results directly proved that the adsorption was successful. Plus, the ratio of Si/Al of clean clinoptilolite was around 4.14, which was similar to reported data 4.84 by Nezamzadeh-Ejhieh and Moeinirad [49,50].

3.3.2. BET

The BET was also conducted for the aforementioned samples, and the results are shown in Table 3 below. After the adsorption, the surface area of clinoptilolite increased from 30.99 m2 g−1 to 44.17 m2 g−1. Similarly, the surface area of zeolite A drastically increased from 40.69 m2 g−1 to 301.48 m2 g−1, indicating that a more complex structure was formed.

3.3.3. XRD

The lattice parameters of the zeolites are also listed in the Table 4 below, where a, b and c represent the size of the unit cells, while α, β and γ represent the angles of these unit cells. The tested samples include: 1. clean clinoptilolite; 2. Zn2+-contaminated clinoptilolite; 3. clean zeolite A; 4. Zn2+-contaminated zeolite A. Clinoptilolite unit cell has a wider β angle, while the α and γ angles are 90°. The heavy metal removal procedure reduced the overall size of the unit cell and transferred the lattice structure of clinoptilolite from the HEU phase-type to CHA phase-type. This phenomenon also happened to zeolite A. The lattice size of the zeolite A was decreased by half from that of the previous state. The detailed matching spectra was shown in the Supporting Information, where typical XRD peaks of ZnO were assigned at 31.84 (100), 34.51 (002), 47.63 (102), 56.71 (110), 62.69 (103), 69.18 (201) which agree with JCPDS PDF No. 88-0287 [51].

3.4. Leaching Mechanism

The charge deficit, which was resulted from the replacement of Si4+ by Al3+, enabled the tridymite structure to adsorb cations. Hence, both the adsorption and ion exchange happened simultaneously during the leaching process [52,53,54]. The leaching mechanisms can be summarized by the following equations:
nSi OH + M n +       ( Si O ) n M + nH +
SiO + MOH +     SiOMOH
nSiO + M n +       ( Si O ) n M
nMOH + + M ( z )       M n ( z ) + M n + + nOH
where Mn+ represents the heavy metal ions in the bulk solutions, while M’(z) represents the free ions in the zeolites matrix. Among those equations, Equations (2)–(4) illustrate the adsorption process, while Equation (5) represents the ion exchange mechanism. A similar phenomenon was also found in using clinoptilolite-based electrodes for bromate determination. Electroactive cations in zeolite were exchanged by zeolitic compounds, hence the ion exchange was achieved through replacing the modifier agent to the cations of the supporting electrolyte in solution [55].
During the leaching process, more negatively charged surfaces became available, therefore leading to greater Zn2+ uptake [56]. According to the analysis of the characterization, the results of the experiments as well as the literature studies, the leaching procedure mainly consists of 3 steps. Take zeolite A leaching Zn2+ as an example; the details are shown in Figure 8.
In Figure 8a, at the beginning of the adsorption, the pH of the bulk solution was 4. According to the cation activity series, H+ was also involved in the exchange process. At this time, monovalent H+ entered the zeolite matrix and led to the rapid pH rise of the solution. Meanwhile, ions including K+ and Ca2+ were also substituted, and therefore, a huge increase of conductivity from 915 μS (raw Zn2+ solution conductivity) to around 3300 μS under 5 BV h−1 was observed. As the pH was elevated in Figure 8b, more negatively charged surfaces become available, therefore leading to greater metal uptake, even at precipitation pH value [56]. This explained the remarkable leaching behavior of zeolites at the beginning of the experiments. During this stage, some ions were precipitated, while other partial ions entered the zeolite matrix. Both of the activities resulted in a dramatic drop in conductivity. As shown in Figure 8c, after the Zn2+ ions entered the lattice structure, the hydrolysis began as:
[ M ( H 2 O ) x ] c + + H 2 O       [ M ( H 2 O ) x 1 ( OH ) ] ( c 1 ) + + H 3 O +
Zinc ions combined with water molecules created larger complexes and therefore produced H+. This process lowered the pH of the solution back to near neutral. This large complex gathered on the surface of the zeolite, increasing the specific surface area as shown in the BET results. Over time, the leaching process will reach saturation, that is, adsorbents will be exhausted. From then on, the conductivity in the solution will remain at a certain level.

4. Conclusions

In this work, the heavy metal ions leaching performances of natural clinoptilolite and synthesized zeolite A at different flow rates were comprehensively studied in a fixed-bed adsorption column. The removal preference of the two types of zeolites was found to be Pb2+ > Cu2+ > Zn2+ when using clinoptilolite, while it was Zn2+ > Cu2+ > Pb2+ for zeolite A. Among the two zeolites used for adsorption experiments, zeolite A is better for the adsorption of heavy metal ions in water than clinoptilolite. Namely, zeolite A leached 4 times higher Zn2+ ions under 5 BV h−1 in comparison to clinoptilolite (83.45% for zeolite A and 21.55% for clinoptilolite). A novel regeneration method was applied by combining the ultrasonication and NH4Cl solution soaking. The regeneration method was effective without influencing the intrinsic zeolite properties. The unit cell stayed the same before and after the regeneration treatment, while the adsorption process created smaller secondary structures, especially on zeolite A. The surface area of zeolite A before adsorption was 40.69 m2 g−1 and it was 301.48 m2 g−1 after the adsorption. This was caused by the hydrolysis phenomenon. This study extends the use of zeolite beyond electrical-chemical wastewater treatment and agriculture: our results encourage testing the mentioned mechanisms when the novel regeneration method is applied.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en15010347/s1. Figure S1: XRD matching pattern of clean zeolite A; Figure S2: XRD matching pattern of clean clinoptilolite; Figure S3: XRD matching pattern of Zinc contami-nated zeolite A; Figure S4: XRD matching pattern of Zinc contaminated clinop-tilolite; Figure S5: BET results of clean zeolite A (a), Cu2+ contaminated zeolite A (b), clean clinoptilolite (c) and Cu2+ contaminated clinoptilolite (d).

Author Contributions

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

Funding

This research received no external funding.

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.

Acknowledgments

The research reported in this publication was supported by the University of Manchester. The authors thank Desmond Doocey for his kind help with the ICP training.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Al-Juboori, O.; Sher, F.; Khalid, U.; Niazi, M.B.K.; Chen, G.Z. Electrochemical Production of Sustainable Hydrocarbon Fuels from CO2 Co-electrolysis in Eutectic Molten Melts. ACS Sustain. Chem. Eng. 2020, 8, 12877–12890. [Google Scholar] [CrossRef]
  2. Al-Juboori, O.; Sher, F.; Hazafa, A.; Khan, M.K.; Chen, G.Z. The effect of variable operating parameters for hydrocarbon fuel formation from CO2 by molten salts electrolysis. J. CO2 Util. 2020, 40, 101193. [Google Scholar] [CrossRef]
  3. Al-Shara, N.K.; Sher, F.; Yaqoob, A.; Chen, G.Z. Electrochemical investigation of novel reference electrode Ni/Ni(OH)₂ in comparison with silver and platinum inert quasi-reference electrodes for electrolysis in eutectic molten hydroxide. Int. J. Hydrogen Energy 2019, 44, 27224–27236. [Google Scholar] [CrossRef]
  4. Al-Shara, N.K.; Sher, F.; Iqbal, S.Z.; Sajid, Z.; Chen, G.Z. Electrochemical study of different membrane materials for the fabrication of stable, reproducible and reusable reference electrode. J. Energy Chem. 2020, 49, 33–41. [Google Scholar] [CrossRef]
  5. Dunn, B.; Kamath, H.; Tarascon, J.-M.J.S. Electrical energy storage for the grid: A battery of choices. Science 2011, 334, 928–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Huang, J.; Guo, Z.; Ma, Y.; Bin, D.; Wang, Y.; Xia, Y. Recent Progress of Rechargeable Batteries Using Mild Aqueous Electrolytes. Small Methods 2019, 3, 1800272. [Google Scholar] [CrossRef] [Green Version]
  7. Leong, K.W.; Wang, Y.; Ni, M.; Pan, W.; Luo, S.; Leung, D.Y.C. Rechargeable Zn-air batteries: Recent trends and future perspectives. Renew. Sustain. Energy Rev. 2022, 154, 111771. [Google Scholar] [CrossRef]
  8. Heidari-Chaleshtori, M.; Nezamzadeh-Ejhieh, A. Clinoptilolite nano-particles modified with aspartic acid for removal of Cu(II) from aqueous solutions: Isotherms and kinetic aspects. New J. Chem. 2015, 39, 9396–9406. [Google Scholar] [CrossRef]
  9. Enya, O.; Lin, C.; Qin, J. Heavy metal contamination status in soil-plant system in the Upper Mersey Estuarine Floodplain, Northwest England. Mar. Pollut. Bull. 2019, 146, 292–304. [Google Scholar] [CrossRef]
  10. Tamiji, T.; Nezamzadeh-Ejhieh, A. A comprehensive study on the kinetic aspects and experimental design for the voltammetric response of a Sn(IV)-clinoptilolite carbon paste electrode towards Hg(II). J. Electroanal. Chem. 2018, 829, 95–105. [Google Scholar] [CrossRef]
  11. Nosuhi, M.; Nezamzadeh-Ejhieh, A. Voltammetric determination of trace amounts of permanganate at a zeolite modified carbon paste electrode. New J. Chem. 2017, 41, 15508–15516. [Google Scholar] [CrossRef]
  12. Luo, X.; Yuan, J.; Liu, Y.; Liu, C.; Zhu, X.; Dai, X.; Ma, Z.; Wang, F. Improved Solid-Phase Synthesis of Phosphorylated Cellulose Microsphere Adsorbents for Highly Effective Pb2+ Removal from Water: Batch and Fixed-Bed Column Performance and Adsorption Mechanism. ACS Sustain. Chem. Eng. 2017, 5, 5108–5117. [Google Scholar] [CrossRef]
  13. Shirzadi, H.; Nezamzadeh-Ejhieh, A. An efficient modified zeolite for simultaneous removal of Pb(II) and Hg(II) from aqueous solution. J. Mol. Liq. 2017, 230, 221–229. [Google Scholar] [CrossRef]
  14. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef]
  15. Haddad, S.A.; Lemanowicz, J. Benefits of Corn-Cob Biochar to the Microbial and Enzymatic Activity of Soybean Plants Grown in Soils Contaminated with Heavy Metals. Energies 2021, 14, 5763. [Google Scholar] [CrossRef]
  16. Ku, Y.; Jung, I.-L. Photocatalytic reduction of Cr(VI) in aqueous solutions by UV irradiation with the presence of titanium dioxide. Water Res. 2001, 35, 135–142. [Google Scholar] [CrossRef]
  17. Vepsäläinen, M.; Sillanpää, M. Chapter 1—Electrocoagulation in the treatment of industrial waters and wastewaters. In Advanced Water Treatment; Sillanpää, M., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 1–78. [Google Scholar]
  18. Kurniawan, T.A.; Chan, G.Y.S.; Lo, W.-H.; Babel, S. Physico-chemical treatment techniques for wastewater laden with heavy metals. Chem. Eng. J. 2006, 118, 83–98. [Google Scholar] [CrossRef]
  19. Pessoa, M.E.; de Sousa, K.S.; Clericuzi, G.Z.; Ferreira, A.L.; Soares, M.C.; Neto, J.C. Adsorption of Reactive Dye onto Uçá Crab Shell (Ucides cordatus): Scale-Up and Comparative Studies. Energies 2021, 14, 5876. [Google Scholar] [CrossRef]
  20. Yang, C.; Wang, L.; Yu, Y.; Wu, P.; Wang, F.; Liu, S.; Luo, X. Highly efficient removal of amoxicillin from water by Mg-Al layered double hydroxide/cellulose nanocomposite beads synthesized through in-situ coprecipitation method. Int. J. Biol. Macromol. 2020, 149, 93–100. [Google Scholar] [CrossRef] [PubMed]
  21. Burakov, A.E.; Galunin, E.V.; Burakova, I.V.; Kucherova, A.E.; Agarwal, S.; Tkachev, A.G.; Gupta, V.K. Adsorption of heavy metals on conventional and nanostructured materials for wastewater treatment purposes: A review. Ecotoxicol. Environ. Saf. 2018, 148, 702–712. [Google Scholar] [CrossRef] [PubMed]
  22. Huang, H.; Li, J.-R.; Wang, K.; Han, T.; Tong, M.; Li, L.; Xie, Y.; Yang, Q.; Liu, D.; Zhong, C. An in situ self-assembly template strategy for the preparation of hierarchical-pore metal-organic frameworks. Nat. Commun. 2015, 6, 8847. [Google Scholar] [CrossRef] [Green Version]
  23. Haw, J.F. Zeolite acid strength and reaction mechanisms in catalysis. Phys. Chem. Chem. Phys. 2002, 4, 5431–5441. [Google Scholar] [CrossRef]
  24. Perić, J.; Trgo, M.; Vukojević Medvidović, N. Removal of zinc, copper and lead by natural zeolite—A comparison of adsorption isotherms. Water Res. 2004, 38, 1893–1899. [Google Scholar] [CrossRef]
  25. Vaca Mier, M.; López Callejas, R.; Gehr, R.; Jiménez Cisneros, B.E.; Alvarez, P.J.J. Heavy metal removal with mexican clinoptilolite: Multi-component ionic exchange. Water Res. 2001, 35, 373–378. [Google Scholar] [CrossRef]
  26. Alidusty, F.; Nezamzadeh-Ejhieh, A. Considerable decrease in overvoltage of electro-catalytic oxidation of methanol by modification of carbon paste electrode with Cobalt(II)-clinoptilolite nanoparticles. Int. J. Hydrogen Energy 2016, 41, 6288–6299. [Google Scholar] [CrossRef]
  27. Nosuhi, M.; Nezamzadeh-Ejhieh, A. High catalytic activity of Fe(II)-clinoptilolite nanoparticales for indirect voltammetric determination of dichromate: Experimental design by response surface methodology (RSM). Electrochim. Acta 2017, 223, 47–62. [Google Scholar] [CrossRef]
  28. Wang, S.; Peng, Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chem. Eng. J. 2010, 156, 11–24. [Google Scholar] [CrossRef]
  29. Zanin, E.; Scapinello, J.; de Oliveira, M.; Rambo, C.L.; Franscescon, F.; Freitas, L.; de Mello, J.M.M.; Fiori, M.A.; Oliveira, J.V.; Dal Magro, J. Adsorption of heavy metals from wastewater graphic industry using clinoptilolite zeolite as adsorbent. Process Saf. Environ. Prot. 2017, 105, 194–200. [Google Scholar] [CrossRef]
  30. Kocasoy, G.; Şahin, V. Heavy metal removal from industrial wastewater by clinoptilolite. J. Environ. Sci. Health Part A 2007, 42, 2139–2146. [Google Scholar] [CrossRef] [PubMed]
  31. Çoruh, S. The removal of zinc ions by natural and conditioned clinoptilolites. Desalination 2008, 225, 41–57. [Google Scholar] [CrossRef]
  32. Nezamzadeh-Ejhieh, A.; Banan, Z. A comparison between the efficiency of CdS nanoparticles/zeolite A and CdO/zeolite A as catalysts in photodecolorization of crystal violet. Desalination 2011, 279, 146–151. [Google Scholar] [CrossRef]
  33. Nezamzadeh-Ejhieh, A.; Mirzaeyan, E. Hexadecylpyridinium surfactant modified zeolite A as an active component of a polymeric membrane sulfite selective electrode. Mater. Sci. Eng. C 2013, 33, 4751–4758. [Google Scholar] [CrossRef]
  34. Fareed, B.; Sher, F.; Sehar, S.; Rasheed, T.; Zafar, F.; Ameen, M.; Lima, E.C. Tailor made Functional Zeolite as Sustainable Potential Candidates for Catalytic Cracking of Heavy Hydrocarbons. Catal. Lett. 2021, 1–13. [Google Scholar] [CrossRef]
  35. Turan, M.; Mart, U.; Yüksel, B.; Çelik, M.S. Lead removal in fixed-bed columns by zeolite and sepiolite. Chemosphere 2005, 60, 1487–1492. [Google Scholar] [CrossRef]
  36. Netzer, A.; Hughes, D.E. Adsorption of copper, lead and cobalt by activated carbon. Water Res. 1984, 18, 927–933. [Google Scholar] [CrossRef]
  37. Arulanantham, A.; Balasubramanian, N.; Ramakrishna, T.V. Coconut shell carbon for treatment of cadmium and lead containing wastewater. Met. Finish. 1989, 87, 51–55. [Google Scholar]
  38. Stylianou, M.A.; Hadjiconstantinou, M.P.; Inglezakis, V.J.; Moustakas, K.G.; Loizidou, M.D. Use of natural clinoptilolite for the removal of lead, copper and zinc in fixed bed column. J. Hazard. Mater. 2007, 143, 575–581. [Google Scholar] [CrossRef]
  39. Demirel, S.; Uyanik, I. Simultaneous fluoride and nitrate removal from drinking water using mixotrophic denitrification processes in a fixed bed column reactor. Desalination Water Treat. 2019, 164, 56–61. [Google Scholar] [CrossRef]
  40. Liu, H.; Peng, S.; Shu, L.; Chen, T.; Bao, T.; Frost, R.L. Magnetic zeolite NaA: Synthesis, characterization based on metakaolin and its application for the removal of Cu2+, Pb2+. Chemosphere 2013, 91, 1539–1546. [Google Scholar] [CrossRef]
  41. Wieland, E.; Stumm, W. Dissolution kinetics of kaolinite in acidic aqueous solutions at 25 °C. Geochim. Cosmochim. Acta 1992, 56, 3339–3355. [Google Scholar] [CrossRef]
  42. Ames, L.L., Jr. Cation sieve properties of the open zeolites chabazite, mordenite, erionite and clinoptilolite. Am. Mineral. 1961, 46, 1120–1131. [Google Scholar]
  43. Neveu, A.; Gaspard, M.; Blanchard, G.; Martin, G. La diffusion intraparticulaire dans la clinoptilolite application aux ions Na+ et NH4+. Water Res. 1985, 19, 611–618. [Google Scholar] [CrossRef]
  44. Kurtoğlu, A.E.; Atun, G. Determination of kinetics and equilibrium of Pb/Na exchange on clinoptilolite. Sep. Purif. Technol. 2006, 50, 62–70. [Google Scholar] [CrossRef]
  45. Inglezakis, V.J.; Stylianou, M.A.; Gkantzou, D.; Loizidou, M.D. Removal of Pb(II) from aqueous solutions by using clinoptilolite and bentonite as adsorbents. Desalination 2007, 210, 248–256. [Google Scholar] [CrossRef]
  46. Mehrali-Afjani, M.; Nezamzadeh-Ejhieh, A. Efficient solid amino acid-clinoptilolite nanoparticles adsorbent for Mn(II) removal: A comprehensive study on designing the experiments, thermodynamic and kinetic aspects. Solid State Sci. 2020, 101, 106124. [Google Scholar] [CrossRef]
  47. Nezamzadeh-Ejhieh, A.; Afshari, E. Modification of a PVC-membrane electrode by surfactant modified clinoptilolite zeolite towards potentiometric determination of sulfide. Microporous Mesoporous Mater. 2012, 153, 267–274. [Google Scholar] [CrossRef]
  48. Luo, X.; Lei, X.; Cai, N.; Xie, X.; Xue, Y.; Yu, F. Removal of Heavy Metal Ions from Water by Magnetic Cellulose-Based Beads with Embedded Chemically Modified Magnetite Nanoparticles and Activated Carbon. ACS Sustain. Chem. Eng. 2016, 4, 3960–3969. [Google Scholar] [CrossRef]
  49. Nezamzadeh-Ejhieh, A.; Moeinirad, S. Heterogeneous photocatalytic degradation of furfural using NiS-clinoptilolite zeolite. Desalination 2011, 273, 248–257. [Google Scholar] [CrossRef]
  50. Nezamzadeh-Ejhieh, A.; Shirvani, K. CdS loaded an Iranian clinoptilolite as a heterogeneous catalyst in photodegradation of p-aminophenol. J. Chem. 2013, 2013, 541736. [Google Scholar] [CrossRef]
  51. Derikvandi, H.; Nezamzadeh-Ejhieh, A. Increased photocatalytic activity of NiO and ZnO in photodegradation of a model drug aqueous solution: Effect of coupling, supporting, particles size and calcination temperature. J. Hazard. Mater. 2017, 321, 629–638. [Google Scholar] [CrossRef]
  52. Busca, G. Acidity and basicity of zeolites: A fundamental approach. Microporous Mesoporous Mater. 2017, 254, 3–16. [Google Scholar] [CrossRef]
  53. Keane, M.A. The removal of copper and nickel from aqueous solution using Y zeolite ion exchangers. Colloids Surf. A Physicochem. Eng. Asp. 1998, 138, 11–20. [Google Scholar] [CrossRef]
  54. Cabrera, C.; Gabaldón, C.; Marzal, P. Sorption characteristics of heavy metal ions by a natural zeolite. J. Chem. Technol. Biotechnol. 2005, 80, 477–481. [Google Scholar] [CrossRef]
  55. Tamiji, T.; Nezamzadeh-Ejhieh, A. Sensitive voltammetric determination of bromate by using ion-exchange property of a Sn(II)-clinoptilolite-modified carbon paste electrode. J. Solid State Electrochem. 2019, 23, 143–157. [Google Scholar] [CrossRef]
  56. Ijagbemi, C.O.; Baek, M.H.; Kim, D.S. Montmorillonite surface properties and sorption characteristics for heavy metal removal from aqueous solutions. J. Hazard. Mater. 2009, 166, 538–546. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The fixed-bed reactor: (1) represents the heavy metal solution feed tank, then the solution is pumped up through the peristaltic pump (2) to the plexiglass column (4) equipped with a plastic sieve (3 and 5) at the bottom and the top which is stuffed with zeolites and sent out (6). The optical graph is also presented on the right.
Figure 1. The fixed-bed reactor: (1) represents the heavy metal solution feed tank, then the solution is pumped up through the peristaltic pump (2) to the plexiglass column (4) equipped with a plastic sieve (3 and 5) at the bottom and the top which is stuffed with zeolites and sent out (6). The optical graph is also presented on the right.
Energies 15 00347 g001
Figure 2. Breakthrough curves of clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f) under 5 BV h−1, 7 BV h−1 and 10 BV h−1.
Figure 2. Breakthrough curves of clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f) under 5 BV h−1, 7 BV h−1 and 10 BV h−1.
Energies 15 00347 g002
Figure 3. pH studies using clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f).
Figure 3. pH studies using clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f).
Energies 15 00347 g003
Figure 4. Conductivity studies of using clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f).
Figure 4. Conductivity studies of using clinoptilolite leaching Cu2+ (a); Zn2+ (b); Pb2+ (c) and zeolite A leaching Cu2+ (d); Zn2+ (e); Pb2+ (f).
Energies 15 00347 g004
Figure 5. The removal percentages of clinoptilolite and zeolite A applied in different ions under 5 BV h−1 at the end of the experiment.
Figure 5. The removal percentages of clinoptilolite and zeolite A applied in different ions under 5 BV h−1 at the end of the experiment.
Energies 15 00347 g005
Figure 6. The breakthrough curves of the clinoptilolite (a) and zeolite A (b) leach Zn2+ under 5 BV h−1; the black data points represented pristine zeolites and the red data points stand for recycled zeolites.
Figure 6. The breakthrough curves of the clinoptilolite (a) and zeolite A (b) leach Zn2+ under 5 BV h−1; the black data points represented pristine zeolites and the red data points stand for recycled zeolites.
Energies 15 00347 g006
Figure 7. SEM results of clean clinoptilolite (a); Zn2+-contaminated clinoptilolite (b); clean zeolite A (c) and Zn2+-contaminated zeolite A (d).
Figure 7. SEM results of clean clinoptilolite (a); Zn2+-contaminated clinoptilolite (b); clean zeolite A (c) and Zn2+-contaminated zeolite A (d).
Energies 15 00347 g007
Figure 8. The leaching mechanism of the process: at the adsorption beginning stage: (a) The porous green cubic matrix of zeolite A were surrounded with yellow spheres Zn2+ ions, blue spheres the H+ ions. In the pH value elevation stage (b), the precipitates formed as green hexagonal plates. After the hydrolysis happened in stage (c), dark green cubic material was the hydrolysis complex formed at the surface of the pale blue matrix of the zeolite.
Figure 8. The leaching mechanism of the process: at the adsorption beginning stage: (a) The porous green cubic matrix of zeolite A were surrounded with yellow spheres Zn2+ ions, blue spheres the H+ ions. In the pH value elevation stage (b), the precipitates formed as green hexagonal plates. After the hydrolysis happened in stage (c), dark green cubic material was the hydrolysis complex formed at the surface of the pale blue matrix of the zeolite.
Energies 15 00347 g008
Table 1. The removal performances of the clean zeolites used to remove the heavy metal ions for the first time (Original experiment) and the regenerated zeolites used to remove the heavy metal ions (Regeneration experiment).
Table 1. The removal performances of the clean zeolites used to remove the heavy metal ions for the first time (Original experiment) and the regenerated zeolites used to remove the heavy metal ions (Regeneration experiment).
Removal Percentage (%)
Clinoptilolite leach Zn2+ at 5 BV h−1Original experiment23.91
Regeneration experiment16.63
Zeolite A leach Zn2+ at 5 BV h−1Original experiment84.44
Regeneration experiment74.99
Table 2. Elemental analysis of clean clinoptilolite, zinc-contaminated clinoptilolite, clean zeolite A and zinc-contaminated zeolite A.
Table 2. Elemental analysis of clean clinoptilolite, zinc-contaminated clinoptilolite, clean zeolite A and zinc-contaminated zeolite A.
Element Content (%)SiAlKNaCaMgOZn
Clean clinoptilolite62.5915.117.0103.593.388.310
Zn2+-contaminated clinoptilolite61.6912.265.3105.212.6511.551.33
Clean zeolite A46.9725.573.885.432.123.712.330
Zn2+-contaminated zeolite A45.5324.773.276.173.053.4911.692.03
Table 3. BET results of 4 samples (1: clean clinoptilolite; 2: Zn2+-contaminated clinoptilolite; 3: clean zeolite A; 4: Zn2+-contaminated zeolite A).
Table 3. BET results of 4 samples (1: clean clinoptilolite; 2: Zn2+-contaminated clinoptilolite; 3: clean zeolite A; 4: Zn2+-contaminated zeolite A).
1234
BET surface area (m2 g−1)30.9944.1740.69301.48
Micropore Volume
(×10−2 cm3 g−1 STP)
0.170.311.071.40
Micropore Area (m2 g−1)2.745.2318.68244.28
External Surface Area (m2 g−1)28.2638.9522.0157.20
Table 4. The lattice parameters of zeolites before and after the adsorption of Zn2+.
Table 4. The lattice parameters of zeolites before and after the adsorption of Zn2+.
1234
Major Phase HEU
(Heulandite)
CHA
(Chabazite)
Sodium Aluminum SilicateSodium Zinc Aluminum Silicate
a (Å)17.5368.84524.950612.152
b (Å)17.27716.60724.950612.152
c (Å)7.4099.74624.950612.152
α90°90°90°90°
β116.62°123.19°90°90°
γ90°90°90°90°
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, C.; Wang, Y.; Alfutimie, A. Comparison of Nature and Synthetic Zeolite for Waste Battery Electrolyte Treatment in Fixed-Bed Adsorption Column. Energies 2022, 15, 347. https://doi.org/10.3390/en15010347

AMA Style

Yang C, Wang Y, Alfutimie A. Comparison of Nature and Synthetic Zeolite for Waste Battery Electrolyte Treatment in Fixed-Bed Adsorption Column. Energies. 2022; 15(1):347. https://doi.org/10.3390/en15010347

Chicago/Turabian Style

Yang, Cong, Yifei Wang, and Abdullatif Alfutimie. 2022. "Comparison of Nature and Synthetic Zeolite for Waste Battery Electrolyte Treatment in Fixed-Bed Adsorption Column" Energies 15, no. 1: 347. https://doi.org/10.3390/en15010347

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