Next Article in Journal
Structural Characteristics, Stability, and Electronic Properties of 001 Surface with Point Defects of Zinc Stannate: A First-Principle Study
Previous Article in Journal
Efficient and Rapid Removal of Pb(II) and Cu(II) Heavy Metals from Aqueous Solutions by MgO Nanorods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Peppermint-Mediated Green Synthesis of Nano ZrO2 and Its Adsorptive Removal of Cobalt from Water

by
Ibrahem Mohamed Abouzeid Hasan
*,
Hanan Salah El-Din
and
Ahmed A. AbdElRaady
Chemistry Department, Faculty of Science, South Valley University, Qena 83523, Egypt
*
Author to whom correspondence should be addressed.
Inorganics 2022, 10(12), 257; https://doi.org/10.3390/inorganics10120257
Submission received: 22 October 2022 / Revised: 20 November 2022 / Accepted: 28 November 2022 / Published: 12 December 2022

Abstract

:
Zirconium oxide nanoparticles (ZrO2NPs) were green synthesized for the first time using an aqueous peppermint extract as a precipitating and capping agent. Addition of the extract to Zr4+ solution was followed by calcination of the resulting precipitate at 570 and 700 °C to form ZrO2NPs570 and ZrO2NPs700, respectively. These oxides were characterized using X-ray diffraction, transmission electron microscopy, and BET surface area analysis, and used as adsorbents for cobalt ions (Co2+) in water. The effects of pH, initial Co2+ concentration, ZrO2NPs mass, and contact time on adsorption efficiency were studied. Characterization results showed formation of cubic ZrO2 with average crystallite sizes (XRD data) of 6.27 and 7.26 nm for ZrO2NPs570 and ZrO2NPs700, respectively. TEM images of the two oxides exhibited nearly spherical nanoparticles and BET surface area measurements indicated the formation of mesoporous oxides having surface areas of 94.8 and 62.4 m2/g, respectively. The results of the adsorption study confirmed that the synthesized ZrO2NPs can be efficiently used for the adsorption of Co2+ from water. The uptake of Co2+ from the treated solution is favored at pH values higher than its point of zero charge (6.0). In addition, the adsorption of Co2+ by ZrO2 follows a pseudo-second order kinetics (R2 = 1.0) and can be explained by the Langmuir adsorption isotherm (R2 = 0.973).

Graphical Abstract

1. Introduction

Zirconium oxide (ZrO2) is a wide direct band gap semiconductor and considered a multifunctional material due to its unique physical and chemical properties, such as high thermal stability, excellent mechanical characteristics, non-toxicity, and adsorption ability. It was used in many scientific and technological fields, such as catalysis, adsorption, biology, optics, electronics, magnetism, and dental medicine [1,2,3,4,5,6,7,8].
Adsorption is a well-known technique for removing various organic and inorganic contaminants from aqueous media [9,10]; as a surface phenomenon, it involves the formation of physical or chemical bonds between the adsorbate molecules and the adsorbent surface. Accordingly, adsorbent surface properties such as morphology, surface area, pore size distribution, charge, nature of active sites, and their distribution and density determine their efficiency. These properties are greatly influenced by preparation method and conditions.
Many substances are used as adsorbents [11,12]; however, using metal oxides such as aluminum oxide (Al2O3), zinc oxide (ZnO), magnetic iron oxide (Fe3O4), and zirconium oxide (ZrO2) as adsorbents is promising for removing different organic and inorganic pollutants from water because they have several advantages such as the simplicity of operation and the cost effectiveness [13,14,15]. Among these oxides, ZrO2 is a preferred choice for removing different contaminants from water owing to its non-toxicity, high specific surface area, and ease of synthesis. The presence of both surface acidic and basic centers also makes it a promising adsorbent for the removal of many kinds of pollutants, and its adsorption capability can be modified significantly when combined with other substances [16,17].
ZrO2 was synthesized by different physicochemical methods [18,19,20,21]. Using non-toxic and environmentally benign reagents and solvents are a key issue in green synthesis routes. In this context, using plant extracts in particular has many advantages since this reduces the number of hazardous chemicals utilized in the synthesis processes of such adsorbents [22,23,24,25,26]. In addition, phytochemicals in the plant extract act as reducing, precipitating, and capping agents and thus play an important role in controlling particle size, shape, phase stability, and other characteristics of nanomaterials [27,28].
Peppermint (Menthapiperita L.) is a medicinal herb used as flavoring agent in pharmaceutics and cosmetics, etc. In addition, it is a natural hybrid herb belonging to the Lamiaceae family, its major components being menthol, menthone, neomenthol, and iso-menthone [29,30].
Trace levels of some heavy metals, such as cobalt, are important for the structure of various biological species [31]. However, a high concentration of these heavy metals is toxic [32,33,34] principally due to its accumulation in biological systems. As some heavy metals are essential in the industries of pigments, electronics, paints, and metallurgy, the discharges from these industries contain considerable amounts of these persistent and non-biodegradable species that result in serious environmental pollution [35]. The adsorption of Co2+ by different types of adsorbents was reported [36,37]. Some of these methods reported that the efficiency of adsorption is dependent on Co2+ concentration—being high at low Co2+ concentrations but decreasing significantly at high Co2+ concentrations [37].
To the best of our knowledge, there is no work on using peppermint extract as a precipitating and capping agent in the literature for the synthesis of ZrO2. Thus, the aim of this study is to green synthesize ZrO2NPs using an aqueous peppermint extract as the precipitant and stabilizing agent, and to investigate its characteristics and efficiency as an adsorbent for Co2+ removal from water.

2. Results and Discussion

2.1. XRD Analysis

The XRD patterns of ZrO2NPs570 and ZrO2NPs700 are shown in Figure 1. These patterns show diffraction peaks at 2 theta values corresponding to the (111), (200), (220), (311), (400), (331), and (422) planes of cubic ZrO2 Fm-3m (225) (COD 9009051) [38,39]. It is also clear from Figure 1 that the peak intensities increase little with the increasing calcination temperature from 570 to 700 °C. On the other hand, the broadening of the XRD peaks indicates that the crystallite sizes of the samples are in the nanoscale [40]. The crystallite sizes of the ZrO2NPs estimated using the Scherrer equation [41] from peaks 111, 220, and 311 averaged 6.27 and 7.26 nm for ZrO2NPs570 and ZrO2NPs700, respectively, as can be seen in Table 1.

2.2. Transmission Electron Microscopy

TEM provides valuable information about the particle size and shape. Therefore, the changes in particle size and shape accompanying calcination at different temperatures can be examined. The TEM images of the two ZrO2570 and ZrO2700 samples are given in Figure 2a,b and their particle size distributions are in Figure 2c,d, respectively. It is clear that the particles of the two samples are nearly spherical, with diameters in the range of 7–8 nm, which are comparable to the crystallite sizes calculated from the XRD data. This small size of the nanoparticles may be attributed to the capping effect of the phytochemicals in PAE which hinders growth of nanoparticles. It is also evident that neither the shape nor the size of the particles is largely affected by increasing the calcination temperature from 570 to 700 °C.

2.3. BET Surface Area

The nitrogen sorption isotherms and the corresponding pore distribution of ZrO2NPs570 and ZrO2NPs700 are presented in Figure 3a–d. The respective specific surface areas were measured to be 94.805 and 62.384 m2·g−1 for ZrO2NPs570 and ZrO2NPs700 as can be seen in Table 1. According to the values in this table, the calculated pore diameter doubles due to the increasing calcination temperature, while the pore volume is little affected. This may be due to the pores becoming less deep for the sample calcined at the higher temperature. The N2 adsorption/desorption isotherms in Figure 3, and the values of pore diameter calculated in Table 1, suggest that ZrO2 synthesized by this method is mesoporous and the isotherms with their hysteresis loops can be classified as type IV [42].

2.4. Mechanism of ZrO2NPs Green Synthesis

The main phytochemicals in peppermint are menthol and menthone [29,30]. The possible mechanism for ZrO2NPs formation is shown in Scheme 1. A complex is formed between Zr4+ ions and menthol molecules in aqueous medium. The complex is capped by other organic molecules, terminating further growth of the particles. Calcination of the complex then results in the formation of ZrO2NPs via oxidation by oxygen in air [27,28].

2.5. Adsorption of Co2+ by ZrO2NPs and Optimization of Adsorption Parameters

Preliminary experiments on cobalt ion adsorption by the synthesized ZrO2NPs showed that ZrO2NPs570 had a significantly higher adsorption efficiency than ZrO2NPs700. ZrO2NPs570 was, therefore, used for further adsorption studies.

2.5.1. Effect of pH

The adsorption phenomenon involves concentration of the adsorbate molecules onto the adsorbent surface. The electrostatic attraction between the surface and the adsorbate species is a major mechanism by which adsorption occurs. The nature and magnitude of the adsorbent surface charge, and also the charge on the adsorbate molecules, are affected by the pH value of the medium and, as a result, the adsorption process is pH dependent [43]. To determine the optimum pH for the uptake of Co2+ by the ZrO2 surface, the pH value of the Co2+ solution was adjusted at 3, 5, 6.5 and 7.5. pH values equal to or higher than 8.5, resulting in the precipitation of Co2+.
In this study, 6 g·L−1 of the adsorbent was used with 50 mL of 150 ppm Co2+ aqueous solution. The pH values were adjusted by the addition of NaOH or HCl (0.1 M each) at 25 °C and the adsorption lasted for 30 min. It is clear from Figure 4a that the adsorption of cobalt (R% and qe) increases with increasing pH. The R% reaches about 45% at pH 7.5. To understand this effect of the pH on the efficiency of ZrO2NPs as an adsorbent, the value of the point of zero charge (PZC) of the adsorbent should be considered [44]. The PZC value of ZrO2 used in this study was determined to be 6.0 using a simple method [45] and is shown Figure 4b. The negatively charged surface of ZrO2 at the higher pH values (6.5 and 7.5) favors the adsorption of the positively charged Co2+. Protonation of the active surface sites at pH values lower than the PZC decreases the tendency of interaction of the positively charged cobalt with the surface [46].

2.5.2. Effect of Contact Time

It is important to determine the equilibrium contact time under the experimental conditions used. Equilibrium concentration values are required for estimation of different adsorption parameters and also for consideration of the adsorption process for application on the industrial scale. Contact time was investigated with different adsorbate concentrations, using adsorbent masses of 6 g·L−1. In all cases, as indicated in Figure 5, the incremental increase in the R% decreases gradually with time where the lines tend to flatten down with the progress of adsorption. The initial relatively fast uptake of cobalt is accounted for by the availability of surface adsorption sites in the initial adsorption stage. At longer contact times, the tendency of surface sites to bind Co2+ decreases due to the depletion of cobalt concentration with time, and also due to the development of surface electrostatic repulsion between adsorbed cobalt ions for a contact time of about 30 min, where equilibrium is assumed to be reached under the experimental conditions. This behavior is frequently encountered in adsorption studies [47,48]. Figure 6, on the other hand, indicates that qt increases with increasing time until equilibrium is reached, as is expected.

2.5.3. Effect of Adsorbent Mass

In adsorption studies, the achievement of the maximum degree of adsorption is required. Accordingly, an important factor to study is the optimum mass of the adsorbent under experimental conditions. The effect of the amount of adsorbent on its efficiency for the removal of Co2+ was studied using ZrO2NPs masses in the range 2–10 g·L−1. As can be seen from Figure 7a, the Co2+ % removal increases with increasing ZrO2NPs mass. This direct proportionality between adsorbent mass and %R at a fixed adsorbate concentration of 100 ppm can be explained by the availability of adsorption sites responsible for the uptake of Co2+ with increasing ZrO2 mass [49]. On the other hand, the decrease in the equilibrium adsorption capacity from about 6.4 mg·g−1 at the beginning to about 5.2 mg·g−1 at the end is mainly due to the fact that an increasing adsorbent mass for a fixed Co2+ concentration results in a larger degree of unsaturation of adsorption sites.

2.5.4. Effect of Initial Co2+ Concentration

The effect of initial Co2+ concentration on its equilibrium adsorption % removal by ZrO2NPs was investigated at different initial Co2+ concentrations in the range of 50–200 mg·L−1 at a fixed adsorbent mass of 6 g·L−1. The results of this study are summarized in Figure 7b. It is clear that the % removal of Co2+ by ZrO2NPs decreases with the increase in initial cobalt concentration. The percent removal of Co2+ decreases from 50% to 28.6% when the initial Co2+ concentration is increased from 50 to 200 mg·L−1 at a fixed adsorbent mass. However, the amount of Co2+ adsorbed by the fixed amount of adsorbent (qt) nearly doubles (from about 4 to about 9 mg·g−1) when the cobalt concentration is increased in the same range (from 50 to 200 mg·L−1). An increase in the cobalt ion concentration is accompanied by an increase in the degree of interaction between Co2+ ions in the aqueous phase and surface of the adsorbent, according to the law of mass action. In addition, the increase in the amount of Co2+ adsorbed when the adsorbate concentration is increased can be explained by the fact that a high concentration of Co2+ creates a higher driving force for mass transfer between the adsorbent surface and adsorbate [50]. It is worth noting that the adsorption capacity value obtained in our experiments (~9 mg·g−1) approximately equals a previously published value (9.43 mg·g−1) in spite of the 10-fold higher temperature (250 °C) in the latter [51]. This may be attributed to the 4-fold higher surface area (94 m2·g−1) of the peppermint-mediated ZrO2NPs than that of the chemically synthesized ZrO2NPs (24 m2·g−1).

2.6. Adsorption Isotherms

Adsorption isotherms are theoretical relationships that correlate the adsorbate concentration in a medium to its equilibrium concentration onto the adsorbent surface at a certain constant temperature [52]. In addition, they play an important role in determining the adsorption capacities of the adsorbents. Furthermore, they give an insight into the degree of affinity of the adsorbent to the adsorbate species and the adsorption mechanism. The adsorption capacity was calculated from different equilibrium concentrations of the adsorbate and the experimental data obtained were fitted to Freundlich and Langmuir models at 298 K for initial Co2+ concentrations of 50, 100, 150, and 200 mg·L−1. The linearized forms of the Langmuir and Freundlich isotherms are written in Equations (1) and (2), respectively [53,54]:
Ce/qe = (1/KL·qm) + (Ce/qm)
log qe = (1/n) log Ce + log KF
where qe (mg·g−1): equilibrium adsorption capacity, qm (mg·g−1): Langmuir maximum adsorption capacity, KL (L·mg−1): Langmuir constant, Ce (mg·L−1): equilibrium adsorbate concentration in solution, KF (mg·g−1)/(mg·L−1)1/n: Freundlich adsorption capacity, and 1/n (dimensionless): Freundlich constant which provides adsorption intensity.
The Langmuir and Freundlich isotherms are presented at 298 K and an initial Co2+ concentration in the range 50–200 mg·L−1 in Figure 8a,c, respectively. The values of constants for each model were calculated to evaluate the adsorption characteristics and affinity of the adsorbent for Co2+ (Table 2). The results in Table 2 suggest that ZrO2NPs are effective adsorbents for Co2+ uptake from the aqueous solution to form a monolayer of adsorbate on the surface of the adsorbent. According to the Langmuir model, the high value of KL at room temperature indicates the strong affinity of Co2+ to the ZrO2NPs surface and reflects the ability of the adsorbent to adsorb large amount of Co2+. This agrees with previous studies that suggested the Langmuir adsorption isotherm for the adsorption of Co2+ by different adsorbents, including metal oxides [55,56,57,58]. The Langmuir equation also can be used to predict whether an adsorption system is favorable or unfavorable [53]. The adsorption intensity (RL) can be defined by Equation (3):
RL = 1/(1 + KLCo)
The value of RL indicates the favorability of the adsorption isotherm; irreversible (RL = 0), favorable (0 < RL < 1), linear (RL = 1), or unfavorable (RL > 1) [53]. The variation in adsorption intensity RL with the initial concentration of the solution is presented in Figure 8b. The calculated RL values are 0.337, 0.253, 0.202, 0.145, and 0.113 for the respective Co2+ initial concentrations of 50, 75, 100, 150, and 200 mg·L−1. These values indicate a favorable and efficient adsorption process across the entire concentration range.
On the other hand, the empirical equation of the Freundlich model can be used to describe adsorption from a solution [53]. The linear equation for this model allows calculation of the values of the Freundlich isotherm constants (Kf and n). The high values of these constants show an easy uptake and good adsorption process of adsorbate from aqueous solutions, and also high adsorptive capacities at the studied conditions. The values of the Freundlich isotherm constants n and Kf for the adsorption of Co2+ by ZrO2 are given in Table 2. Comparing the correlation coefficients, the R2 values of both isotherms indicate that the results are best fitted by the Langmuir model.

2.7. Kinetics of the Adsorption of Co2+ onto ZrO2NPs

Investigation of adsorption kinetics provides valuable information about the mechanism of the adsorption process. The kinetics of adsorbate uptake by the adsorbent was described by various models [59,60]. Of these, the pseudo-first order, pseudo-second order, the Elovich, and the intraparticle diffusion models were used in this study to test the experimental data. The respective linear equations for these four models are shown in Equations (4)–(7) [59,60].
log (qe − qt) = logqe − (k1/2.303) t
qt = (1/k2qe2) + (1/qe) t
qt = (1/β) ln αβ + (1/β) lnt
qt = ki t0.5 + C
where qt (mg·g−1): amount of adsorbate adsorbed at time t, k1 (min−1): pseudo-first order rate constant, k2 (g·mg−1·min−1): pseudo-second order rate constant, α (mg·g−1·min−1): the initial adsorption rate, β (g·mg−1): parameter related to the extent of surface coverage and activation energy of the Elovich equation, ki (mg·g−1·min−0.5): the intraparticle diffusion rate constant, and C: the width of the boundary layer.
Figure 9a–d presents the different models. The values of R2 as well as the constants of the models are summarized in Table 3. As can be seen from this table, the obtained results are fitted well by the pseudo-second order model. This suggests that the uptake of Co2+ is a chemisorption process involving a physicochemical interaction between the Co2+ and the ZrO2 surface.
The Elovich equation posits that the active sites on the solid surface are diverse in character so they exhibit varied properties for chemisorption activation energies [61]. A linear characteristic is revealed by the values of (R2) calculated from Figure 9c. Table 3 presents the α and β coefficients. In the Elovich model, α is proportional to the rate of change in chemisorption (initial adsorption rate), and β is related to surface amplification (desorption constant). High values of α and β indicate the rapid rate of chemisorption and the increase in the available adsorption surface for Co2+ [61]. Figure 9d shows the effect of intraparticle diffusion on adsorption. Fluid flow, film diffusion, and the plateau region are all represented by the three portions of this curve. The straight lines do not pass through the origin under the conditions tested, indicating that intraparticle diffusion is not the determining factor in the sorption process.

3. Materials and Methods

3.1. Synthesis of ZrO2NPs

About 100 g of fresh green peppermint (whole plant) was washed with tap water, rinsed with distilled water (DW), dried in air for 48 h, cut into small pieces and then impregnated in DW for 24 h. The peppermint aqueous extract (PAE) was collected by filtration and was kept in refrigerator. Fifteen grams of zirconium oxychloride were dissolved in about 50 mL distilled water in a 500 mL Pyrex glass beaker. To this solution, the PAE (about 200 mL) was added dropwise while the mixture was subjected to continuous stirring. A pale brown precipitate was formed and was magnetically stirred for 2 h, filtered off, washed with DW several times, dried in air for 24 h, and then dried in an oven at 100 °C for 5 h to form a black solid [27], as shown in Scheme 2. The latter solid was carefully milled and calcined, in air, for 3 h at 570 and 700 °C to form ZrO2NPs570 and ZrO2NPs700, respectively.

3.2. Characterization

The phase compositions of the samples were determined using a powder X-ray diffractometer (X’Pert3 Powder, PAN Anlytical, Almelo, The Netherlands) operating at 40 KV and 30 mA using Cu-Kα = 1.54056 Å as a radiation source and a nickel filter. The crystallite size of different samples was calculated from the XRD data using the Scherrer equation: D = Kλ/βcosθ, where D is the average crystallite size (nm), K is the full width at half maximum, θ is the Bragg angle, β is the shape factor, and λ is the wavelength = 1.54056 Å. A BEL SORP-MAX analyzer (Microtrac BEL, Osaka, Japan) was used to determine the adsorbent specific surface areas. A UV-vis spectrophotometer (PG Instruments, model T80, Leicestershire, UK) was used to measure the concentrations of Co2+ using quartz cells with a 1 cm path length. Transmission electron microscope (TEM) images were taken using a JEM-1010 transmission electron microscope operating at 70 KV and 58 μA. A portion of the sample was dispersed in absolute ethanol, sonicated, and a drop of the resulting suspension was taken on a carbon grid (200 mesh). Before being admitted to the microscope, the grid was left in air to evaporate the ethanol.

3.3. Adsorption Experiments

The adsorption experiments were performed in clean reagent bottles at room temperature. In these experiments, ZrO2NPs were dispersed in a specific volume of Co2+ solution of known concentration, and the mixture was continuously agitated using a mechanical shaker at room temperature (25 ± 1 °C). The effects of different parameters influencing adsorption such as pH, initial Co2+ concentration, ZrO2NPs mass, and contact time were studied. Adsorption conditions were optimized by changing one adsorption parameter while keeping the other parameters fixed [62]. After suitable treatment times, the adsorbents were separated from the treated solutions by centrifugation and the remaining Co2+ concentration in the solutions was determined by UV-vis spectroscopy [63]. The concentration of adsorbed Co2+ was calculated from the difference between the initial and final concentrations (Co and Cf, respectively) of Co2+. The removal percentage (R%) and the adsorption capacity (qt) were estimated using the following Equations (8) and (9), respectively:
R% = [(Co − Cf)/Co] × 100
qt = [(Co − Cf) × V]/W
In Equation (2), V is the volume of the treated solution in liters and W is the weight of ZrO2NPs utilized in the experiment in grams.

4. Conclusions

Peppermint-mediated green synthesis of ZrO2NPs was described in this paper. The oxides calcined at 570 and 700 °C, ZrO2570 and ZrO2700, were characterized and used as adsorbents for the removal of cobalt ions from water. Different parameters affecting adsorption (pH, initial adsorbate concentration, adsorbent mass, and agitation time) were studied. Results show that ZrO2NPs samples are mesoporous and can be used efficiently for the removal of cobalt from water. The adsorption follows a pseudo-second order kinetics and can be explained by the Langmuir adsorption isotherm.

Author Contributions

I.M.A.H.: conceptualization, methodology, preparation, investigation, writing—original draft; H.S.E.-D.: validation, writing—review and editing; A.A.A.: conceptualization, methodology, preparation, investigation, writing—original draft. All authors have read and agreed to the published version of the manuscript.

Funding

No funds were received for this study.

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. Trejo-Arroyo, D.L.; Acosta, K.E.; Cruz, J.C.; Valenzuela-Muñiz, A.M.; Vega-Azamar, R.E.; Jiménez, L.F. Influence of ZrO2 Nanoparticles on the Microstructural Development of Cement Mortars with Limestone Aggregates. Appl. Sci. 2019, 9, 598. [Google Scholar] [CrossRef] [Green Version]
  2. Matt, S.B.; Shivanna, M.; Manjunath, S.; Siddalinganahalli, M.; Siddalingappa, D.M. Electrochemical Detection of Serotonin Using t-ZrO2 Nanoparticles Modified Carbon Paste Electrode. J. Electrochem. Soc. 2020, 167, 155512. [Google Scholar] [CrossRef]
  3. Mok, Z.H.; Proctor, G.; Thanou, M. Emerging nanomaterials for dental treatments. Emerg. Top. Life Sci. 2020, 4, 613–625. [Google Scholar]
  4. Khan, M.; Shaik, M.R.; Khan, S.T.; Adil, S.F.; Kuniyil, M.; Khan, M.; Al-Warthan, A.A.; Siddiqui, M.R.H.; Tahir, M.N. Enhanced Antimicrobial Activity of Biofunctionalized Zirconia Nanoparticles. ACS Omega 2020, 5, 1987–1996. [Google Scholar] [CrossRef]
  5. Ashiq, M.N.; Aman, A.; Alshahrani, T.; Iqbal, M.F.; Razzaq, A.; Najam-Ul-Haq, M.; Shah, A.; Nisar, J.; Tyagi, D.; Ehsan, M.F. Enhanced electrochemical properties of silvercoated zirconia nanoparticles for supercapacitor application. J. Taibah Univ. Sci. 2021, 15, 10–16. [Google Scholar] [CrossRef]
  6. Shelke, G.B.; Patil, D.R. Gas Sensing Performance of Pure and Modified Nanostructured Screen Printed Zirconia Thick Films. Int. J. Eng. Res. Technol. 2019, 8, 198–202. [Google Scholar]
  7. Chikere, C.O.; Faisal, N.H.; Kong-Thoo-Lin, P.; Fernandez, C. Interaction between Amorphous Zirconia Nanoparticles and Graphite: Electrochemical Applications for Gallic Acid Sensing Using Carbon Paste Electrodes in Wine. Nanomater 2020, 10, 537. [Google Scholar] [CrossRef] [Green Version]
  8. Annu, A.; Sivasankari, C.; Krupasankar, U. Synthesis and characterization of ZrO2 nanoparticle by leaf extract bioreduction process for its biological studies. Mater. Today Proc. 2020, 33, 5317–5323. [Google Scholar] [CrossRef]
  9. Demirbas, E. Adsorption of Cobalt(II) Ions from Aqueous Solution onto Activated Carbon Prepared from Hazelnut Shells. Adsor. Sci. Technol. 2003, 21, 951–963. [Google Scholar] [CrossRef] [Green Version]
  10. Patel, H. Charcoal as an adsorbent for textile wastewater Treatment. Sep. Sci. Technol. 2018, 53, 2797–2812. [Google Scholar] [CrossRef]
  11. Rajendran, S.; Priya, A.K.; Kumar, P.S.; Hoang, T.K.A.; Sekar, K.; Chong, K.Y.; Khoo, K.S.; Ng, H.S.; Show, P.L. A critical and recent developments on adsorption technique for removal of heavy metals from wastewater-A review. Chemosphere 2022, 303, 135146. [Google Scholar] [CrossRef]
  12. Zou, W.; Bai, H.; Zhao, L.; Li, K.; Han, R. Characterization and properties of zeolite as adsorbent for removal of uranium (VI) from solution in fixed bed column. J. Nucl. Chem. 2011, 288, 779–788. [Google Scholar] [CrossRef]
  13. Gua, M.; Haoa, L.; Wang, Y.; Li, X.; Chen, Y.; Li, W.; Jiang, L. The selective heavy metal ions adsorption of zinc oxide nanoparticles from dental wastewater. Chem. Phys. 2020, 534, 110750. [Google Scholar] [CrossRef]
  14. Roto, R. Surface Modification of Fe3O4 as Magnetic Adsorbents for Recovery of Precious Metals. In Advanced Surface Engineering Research; Chowdhury, M.A., Ed.; Intechopen: London, UK, 2018; pp. 127–145. [Google Scholar] [CrossRef] [Green Version]
  15. Viana da Silva, A.F.; Fagundes, A.P.; Macuvele, D.L.P.; Urano de Carvalho, E.F.; Durazzo, M.; Padoin, N.; Soares, C.; Riella, H.G. Green synthesis of zirconia nanoparticles based on Eucleanatalensis plant extract: Optimization of reaction conditions and evaluation of adsorptive properties. Colloids Surf. A Physicochem. Eng. Asp. 2019, 583, 123915. [Google Scholar] [CrossRef]
  16. Navı´o, J.A.; Hidalgo, M.C.; Colo´n, G.; Botta, S.G.; Litter, M.I. Preparation and Physicochemical Properties of ZrO2 and Fe/ZrO2 Prepared by a Sol-Gel Technique. Langmuir 2001, 17, 202–210. [Google Scholar] [CrossRef]
  17. Gurushantha, K.; Anantharaju, K.S.; Nagabhushana, H.; Sharma, S.C.; Vidya, Y.S.; Shivakumara, C.; Nagaswarupa, H.P.; Prashantha, S.C.; Anilkumar, M.R. Facile green fabrication of iron-doped cubic ZrO2 nanoparticles by Phyllanthusacidus: Structural, photocatalytic and photoluminescent properties. J. Mol. Catal. A Chem. 2015, 397, 36–47. [Google Scholar] [CrossRef]
  18. Behbahani, A.; Rowshanzamir, S.; Esmaeilifar, A. Hydrothermal synthesis of zirconia nanoparticles from commercial zirconia. Procedia Eng. 2012, 42, 908–917. [Google Scholar] [CrossRef] [Green Version]
  19. Singh, A.K.; Nakate, U.T. Microwave Synthesis, Characterization and Photoluminiscence Properties of Nanocrystalline Zirconia. Sci. World J. 2014, 2014, 349457. [Google Scholar] [CrossRef] [Green Version]
  20. Dang, Y.Y.; Bhuiyan, M.S.; Paranthaman, M.P. Zirconium oxide nanostructures prepared by anodic oxidation. U.S. Dep. Energy J. Undergrad. Res. 2008, 8, 48–53. [Google Scholar]
  21. Zhu, H.Y.; Liu, B.; Shen, M.M.; Kong, Y.; Hong, X.; Hu, Y.H.; Ding, W.P.; Dong, L.; Chen, Y. Effect of maltose for the crystallization of tetragonal zirconia. Mater. Lett. 2004, 58, 3107–3110. [Google Scholar] [CrossRef]
  22. Nimare, P.; Koser, A.A. Biological Synthesis of ZrO2 Nanoparticle Using Azadirachta Indica Leaf Extract. Int. Res. J. Eng. Technol. 2016, 3, 1910–1912. [Google Scholar]
  23. Al-Zaqri, N.; Muthuvel, A.; Jothibas, M.; Alsalme, A.; Alharthi, F.A.; Mohana, V. Biosynthesis of zirconium oxide nanoparticles using Wrightiatinctoria leaf extract: Characterization, photocatalytic degradation and antibacterial activities. Inorg. Chem. Commun. 2021, 127, 108507. [Google Scholar] [CrossRef]
  24. Shinde, H.M.; Bhosale1, T.T.; Gavade, N.L.; Babar, S.B.; Kamble, R.J.; Shirke, B.S.; Garadkar, K.M. Biosynthesis of ZrO2 nanoparticles from Ficusbenghalensis leaf extract for photocatalytic activity. J. Mater. Sci. Mater. Electron. 2018, 29, 14055–14064. [Google Scholar]
  25. Majedi, A.; Abbasi, A.; Davar, F. Green synthesis of zirconia nanoparticles using the modified Pechini method and characterization of its optical and electrical Properties. J. Sol-Gel Sci. Technol. 2016, 77, 542–552. [Google Scholar] [CrossRef]
  26. Shanthi, S.; Tharani, S.S.N. Green Synthesis of Zirconium Dioxide (ZrO2) Nano Particles Using Acalypha Indica Leaf Extract. Int. J. Eng. Appl. Sci. 2016, 3, 23–25. [Google Scholar]
  27. Hasan, I.M.A.; Tawfik, A.R.; Assaf, F.H. GC/MS screening of buckthorn phytochemicals and their use to synthesize ZnO nanoparticles for photocatalytic degradation of malachite Green dye in water. Water Sci. Technol. 2021, 85, 666. [Google Scholar] [CrossRef]
  28. Huang, Y.; Haw, C.; Zheng, Z.; Kang, J.; Zheng, J.; Wang, H. Biosynthesis of zinc oxide nanomaterials from plant extracts and future green prospects: A topical review. Adv. Sustain. Syst. 2021, 5, 2000266. [Google Scholar] [CrossRef]
  29. Zhao, H.; Ren, S.; Yang, H.; Tang, S.; Guo, C.; Liu, M.; Tao, Q.; Ming, T.; Xu, H. Peppermint essential oil: Its phytochemistry, biological activity, pharmacological effect and application. Biomedi. Pharmacother. 2022, 154, 113559. [Google Scholar] [CrossRef]
  30. Marwa, C.; Fikri-Benbrahim, K.; Ou-Yahia, D.; Farah, A. African peppermint (Menthapiperita) from Morocco: Chemical composition and antimicrobial properties of essential oil. J. Adv. Pharm. Technol. Res. 2017, 8, 86–90. [Google Scholar]
  31. González-Montaña, J.R.; Escalera, F.; Alonso, A.J.; Lumillos, J.M.; Robles, R.; Alonso, M. Relationship between Vitamin B12 and Cobalt Metabolism in Domestic Ruminant: An Update. Animal 2020, 10, 1855. [Google Scholar] [CrossRef]
  32. Jaiswal, A.K.; Kumar, R.; Thakur, A.; Pori, P.; Kumara, R.; Kanojia, R. Cobalt Toxicity/Poisoning with analytical aspects and its Manaagment. Int. J. Med. Lab. Res. 2019, 4, 29–36. [Google Scholar] [CrossRef]
  33. Czarnek, K.; Terpiłowska, S.; Siwicki, A.K. Selected aspects of the action of cobalt ions in the human body. Cent. Eur. J. Immunol. 2015, 40, 236–242. [Google Scholar] [CrossRef] [PubMed]
  34. Gómez-Arnaiz, S.; Tate, R.J.; Grant, M.H. Cobalt neurotoxicity: Transcriptional Effect of elevated cobalt blood levels in the rodent brain. Toxics 2022, 10, 59. [Google Scholar] [CrossRef] [PubMed]
  35. Alloway, B.J.; Ayres, D.C. Chemical Principles of Environmental Pollution; Chapman and Hall: Oxford, UK, 1993. [Google Scholar]
  36. Caramalau, C.; Bulgariu, L.; Macobeanu, M. Kinetic study of cobalt adsorption on peat activated by simple chemical treatment. Envron. Eng. Manag. J. 2009, 8, 1351–1358. [Google Scholar]
  37. Caramalau, C.; Bulgariu, L.; Macobeanu, M. Adsorption Characteristics of Cobalt(II) ions from aqueous solution on Romanian peat moss. Environ. Eng. Manag. J. 2009, 8, 1089–1095. [Google Scholar]
  38. Basahel, S.N.; Ali1, T.T.; Mokhtar, M.; Narasimharao, K. Influence of crystal structure of nanosized ZrO2 on photocatalytic degradation of methyl orange. Nanoscale Res. Lett. 2015, 10, 73. [Google Scholar] [CrossRef] [Green Version]
  39. Manivasakan, P.; Rajendran, V.; Rauta, P.R.; Sahu, B.B.; Panda, B.K. Synthesis of Monoclinic and Cubic ZrO2 Nanoparticles from Zircon. J. Am. Ceram. Soc. 2011, 94, 1410–1420. [Google Scholar] [CrossRef]
  40. Andrade, A.B.; Ferreira, N.S.; Valerio, M.E.G. Particle size effects on structural and optical properties of BaF2 nanoparticles. RSC Adv. 2017, 7, 26839. [Google Scholar] [CrossRef]
  41. Bokuniaeva, A.O.; Vorokh, A.S. Estimation of particle size using the Debye equation and the Scherrer formula for polyphasic TiO2 powder. J. Phys. Conf. Ser. 2019, 1410, 012057. [Google Scholar] [CrossRef]
  42. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC technical report). Pure. Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  43. Cruz-Lopes, L.P.; Macena, M.; Esteves, B.; Guiné, R.P.F. Ideal pH for the adsorption of metal ions Cr6+, Ni2+, Pb2+ in aqueous solution with different adsorbent materials. Open Agric. 2021, 6, 115–123. [Google Scholar] [CrossRef]
  44. Kosmulski, M. The pH dependent surface charging and points of zero charge VIII. Update. Adv. Colloid Interface Sci. 2020, 275, 102064. [Google Scholar] [CrossRef] [PubMed]
  45. Al-Maliky, E.A.; Gzar, H.A.; Al-Azawy, M.G. Determination of Point of Zero Charge (PZC) of Concrete Particles Adsorbents. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1184, 012004. [Google Scholar] [CrossRef]
  46. Abate, G.Y.; Alene, A.N.; Habte, A.T.; Addis, Y.A. Adsorptive removal of basic green dye from aqueous solution using humic acid modified magnetite nanoparticles: Kinetics, equilibrium and thermodynamic studies. J. Polym. Environ. 2021, 29, 967–984. [Google Scholar] [CrossRef]
  47. Azman, A.; Ngadi, N.; Awg Zaini, D.K.; Jusoh, M.; Mohamad, Z.; Arsad, A. Effect of adsorption parameter on the removal of aspirin using tyre waste adsorbent. Chem. Eng. Trans. 2019, 72, 157–162. [Google Scholar]
  48. Gulipalli, C.H.S.; Prasad, B.; Wasewar, K.L. Bach study, Equilibrium and kinetics of adsorption of selenium using rich husk ash (RHA). J. Eng. Sci. Technol. 2011, 6, 586–605. [Google Scholar]
  49. Altintig, E.; Yenigun, M.; Sari, A.; Altundag, H.; Tuzen, M.; Saleh, T.A. Facile synthesis of zinc oxide nanoparticles loaded activated carbon as an eco-friendly adsorbent for ultra-removal of malachite green from water. Environ. Technol. Innov. 2021, 21, 101305. [Google Scholar] [CrossRef]
  50. Al-Musawi, T.J.; Arghavan, S.M.A.; Allahyari, E.; Arghavan, F.S.; Othmani, A.; Nasseh, N. Adsorption of malachite green dye onto almond peel waste: A study focusing on application of the ANN approach for optimization of the effect of environmental parameters. Biomass Convers. Biorefin. 2022, 1–13. [Google Scholar] [CrossRef]
  51. Kim, Y.-H. Adsorption Characteristics of Cobalt on ZrO2 and Al2O3 Adsorbents in High-Temperature Water. Sep. Sci. Technol. 2000, 35, 2327–2341. [Google Scholar] [CrossRef]
  52. Barrow, G.M. Physical Chemistry, 5th ed.; McGrow-Hill: Singapore, 1988. [Google Scholar]
  53. Tang, R.; Hong, W.; Srinivasakannan, C.; Liu, X.; Wang, X.; Duan, X. A novel mesoporous Fe-silica aerogel composite with phenomenal adsorption capacity for malachite green. Sep. Purif. Technol. 2022, 281, 119950. [Google Scholar] [CrossRef]
  54. Wang, J.; Guo, X. Adsorption isotherm models: Classification, physical meaning, application and solving method. Chemosphere 2020, 258, 127279. [Google Scholar] [CrossRef] [PubMed]
  55. Essaadaoui, Y.; Lebkiri, A.; Rifi, E.; Kadiri, L.; Ouass, A. Adsorption of cobalt from aqueous solutions onto Bark of Eucalyptus. Mediterranean. J. Chem. 2018, 7, 145–155. [Google Scholar]
  56. Ramos, S.N.C.; Xavier, A.L.P.; Teodoro, F.S.; Gil, L.F.; Gurgel, L.V.A. Removal of cobalt(II), copper(II), and nickel(II) ions from aqueous solutions using phthalate-functionalized sugarcane bagasse: Mono-and multicomponent adsorption in batch mode. Ind. Crops Prod. 2016, 79, 116–130. [Google Scholar] [CrossRef]
  57. Ketsela, G.; Animen, Z.; Talema, A. Adsorption of Lead (II), Cobalt (II) and Iron (II) From Aqueous Solution by Activated Carbon Prepared from White Lupine (GIBITO) HSUK. J. Thermodyn. Catal. 2020, 11, 203. [Google Scholar]
  58. Rengaraj, S.; Moon, S. Kinetics of adsorption of Co(II) removal from water and wastewater by ion exchange resins. Water Res. 2002, 36, 1783–1793. [Google Scholar] [CrossRef]
  59. Wanga, J.; Guo, X. Adsorption kinetic models: Physical meanings, applications, and solving methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  60. Wang, L.; Shi, C.; Wang, L.; Pan, L.; Zhang, X.; Zou, J. Rational design, synthesis, adsorption principles and applications of metal oxide adsorbents: A review. Nanoscale 2020, 12, 4790–4815. [Google Scholar] [CrossRef]
  61. Mahadevan, H.; Nimina, P.V.M.; Krishnan, K.A. An environmental green approach for the effective removal of malachite green from estuarine waters using Pistaciavera L. shell-based active carbon. Sustain. Water Resour. Manag. 2022, 8, 38. [Google Scholar]
  62. Hasan, I.M.A.; Tawfik, A.R.; Assaf, F.H. Biosynthesis of α-MoO3 nanoparticles and its adsorption performance of cadmium from aqueous solutions. Adv. Nat. Sci. Nanosci. Nanotechnol. 2021, 12, 035007. [Google Scholar] [CrossRef]
  63. Khan, M.I.; Kazmi, S.A. Spectrophotometric determination of cobalt(II). J. Chem. Soc. Pak. 1979, 1, 73–74. [Google Scholar]
Figure 1. XRD patterns of ZrO2NPs570 and ZrO2NPs700.
Figure 1. XRD patterns of ZrO2NPs570 and ZrO2NPs700.
Inorganics 10 00257 g001
Figure 2. TEM images of (a) ZrO2NPs570 (scale bar: 100 nm), (b) ZrO2NPs700 (scale bar: 500 nm), and their respective particle-size distributions (c,d).
Figure 2. TEM images of (a) ZrO2NPs570 (scale bar: 100 nm), (b) ZrO2NPs700 (scale bar: 500 nm), and their respective particle-size distributions (c,d).
Inorganics 10 00257 g002
Figure 3. The nitrogen adsorption/desorption isotherms and the corresponding pore size distribution of (a,b) ZrO2NPs570 and (c,d) ZrO2NPs700. (Blue lines are adsorption and red lines are desorption in (a,c).
Figure 3. The nitrogen adsorption/desorption isotherms and the corresponding pore size distribution of (a,b) ZrO2NPs570 and (c,d) ZrO2NPs700. (Blue lines are adsorption and red lines are desorption in (a,c).
Inorganics 10 00257 g003
Scheme 1. Mechanism for green synthesis of ZrO2NPs using peppermint.
Scheme 1. Mechanism for green synthesis of ZrO2NPs using peppermint.
Inorganics 10 00257 sch001
Figure 4. (a) Effect of pH on adsorption of Co2+ by ZrO2NPs (ZrO2 mass: 0.3 g, vol: 50 mL, Co2+ conc.: 150 ppm, time: 30 min, Temp.: 25 °C) and (b) The point of zero charge of ZrO2NPs570.
Figure 4. (a) Effect of pH on adsorption of Co2+ by ZrO2NPs (ZrO2 mass: 0.3 g, vol: 50 mL, Co2+ conc.: 150 ppm, time: 30 min, Temp.: 25 °C) and (b) The point of zero charge of ZrO2NPs570.
Inorganics 10 00257 g004
Figure 5. Effect of contact time on the % removal of cobalt by ZrO2NPs570 (λ = 667, Volume: 50 mL, Conc.: 50–200 ppm, pH: 6.5, ZrO2 mass: 0.3 g).
Figure 5. Effect of contact time on the % removal of cobalt by ZrO2NPs570 (λ = 667, Volume: 50 mL, Conc.: 50–200 ppm, pH: 6.5, ZrO2 mass: 0.3 g).
Inorganics 10 00257 g005
Figure 6. Effect of contact time on qt during the removal of cobalt by ZrO2NPs570 (λ = 667, Volume: 50 mL, Conc.: 50–200 ppm, pH: 6.5, ZrO2 mass: 0.3 g).
Figure 6. Effect of contact time on qt during the removal of cobalt by ZrO2NPs570 (λ = 667, Volume: 50 mL, Conc.: 50–200 ppm, pH: 6.5, ZrO2 mass: 0.3 g).
Inorganics 10 00257 g006
Figure 7. Effect of (a) ZrO2NPs mass (λ = 667, time: 30 min, volume: 50 mL, Co2+ conc.: 100 ppm, pH: 6.5) and (b) initial Co2+ concentration on its % removal (λ = 667, time: 30 min, volume: 50 mL, ZrO2 mass: 0.3 g, pH: 6.5).
Figure 7. Effect of (a) ZrO2NPs mass (λ = 667, time: 30 min, volume: 50 mL, Co2+ conc.: 100 ppm, pH: 6.5) and (b) initial Co2+ concentration on its % removal (λ = 667, time: 30 min, volume: 50 mL, ZrO2 mass: 0.3 g, pH: 6.5).
Inorganics 10 00257 g007
Figure 8. (a) Langmuir adsorption isotherm, (b) variation in RL values and (c) Freundlich adsorption isotherm.
Figure 8. (a) Langmuir adsorption isotherm, (b) variation in RL values and (c) Freundlich adsorption isotherm.
Inorganics 10 00257 g008
Figure 9. (a) Pseudo 1st order, (b) Pseudo 2nd order, (c) Elovich, and (d) Intraparticle diffusion models (50 mL, 50 ppm Co2+, 0.3 g ZrO2, time 30 min, temp 25 °C).
Figure 9. (a) Pseudo 1st order, (b) Pseudo 2nd order, (c) Elovich, and (d) Intraparticle diffusion models (50 mL, 50 ppm Co2+, 0.3 g ZrO2, time 30 min, temp 25 °C).
Inorganics 10 00257 g009
Scheme 2. Green synthesis of ZrO2NPs from peppermint.
Scheme 2. Green synthesis of ZrO2NPs from peppermint.
Inorganics 10 00257 sch002
Table 1. Crystallite size, BET surface area, and pore dimensions of ZrO2NPs570 and ZrO2NPs700.
Table 1. Crystallite size, BET surface area, and pore dimensions of ZrO2NPs570 and ZrO2NPs700.
ZrO2570ZrO2700
Average crystallite size, nm (XRD)6.277.26
Average particle size, nm (TEM)7.58.1
Surface area, m2/g (BET)94.80562.384
Surface area, m2/g (BJH)80.32961.222
Mean pore diameter, nm (BET)3.9227.588
Mean pore diameter, nm (BJH)3.927.32
Total pore volume, cm3/g (BET)0.092960.1183
Pore volume, cm3/g (NLDFT)0.06980.06098
Pore volume, cm3/g (BJH)0.085130.1164
Table 2. The calculated adsorption isotherm parameters for Co2+ adsorption by ZrO2.
Table 2. The calculated adsorption isotherm parameters for Co2+ adsorption by ZrO2.
T, KLangmuirFreundlich
qmKLR2nKFR2
2988.9950.0390.9733.8102.4200.840
Table 3. Calculated kinetic model parameters for Co adsorption by ZrO2.
Table 3. Calculated kinetic model parameters for Co adsorption by ZrO2.
C0Pseudo 1st OrderPseudo 2nd OrderElovichqeexp
K1qecalcR2K2qecalcR2βαR2
500.0786.2070.9550.01259.2250.9760.96131.7490.9935.469
1000.0528.6690.9010.00912.5000.9770.6172.3160.9657.258
1500.0525.7440.6760.01013.6430.9890.6682.0480.9748.523
2000.0846.1870.8180.014114.0450.9801.3480.8320.9998.995
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hasan, I.M.A.; Salah El-Din, H.; AbdElRaady, A.A. Peppermint-Mediated Green Synthesis of Nano ZrO2 and Its Adsorptive Removal of Cobalt from Water. Inorganics 2022, 10, 257. https://doi.org/10.3390/inorganics10120257

AMA Style

Hasan IMA, Salah El-Din H, AbdElRaady AA. Peppermint-Mediated Green Synthesis of Nano ZrO2 and Its Adsorptive Removal of Cobalt from Water. Inorganics. 2022; 10(12):257. https://doi.org/10.3390/inorganics10120257

Chicago/Turabian Style

Hasan, Ibrahem Mohamed Abouzeid, Hanan Salah El-Din, and Ahmed A. AbdElRaady. 2022. "Peppermint-Mediated Green Synthesis of Nano ZrO2 and Its Adsorptive Removal of Cobalt from Water" Inorganics 10, no. 12: 257. https://doi.org/10.3390/inorganics10120257

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