Anastassiya A. Mashentseva*ab,
Nurzhigit Seitzhaparab,
Murat Barsbayc,
Nurgulim A. Aimanovaa,
Assel N. Alimkhanovaab,
Dmitriy A. Zheltova,
Alisher M. Zhumabayevab,
Bakhtiyar S. Temirgazievd,
Alimzhan A. Almanovab and
Daniyar T. Sadyrbekovd
aThe Institute of Nuclear Physics of the Republic of Kazakhstan, 050032 Almaty, Kazakhstan. E-mail: a.mashentseva@inp.kz
bDepartment of Nuclear Physics, New Materials and Technologies, L.N. Gumilyov Eurasian National University, 010008 Astana, Kazakhstan
cDepartment of Chemistry, Hacettepe University, 06800 Ankara, Turkey
dNPJSC E.A. Buketov Karaganda University, 100024 Karaganda, Kazakhstan
First published on 7th September 2023
This study investigates the sorption removal of lead(II) ions using zinc oxide (ZnO) and copper(II) oxide (CuO) nanoparticles synthesized through a wet burning method with the aid of plant extract from Serratula coronata L. The effect of plant collection time on polyphenol content was investigated and optimal conditions were determined. The structural and chemical properties of the nanoparticles were studied by scanning electron microscopy, energy dispersive analysis, X-ray phase analysis, and X-ray photoelectron spectroscopy. A comparative analysis of lead ion sorption on the surface of synthesized nanoparticles was conducted. The kinetic study revealed that the sorption process follows a pseudo-second-order mechanism, and the Freundlich sorption model provides a better fit for the experimental data. ZnO and CuO nanoparticles exhibited significant sorption capacities, with values of 163.6 and 153.8 mg g−1, respectively.
Among the variety of metal nanoparticles (NPs) obtained using plant extracts, nanoparticles of Cu, Ag, Co, Ni and oxides such as CuO, ZnO, NiO, Fe3O4 have garnered significant research attention due to their straightforward synthesis techniques and distinctive dimensional physicochemical properties.5,6 These NPs hold immense potential for biomedical application7,8 as well as various fields of materials science.9–11 Additionally, they find extensive use as effective catalysts and sorbents for wastewater remediation, targeting diverse classes of pollutants such as nitrophenols, organic dyes, pesticides, pharmaceuticals, and others.12–14 Among the aforementioned persistent pollutants, the presence of heavy metals in water poses a substantial global concern due to their significant contribution to environmental degradation.15 Contaminated drinking water containing heavy metals such as arsenic, cadmium, nickel, mercury, chromium, zinc, and lead has emerged as a major health concern for the public and health care professionals alike.16,17 According to the Environmental Protection Agency (EPA), heavy metal ions are designated as priority pollutants, necessitating their elimination or reduction from any water bodies that may or may not interact with the environment.18,19 The accumulation of heavy metals in living systems without degradation leads to harmful levels of exposure, while heavy metal waste pollutes water and soil surfaces, posing adverse effects on the health and continuity of all living species.17,20
A diverse array of conventional techniques has been employed for the removal of heavy metal ions from water, and several of these methods, such as precipitation, ion-exchange, reverse osmosis, membrane separation, have already demonstrated successful results.21,22 Among these approaches, adsorption stands out as one of the most promising and widely used methods due to its simplicity of operation, high efficiency, and economic benefits. Recently, there has been considerable interest in exploring the potential of biogenic nanomaterials as effective sorbents for the removal of heavy metal ions.23,24 Various plant sources, including extracts from leaves, flowers, seeds, fruits, peels, and roots, have not only been extensively utilized for the synthesis of NPs but also hold promise as supporting materials for biogenic composites.25
The Republic of Kazakhstan boasts a rich flora, with over 200 endemic plants. Among these is Serratula coronata L. CR (SCR), a forest-steppe species belonging to the Asteraceae Dumort family, which is widely distributed in Eastern Europe, Western Siberia, Central and Northern Kazakhstan.26 The aerial part (leaves, stems) of SCR, found in Central Kazakhstan (Karaganda region) exhibits a significant enrichment of various phytoecdysteroids and flavonoids.27–29 Notably, the phytoecdysones derived from SCR promise as potential agents for the treatment of different skin diseases,30–32 while flavonoids demonstrate noteworthy antioxidant activity.33 Despite the abundant presence of secondary metabolites in SCR, which are known to be capable of reducing metal ions, this plant remains relatively understudied as a potential precursor for the biogenic synthesis of catalytically active metal oxide nanoparticles.34,35
In recent studies, various biogenic nanomaterials have been explored for the removal of lead (Pb) ions, yet the utilization of indigenous plant sources to develop environmentally friendly wastewater treatment technologies remains a pressing objective for every nation. This study aims to provide essential insights into the optimal conditions for the collection and extraction of SCR, and subsequently, to investigate the efficacy of biogenic ZnO and CuO nanosized adsorbents in the removal lead(II) ions from aqueous media under bath mode conditions. A comprehensive analysis of the physical and chemical properties of the NPs will be undertaken, and a thorough investigation of the kinetics and equilibrium sorption of Pb(II) ions will be conducted.
A pre-weighed sample of the plant material was placed inside a 250 mL round-bottom flask, which was equipped with a reflux condenser. The flask was then filled with the appropriate type of extractant. Subsequently, the flask was thoroughly mixed, and the mixture inside was heated using a mantle heater until the solvent reached its boiling point. The extraction process was allowed to proceed for a maximum duration of 3 hours. After completion of the extraction, the liquid extract was cooled to room temperature and then subjected to thickening through vacuum distillation following the filtration process. To ensure the most effective removal of any remaining extractant residues, the obtained extract was subjected to further evaporation in a water bath at a controlled temperature of 35–40 °C for a period of 6–7 h.
The crystal structure of the nanoparticles was investigated using a D8 Advance diffractometer (Bruker, Germany) in the angular range of 2θ 30–80°, with a step size of 2θ = 0.02° and a measuring time 1 s. The tube mode settings were 40 kV and 40 mA. The average size of crystallites was determined by analyzing the broadening of X-ray diffraction reflections using the Scherer equation.
For the X-ray photoelectron spectroscopy (XPS) measurements, a Thermo Scientific K-Alpha spectrometer (Waltham, MA, USA) was employed equipped with a monochromatized Al Kα X-ray source (1486.6 eV photons). Core-level spectra were recorded with a constant dwell time of 100 ms and pass energy of 30 eV, while survey spectra were obtained with pass energy of 200 eV. The step size for core-level spectra was 0.1 eV, and for survey spectra, it was 1.0 eV. The analysis chamber maintained a pressure of 2 × 10−9 Torr or lower. The binding energy (BE) values were referenced to the C 1s peak at 285 eV. Data processing was performed using Avantage software (version 5.41, 2019, Waltham, MA, USA).
The determination of the surface charge of the adsorbent based on the pH value was carried out by studying the pHPZC (point of zero charge) value within the pH range of 3.0 to 9.0, following the method described in ref. 34. For this, 10 mL of a 0.01 M NaCl solution was adjusted to the desired pH value (pHi) using 0.1 M HCl(aq) or NaOH(aq). Subsequently, 50 mg of ZnO NPs powder was added to each flask and shaken on a shaker for 12 h at room temperature. Afterward, the NPs were separated from the solution through filtration, and the final pH value (pHf) of the resulting filtrate was measured.
The quantity of Pb(II) adsorbed per unit mass of the sorbent (Qe, mg g−1) was determined using the eqn (1):40
(1) |
The influence of pH on Pb(II) adsorption was studied within the pH range of 3 to 9, while keeping other parameters constant (initial Pb(II) concentration: 50 ppm, adsorbent dose: 50 mg, contact time: 300 min). The pH of the solution was adjusted dropwise using 1.0 M HCl(aq) or 1.0 M NaOH(aq). pH was measurements were performed using a digital pH meter, HANNA HI2020-02 (HANNA Instruments, Smithfield, VA, USA). All experiments were conducted in triplicate. Upon completion of the sorption process, the solution was decanted by centrifugation at 12000 rpm, and the NPs were washed with deionized water, dried, and stored in a dry location.
In our studies, we opted to use ethyl alcohol and its 70% aqueous solution as the extractants, considering that ethanol is one of the main universal food extractants utilized for the extraction of biologically active compounds found in plants.44 Additionally, it is important to note that ethanol is comparatively less toxic to the human body compared to other organic solvents. For the purpose of conducting a comparative analysis of the polyphenol extraction efficiency from the studied plants, an alkaline solution of sodium bicarbonate (baking soda) was employed as an extractant. This choice was motivated by the widespread utilization of sodium bicarbonate in both food and household applications, rendering it a viable alternative option for extraction in our research.
As can be seen from the data presented in Table 2, the highest extraction efficiency by weight was achieved with aqueous ethanol solution (EtOH-70), while the lowest extract yield was obtained during extraction with 0.1 M NaHCO3 solution, owing to the limited solubility of many lipophilic compounds in aqueous solutions of inorganic salts. Comparatively, extraction with ethanol resulted in a lower degree of separation of the slurry extract than with a water–ethanol mixture.
Time of plant materials collection | Solvent | Extraction temperature, °C | Weight of final extract, g |
---|---|---|---|
Beginning of vegetation phase | EtOH | 78.0 | 3.91 |
EtOH:water (70:30) | 80.0 | 5.70 | |
0.1 M NaHCO3 | 101.2 | 2.63 | |
Vegetation phase | EtOH | 78.0 | 4.2 |
EtOH:water (70:30) | 80.0 | 5.66 | |
0.1 M NaHCO3 | 101.2 | 1.86 | |
Beginning of budding | EtOH | 78.0 | 3.75 |
EtOH:water (70:30) | 80.0 | 6.27 | |
0.1 M NaHCO3 | 101.2 | 2.58 |
The concentration of flavonoids in the studied plant extracts was determined by HPLC via the method of absolute calibration peak area. This involved comparing the peak area of the analyte with the peak area of a standard analyte sample with a pre-derminated concentration. Although the highest amount of slurry extract (i.e. the total extractable substances) was observed during the processing of plant materials harvested at the beginning of the budding stage, the content of flavonoids in this type of material was found to be the lowest, as shown by HPLC analysis. Previous studies have highlighted that the presence of a large number of flavonoids and phenolic compounds in the composition of plant raw materials influences their reducing properties during the synthesis of metal nanoparticles and their oxides.45–47 Flavonoids, which exhibit proton-donor properties, along with phenolic compounds known for their metal ion chelating capabilities, have been shown to play a crucial role in reducing Cu2+ ions to metallic copper and copper(I) oxide. These compounds can then be oxidized to form CuO by thermal annealing.35,48 A similar influence of flavonoids on the formation of ZnO nanoparticles has been described elsewhere.49,50
To determine the total amount of phenolic compounds and specific phenolic substances, such as phenolic acids, in plant materials, various spectrophotometric methods are utilized. The Folin–Ciocalteu method, initially developed by Folin and Ciocalteu and later modified by Singleton Rossi Jr, is commonly employed for analyzing phenolic compounds. This method relies on the reduction of phenols with a phosphomolybdic acid reagent. In an alkaline medium, phenolic compounds undergo oxidation, leading to the formation of superoxide ions, which, in turn, react with molybdate to produce molybdenum oxide MoO4+, exhibiting intense absorption at 725 nm.38 In this study, the content of phenolic compounds in the extracts was determined in terms of equivalents of specific phenolic acids, namely gallic, protocatechuic, and caffeic acids. Calibration curves were plotted for each standard substance (Fig. S2†). Table 3 presents the data on the content of phenolic compounds in the extracts of SCR expressed in terms of mg per 1.0 g of raw material. As evident from Table 3, the most enriched extract of SCR with flavonoids and phenolic acids was prepared from the plants collected during the beginning of the vegetation phase (mid-May). This specific type of extract will be used in the subsequent experiments concerning the synthesis of ZnO and CuO NPs.
Time of collection of plant materials | Solvent | Flavonoids content, mg g−1 | Phenolic compounds content, mg g−1 | ||||
---|---|---|---|---|---|---|---|
Quercetin | Rutin | Pinostrobin | Gallic acid | Protocatechuic acid | Caffeic acid | ||
Beginning of vegetation phase | EtOH | 2.77 | 0.60 | 0.10 | 0.038 | 0.062 | 0.034 |
EtOH-70.0 | 2.59 | 0.89 | 0.12 | 0.112 | 0.174 | 0.104 | |
NaHCO3 | 1.86 | 0.33 | 0.01 | 0.058 | 0.097 | 0.056 | |
Vegetation phase | EtOH | 2.01 | 0.45 | 0.29 | 0.110 | 0.188 | 0.112 |
EtOH-70.0 | 0.02 | 0.06 | 0.00 | 0.032 | 0.052 | 0.028 | |
NaHCO3 | 1.50 | 0.19 | 0.02 | 0.110 | 0.186 | 0.111 | |
Beginning of budding | EtOH | 0.32 | 0.00 | 0.00 | 0.110 | 0.187 | 0.112 |
EtOH-70.0 | 0.67 | 0.04 | 0.00 | 0.117 | 0.199 | 0.119 | |
NaHCO3 | 0.38 | 0.00 | 0.00 | 0.116 | 0.197 | 0.118 |
Fig. 1 Electron microphotographs of ZnO (a) and CuO (b) nanoparticle powders, TEM images of ZnO (c) and CuO (d) nanoparticles, X-ray diffraction spectra of ZnO (e) and CuO (f) nanoparticles. |
The X-ray diffractogram of ZnO nanoparticles (Fig. 1e) reveals distinct diffraction peaks corresponding to the ZnO phase at 2θ values of 32.54° (110), 35.58° (002), 38.86° (200), 40.43° (112), 48.57° (−202), 53.65° (020), 58.39° (202), 61.57° (−113), and 68.07° (113), in accordance with previous studies.54–56 These identified planes closely match the JCPDS file (JCPDS: 01-007-2551), indicating the hexagonal wurtzite structure (symmetry group P62mc (186)) of the ZnO nanoparticles. Similarly, the X-ray diffraction pattern of CuO nanoparticles (Fig. 1f) exhibits a series of diffraction peaks characteristic of the monoclinic structure of copper(II) oxide, located at 2θ values of 32.54°, 35.58°, 38.86°, 40.43°, 48.573°, 53.65°, 58.40°, 61.57°, 66.05°, 68.07°, 72.36°, and 75.16°, which were assigned to (110), (002), (200), (112), (−202), (020), (202), (−113), (−311), (113), (311) and (004) respectively. No other characteristic peaks were observed except for those for CuO (JCPDS: 45-0937), indicating the as-obtained CuO NPs has high-purity.
The obtained X-ray diffraction spectra indicate the crystalline nature of the nanoparticles. The average size of CuO crystallites, calculated using the Scherer equation, was found to be 25.4 ± 3.9 nm, and for ZnO nanoparticles, it was 44.6 ± 5.9 nm. All synthesized nanoparticles demonstrated a high degree of crystallinity, with 84.0% for CuO nanoparticles and 91.2% for ZnO nanoparticles, as determined by approximating the values of the full width at half maximum (FWHM) lines using symmetric pseudo-Voigt functions.57
The wide energy range X-ray scans of each nanoparticle (Fig. 2) show that their chemical composition primarily consists of two basic elements, i.e. oxygen and the corresponding metal atom (Cu or Zn). In addition to these anticipated components, the presence of K and Cl elements is noteworthy, constituting about 5% of the composition, particularly in the case of CuO, where they are more prominent as impurities. These two elements also appear in the spectrum of ZnO, albeit in smaller quantities. The amount of carbon, a common element in both nanoparticle compositions, was determined as 9.1%, and 8.2% in the survey scans of CuO and ZnO, respectively. Furthermore, the high-resolution C 1s spectra of the as-synthesized CuO and ZnO exhibit a primary peak at 284.5 eV and another peak at a higher energy level around 288.1 eV. These C 1s peaks are commonly observed in XPS spectra on the surface of nearly all samples, attributed surface contaminations, and are referred to as adventitious carbon.58–60 The C 1s peak component with a binding energy (BE) of 284.5 eV corresponds to C–C bonds, while the component at 288.1 eV is assigned to O–CO bonds.60 The detection of the C element in the XPS spectra of nanoparticles is a crucial finding, providing evidence for the involvement of organic residues from plant extracts in both formation and stabilization of nanoparticles. Furthermore, it aligns well with the earlier discussed EDX results.
The Cu 2p core-level XPS spectrum exhibits distinctive features that differentiate it from both Cu2O and metallic Cu. Notably, dominant shake-up peaks are observed at the higher BE side of the Cu 2p3/2 and Cu 2p1/2 peaks, indicating the existence of an unfilled Cu 3d9 shell. This observation confirms the presence of Cu2+ in the CuO sample.61,62 Moreover, the peaks at 953.9 eV and 933.4 eV in the core level spectra of Cu 2p can be assigned to Cu2+ 2p1/2 and Cu2+ 2p3/2 of CuO, respectively.62 Additionally, the core-level XPS spectrum for O 1s in CuO nanoparticles reveals two components at approximately 529.1, 531.0 eV. The more intense peak at 529.1 eV is attributed to the binding energy of lattice oxygen in CuO lattice, consistent with the binding energy of O2− ions in metal oxide sites (Cu2+–O2−).62,63 The second peak at 530.78 eV can be assigned to the binding energy of oxygen defects/vacancies within the matrix of CuO.57,62 These results align well with previous works63,64 and provide strong evidence of the CuO structure. Similarly, the high-resolution XPS spectrum of the Zn 2p region in ZnO nanoparticles displays two fitting peaks located at approximately 1043.6 and 1020.5 eV, corresponding to Zn 2p1/2 and Zn 2p3/2, respectively.34 Furthermore, the XPS spectrum of the O 1s region in ZnO nanoparticles shows an asymmetrical peak centered at about 529.6 eV. These results from the core-level Zn 2p and O 1s spectra are in good agreement with previous data and confirm the ZnO structure.65,66 In conclusion, based on the discussed XPS data, the successful synthesis of both nanoparticle structures (CuO and ZnO) is confirmed, and the desired structures have been achieved.
At a pH of 5.5, the removal efficiency of Pb2+ by ZnO and CuO nanoparticles is found to be 21.9% and 37.4%, respectively. However, at a pH of 7.0, there is a notable increase in the sorption of lead ions by CuO, reaching 78.2% in the case of CuO. Nevertheless, several previous studies have shown that at pH levels above 6.0, additional deposition of Pb2+ ions occurs due to the formation of lead hydroxide, and at pH 9.0, nearly complete disappearance of Pb2+ ions in the solution is observed.69,70 Therefore, conducting sorption tests at pH levels lower than 6.0 is deemed more appropriate.
The parameter known as pH zero charge point (pHPZC) determines the pH at which the adsorbent's total surface charge becomes zero. The study of pHPZC is important for revealing the adsorption mechanism and explaining the interactions between the sorbent and adsorbate.71 Fig. 3b shows the relationship between pHfinal on pHinitial, and the pHPZC value obtained for biogenic ZnO NPs is approximately 7.1. This result indicates a positive surface charge under the conditions studied and agrees well with pHPZC values reported in previous studies.72–74
Regarding the change in the sorption capacity of nanoparticles over time (Fig. 3c), it was observed that biogenic CuO NPs reach saturation after 180 minutes of sorption, while ZnO NPs reach saturation later, after 240 minutes. Subsequently, the equilibrium sorption capacity (qe) for Pb(II) sorption from a solution with a concentration of 50 ppm was deduced based on the experimental data (Fig. 3d).
For a comprehensive understanding of the adsorption mechanism of new sorbents, kinetic studies are of paramount importance kinetic studies of paramount importance. In this study, we used three kinetic models to predict the kinetics of Pb(II) adsorption on bigenic NPs. The models used were the pseudo-first and pseudo-second order, as well as the Elovich model.75–79 The pseudo-first-order rate model, originally proposed by Lagergren,80 is among the early models of sorption kinetics, focusing on sorption capacity and describing sorption from a liquid medium by solid sorbents. The linear form of the pseudo-first-order model is provided in Table 4. From the graphical relationship ln(qe − qt) = f(t) (Fig. 4a), the values of k1 and the correlation coefficient R2 were determined and are listed in Table 1. However, it is important to note that the observed low correlation coefficient value for the pseudo-first-order kinetic model suggests that the kinetics of Pb(II) adsorption by the biogenic nanoparticles of metal oxides obtained do not conform to the pseudo-first-order kinetic model. Therefore, further investigation and application of alternative kinetic models are necessary to elucidate the kinetics of the sorption process accurately.
Fig. 4 Kinetics of Pb(II) sorption by biogenic NPs according to the pseudo-first (a), pseudo-second (b), and the Elovich (c) kinetic models. |
To describe the sorption process, it is believed that the first-order kinetic model is applicable at the initial stage, where the concentration of sorbate ions reaching the functional groups of the sorbent is significantly lower than the active centers of the sorbent.81 Consequently, the order of the reaction can be mathematically reduced by incorporating the concentration of functional groups of the sorbent into the reaction rate constant. However, at later stages of the sorption process, the reaction rate is affected by the product of the concentrations of both components of both components, leading to a formal reaction order of two.82
A pseudo-second-order equation is considered suitable for describing the sorption process if the reaction between the adsorbate and the functional groups of the adsorbent proceeds strictly stoichiometrically, where one metal ion occupies one sorption position.83 This model shares similarities with the pseudo-first-order kinetics model but is more comprehensive in describing the entire sorption process.
The rate constant k2, the value of qe, and the corresponding linear regression correlation coefficient R2 were calculated from the linear plot of the dependence t/qt = f(t) as shown in Fig. 4b and are presented in Table 4. The calculated qe values for ZnO and CuO were 169.45 and 178.57 mg g−1, respectively, and in agreement with the experimental data (Fig. 3g) for the pseudo-second-order kinetics.
The pseudo-second-order equation accounts not only for sorbate–sorbent interactions but also for the intermolecular interactions of the adsorbed substances. On the other hand, the Elovich model considers the contribution of both sorption processes and desorption phenomena to the kinetics of substance extraction, with desorption gaining significance as equilibrium is approached.84 The kinetic parameters of the Elovich model were calculated from the linear dependence of qt on ln(t) (Fig. 4c) and are provided in Table 4. Higher values of α compared to β indicate a faster absorption rate compared to lead ions' desorption.85
The correlation coefficient values (R2) indicate that the most appropriate description of the sorption of lead(II) ions on the studied sorbents is achieved using the pseudo-second order model. This is evident from the linear nature of the graphical dependencies of this model in Fig. 3b, which remains consistent throughout the entire time interval. The applicability of the pseudo-second-order model to the biogenic sorbents suggests that the chemisorption stage governs the limiting step of the process, with the influence of the diffusion stage being insignificant.86
According to Langmuir's adsorption theory, adsorption takes place on equivalent, localized active centers, with each center accommodating only one molecule. Saturation of adsorption occurs as these active centers become filled. The adsorbed molecules do not interact with one another and desorb after a certain time, establishing a dynamic equilibrium.87,88 The linear form of the Langmuir adsorption isotherm equation, derived from molecular-kinetic theory and the concept of the monomolecular nature of the adsorption process, is presented in Table 5. The parameters b and Q0 characterize the adsorbent–adsorbate pair. Fig. 5a illustrates the graph of the dependence Ce = f(Ce/Qe). The values of Q0 and b were calculated from the slope and intersection of the lines on the graph in the corresponding coordinates of the linear isotherm equation, and are listed in Table 5.
Fig. 5 Adsorption isotherms of Pb(II) on biogenic CuO and ZnO nanoparticles in linear coordinates of Langmuir (a) and Freundlich (b) isotherms. |
The Freundlich model isotherm equation describes adsorption on a heterogeneous surface, where the adsorption centers possess varying energy levels. Initially, the active sorption centers with maximum energy become filled. Fig. 4b shows the experimental data on lead(II) ion adsorption in the linear Freundlich equation lnqe = f(lnCe) coordinates. The constant n is an empirical parameter related to the adsorption intensity, and values of ‘n’ in the range of 1–10 indicate favorable adsorption.89 The kF values for the studied nanoparticle samples were 387.5 and 235.4 mg g−1 for CuO and ZnO NPs, respectively, indicating an enhanced adsorption capacity of copper oxide-based nanoparticles. This observation is in line with the data obtained from the Langmuir isotherm. For all studied samples, the values of the Freundlich constant (n) significantly exceed unity, indicating the feasibility of Pb(II) adsorption on the nanoparticle surface. The regression coefficients for the linear plots are close to unity, suggesting that the experimental adsorption data are in good agreement with the Freundlich adsorption isotherm.
Table 6 presents data on the sorption capacity of various nanosized sorbents for lead ions, as reported in earlier works. It is important to note that direct comparison of data from different studies can be challenging due to variations in factors influencing sorbent capacity, such as the mixing rate, pH and temperature. Nevertheless, the presented data indicate that biogenic ZnO and CuO nanoparticles synthesized using plant raw materials from SCR exhibit higher sorption capacity compared to existing analogues, rendering them promising materials for effective sorption removal of lead ions from aqueous media.
Sorbent | Sorbent testing conditions | Qe, mg g−1 | Rate constant k2, g mg−1 min−1 | Ref. | |||
---|---|---|---|---|---|---|---|
Initial sorbate concentration, mg L−1 | Sorbent weight, g | Aliquot volume, mL | pH | ||||
CuO NPs (microwave synthesis) | 50.0 | 0.1 | 50.0 | — | 125.0 | 0.0235 | 90 |
CuO NPs (magnetron sputtering) | 1000.0 | 2.0 | 100.0 | 6.0 | 37.03 | 0.0054 | 91 |
CuO nanorods | 1000.0 | 1.0 | 100.0 | 8.5 | 3.31 | — | 92 |
ZnO NPs (biogenic synthesis) | 25.0 | 0.1 | 25.0 | 5.0 | 19.65 | — | 24 |
Magnetic NPs modified with chitosan | 100.0 | 0.1 | 100.0 | 6.0 | 498.6 | 0.0002 | 93 |
Magnetic NPs Fe3O4, modified with polyethyleneimine | 50.0 | 1.0 | 100.0 | 5.0 | 33.65 | 0.023 | 94 |
CuO NPs (biogenic synthesis) | 50.0 | 0.05 | 15.0 | 5.0 | 163.6 | 0.00019 | This study |
ZnO NPs (biogenic synthesis) | 153.8 | 0.00021 |
X-ray diffraction analysis of the NPs' crystalline structure and phase composition showed the formation of a single metal oxide phase in both cases, with monoclinic ZnO and cubic CuO phases exhibiting high purity and a narrow size distribution. Characterization techniques, such as XRD, SEM, XPS, TEM and EDS, confirmed the desired structure and composition of the synthesized NPs.
The sorption removal of lead ions from aqueous solutions using the biogenic NPs was thoroughly studied, and the impact of pH on the removal efficiency of Pb(II) was investigated. The adsorption behavior of Pb(II) on the surface of ZnO and CuO NPs was analyzed using Langmuir and Freundlich isotherm models, providing insights into the underlying mechanisms of the sorption process. Comparative analysis based on the mean-square deviations (R2) indicated that the Freundlich model best describes the adsorption of Pb(II) on the surface of ZnO and CuO NPs. Furthermore, the adsorption kinetics of Pb(II) ions onto the NPs' surface was investigated using the Elovich rate equation, pseudo-first-order, and pseudo-second-order kinetic models. The results showed that the pseudo-second-order model provided a better fit for both NPs, suggesting that the rate-limiting step in the adsorption of Pb(II) ions can be attributed to chemical interactions between the metal ions and the functional groups present on the surface of NPs.
The findings of this study contribute to the understanding of the potential applications of biogenic NPs in the removal of lead ions from aqueous media, paving the way for further exploration and development of efficient sorption methods for water purification and environmental remediation.
SCR | Serratula coronata L. |
NPs | Nanoparticles |
XRD | X-ray diffraction |
XPS | X-ray photoelectron spectroscopy |
SEM | Scanning electron microscopy |
EDX | Energy dispersive X-ray analysis |
TEM | Transmission electron microscopy |
HPLC | High-performance liquid chromatography |
FWHM | Full width at half maximum |
Qe | Amount of Pb(II) adsorbed by the unit mass of copper (mg g−1) |
C0 | Feed Pb(II) concentration (mg L−1) |
Ce | Concentration of Pb(II) in aliquots (mg L−1) |
DC | Degree of crystallinity (%) |
L | Average crystallite size (nm) |
qt | Adsorption capacity at time t (mg g−1) |
α | Initial rate of the adsorption process, mg g−1 min−1 |
β | Desorption constant (g mmol−1) |
Ra | Roughness of the composite (nm) |
b | Constant related to the energy of adsorption (i.e., Langmuir constant (L μg−1)) |
Ce | Equilibrium concentration of adsorbate (mg L−1) |
Q0 | Maximum monolayer coverage capacity (mg g−1) |
kF | Freundlich isotherm constant related to the adsorption capacity (μg g−1) |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ra05347d |
This journal is © The Royal Society of Chemistry 2023 |