Next Article in Journal
Graphene Family Nanomaterial Reinforced Magnesium-Based Matrix Composites for Biomedical Application: A Comprehensive Review
Next Article in Special Issue
Dilution Ratio and the Resulting Composition Profile in Dissimilar Laser Powder Bed Fusion of AlSi10Mg and Al99.8
Previous Article in Journal
Effect of Pass Strain on the Microstructure, Texture and Mechanical Properties of AZ31 Magnesium Alloy Fabricated by High Strain Rate Multiple Forging
Previous Article in Special Issue
Numerical and Experimental Investigation of Germanium Refining via Fractional Crystallization Based Innovative Rotary Cooling Device
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis, Characterization, and Adsorption Insights of Lanthanum Oxide Nanorods

by
Lakshmi Prasanna Lingamdinne
,
Janardhan Reddy Koduru
*,
Yoon-Young Chang
,
Seon-Hong Kang
and
Jae-Kyu Yang
*
Department of Environmental Engineering, Kwangwoon University, Seoul 01897, Korea
*
Authors to whom correspondence should be addressed.
Metals 2020, 10(8), 1001; https://doi.org/10.3390/met10081001
Submission received: 18 May 2020 / Revised: 16 July 2020 / Accepted: 22 July 2020 / Published: 24 July 2020
(This article belongs to the Special Issue 10th Anniversary of Metals: Metallurgy and Metal Technology)

Abstract

:
This study synthesized lanthanum oxide (La2O3) nanorods to develop a practical approach for the removal of arsenic from groundwater. La2O3 nanorods were synthesized by a simple hydrothermal process followed by calcination at 500 °C and were characterized by spectroscopic and microscopic techniques. To evaluate the adsorption mechanism of La2O3 nanorods, adsorption parameters including solution pH, temperature, equilibrium isotherms, and kinetics for arsenic were studied. The results suggested that the arsenic uptake was a rate-limiting, monolayer adsorption interaction on the La2O3 nanorods homogeneous surface. In addition, it was found that the adsorptive removal behavior of La2O3 for As(V) was sensitive to the initial pH and temperature, and the maximum uptake amount of as prepared La2O3 was found to be 260.56 mg/g of As(V) at pH 6.0 and 25 °C. Furthermore, the uptake capacity of La2O3 nanorods for As(V) increased with temperature. The resultant thermodynamic parameters (ΔG0, ΔH0, and ΔS0) suggested an endothermic adsorption of As(V) on La2O3. The adsorption capacity of La2O3 was higher than that of several reported nanocomposites, suggesting its practical applicability and novelty for As-contaminated wastewater treatment.

1. Introduction

Lanthanides are widely used in hydrogen storage, electrodes, gate insulators, adsorbents, and superconductors, owing to their versatility and multifunctionality [1,2,3,4]. In lanthanide series, lanthanum (La) is one of the lightest elements that has been widely used in various applications, including optoelectronic devices [5], phosphors [6], solid electrolytes [7], catalysts [8,9], sorbents [10], and gas sensors [11] in the form of oxide, carbonate, and phosphate. Currently, lanthanum oxide (La2O3) is used in various applications, including water treatment [12,13], catalytic exhaust gas converters [14], and high-k gate dielectric materials [15]. Various methods have been adopted for the preparation of La2O3 nanostructures, including precipitation, thermal decomposition, hydrothermal, sol gel, and pyrolysis methods [13,15,16,17,18,19,20,21,22,23,24]. The adsorption properties of La2O3 mainly depend on its texture and surface morphology of materials. With this, there is still a research gap in developing a novel surface texture and morphology for La2O3 in water treatment. Thus, this article focuses on the synthesis of nanostructured La2O3 through simple precipitation followed by calcination at 500 °C and its adsorptive removal ability for As(V) in groundwater.
Arsenic is a highly toxic pollutant in the environment, occurring through natural and anthropogenic sources. Among existing arsenic forms in the environment, inorganic arsenic is more prevalent in +3 or +5 oxidation states as arsenate (H2AsO4, HAsO42−) or arsenite (AsO33−), respectively, under normal conditions of natural water [25,26]. The species H2AsO4 and HAsO42− are of great concern due to their high toxicity in natural water sources including drinking water [27]. Arsenic contamination of drinking water has been a global issue occurring in many countries including India, Bangladesh, Malaysia, Mexico, Hungary, New Zealand, USA, Spain, Japan, Canada, Taiwan, and Mainland China [27,28,29]. The exposure of arsenic for long-term periods may lead to lung, bladder, kidney, and skin cancer as well as changes in pigmentation [25,30]. The treatment of arsenic-contaminated wastewater is necessary. Some of researchers has been used lanthanum oxides for arsenic removal and those reported that the surface morphology, surface functions along with surface area and oxyanions are influence the adsorptive removal of lanthanum oxides [31,32]. Hence, we prepared lanthanum oxide with a specific morphology in this study.
Herein, we demonstrate a simple synthesis of La2O3 nanorods 18 nm wide and approximately 160–200 nm long. The physical and chemical properties and purity of the obtained products were evaluated using advanced instrumental techniques including XRD, SEM, TEM, XPS, and FT-IR. Influencing parameters, including pH, kinetics, isotherms, and coexisting anions and cations, on the adsorptive removal of As(V) by La2O3 were studied. Further, we carried out thermodynamic studies. The results suggest that the adsorptive removal of As(V) was an endothermic rate-limiting monolayer interaction on the homogeneous surface of La2O3 nanorods. The maximum uptake of as prepared La2O3 is 260.56 mg arsenic/g-La2O3 at 25 °C, and pH 6.0 was higher than that of several reported nanocomposites, suggesting its novelty and practical applicability for the treatment of As-contaminated water.

2. Materials and Methods

2.1. Materials

The materials used were of analytical reagent grade unless stated here. Junsei Chemicals Co., Ltd. (Tokyo, Japan) supplied the reagent Na2HAsO·7H2O. La(NO3)3·6H2O was purchased from Daejung chemicals Co., Ltd. (Siheung-Si, Korea). Samchun Pure Chemicals Co., Ltd. (Pyeongtaek, Korea) supplied the reagents HCl and NaOH. HCl and NaOH were used to adjust the pH of the aqueous solutions.

2.2. Synthesis of La2O3 Nanorods

All chemicals were of analytical grade and used without further purification. La2O3 was synthesized by simple hydrothermal precipitation followed by calcination. La(NO3)3·6H2O powder was dissolved in water and stirred vigorously to obtain a clear solution. The solution was adjusted to pH 10 using 5 M NaOH solution and stirred for 30 min and filtered. The obtained precipitate was then collected. The precipitate was evaporated to dryness at 100 °C and calcined at 500 °C for 4 h. The calcined products were stored and used for further studies. The calcination process was carried out at different time (1–6 h) intervals and different calcination temperatures (300, 400, 500, and 600 °C), but the morphology and surface area of the material remained the same at the above calcination at 500 °C for 4 h. Hence, the calcination condition 500 °C for 4 h was chosen as the optimum condition in the preparation of La2O3 in this study.

2.3. Analytical Methods

A D/Max-2500 X-ray diffractometer (Rigaku, Tokyo, Japan) was used to evaluate the crystallinity and textural properties of the prepared adsorbents. The elemental composition was analyzed using a PHI Quantera-II XPS (Ulvac-PHI, Kanagawa, Japan). A scanning electron microscope (S-4300 and EDX-350, Hitachi, Tokyo, Japan) was used to investigate the surface morphology of the adsorbents. HR-TEM (JEM-4010, JEOL, Peabody, MA, USA) was used to measure the shape and particle size of the adsorbents. The samples for TEM were prepared by dispersing the final samples in distilled water, and this dispersant was then dropped on carbon–copper grids covered by an amorphous carbon film. To prevent agglomeration of the adsorbent, the copper grid was placed on a filter paper at the bottom of a Petri dish. N2 adsorption–desorption isotherms of the prepared materials were constructed using an Autosorb-1 (Quantachrome Instruments, Boynton Beach, FL, USA) instrument that was also used to measure the surface area, pore volume, and pore diameter. FT-Raman spectroscopy was carried out with BRUKER OPTICKGMBH and ESCALAB–210 (Barcelona, Spain) instruments. The pH at the potential of zero-point charge (pHzpc) of the sample was measured using the pH drift method. The pH of a solution of 0.01 M NaCl was adjusted between 2 and 12 by adding either HCl or NaOH. Absorbent (0.1 g) was added to 50 mL of the solution. After the pH had stabilized (after 24 h), the final pH was recorded. The graph of the final pH vs. initial pH was used to determine the point at which the initial pH and final pH values were equal. This was taken as the pHzpc of the sample.

2.4. Adsorption Process

As(V) stock solutions were freshly prepared by dissolving the desired amount of Na2HAsO4·7H2O (99% purity, Junsi, Japan). The required arsenic experimental solutions were prepared by diluting the stock solution. The batch studies were carried out using 0.1 g/L La2O3 in a 50 mL Falcon tube contain metal ion solution (10 and 100 mg/L) at 25 °C and pH 6.0. The adsorption isotherms of As(V) on the adsorbent were studied at different temperatures (25, 35, and 50 °C) at pH 6.0 and a predetermined equilibrium time of 60 min. The solution pH influence on arsenic adsorption was estimated under pH ranges from 2.0 to 12.0 that were adapted using 0.1 M of HCl or NaOH solution under constant conditions. To evaluate the influence of associate ions on the adsorption efficiency of La2O3 for As(V), NaCl, KCl, CaCl2, MgCl2, Na2SO4, Na3PO4, NaNO3, and NaHCO3 salts were used to adapt the ionic concentration (0.05–0.5 mg/L), respectively. The adsorbed amount (qe) of As(V) was calculated using
qe = (C0Ce) × V/m,
where the initial and equilibrium concentrations of As(V) (mg/L) are indicated by C0 and Ce, respectively, V is the aqueous phase volume (L), and m is the adsorbent weight (g). To increase the precision and accuracy of the data, duplicates, blanks, and reference standards were used in this study.

3. Results and Discussion

3.1. Characterization

Figure 1a indicates the FT-IR spectra of the prepared nanomaterial. Figure 1a clearly shows a sharp band at 3576 cm−1, suggesting the O–H bond vibrates from the water molecules on the external surface of the material. The sharp bands at 1462 and 1318 cm−1 were assigned to the asymmetric stretching mode of the CO32− group. It is highly probable that all CO32 groups occupy identical sites in the crystal lattice, so the lanthanum oxide precursor might be lanthanum carbonate due to the basic properties of La2O3 which can further react with atmospheric CO2 and water [33,34]. The sharp bands at 498, 672, 846, and 1050 cm−1 were assigned to the stretching vibration of the La-O bond in La2O3 [12,35]. The XRD pattern in Figure 1b shows the phases of La2O3. The 2θ values of 26.16°, 29.16°, 30.02°, 39.57°, 52.24°, and 55.53° denote La2O3 [36]. The remaining diffraction peaks showed a small amount of La(OH)3 in the La2O3 product [23,36]. The crystallite size was calculated by Debye–Scherer’s formula, D = k λ / β   c o s θ , where D is the average crystalline size of the nanorod, λ is the wave length of radiation, β is the full width at a half maximum of the peak at 30.02°, and θ is Bragg’s angle. The resultant mean crystalline size of the nanorods is 19.3 nm. The full-range XPS spectra of La2O3 are displayed in Figure 1c, exhibiting the characteristic La and O1s peaks. The distinguished sharp peaks centered at 838.75 and 853.06 eV ascribed to La 3d5/2 and 3d3/2, respectively, suggest La(III) [37]. Further, the high-resolution XPS of La 3d and O 1s (Figure 1d,e) confirmed the existence of La2O3. The La3d high-resolution XPS clearly showed that the bonding energy peaks at 832.68 and 849.84 eV are attributed to the corresponding two spin−orbits of 3d5/2 and 3d3/2 respectively, and the same is also authenticated from the reported literature [38,39]. Further, the peaks at 836.32 and 853.57 eV can be attributed to the satellite peaks of the 3d5/2 and 3d3/2 spin−orbits, respectively [38], while the O 1s spectrum demonstrated a fine deconvolution with the two peaks. The peaks at 528.11 and 531.50 eV are designated to the oxygen anions from metallic oxygen (La−O) [37,38]. Further the SEM-EDX (Figure S1a) found La and oxygen along with traces of carbon from carbonate around 0.5 k eV due to exposure to the natural atmosphere. These results confirmed the formation of La2O3 with traces of carbonate at the surface, which is authenticated by the reported results [33,34]. Further, the SEM-EDX (Figure S1b) of La2O3, which was prepared under N2 atmosphere, is not shown with authentic carbon peaks, which are responsible for carbonate functions.
The morphologies of the La2O3 samples were evaluated using SEM and TEM images. Figure 2 shows a typical SEM and TEM image of La2O3 nanorods. The SEM images of the prepared compound confirmed the formation of nanoparticles. The nanorod texture from the TEM (Figure 2a) observations revealed rod-like nanostructures. The nanostructures of the synthesized La2O3 exhibit uniform morphology of the rods with a width of 18 nm and length of 160–200 nm. The resultant width of the prepared nanorod was close in agreement with Scherer’s crystalline size. A Brunauer–Emmett–Teller (BET) measurement was conducted to find the pore volume, pore diameter, and surface area of the prepared La2O3 nanorods. From the BET results, we found that the surface area of La2O3 nanorods was 18.76 m2/g with a pore volume of 0.378 cm3/g and diameter of 1.785 nm. The overall microscopic results suggest that the prepared La2O3 is classified as a nano-sized rod-shaped material.

3.2. Adsorption Studies

Figure 3a indicates the influence of the solution initial pH on the adsorptive removal of As(V) by the La2O3 nanorods. The uptake of As(V) was rapidly increased from pH 2.0–3.0, reaching a plateau around 99.8% removal at pH 3.0–7.0. As pH is further increased (7.0–11.0), As(V) adsorptive removal slowly decreased. The slow increase in As(V) removal at pH 2.0–3.0 may be attributed to the predominance of As(V) in the molecular state of H3AsO4 below pH 2.3 [40]. The maximum removal efficiency remained constant when As(v) pH increased to 3.0–7.0, possibly due to the more negatively charged As(V) species, leading to high adsorption on the positive surface of the La2O3 nanorods (pH < pHPZC (6.0) (Figure S2). A rapid decline emerged in the alkali pH 8.0–12.0 due to the presence of As(V) as AsO43− in aqueous media, leading to electrostatic repulsion with the negative surface charge of the La2O3 nanorods (pH > pHPZC (6.0)). However, there is significant adsorption removal under alkali conditions. Hence, the influence of the solution pH on As(V) adsorptive removal may be explained by two mechanisms: (1) electrostatic interaction and (2) surface complexation through ligand exchange [41]. As confirmed by FT-IR and XPS, there are active surface functional groups, lanthanum oxide, hydroxides, and traces of nitrate (Figure 1a) on the La2O3 surface. At low pH solution (3.0–7.0), La2O3 nanorods protonated the surface functions and formed positively charged surface functions that could adsorb negatively charged arsenic species (H2AsO4 and HAsO42−) through electrostatic attraction and surface complexation [42]. At pH 8.0–11.0, the significant adsorption of As(V) may be due to the surface complexation of arsenic species through surface functional groups. However, the adsorption removal above pH > 10 was not significant. As shown in Figure 3b, the adsorptive removal of As(V) was efficient (99.30%) even at a low La2O3 dosage (0.1 g/L). However, as the La2O3 concentration increased from 0.1 to 1.0 g/L, the uptake capacity remained constant with an average removal efficiency of 99.59 ± 0.12%.
Adsorption kinetic studies were carried out to determine the removal of As(V) (10–100 mg/L) using the as-prepared La2O3 nanorods (Figure 3b). The adsorptive removal of As(V) on La2O3 increased rapidly as the contact time increased from 0–60 min. After 60 min, increasing the contact time caused the adsorption removal to remain constant. Pseudo first-order (PFO) (Equation (2)) and pseudo second-order (PSO) (Equation (3)) kinetics are used to investigate the adsorption kinetics of As(V) onto La2O3:
PFO   kinetic   equation :   q = q e ( 1 10 k 1 × t / 2.303 ) ,
PSO   kinetic   equation :   q = k 2 × t × q e 2 / ( 1 + k 2 × t × q e ) ,
where the PFO and PSO rate constants are indicated by k1 and k2, respectively. The resultant kinetic model fitting linear plots are shown in Figure S3.
The calculated parameters for the adsorption kinetics of As(V) onto La2O3 nanorods at various metal concentrations (10–100 mg/L) are tabulated in Table 1. From the results, the PSO equation was found to be the best-fitting model considering their correlation coefficients R2 (PSO correlation coefficient) approximately equal to one. This was further confirmed by the close agreement of the theoretical qe (mg/g) and predicted qe (mg/g) of the PSO equation. Thus, the adsorption of As(V) on La2O3 was a rate-limiting PSO reaction.
Equilibrium adsorption isotherms are an important parameter for understanding adsorption mechanisms and describing the interaction between the adsorbate and adsorbent. An As(V) adsorption isotherm was obtained by varying the initial As(V) concentration (10–100 mg/L) (Figure 3d) and fitting the experimental data using Langmuir (Equation (4)), Freundlich (Equation (5)), and Temkin (Equation (6)) isotherm models to evaluate the adsorption process mechanism:
C e q e = 1 q m K L + C e q m ,
l o g   q e = l o g   K F + n 1   l o g   C e ,
q e = B   l o g   K T + B   l o g   C e
where the constants qm, KL and n, and KF are predicted from the intercept and slope of the plots of Ce/qe vs. Ce (Equation (3)), logCe vs. logqe (Equation (4)), and qe vs. logCe (Equation (5)) (Figure S4), respectively. These adsorption models represent the adsorption equilibrium at interfaces of the adsorbent with metal ions in a solution.
To determine the isotherm model most suitable in describing the As(V) adsorption process, data analysis was conducted using linear fitting isotherm models (Figure S3) and was considered based on the value of the correlation coefficients (R2) tabulated in Table 2.
Langmuir isotherms produced higher and more precise R2 values (close to 1) than the other tested isotherms. This was further confirmed from Pearson’s chi-square (χ2) analysis, which compared all isotherms on the same abscissa and ordinate to evaluate the best fitted model for the equilibrium adsorption isotherm data. The χ2 is calculated using Equation (7)
χ 2 = ( q e x p q m o d e l ) 2 q m o d e l
where qmodel and qexp (mg/g) are the adsorption capacities determined from the model fitting and the experimental data, respectively. The χ2 values for each model are shown in Table 2. From Table 2, it can be found that the resulting Pearson’s chi-square (χ2) errors have are lower for the Langmuir model fitting than others at three temperatures suggesting the Langmuir isotherm well explains the adsorption equilibrium data. Thus, the results of the isotherm suggest that monolayer adsorption occurred on La2O3’s energetically homogeneous surface. The maximum uptake capacity (qe) was found to be 260.56 mg/g at pH 6.0 and 25 °C. Further, the adsorption capacity of La2O3, prepared at N2 atmosphere, has been found to be 272.12 mg/g. This higher adsorption of La2O3 (at N2 atmosphere) may be caused by the absence of carbonate on the La2O3 surface (Figure S1b). Further, the adsorption capacity of La2O3 nanorods for As(V) was comparable or more than the reported hybrid composites (Table 3), though there are similar reports offering comparable or lower As(V) removal efficiencies [25,40,42,43,44,45] as compared with the present material. Thus, these results prompted the potential of as prepared La2O3 nanorods for arsenic-contaminated wastewater treatment.

3.3. Effects of Coexisting Ions

Real environmental water contains several ions that may influence the adsorptive removal of As(V). To assess the influence of competitive ions on As(V) adsorptive removal, 0.05–0.5 mg/L salt solutions of K+, Na+, Mg2+, Ca2+, CO32−, SO42−, PO43−, and NO3 ions, respectively, were used and the results are shown in Figure 4. It was clearly observed from Figure 4, as salt concentration increased from 0.05 to 0.5 mg/L, that the adsorptive removal of As(V) decreased slightly from 99% to 90% and 83% with the respective cations, Mg2+ and Ca2+. These results suggested that the tested cations were not significantly influenced by the adsorptive removal of As(V), where 7.7%, 92.7%, 39.5%, and 96.7% of the adsorptive removal of As(V) are shown in the presence of CO32−, SO42−,PO43−, and NO3 anions, respectively. The adsorptive removal of As(V) on the adsorbent occurs between As(V) ions and La2O3 ions at the adsorbent interface. K+ and Na+ exhibit a positive influence on the adsorption affinity of La2O3 for As(V) than the other investigated cations and anions. However, CO32− and PO43− are significantly influenced by the adsorption affinity of La2O3. The significant effects of CO32 on the adsorptive removal of arsenic might be due to an inner-sphere surface complexation [46]. That PO43− decreased the adsorptive removal of As(V) significantly might be because As(V) and PO43 are in the same symmetry (tetrahedral) and compete with each other for the surface sites at the nanocomposite surface [47,48]. Further, the results of the adsorption capacity of La2O3, prepared at N2 atmosphere, confirmed that the carbonate species predominately influence the adsorptive removal of As(V) (Figure S1b).

3.4. Thermodynamic Parameters

The thermodynamic parameters, enthalpy (ΔH0), entropy (ΔS0), and Gibbs free energy (ΔG0) of As(V) adsorption using La2O3 can be calculated from RT logKc vs. T (Figure 5). The values of the free energy change (ΔG0) are calculated using
Δ G = R T l n K C ,
Δ G = Δ H T Δ S ,
where the ideal gas constant is indicated by R in 8.314 kJ/mol, and temperature is indicated by T in “°C”. Thus, the ΔH0 and ΔS0 can be measured from the slope and intercept of the RT logKc vs. T plot (Figure 5).
Based on Table 4, the negative ΔG0 values suggest that the process of As(V) adsorptive removal by La2O3 is spontaneous. The high ΔG0 at 50 °C confers a favorable sorption at higher temperatures. The positive ΔH0 suggests that the process of As(V) adsorptive removal by La2O3 is endothermic. This is attributed to the assumption that the exothermic energy of the ions attached to the solid surface is lower than the endothermic energy of dehydration [49]. The positive value of ΔS0 suggest increased randomness at the solid–liquid interface during the process of As(V) adsorptive removal by La2O3. Therefore, from the overall thermodynamic results, it is suggested that the uptake of As(V) on La2O3 is an endothermic and spontaneous process.

3.5. As(V) Adsorption Mechanism Using La2O3

Based on batch adsorption studies, including pH, kinetics, and equilibrium isotherms results, the adsorptive removal of As(V) with La2O3 involves electrostatic interaction and surface complexation. From the pH experiment, high adsorption was obtained between pH 3.0–7.0, suggesting an electrostatic attraction between anionic arsenic species and the positive surface charge of the nanorods (Figure 6). However, arsenic removal decreased when pH was increased to 7.0–12.0, implying an electrostatic repulsion between the negative surface charge of the adsorbent and anionic species of As(V). Regardless, there is still significant adsorptive removal, revealing that the adsorption process involved chemical surface complexation predominantly at higher pH values (3.0–7.0). Thus, both electrostatic interaction and chemical surface complexation led to the adsorption process of As(V) at pH 3.0–7.0, which is further confirmed by FTIR results (Figure 7). Figure 7 clearly showed –OH (~3576 cm−1) and La–O (~498 cm−1) peaks altered by the adsorption of As(V), suggesting an adsorption process through surface complexation along with electrostatic interactions. The overall results of the adsorption process suggest that the electrostatic attraction of arsenic species through the protonated surface of La2O3 and surface complexation are predominant in acidic conditions, while electrostatic repulsion and surface complexation of arsenic species with the La2O3 surface are predominant in basic conditions. These representations are clearly shown in Figure 6 for better understanding.

4. Conclusions

La2O3 was successfully synthesized to remove As(V) from contaminated water. The FT-IR image showed nanoparticles with an elemental composition of La and O as La2O3. Using TEM, the La2O3 nanorod dimensions were measured, obtaining a width of 18 nm and length of 160–200 nm. The batch studies conferred that the adsorptive removal efficiency of La2O3 nanorods for As(V) was sensitive to pH, temperature, and initial metal concentration. As(V) was most effectively removed at pH 3.0–7.0. The kinetic data were well explained by the PSO kinetics and the equilibrium isotherm data agreed with the Langmuir isotherm model fitting. These results allow one to conclude that the uptake efficiency of La2O3 for As(V) was rate-limiting kinetics with monolayer sorption on the homogeneous energetic surface of the nanorods. The maximum adsorptive removal capacity of La2O3 has been found to be 260.56 mg/g at pH 6.0 and 25 °C, which is comparable with or more than the reported amounts of some hybrid materials. Therefore, La2O3 is a potential material for As-contaminated wastewater treatment.

Supplementary Materials

The following are available online at https://www.mdpi.com/2075-4701/10/8/1001/s1, Figure S1. SEM-EDX spectrum of La2O3 nanorods without N2 atmosphere (a) and of La2O3 nanorods with N2 atmosphere (b). Figure S2. pH initial versus pH final graph and initial difference for pH pzc calculation for La2O3. Figure S3. PFO (a) and PSO (b) kinetcis model linear fitting plots for As(V) adsorption on to La2O3 nanorods. Metal initial concentration: 10 mg/L, Adsorbent dosage: 0.1 g/L, at pH 6, and 25 °C. Figure S4. Langmuir (a), Freundlich (b) and Temkin (c) isotherms on adsorptive removal of As(V) on to La2O3 nanorods. Metal initial concentration: 10–100 mg/L, Adsorbent dosage: 0.1 g/L, at pH 6, and 25 °C for 60 min equilibrium.

Author Contributions

L.P.L. and J.R.K. conceived and designed the experiments; L.P.L. contributed to conducting experiments and contributed analysis tools; L.P.L., and J.R.K. analyzed the data and wrote the manuscript; J.-K.Y., S.-H.K., and Y.-Y.C. reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation (NRF) of Korea, funded by the Ministry of Education (2019R1I1A1A01061487). The APC was funded by the Kwangwoon University, Seoul, Korea through Research Grant–2019.

Acknowledgments

This research was financially supported by the National Research Foundation (NRF) of Korea, funded by the Ministry of Education (2019R1I1A1A01061487) of the Korean Government. In addition, this study partially supported by the Kwangwoon University, Seoul, Korea through Research Grant–2019.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boldish, S.I.; White, W.B. Vibrational spectra of crystals with the A-type rare earth oxide structure. La2O3 and Nd2O3. Spectrochim. Acta Part A 1979, 35, 1235. [Google Scholar] [CrossRef]
  2. Gouteron, J.; Michel, D.; Lejus, A.M.; Zarembowitch, J. Raman spectra of lanthanide sesquioxide single crystals: Correlation between A and B-type structures. J. Solid State Chem. 1981, 38, 288. [Google Scholar] [CrossRef]
  3. Kang, S.W.; Rhee, S.W. Deposition of La2O3 Films by Direct Liquid Injection Metallorganic Chemical Vapor Deposition. J. Electrochem. Soc. 2002, 149, 345. [Google Scholar] [CrossRef]
  4. De Asha, A.M.; Critchley, J.T.S.; Nix, R.M. Molecular adsorption characteristics of lanthanum oxide surfaces: The interaction of water with oxide overlayers grown on Cu. Surf. Sci. 1998, 405, 201. [Google Scholar] [CrossRef]
  5. Dakhel, A.A. Preparation and optical study of Au nanograins in amorphous La-oxide medium. Colloids Surf. A Physicochem. Eng. Asp. 2009, 332, 9–12. [Google Scholar] [CrossRef]
  6. Ferhi, M.; Horchani-Naifer, K.; Férid, M. Hydrothermal synthesis and photoluminescence of the monophosphate LaPO4:Eu(5%). J. Lumin. 2008, 128, 1777–1782. [Google Scholar] [CrossRef]
  7. Imanaka, N.; Okamoto, K.; Adachi, G.Y. Water-insoluble lanthanum oxychloride-based solid electrolytes with ultra-high chloride ion conductivity. Angew. Chem. Int. Ed. 2002, 41, 3890–3892. [Google Scholar] [CrossRef]
  8. Slagtern, Y.; Schuurman, C.; Leclercq, X.; Verykios, C. Mirodatos, Specific features concerning the mechanism of methane reforming by carbon dioxide over Ni/La2O3 catalyst. J. Catal. 1997, 172, 118–126. [Google Scholar] [CrossRef]
  9. Ordonez-Regil, E.; Drot, R.; Simoni, E.; Ehrhardt, J.J. Sorption of uranium (VI) onto lanthanum phosphate surfaces. Langmuir 2002, 18, 7977–7984. [Google Scholar] [CrossRef]
  10. Tokunaga, S.; Wasay, S.A.; Park, S.W. Removal of arsenic (V) ion from aqueous soloutions by lanthanum compounds. Water Sci. Technol. 1997, 35, 71–78. [Google Scholar] [CrossRef]
  11. Imanaka, N.; Okamoto, K.; Adachi, N.G. A chlorine gas sensor based on the combination of Mg2+ cation conducting and O2− anion conducting solid electrolytes with lanthanum oxychloride as an auxiliary electrode. Electrochem. Commun. 2001, 3, 49–51. [Google Scholar] [CrossRef]
  12. Moothedan, M.; Sherly, K.B. Synthesis, characterization and sorption studies of nano lanthanum oxide. J. Water Process Eng. 2016, 9, 29–37. [Google Scholar] [CrossRef]
  13. Yulin, L.; Jun, Y.; Xiaoci, L.; Huang, W.; Yu, T.; Zhang, Y. Combustion synthesis and stability of nanocrystalline La2O3 via ethanolamine-nitrate process. J. Rare Earths 2012, 30, 48–52. [Google Scholar]
  14. Keav, S.; Matam, S.K.; Ferri, D.; Weidenkaff, A. Structured perovskite-based catalysts and their application as three-way catalytic converters—A review. Catalysis 2014, 4, 226–255. [Google Scholar] [CrossRef] [Green Version]
  15. Zhao, C.; Zhao, C.Z.; Werner, M.; Taylor, S.; Chalker, P. Dielectric relaxation of high-k oxides. Nanoscale Res. Lett. 2013, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  16. Tang, L.; Li, Y.M.; Xu, K.L.; Hou, X.D.; Lv, Y. Sensitive and selective acetone sensor based on its cataluminescence from nano-La2O3 surface. Sens. Actuators B 2008, 132, 243. [Google Scholar] [CrossRef]
  17. Cao, J.M.; Ji, H.M.; Liu, J.S.; Zheng, M.B.; Chang, X.; Ma, X.J.; Zhang, A.M.; Xu, Q.H. Controllable syntheses of hexagonal and lamellar mesostructured lanthanum oxide. Mater. Lett. 2005, 59, 408. [Google Scholar] [CrossRef]
  18. Wang, S.Y.; Wang, W.; Qian, Y.T. Preparation of La2O3 thin films by pulse ultrasonic spray pyrolysis method. Thin Solid Films 2000, 372, 50–53. [Google Scholar] [CrossRef]
  19. Imanaka, N.; Masui, T.; Kato, Y. Preparation of the cubic-type La2O3 phase by thermal decomposition of LaI3. J. Solid State Chem. 2005, 178, 395–398. [Google Scholar] [CrossRef]
  20. Sheng, J.; Zhang, S.; Lv, S.; Sun, W.D. Surfactant-assisted synthesis and characterization of lanthanum oxide nanostructures. J. Mater. Sci. 2007, 42, 9565–9571. [Google Scholar] [CrossRef]
  21. Wu, Q.Z.; Shen, Y.; Liao, J.F.; Li, Y.G. Synthesis and characterization of three-dimensionally ordered macroporous rare earth oxides. Mater. Lett. 2004, 58, 2688–2691. [Google Scholar] [CrossRef]
  22. Hu, C.G.; Liu, H.; Dong, W.T.; Zhang, Y.Y.; Bao, G.; Lao, C.S.; Wang, Z.L. La(OH)3 and La2O3 nanobelts-synthesis and physical properties. Adv. Mater. 2007, 19, 470–474. [Google Scholar] [CrossRef]
  23. Bazzi, R.; Flores-Gonzalez, M.A.; Louis, C.; Lebbou, K.; Dujardin, C.; Brenier, A.; Zhang, W.; Tillement, O.; Bernstein, E.; Perriat, P. Synthesis and luminescent properties of sub-5-nm lanthanide oxides nanoparticles. J. Lumin. 2003, 102–103, 445–450. [Google Scholar] [CrossRef]
  24. Vadivel Murugan, A.; Navale, S.C.; Ravi, V. Synthesis of nano-crystalline La2O3 powder at 100 °C. Mater. Lett. 2006, 60, 848–849. [Google Scholar] [CrossRef]
  25. Lingamdinne, L.P.; Choi, Y.L.; Kim, I.S.; Chang, Y.Y.; Koduru, J.R.; Yang, J.K. Porous graphene oxide based inverse spinel nickel ferrite nanocomposites for the enhanced adsorption removal of arsenic. RSC Adv. 2016, 6, 73776–73789. [Google Scholar] [CrossRef]
  26. Tuutijärvi, T.; Lu, J.; Sillanpää, M.; Chen, G. As (V) adsorption on maghemite nanoparticles. J. Hazard. Mater. 2009, 66, 1415–1420. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, S.L.; Dzeng, S.R.; Yang, M.H.; Chiu, K.H.; Shieh, G.M.; Wai, C.M. Arsenic species in groundwaters of the black foot disease area, Taiwan. Environ. Sci. Technol. 1994, 28, 877–881. [Google Scholar] [CrossRef]
  28. Karim, M. Arsenic in groundwater and health problems in Bangladesh. Water Res. 2000, 34, 304–310. [Google Scholar] [CrossRef]
  29. Chakraborti, D.; Rahman, M.M.; Paul, K.; Chowdhury, U.K.; Sengupta, M.K.; Lodh, D.; Chanda, C.R.; Saha, K.C.; Mukherjee, S.C. Arsenic calamity in the Indian subcontinent. What lessons have been learned? Talanta 2002, 58, 3–22. [Google Scholar] [CrossRef]
  30. Khosrow-pour, F.; Aghazadeh, M.; Sabour, B.; Dalvand, S. Large-scale synthesis of uniform lanthanum oxide nanowires via template-free deposition followed by heat-treatment. Ceram. Int. 2013, 39, 9491–9498. [Google Scholar] [CrossRef]
  31. Misra, M.; Deba, C.N. Process for Removal of Selenium and Arsenic from Aqueous Streams. U.S. Patent 5,603,838, 18 February 1997. [Google Scholar]
  32. Davis, S.A.; Misra, M. Transport model for the adsorption of oxyanions of selenium (IV) and arsenic (V) from water onto lanthanum-and aluminum-based oxides. J. Colloid Interface Sci. 1997, 188, 340–350. [Google Scholar] [CrossRef]
  33. Karthikeyan, S.; Selvapandiyan, M. Effect of annealing temperature on the properties of lanthanum oxide (La2O3) nanoplates by reflux routes. Int. J. Eng. Sci. Res. Technol. 2018, 7, 559. [Google Scholar]
  34. Mu, Q.; Wang, Y. Synthesis, characterization, shape-preserved transformation, and optical properties of La(OH)3, La2O2CO3, and La2O3 nanorods. J. Alloys Compd. 2011, 509, 396–401. [Google Scholar] [CrossRef]
  35. Xiao, Y.; Zongyu, F.; Xiaowei, H.; Li, H.; Zhiqi, L.; Qiang, W.; Yongke, H. Synthesis of lanthanum oxide nanosheets by a green carbonation process. Chin. Sci. Bull. 2014, 59, 1864–1867. [Google Scholar] [CrossRef]
  36. Wang, K.; Wu, Y.; Li, H.; Li, M.; Guan, F.; Fan, H. A hybrid antioxidizing and antibacterial material based on Ag–La2O3 nanocomposites. J. Inorg. Biochem. 2014, 141, 36–42. [Google Scholar] [CrossRef]
  37. Hu, S.; Chi, B.; Pu, J.; Jian, L. Novel heterojunction photocatalysts based on lanthanum titanate nanosheets and indium oxide nanoparticles with enhanced photocatalytic hydrogen production activity. J. Mater. Chem. A 2014, 2, 19260–19267. [Google Scholar] [CrossRef]
  38. Sanivarapu, S.R.; Lawrence, J.B.; Sreedhar, G. Role of surface oxygen vacancies and lanthanide contraction phenomenon of Ln(OH)3 (Ln= La, Pr, and Nd) in sulfide-mediated photoelectrochemical water splitting. ACS Omega 2018, 3, 6267–6278. [Google Scholar] [CrossRef]
  39. Sunding, M.F.; Hadidi, K.; Diplas, S.; Løvvik, O.M.; Norby, T.E.; Gunnæs, A.E. XPS Characterization of In Situ Treated Lanthanum Oxide and Hydroxide Using Tailored Referencing and Peak Fitting Procedures. J. Electron Spectrosc. Relat. Phenom. 2011, 184, 399–409. [Google Scholar] [CrossRef]
  40. Chen, B.; Zhu, Z.; Ma, J.; Qiu, Y.; Chen, J. Surfactant assisted Ce–Fe mixed oxide decorated multiwalled carbon nanotubes and their arsenic adsorption performance. J. Mater. Chem. A 2013, 1, 11355–11367. [Google Scholar] [CrossRef] [Green Version]
  41. Sikder, M.T.; Tanaka, S.; Saito, T.; Kurasaki, M. Application of zero-valent iron impregnated chitosan-caboxymethyl β-cyclodextrin composite beads as arsenic sorbent. J. Environ. Chem. Eng. 2014, 2, 370–376. [Google Scholar] [CrossRef]
  42. Kyzas, G.Z.; Siafaka, P.I.; Kostoglou, M.; Bikiaris, D.N. Adsorption of As(III) and As(V) onto colloidal microparticles of commercial cross-linked polyallylamine (sevelamer) from single and binary ion solutions. J. Colloid Interface Sci. 2016, 474, 137–145. [Google Scholar] [CrossRef] [PubMed]
  43. Panda, A.P.; Rout, P.; Sanjukta, A.K.; Jha, U.; Swain, S.K. Enhanced performance of a core–shell structured Fe@Fe oxide and Mn@Mn oxide (ZVIM) nanocomposite towards remediation of arsenic contaminated drinking water. J. Mater. Chem. A 2020, 8, 4318–4333. [Google Scholar] [CrossRef]
  44. Shi, Q.; Yan, L.; Chan, T.; Jing, C. Arsenic adsorption on lanthanum-impregnated activated alumina: Spectroscopic and DFT study. ACS Appl. Mater. Interfaces 2015, 7, 26735–26741. [Google Scholar] [CrossRef] [PubMed]
  45. Jang, M.; Park, J.K.; Shin, E.W. Lanthanum functionalized highly ordered mesoporous media: Implications of arsenate removal. Microporous Mesoporous Mater. 2004, 75, 159–168. [Google Scholar] [CrossRef] [Green Version]
  46. Wang, C.; Luo, H.; Zhang, Z.; Wu, Y.; Zhang, J.; Chen, S. Removal of As (III) and As (V) from aqueous solutions using nanoscale zero valent iron-reduced graphite oxide modified composites. J. Hazard. Mater. 2014, 268, 124–131. [Google Scholar] [CrossRef]
  47. Yu, F.; Sun, S.; Ma, J.; Han, S. Enhanced removal performance of arsenate and arsenite by magnetic graphene oxide with high iron oxide loading. Phys. Chem. Chem. Phys. 2015, 17, 4388–4397. [Google Scholar] [CrossRef]
  48. Sigdel, A.; Park, J.W.; Kwak, H.; Park, P.K. Arsenic removal from aqueous solutions by adsorption onto hydrous iron oxide-impregnated alginate beads. J. Ind. Eng. Chem. 2016, 35, 277–286. [Google Scholar] [CrossRef]
  49. Huang, J.Y.; Wu, Z.W.; Chen, L.W.; Sun, Y.B. The sorption of Cd (II) and U (VI) on sepiolite: A combined experimental and modeling studies. J. Mol. Liq. 2015, 209, 706–712. [Google Scholar] [CrossRef]
Figure 1. Spectroscopic characterization: (a) FT-IR spectra, (b) XRD patterns, and (c) XPS patterns of La2O3 nanorods, and (d) high-resolution XPS of La 3d and (e) O 1s of prepared La2O3.
Figure 1. Spectroscopic characterization: (a) FT-IR spectra, (b) XRD patterns, and (c) XPS patterns of La2O3 nanorods, and (d) high-resolution XPS of La 3d and (e) O 1s of prepared La2O3.
Metals 10 01001 g001
Figure 2. (a) SEM and (b,c) TEM images of La2O3 produced by alkaline precipitation followed by calcination at 500 °C.
Figure 2. (a) SEM and (b,c) TEM images of La2O3 produced by alkaline precipitation followed by calcination at 500 °C.
Metals 10 01001 g002
Figure 3. As(V) adsorption onto La2O3 (10–100 mg/L, 0.1 g/L, pH 6.0) and the effects of (a) pH, (b) dosage (20 mg/L, pH 6.0) (c) kinetics, and (d) isotherms at 25 °C for 60 min equilibrium.
Figure 3. As(V) adsorption onto La2O3 (10–100 mg/L, 0.1 g/L, pH 6.0) and the effects of (a) pH, (b) dosage (20 mg/L, pH 6.0) (c) kinetics, and (d) isotherms at 25 °C for 60 min equilibrium.
Metals 10 01001 g003
Figure 4. Co-existing ions: (a) cations’ and (b) anions’ influences on La2O3 adsorptive removal of As(V) (10 mg/L) with 0.1 g/L La2O3 dosage, at pH 6, and 25 °C for 60 min equilibrium.
Figure 4. Co-existing ions: (a) cations’ and (b) anions’ influences on La2O3 adsorptive removal of As(V) (10 mg/L) with 0.1 g/L La2O3 dosage, at pH 6, and 25 °C for 60 min equilibrium.
Metals 10 01001 g004
Figure 5. Temperature vs.   Δ G plot to understand the influence of temperature on the adsorption efficiency of La2O3 (0.1 g/L) for As(V) (10 mg/L), at pH 6, and at 25 °C for 60 min equilibrium.
Figure 5. Temperature vs.   Δ G plot to understand the influence of temperature on the adsorption efficiency of La2O3 (0.1 g/L) for As(V) (10 mg/L), at pH 6, and at 25 °C for 60 min equilibrium.
Metals 10 01001 g005
Figure 6. Adsorption mechanism of As(V) on La2O3 nanorods at pH range 3.0–12.0.
Figure 6. Adsorption mechanism of As(V) on La2O3 nanorods at pH range 3.0–12.0.
Metals 10 01001 g006
Figure 7. FT-IR spectra of La2O3 nanorods before and after As(V) adsorption.
Figure 7. FT-IR spectra of La2O3 nanorods before and after As(V) adsorption.
Metals 10 01001 g007
Table 1. La2O3 nanorods adsorptive removal kinetics parameters for As(V) (10.0–100 mg/L) at pH 6.0, 25 °C for 60 min equilibrium time with 0.1 g/L nanorods.
Table 1. La2O3 nanorods adsorptive removal kinetics parameters for As(V) (10.0–100 mg/L) at pH 6.0, 25 °C for 60 min equilibrium time with 0.1 g/L nanorods.
Metal Concentration, mg/Lqe, Th,
mg/g
PFOPSO
qe, Cal, mg/gk1R2qe, Cal, mg/gk2R2
1086.9834.570.05290.80184.560.002150.999
20118.9888.970.03700.753119.560.000490.999
40145.68116.970.01980.867142.560.000270.998
60185.69135.690.03960.784186.310.000180.995
80205.69158.950.00690.472203.560.000210.991
100245.68185.980.03130.589244.560.000440.992
Table 2. Isotherm predicted values for the adsorptive removal of As(V) on La2O3 (0.1 g/L) nanorods at pH 6.0 and 25 °C for 60 min equilibrium.
Table 2. Isotherm predicted values for the adsorptive removal of As(V) on La2O3 (0.1 g/L) nanorods at pH 6.0 and 25 °C for 60 min equilibrium.
Temperature, °CLangmuirFreundlichTemkin
qmax, mg/gKLR2χ2KF,mg/g*(L/mg)1/nnR2χ2BKTRχ2
25260.561.880.9952.16185.055.300.9459.8798.60127.960.97113.45
35285.982.560.9991.65135.086.880.91712.6999.88212.930.94818.97
50305.693.120.9981.98131.054.060.96320.7656.87306.120.90917.89
Table 3. Comparison of As(V) adsorption capacities of various adsorbents.
Table 3. Comparison of As(V) adsorption capacities of various adsorbents.
Type of Adsorbentqmax, (mg/g)ConditionsReference
Initial Concentration, Dosage and pH
Reduced graphene oxide-nickel ferrite (RGONF) nanocomposite106.402–30 mg/L, 0.3 g/L
and pH 6.5
[25]
Ce–Fe oxide decorated multiwalled carbon nanotubes30.961–20 mg/L, 0.2 g/L
and pH 4.0.
[40]
Sevelamer 133.000–250 mg/L, 1 g/L
and pH 6
[42]
Core–shell structured Fe(0)@Fe oxide and Mn(0)@Mn oxide (ZVIM) nanocomposite101.950.2–1.0 mg/L, 0.2 g/L
and pH 6.5
[43]
Lanthanum-impregnated activated alumina26.300–250 mg/L, 1.0 g/L
and pH 7.0
[44]
Lanthanum-impregnated SBA-15 123.7020–60 mg/L and pH 7.2[45]
La2O3260.5610–100 mg/L, 0.1 g/L
and pH 6.0
This work
Table 4. Thermodynamics for As(V) adsorption on La2O3 (0.1 g/L) at pH 6.0.
Table 4. Thermodynamics for As(V) adsorption on La2O3 (0.1 g/L) at pH 6.0.
Temperature, °C Δ G 0   ( KJ / mol ) Δ H   ( KJ / mol ) Δ S   ( KJ / mol   ° C ) log Kc
25−601.291103.9328.202.892
35−883.1343.034
50−1306.4003.142

Share and Cite

MDPI and ACS Style

Lingamdinne, L.P.; Koduru, J.R.; Chang, Y.-Y.; Kang, S.-H.; Yang, J.-K. Facile Synthesis, Characterization, and Adsorption Insights of Lanthanum Oxide Nanorods. Metals 2020, 10, 1001. https://doi.org/10.3390/met10081001

AMA Style

Lingamdinne LP, Koduru JR, Chang Y-Y, Kang S-H, Yang J-K. Facile Synthesis, Characterization, and Adsorption Insights of Lanthanum Oxide Nanorods. Metals. 2020; 10(8):1001. https://doi.org/10.3390/met10081001

Chicago/Turabian Style

Lingamdinne, Lakshmi Prasanna, Janardhan Reddy Koduru, Yoon-Young Chang, Seon-Hong Kang, and Jae-Kyu Yang. 2020. "Facile Synthesis, Characterization, and Adsorption Insights of Lanthanum Oxide Nanorods" Metals 10, no. 8: 1001. https://doi.org/10.3390/met10081001

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