Frontiers of Environmental Science & Engineering, 2020, 14(1): 15-
doi: 10.1007/s11783-019-1194-4
Removal and recovery of toxic nanosized Cerium Oxide using eco-friendly Iron Oxide Nanoparticles
Kanha Gupta1, Nitin Khandelwal2, Gopala Krishna Darbha2,3,,
. Department of Chemical Sciences, IISER Kolkata, Mohanpur, West Bengal, Pincode 741246, India
. Environmental Nanoscience Laboratory, Department of Earth Sciences, IISER Kolkata, Mohanpur, West Bengal, Pincode 741246, India
. Center for Climate & Environmental Studies, IISER Kolkata, Mohanpur, West Bengal, Pincode 741246, India
Abstract:

Eco-friendly IONPs were synthesized through solvothermal method.

IONPs show very high removal efficiency for CeO2 NPs i.e. 688 mg/g.

Removal was >90% in all synthetic and real water samples.

>80% recovery of CeO2 NPs through sonication confirms reusability of IONPs.

Increasing applications of metal oxide nanoparticles and their release in the natural environment is a serious concern due to their toxic nature. Therefore, it is essential to have eco-friendly solutions for the remediation of toxic metal oxides in an aqueous environment. In the present study, eco-friendly Iron Oxide Nanoparticles (IONPs) are synthesized using solvothermal technique and successfully characterized using scanning and transmission electron microscopy (SEM and TEM respectively) and powder X-Ray diffraction (PXRD). These IONPs were further utilized for the remediation of toxic metal oxide nanoparticle, i.e., CeO2. Sorption experiments were also performed in complex aqueous solutions and real water samples to check its applicability in the natural environment. Reusability study was performed to show cost-effectiveness. Results show that these 200 nm-sized spherical IONPs, as revealed by SEM and TEM analysis, were magnetite (Fe3O4) and contained short-range crystallinity as confirmed from XRD spectra. Sorption experiments show that the composite follows the pseudo-second-order kinetic model. Further R2>0.99 for Langmuir sorption isotherm suggests chemisorption as probable removal mechanism with monolayer sorption of CeO2 NPs on IONP. More than 80% recovery of adsorbed CeO2 NPs through ultrasonication and magnetic separation of reaction precipitate confirms reusability of IONPs. Obtained removal % of CeO2 in various synthetic and real water samples was>90% signifying that IONPs are candidate adsorbent for the removal and recovery of toxic metal oxide nanoparticles from contaminated environmental water samples.

Keywords: Adsorption; toxic metal oxide remediation; eco-friendly IONP; Iron oxide; CeO2 removal;
Introduction

Nanoparticles play a salient role in our life by their wide range of applications. They generally have a size range of 1–100 nm, making them unique, possessing characteristics like surface tailorability (Darbha et al., 2008), broad spectral range, multi-functionality, and improved characteristic solubility (Ranjit and Ahmed, 2013). However, the increasing use of nanoparticles leads to their increasing release as a pollutant and contaminates the aqueous environments. The environmental fate of these nanoparticles is complex, as they can be transported or co-transported or may be retained on rocks and mineral surfaces depending on their persistent hydrodynamic, physical and geochemical factors (Darbha et al., 2010; Darbha et al., 2012). Moreover, the water treatment processes that include reverse osmosis (RO) are also incapable of removing these tiny toxic nanoparticles, making them available in drinking water (Kaegi, 2009). Toxicity of these metal oxides follows the order ZnO>CeO2>CNTs>TiO2 (Juganson et al., 2015). Cerium oxide remains one of the most toxic nanoparticles despite having great application in polishing, optics(Teng et al., 2008), cosmetics(Wiechers and Musee, 2010), mixed conduction(Suzuki et al., 2002) and biomedical fields and have emerged with high potential in the research of fuel cells(Omar, 2019), water splitting(Dvořák et al., 2018), and catalytic application(Trovarelli, 1996). There are several studies regarding the toxicity of CeO2 to the ecosystem/biota. The acute exposure of CeO2 NPs through the inhalation route may induce cytotoxicity via oxidative stress and may lead to a chronic inflammatory response (Srinivas et al., 2011). Elevated oxidative stress increases the production of malondialdehyde and lactate dehydrogenase, which are indicators of lipid peroxidation and cell membrane damage, respectively (Kumari et al., 2014).Cell viability decreased significantly as a function of nanoparticle dose and exposure time (Lin et al., 2006).Studies have shown that CeO2 nanoparticles (NPs) can be accumulated in plants without modification, which could pose a threat to human health (Morales et al., 2013). In addition, due to their small size, they have the potential to pass through the various biological barriers and more likely to be dispersed in the bloodstream or even the central nervous system (Masserini, 2013). Toxicity of nanoparticles is one of the emerging research areas of this decade.

Several approaches have been made for the removal of nanoparticles, including membrane filtration (Lin et al., 2007; Springer et al., 2013), biological processes (Ganesh et al., 2010), chemical coagulation (Chang et al., 2007), and cloud point extraction (Liu et al., 2010). However, adsorption remains one of the most favorable separation processes as it involves less energy, cost-effective and straightforward, as investigated for the removal of various toxic metal ions (Sen Gupta and Bhattacharyya, 2011; Khandelwal et al., 2019). However, there are very few studies available for the removal of metal oxide nanoparticles using adsorption (Zhou et al., 2017). Over several other adsorbents like clays, zeolite, ZnO, Charcoal, etc., IONPs are superior, due to their environmental friendliness, fast interaction with adsorbates, reusability and easy magnetic separation (Hua et al., 2012; Zargoosh et al., 2013; Ali et al., 2016).

In this paper, for the remediation of toxic CeO2 NPs, IONPs have been synthesized by the solvothermal method. These synthesized IONPs were characterized by SEM, TEM-EDS, XRD, FT-IR, and Dynamic Light Scattering (DLS), etc. Then, these IONPs have been used for the removal of CeO2 from aqueous solutions by using magnetic separation technique following the batch sorption process. Sorption capacity was calculated under various physical and environmental conditions like pH, time, contaminant concentration, humic acid, and bicarbonates. Further kinetic and sorption isotherm modeling was performed to get insights about removal mechanisms. To get the detailed insights about the removal mechanisms, reaction precipitate after CeO2-IONP interaction was characterized using TEM/Scanning transmission electron microscopy (STEM) along with elemental mapping and FT-IR.The reusability of these IONPs was also investigated using the ultrasonication method. Further, the efficiency of the adsorbent was also evaluated in various real water matrices.

Material and methods
Chemicals and Materials

Ferric chloride hexahydrate (FeCl3·6H2O) and Ethanol were obtained from Merck, Germany. Sodium acetate trihydrate (CH3COONa·3H2O), ethylene glycol, HCl, NaOH, and NaHCO3 were purchased from Merck, India. Humic acid and Cerium oxide nanoparticles dispersion (10 wt.%) were purchased from Sigma Aldrich (St. Louis, MO, USA). All chemicals were of analytical grade and were used without further purification.18.2 MMilli-Q water (Millipore) was used throughout the experiments.

Synthesis of Iron Oxide Nanoparticles

The IONPs were synthesized through a solvothermal reaction with some modifications (Deng et al., 2008; Pang and Liu, 2013; Zhou et al., 2017). Briefly, 2.7 g Ferric chloride hexahydrate (FeCl3·6H2O) was mixed with 7.2 g sodium acetate in 100 mL ethylene glycol solution at 65°C. Sodium acetate was used as a structure-directing and electrostatic stabilizing agent, while ethylene glycol was used as a reducing agent (He et al., 2018). Besides, ethylene glycol has a characteristic of high boiling point and dielectric constant, which makes it efficient for the solvothermal process (He et al., 2018). The mixture was stirred at 1300 rpm for 20 min and the obtained solution was sealed and heated at 200°C overnight. The product was collected by using a strong permanent magnet and was washed continuously with ethanol and Milli-Q water several times to remove the non-magnetic bi-products. The collected product was vacuum dried at 85°C temperature overnight. Finally, dark brown colored powder was obtained and stored for further characterization and sorption experiments.

Morphological and compositional characterization

To evaluate the phase and crystalline nature of the nanoparticles, powder X-ray diffraction (PXRD) measurements were carried out using a Rigaku (mini flex II, Japan) with Cu-K radioactive source (wavelength= 0.154 nm) at 40 kV/70 mA at a scanning rate of 2° per minute in the range of 20°‒70°. The morphology of the particles was investigated by field emission scanning electron microscopy (FESEM). Images were taken with a Carl Zeiss SUPRA 55VP FESEM. TEM images of particles were obtained using a JEOL, JEM 2100 HR model, along with an elemental line scan. Zeta potentials, Hydrodynamic diameter, Point of Zero charge (pHPZC) in aqueous suspension solution were measured with a Malvern Zetasizer ZS-90 using dynamic light scattering (DLS) technique. Fourier-transform infrared (FT-IR) spectra were taken to check the presence of various surface functional groups on IONPs and CeO2, using Thermoscie Nicolet iS5 coupled with iD1 transmission accessory by making KBr pellets. After the interaction of CeO2 with IONPs, reaction precipitate was separated and characterized using TEM/STEM along with elemental distribution and FT-IR spectra to get detailed information about various interaction mechanisms.

Batch sorption experiment

For the good dispersion of particle, CeO2 and IONPs stock solutions were sonicated for 15 min. CeO2 working solutions (10 mL) in the range of 20‒120 mg/L initial concentrations were prepared by diluting a stock solution. Finally,0.5 mL of 2 mg/mL aqueous dispersions was added in CeO2 aqueous solutions. These reaction mixtures were shaken at 300 RPM for 1 h in an amber glass vial at room temperature. Measured pH of the solutions was 5.8. The IONPs were separated through a permanent magnet, and the supernatant was collected. The concentration of CeO2 in the supernatant was determined by using a UV-VIS spectrophotometer (Evolution 201, Thermo) at a wavelength of 300 nm. The adsorption capacity (Qe, mg/g) was calculated as follows.

Qe= (CoCe)V M,

where Co (mg/L) and Ce (mg/L) are the initial concentration of CeO2 and the remaining concentration of CeO2, respectively. Qe (mg/g) is the adsorption capacity of IONPs, V (mL) is the volume of solution, and M (mg) is the mass of IONPs. All experiments were done doubly to confirm Quality assessment/Original content (QA/OC) of data, error was within 10%, and therefore, error bars have been eliminated.

Results and discussion
Characterization

Both SEM and TEM images of IONPs (Fig. 1(a)) show the spherical nature of the particles with nearly 200 nm diameter. Fig. 1b shows the morphology of highly dispersed CeO2 NPs, which are nearly of 25 nm in size and are cubic. To obtain the details of composition, Energy Dispersive X-ray (EDS) data was obtained with a TEM-EDS line scan (Fig. 1(c)) along an IONP sphere, and the result shows Fe/O= 3/4 which is for magnetite. To confirm the suspension stability of CeO2 NPs, change in hydrodynamic diameter of these particles was observed for 180 min, and data in Fig. 1(d) clearly shows no significant change suggesting that these particles have no self-aggregation and are stable in suspension.

Fig.1 (a) SEM image of IONPs and inside TEM image of IONPs (b) TEM image of CeO2 (c) Line-Scan of IONPs (d) change in hydrodynamic diameter of blank CeO2 suspension with time.

PXRD spectra of IONPs (Fig. 2(a)) shows six characteristic peaks at 30.1°, 35.4°, 43.1°, 53.4°, 56.9°, and 62.5°, corresponding to planes (220), (311), (400), (422), (511), and (440) of magnetite (JCPDS File No. 19-0629), respectively, were observed. Further, broad and smaller peaks signify short-range order. Zeta potential with varying pH was measured, and point of zero charges (PZC) for IONPs and CeO2 were determined to be 5.8 and 8.1, respectively as shown in Fig. 2(b). IONPs tend to show negative zeta potential at circum-neutral pH because oxygen atoms on the surface are not fully coordinated with Fe3+ (Chen et al., 2017). In CeO2, there are oxygen vacancies acting as point defects due to which cerium oxide was represented as CeO2—γ (Gunawan et al., 2019). Therefore, cerium oxide shows positive zeta potential at similar pH values. DLS characterization remains a critical factor for physisorption as the particles possess opposite charges between specific pH (5.8‒8.1),which is essential for electrostatic interaction.

Fig.2 (a) XRD pattern of IONPs (b) Zeta potentials and pHPZC of IONPs and CeO2.

Adsorption kinetics

Several solutions with a constant concentration of CeO2 (70 mg/L) and IONPs (100 mg/L) were prepared and kept for different time intervals (0‒480 min) on a shaker to achieve the equilibrium time. It was observed that there are abundant sites available on IONPs, and the majority of the sites available were occupied rapidly by CeO2NPs in the first 30 min of reaction. Then the active sites gradually decreased, and adsorption equilibrium was achieved, which shows the faster removal of CeO2 in the system. Further data was interpreted with various kinetic models like the pseudo-first-order, pseudo-second-order, intra-particle model, and Elovich model(Largitte and Pasquier, 2016).

The linear form of the pseudo-first equation is as follows

ln(Qe Qt)=lnQek1t,

where Qe and Qt are the amount of CeO2 (mg) adsorbed per unit gram of IONPs at equilibrium time and at time t, respectively. k1 (min1) is the pseudo-first-order rate constant. Fitting the data with pseudo-first-order gives the R2 value of 0.75 (Fig. 3(a)) and calculated sorption capacity value Qe (Table 1) was very different from the experimental value, which suggests that the sorption does not follow pseudo-first-order.The equation for second-pseudo order is (Ho and McKay, 1999)

tQt= 1 k2 Qe2+ t Qe,

where abbreviation for Qe, Qt, and t remains the same while k2 is the rate constant for pseudo-second-order. Fitting the data with this model gives the R2 value of 0.99 (Fig. 3(b)), which is high, and calculated sorption capacity value Qe (Table 1) is matching the experimental value, suggesting that the sorption of CeO2 into the IONPs agrees strongly to the pseudo-second-order, further supporting chemisorption as probable removal mechanism.

The Elovich equation is typically used to describe the kinetics of chemisorption on highly heterogeneous sorbents and linearly expressed as(Chien et al., 1980)

Qt= lnαkk+lntk,

where α and k are the initial adsorption rate and extent of surface coverage and activation energy for chemisorption, respectively. Fitting the data with the Elovich equation gives the R2 value of 0.93 (Table 1 and Fig. 3(c)) suggesting that the adsorption process was controlled by chemisorption.

The intraparticle diffusion model was used to interpret the movement of the particles from the surface of the adsorbent to the inner pores. It is based on the principle that the sorption on any material follows three steps process which is (1) film or surface diffusion, where the sorbate is transported from the bulk solution to the external surface of sorbent, (2) intraparticle or pore diffusion, where sorbate molecules move into the pores of sorbent particles, and (3) adsorption on the interior sites of the sorbent (Khandelwal et al., 2019). The model is linearly expressed as(Srivastava et al., 1989)

Qt=kd t 12+C,

where kd is intraparticle diffusion rate constant or rate factor, and c is resistance in mass transfer due to the boundary layer. By comparing the data with the intraparticle diffusion model, it was observed in Fig. 3(d) that the sorption kinetics is very rapid, and diffusion does not affect the overall removal kinetics. Therefore, bulk transport was the dominant sorption process.

Tab.1 Obtained parameters for various kinetic models
Model K Q (mg/g) R2 α (mg/(g·min))
Pseudo first order 9.79103 min1 171 0.75
Pseudo second order 1.41103 g/(mg·min) 671 0.99
Elovich 0.0199 mg/g 0.93 120201
Intraparticle 8.8539 mg/(g·min1/2)

Tab.1 Obtained parameters for various kinetic models

Fig.3 Fitting of different kinetic models to CeO2 adsorption by IONPs: (a) pseudo-first-order, (b) pseudo-second-order, (c) Elovich, and (d) Intraparticle diffusion model.

Adsorption isotherm studies

Sorption studies of CeO2 on IONPs was carried out and compared with the two most prominent isotherm models Langmuir and Freundlich. Langmuir isotherm (Langmuir, 1918)

CeQ e=CeQm+ 1Qmk1 ,

and the Freundlich model is expressed as

lnQ e=lnKf+nlnCe,

where Qe, Qm, Ce, and kl are the adsorption capacity of IONPs, the maximum adsorption capacity of IONPs, remaining concentration, and Langmuir isotherm constant respectively while Kf and n are the Freundlich constant and adsorption intensity respectively. Langmuir model suggests about the homogeneity of the sorption process and accounts for the surface coverage by balancing the relative rates of adsorption and desorption (dynamic equilibrium). While Freundlich isotherm assumes heterogeneous adsorption of pollutants onto the adsorbent, and it gives an expression defining the surface heterogeneity and the exponential distribution of active sites and their energies. On fitting the data with Langmuir and Freundlich isotherm model as shown in Fig. 4, correlation factor R2 comes out to be 0.99 and 0.95 for Langmuir and Freundlich respectively (Table 2) suggesting that the sorption of CeO2 into IONPs follows better Langmuir isotherm model than Freundlich isotherm model supporting monolayer sorption with having uniform sorption energy of the surface of IONP. Adsorption intensity constant (n) value 0.33 (less than 1) suggests adsorption was favorable in the concentration range studied(Zhou et al., 2017).

Tab.2 Obtained parameters for various sorption isotherms
Sorption isotherm model Parameters
Langmuir
Freundlich
Qmax = 688 mg/g, Kl = 1.57 l mg1, R2 = 0.995
Kf = 12.18, n = 0.33, R2 = 0.92

Tab.2 Obtained parameters for various sorption isotherms

Fig.4 (a) Langmuir and (b) Freundlich adsorption isotherms.

Effect of bicarbonate ions

Bicarbonate is one of the commonly present ions in a natural environment and can affect the sorption; therefore, the effect of bicarbonate ions on the sorption capacity of IONPs for CeO2 was studied in wide concentration range (0.1‒1 mM).It was observed in Fig. 5(a) that the sorption capacity decreases with the increasing concentration of bicarbonate, which can be attributed to the fact that bicarbonate ions can act as direct competitors for available binding sites on IONPs(Zachara et al., 1987). A similar phenomenon was also observed in case of As(III) and As(V) sorption on Goethite(Stachowicz et al., 2007).

Fig.5 (a) Effect of bicarbonate ions (HCO3) concentration on CeO2 sorption and (b) adsorption capacity with varying pH.

Effect of solution pH

Solution pH is the judgmental factor for the sorption experiment as it influences the form of functional groups on the surface of adsorbate or adsorbents. Predominantly, sorption capacity changes with pH due to the change in their electrostatic energy. In DLS measurement, change in zeta potential with pH for CeO2 and IONPs has been observed resulting in change in electrostatic energy between the particles. Results in Fig. 5(b) show maximum sorption capacity near pH 6.5, which can be supported by zeta potentials suggesting maximum electrostatic attraction at the same pH due to drastically opposite charges, as shown in Fig. 2(b). Sorption decreased in acidic pH range due to electrostatic repulsion, but some amount of CeO2, which was adsorbed, can be attributed to Fe-O-Ce-O direct interaction, as discussed in section 3.8.

Sorption in real and synthetic waters

To evaluate the practicality of sorption methodology of IONPs in the natural aqueous environment, for that waters with three different chemical composition, i.e. synthetic soft, hard and groundwater were prepared in the laboratory by mixing respective salts to carry out CeO2 sorption experiments(Smith et al., 2002; Xu et al., 2016). Besides, environmental samples of groundwater and river water were also utilized for the same to compare the change in the sorption capacity of IONPs. Similar CeO2 concentrations, i.e., 70 mg/L was spiked in all solutions with 100 mg/L of IONPs and results in Fig. 6 show that removal was more than 90% suggesting that the matrix have negligible influence on the sorption capacity of IONPs. Therefore, IONPs could be utilized for the removal of CeO2 NPs from contaminated environmental water samples.

Fig.6 Removal percentage of CeO2 NPs from complex matrix.

Adsorption and recovery

For the removal of contaminants from the environment, cost factor remains one of the important criteria for the selection of the removal process and adsorbents.Therefore, the reusability of IONPs was evaluated by subjecting the reaction precipitate to ultrasonication, followed by magnetic separation. Sorption process shows the removal efficiency of nearly 88% of CeO2 NPs. While the recovery efficiency was nearly 80% of CeO2 NPs through ultrasonication. It suggests that the pores of IONPs are available again for adsorption and can be reused for the sorption of CeO2 NPs. A schematic is given in Fig. 7 for adsorption-desorption experiments, and Fig. 8(f) provides a visual image of recovered CeO2.

Fig.7 CeO2 adsorption and recovery methodology for IONPs.

CeO2 and IONPs interactions: detailed insights about removal mechanisms

To confirm various removal mechanisms stated and interpreted through kinetic and isotherm modeling, CeO2-IONPs reaction precipitate was collected and characterized using TEM and STEM. As shown in Fig. 8(a), CeO2 NPs can be seen all along the circumference of IONPs, while STEM image in Fig. 8(b) confirms that these small particles were all along the surface of IONPs. TEM elemental mapping was performed to confirm the presence of CeO2 on the IONP surface, and data in Fig. 8(c) confirms the STEM observations, as Ce was distributed all along with IONPs. Significant differences in hydrodynamic diameters of CeO2 NPs and IONPs, i.e.,>100 nm for CeO2 and>200 nm for IONPs, allowed us to monitor the change in hydrodynamic diameter with time in the system containing IONPs and CeO2. Results show in Fig. 8(d) that separate intensity peak for CeO2 at<100 nm, which was present till 100 s of interaction, completely disappears after 600 s and simultaneously single peak dominates with broad size fractions, suggesting attachment of CeO2NPs with IONPs. To confirm whether the sorption was only due to electrostatic attraction or surface complexation was also involved in the removal process, reaction precipitate was analyzed using FT-IR to see the changes in functional groups stretching after sorption of CeO2. Spectra provided in Fig. 8(e) shows that the intense band at 500 cm1 which corresponds to the Ce-O stretching vibration and available in blank CeO2 (Zamiri et al., 2015); Pujar et al., 2018) has also generated in CeO2-IONP reaction precipitate confirming the electrostatic attachment of CeO2. Further, sharp peak at 580 cm1 which can be ascribed to the Fe-O vibration from the magnetite phase (Yang et al., 2015) in case of IONP was disturbed and flattened after interaction with CeO2 NPs suggesting that apart from electrostatic attraction, Fe-O-Ce-O type surface complexation was also helping in CeO2 removal(Gan et al., 2017). This observation also confirms why CeO2 removal was also observed in acidic pH range where both the particles were positively charged.

Fig.8 (a, b) TEM and STEM image of CeO2-IONPs reaction precipitate respectively (c) elemental distribution along spherical IONPs in reaction mixture (d) DLS analysis of CeO2 and IONPs mixture suspension with time (e) FT-IR spectra of IONP, CeO2 and IONP-CeO2 reaction precipitate and (f) experimental images showing blank CeO2 suspension, magnetic separation after sorption, redispersion using sonication and recovered CeO2.

Conclusions

Synthesized IONPs show very high sorption efficiency for toxic metal oxides like CeO2, i.e., 688 mg/g. It follows the pseudo-second-order reaction kinetic model suggesting chemisorption as a probable sorption mechanism. Following Langmuir sorption isotherm (R2>0.99) further supports that the sorption process was chemical and monolayer in nature. CeO2 recovery through ultrasonication after sorption was >80% confirming their reusability. TEM/STEM-elemental mapping along with FT-IR of IONP-CeO2 reaction precipitate combinedly confirm that the sorption was due to both electrostatic attraction and surface complexation. Results also show that there is no significant decrease in the sorption capacity of IONPs in the presence of synthetic and natural water samples (>90% removal); therefore, IONPs could be utilized for the removal of CeO2 NPs from contaminated environmental water samples.

Acknowledgements

We recognize financial support from the SERB-Ramanujan Fellowship grant (SB/S2/RJN-006/2016) and SERB-ECR project grant (ECR/2017/000707) from Department of Science and Technology (DST), India. We are also thankful to the Indian Institute of Science Education and Research Kolkata’s central instrumentation facility for TEM, FESEM, and PXRD analysis.

References
1 A Ali, H Zafar, M Zia, I Ul Haq, A R Phull, J S Ali, A Hussain (2016). Synthesis, characterization, applications, and challenges of iron oxide nanoparticles. Nanotechnology, Science and Applications, 9: 49–67
DOI:10.2147/NSA.S99986      PMID:27578966     
2 M R Chang, D J Lee, J Y Lai (2007). Nanoparticles in wastewater from a science-based industrial park: Coagulation using polyaluminum chloride. Journal of Environmental Management, 85(4): 1009–1014
DOI:10.1016/j.jenvman.2006.11.013      PMID:17202026     
3 Y Chen, E J Bylaska, J H Weare (2017). Weakly bound water structure, bond valence saturation and water dynamics at the goethite (100) surface/aqueous interface: ab initio dynamical simulations. Geochemical Transactions, 18(1): 3
DOI:10.1186/s12932-017-0040-5      PMID:29086806     
4 S H Chien, W R Clayton, G H Mcclellan (1980). Kinetics of dissolution of phosphate rocks in soils. Soil Science Society of America Journal, 44(2): 260–264
5 G K Darbha, C Fischer, J Luetzenkirchen, T Schäfer (2012). Site-specific retention of colloids at rough rock surfaces. Environmental Science & Technology, 46(17): 9378–9387
DOI:10.1021/es301969m      PMID:22861645     
6 G K Darbha, T Schäfer, F Heberling, A Lüttge, C Fischer (2010). Retention of latex colloids on calcite as a function of surface roughness and topography. Langmuir, 26(7): 4743–4752
DOI:10.1021/la9033595      PMID:20201604     
7 G K Darbha, A K Singh, U S Rai, E Yu, H Yu, P Chandra Ray (2008). Selective detection of mercury (II) ion using nonlinear optical properties of gold nanoparticles. Journal of the American Chemical Society, 130(25): 8038–8043
DOI:10.1021/ja801412b      PMID:18517205     
8 Y Deng, D Qi, C Deng, X Zhang, D Zhao (2008). Superparamagnetic high-magnetization microspheres with an Fe3O4@SiO2 core and perpendicularly aligned mesoporous SiO2 shell for removal of microcystins. Journal of the American Chemical Society, 130(1): 28–29
DOI:10.1021/ja0777584      PMID:18076180     
9 F Dvořák, L Szabová, V Johánek, M Farnesi Camellone, V Stetsovych, M Vorokhta, A Tovt, T Skála, I Matolínová, Y Tateyama, J Mysliveček, S Fabris, V Matolín (2018). Bulk hydroxylation and effective water splitting by highly reduced cerium oxide: The role of O vacancy coordination. ACS Catalysis, 8(5): 4354–4363
10 G Gan, J Liu, Z Zhu, Z Yang, C Zhang, X Hou (2017). A novel magnetic nanoscaled Fe3O4/CeO2 composite prepared by oxidation-precipitation process and its application for degradation of orange G in aqueous solution as Fenton-like heterogeneous catalyst. Chemosphere, 168: 254–263
DOI:10.1016/j.chemosphere.2016.10.064      PMID:27788364     
11 R Ganesh, J Smeraldi, T Hosseini, L Khatib, B H Olson, D Rosso (2010). Evaluation of nanocopper removal and toxicity in municipal wastewaters. Environmental Science & Technology, 44(20): 7808–7813
DOI:10.1021/es101355k      PMID:20853883     
12 C Gunawan, M S Lord, E Lovell, R J Wong, M S Jung, D Oscar, R Mann, R Amal (2019). Oxygen-vacancy engineering of cerium-oxide nanoparticles for antioxidant activity. Acs Omega, 4(5): 9473–9479
13 Q G He, J Liu, J Liang, X P Liu, Z Y Ding, D Tuo, W Li (2018). Sodium acetate orientated hollow/mesoporous magnetite nanoparticles: Facile synthesis, characterization and formation mechanism. Applied Sciences-Basel, 8(2): 10.3390/app8020292 
14 Y S Ho, G Mckay (1999). Pseudo-second order model for sorption processes. Process Biochemistry, 34(5): 451–465
15 M Hua, S Zhang, B Pan, W Zhang, L Lv, Q Zhang (2012). Heavy metal removal from water/wastewater by nanosized metal oxides: A review. Journal of Hazardous Materials, 211– 212: 317–331
DOI:10.1016/j.jhazmat.2011.10.016      PMID:22018872     
16 K Juganson, A Ivask, I Blinova, M Mortimer, A Kahru (2015). NanoE-Tox: New and in-depth database concerning ecotoxicity of nanomaterials. Beilstein Journal of Nanotechnology, 6: 1788–1804
DOI:10.3762/bjnano.6.183      PMID:26425431     
17 R Kaegi (2009). Nanoparticles in drinking water. EAWAG News, 66
18 N Khandelwal, N Singh, E Tiwari, G K Darbha (2019). Novel synthesis of a clay supported amorphous aluminum nanocomposite and its application in removal of hexavalent chromium from aqueous solutions. RSC Advances, 9(20): 11160–11169
DOI:10.1039/C9RA00742C     
19 M Kumari, S P Singh, S Chinde, M F Rahman, M Mahboob, P Grover (2014). Toxicity study of cerium oxide nanoparticles in human neuroblastoma cells. International Journal of Toxicology, 33(2): 86–97
DOI:10.1177/1091581814522305      PMID:24510415     
20 I Langmuir (1918). The adsorption of gases on plane surfaces of glass, mica and platinum. Journal of the American Chemical Society, 40(9): 1361–1403
DOI:10.1021/ja02242a004     
21 L Largitte, R Pasquier (2016). A review of the kinetics adsorption models and their application to the adsorption of lead by an activated carbon. Chemical Engineering Research & Design, 109: 495–504
22 W Lin, Y W Huang, X D Zhou, Y Ma (2006). Toxicity of cerium oxide nanoparticles in human lung cancer cells. International Journal of Toxicology, 25(6): 451–457
DOI:10.1080/10915810600959543      PMID:17132603     
23 Y T Lin, M Sung, P F Sanders, A Marinucci, C P Huang (2007). Separation of nano-sized colloidal particles using cross-flow electro-filtration. Separation and Purification Technology, 58(1): 138–147
24 J F Liu, J Sun, G B Jiang (2010). Use of cloud point extraction for removal of nanosized copper oxide from wastewater. Chinese Science Bulletin, 55(4–5): 346–349
25 M Masserini (2013). Nanoparticles for brain drug delivery. ISRN Biochemistry, 2013: 238428–238428
DOI:10.1155/2013/238428      PMID:25937958     
26 M I Morales, C M Rico, J A Hernandez-Viezcas, J E Nunez, A C Barrios, A Tafoya, J P Flores-Marges, J R Peralta-Videa, J L Gardea-Torresdey (2013). Toxicity assessment of cerium oxide nanoparticles in cilantro (Coriandrum sativum L.) plants grown in organic soil. Journal of Agricultural and Food Chemistry, 61(26): 6224–6230
DOI:10.1021/jf401628v      PMID:23799644     
27 S Omar (2019). Doped ceria for solid oxide fuel cells. Intech Open, Chapter 4:43– 59
28 L Pang, J F Liu (2013). Use of Fe3O4@nSiO2@mSiO2 magnetic mesoporous microspheres for fast determination of the sorption coefficients of polycyclic aromatic hydrocarbons to bovine serum albumin in aqueous phase. Acta Chimica Sinica, 71(3): 339–342
DOI:10.6023/A13010089     
29 M S Pujar, S M Hunagund, V R Desai, S Patil, A H Sidarai (2018). One-step synthesis and characterizations of cerium oxide nanoparticles in an ambient temperature via Co-precipitation method. AIP Conference Proceedings, 1942(1): 050026
DOI:10.1063/1.5028657     
30 K Ranjit, A Ahmed (2013). Nanoparticle: An overview of preparation, characterization, and application. Journal of Applied Pharmaceutical Science, 1(6):228– 234
31 S Sen Gupta, K G Bhattacharyya (2011). Kinetics of adsorption of metal ions on inorganic materials: A review. Advances in Colloid and Interface Science, 162(1–2): 39–58
DOI:10.1016/j.cis.2010.12.004      PMID:21272842     
32 E J Smith, W Davison, J Hamilton-Taylor (2002). Methods for preparing synthetic freshwaters. Water Research, 36(5): 1286–1296
DOI:10.1016/S0043-1354(01)00341-4      PMID:11902783     
33 F Springer, S Laborie, C Guigui (2013). Removal of SiO2 nanoparticles from industry wastewaters and subsurface waters by ultrafiltration: Investigation of process efficiency, deposit properties and fouling mechanism. Separation and Purification Technology, 108: 6–14
34 A Srinivas, P J Rao, G Selvam, P B Murthy, P N Reddy (2011). Acute inhalation toxicity of cerium oxide nanoparticles in rats. Toxicology Letters, 205(2): 105–115
DOI:10.1016/j.toxlet.2011.05.1027      PMID:21624445     
35 S K Srivastava, R Tyagi, N Pant (1989). Adsorption of heavy-metal ions on carbonaceous material developed from the waste slurry generated in local fertilizer plants. Water Research, 23(9): 1161–1165
36 M Stachowicz, T Hiemstra, W H van Riemsdijk (2007). Arsenic-bicarbonate interaction on goethite particles. Environmental Science & Technology, 41(16): 5620–5625
DOI:10.1021/es063087i      PMID:17874764     
37 T Suzuki, I Kosacki, H U Anderson (2002). Defect and mixed conductivity in nanocrystalline doped cerium oxide. Journal of the American Ceramic Society, 85(6): 1492–1498
38 W Y Teng, S C Jeng, C W Kuo, Y R Lin, C C Liao, W K Chin (2008). Nanoparticles-doped guest-host liquid crystal displays. Optics Letters, 33(15): 1663–1665
DOI:10.1364/OL.33.001663      PMID:18670496     
39 A Trovarelli (1996). Catalytic properties of ceria and CeO2-containing materials. Catalysis Reviews, 38(4): 439–520
40 J W Wiechers, N Musee (2010). Engineered inorganic nanoparticles and cosmetics: facts, issues, knowledge gaps and challenges. Journal of Biomedical Nanotechnology, 6(5): 408–431
DOI:10.1166/jbn.2010.1143      PMID:21329039     
41 H Xu, Y Sun, J Li, F Li, X Guan (2016). Aging of zerovalent iron in synthetic groundwater: X-ray photoelectron spectroscopy depth profiling characterization and depassivation with uniform magnetic field. Environmental Science & Technology, 50(15): 8214–8222
DOI:10.1021/acs.est.6b01763      PMID:27384928     
42 S Yang, T Zeng, Y Li, J Liu, Q Chen, J Zhou, Y Ye, B Tang (2015). Preparation of graphene oxide decorated Fe3O4@SiO2 nanocomposites with superior adsorption capacity and SERS detection for organic dyes. Journal of Nanomaterials, 2015: 8
DOI:10.1155/2015/817924     
43 J M Zachara, D C Girvin, R L Schmidt, C T Resch (1987). Chromate adsorption on amorphous iron oxyhydroxide in the presence of major groundwater ions. Environmental Science & Technology, 21(6): 589–594
DOI:10.1021/es00160a010      PMID:19994980     
44 R Zamiri, H Abbastabar Ahangar, A Kaushal, A Zakaria, G Zamiri, D Tobaldi, J M F Ferreira (2015). Dielectrical properties of CeO2 nanoparticles at different temperatures. Plos One, 10(4): e0122989
45 K Zargoosh, H Abedini, A Abdolmaleki, M R Molavian (2013). Effective removal of heavy metal ions from industrial wastes using thiosalicylhydrazide-modified magnetic nanoparticles. Industrial & Engineering Chemistry Research, 52(42): 14944–14954
DOI:10.1021/ie401971w     
46 X X Zhou, Y J Li, J F Liu (2017). Highly efficient removal of silver-containing nanoparticles in waters by aged iron oxide magnetic particles. ACS Sustainable Chemistry & Engineering, 5(6): 5468–5476
Downloaded    
RichHTML Viewed    
Abstract Viewed    

Share
Export

Related articles:
Key words
Adsorption
toxic metal oxide remediation
eco-friendly IONP
Iron oxide
CeO2 removal

Authors
Kanha Gupta
Nitin Khandelwal
Gopala Krishna Darbha