Next Article in Journal
Blockchain Technology for Oil and Gas: Implications and Adoption Framework Using Agile and Lean Supply Chains
Previous Article in Journal
How to Choose the Suitable Steel of Wellhead, Wellbore, and Downhole Tools for Acid Gas Reinjection Flooding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Catalytic and Photocatalytic Degradation of Methyl Orange Using Doped LaMnO3 Compounds

by
Paula Sfirloaga
1,
Madalina-Gabriela Ivanovici
1,2,
Maria Poienar
1,*,
Catalin Ianasi
3 and
Paulina Vlazan
1
1
National Institute for Research and Development in Electrochemistry and Condensed Matter, 300569 Timişoara, Romania
2
Faculty of Industrial Chemistry and Environmental Engineering, Politehnica University of Timisoara, Bv. Vasile Parvan No. 6, 300223 Timisoara, Romania
3
“Coriolan Drăgulescu” Institute of Chemistry, Bv. Mihai Viteazul, No. 24, 300223 Timişoara, Romania
*
Author to whom correspondence should be addressed.
Processes 2022, 10(12), 2688; https://doi.org/10.3390/pr10122688
Submission received: 4 November 2022 / Revised: 30 November 2022 / Accepted: 9 December 2022 / Published: 13 December 2022
(This article belongs to the Section Separation Processes)

Abstract

:
LaMnO3 and 1% Pd-, Ag-, or Y-doped perovskite type nanomaterials were prepared by the sol-gel method, followed by heat treatment at a low temperature (600 °C for 6 h). The investigation through X-ray diffraction and FT-IR spectroscopy indicated that all samples were well crystallized, without secondary phases, and that the transition metal doping changed the crystal structure from the R-3c space group for the undoped LaMnO3 to the Pm-3m space group for the doped perovskite compounds. In this research paper, the efficiencies of the perovskite LaMnO3 materials for methyl orange removal were analyzed, wherein the effect of the doping ions and of the pH on the catalytic activity were studied together with a kinetic approach for the LaMnO3 materials at different values of the pH. Moreover, in the catalytic activity, it should be noted that a slightly better performance was obtained for the Ag-doped materials compared to the Y- and Pd-doped perovskite samples. The results presented for the perovskite LaMnO3 nanomaterials reinforce the interest in these multifunctional materials to be used in industrial applications; e.g., in water treatment.

1. Introduction

In recent years, due to more pollution in the environment, intensive research has focused on effluent water treatment, as the availability of clean water is a major issue. Advanced oxidation systems have been widely studied for these applications with very good results, based on which the generated radical intermediates oxidize the organic chemicals found in wastewater [1]. In this context, scientists have been searching for promising visible-light photocatalysts compounds and, among them, perovskite ABO3-type materials have presented interesting results [2,3,4,5,6].
One of the main agents causing water pollution are organic-based compounds, especially resulting from industries which have harmful effects on the environment and for human health. Perovskite-type materials have been reported to have promising results, as they can be applied for the degradation of organic pollutants [7,8,9]. LaMnO3 has attracted particular interest, as it has been reported as a green catalyst which can be applied for the degradation of methyl orange (MO) in aqueous solution under different conditions [7,10], the degradation of other organic pollutants such as Rhodamine B (RhB) [11], and presents photocatalytic activity for the removal of methylene violet [8]. On the other hand, the doping influence of photocatalytic activity on the photodegradation of Direct Green BE was found to be higher in Ag-modified LaMnO3-graphene than in pure LaMnO3 and LaMnO3-graphene [12]. Furthermore, Fe-substituted LaMnO3 materials were investigated as heterogeneous catalysts in the H2O2—assisted degradation of anionic dyes (Remazol Turquoise Blue, Remazol Brilliant Yellow) and cationic dyes (methylene blue, Safranine-O) in both the absence and presence of visible light irradiation [13]. Recently, the sustainable removal of 17α-ethynylestradiol (EE2) from aqueous environment has been shown using rare-earth-doped LaMnO3 nanomaterials in [14], and the best removal efficiency was reached in the presence of Ho in LaMnO3, when 77% of endocrine disruptors EE2 were degraded after 30 min of UV irradiation.
In order to provide more insight on and find new candidates for semiconductor photocatalysts, the organic pollutant degradation activity of undoped and 1% Pd-, Ag-, or Y-doped LaMnO3 perovskite materials is investigated and presented in this study; in the case of 1% Pd-, Ag- or Y-doped perovskite compounds, for the first time. Furthermore, a kinetic study was employed for the degradation of MO over the undoped LaMnO3 in dark conditions in the 2−4 pH range, and, to the best of our knowledge, no kinetics have been reported in the scientific literature for the catalytic activity of LaMnO3 to degrade MO. Our results show that these materials with a perovskite-type structure could be considered as viable solar-driven photocatalysts due to their efficiency for MO removal under solar irradiation. The 1% Pd, Ag, or Y doping affects the perovskite crystal structure, and the materials showed catalytic or photocatalytic activity in the acidic medium. A slightly better performance for MO degradation under dark conditions is registered for the Ag-doped LaMnO3 samples compared to the Y- and Pd-doped LaMnO3 materials.

2. Materials and Method

Pristine LaMnO3 and 1% Pd-, Ag-, or Y-doped samples were prepared using the sol-gel method, followed by heat treatment at a low temperature (600 °C for 6 h). The following reagents were used for the synthesis of the studied materials: La(NO3)3·6H2O, Mn(NO3)2·4H2O and 2M NaOH solution, used as starting materials. To study the influence of doping on the catalytic and photocatalytic properties of perovskite-type LaMnO3 materials, the LaMnO3 compounds were doped with 1% Pd, Ag, or Y by adding the following reactants: Pd(NO3)2·2H2O, AgNO3, and Y(NO3)·6H2O. All chemicals were analytical grade from Sigma-Aldrich, and were used as received without any further purification.
The abovementioned precursors were dissolved in water, alcohol mixture, and then, in order to homogenize the mixture, it was stirred at room temperature for 30 min followed by the addition of citric acid under continuous stirring, while the temperature was raised and kept at 140 °C until the gels were obtained. The resulting gels were dried in an oven at 80 °C for 4 h and then calcined at 600 °C for 6 h to obtain the final perovskite samples. Perovskite manganite synthesis by sol-gel technique has been previously reported for undoped [15,16,17,18,19,20,21,22,23], doped (for example with Fe [24], N [16], Ag [25], or Sr [21]), or substituted LaMnO3 materials (with Pd [26], Ca [15], Fe, Co, or Ni [23]), and is considered to be a successful and flexible synthesis route to obtain well-crystallized micro- or nanomaterials from this family of compounds [27].
Room temperature (RT) X-ray powder diffraction (XRD) data were collected using a PANalytical X’Pert PRO MPD diffractometer with Cu-Kα radiation (λ = 1.5418 Å) in the 2θ = 10–80° range. For FT-IR analysis, a Shimadzu Prestige FT-IR spectrometer was used in the 400–4000 cm−1 range using KBr pellets. The morphology of the investigated samples was studied using scanning electron microscopy (SEM) (Inspect S) equipped with an energy dispersive X-ray (EDX).
Investigation of the catalytic and photocatalytic properties of the pristine and doped LaMnO3 with transition metals, Pd, Ag, or Y, were carried out by monitoring the degradation capacity of MO in aqueous solutions considering different experimental conditions, which were selected based on factors that known in the literature to affect catalytic and photocatalytic behavior. Therefore, two aspects were considered for this study: evaluation of the pH effect on the LaMnO3 catalytic activity (selected in the range between 2 and 5) and evaluation of the doping in the perovskite structure on the photocatalytic activity. The pH was modified using small volumes of 0.1M HNO3 (65% reagent Ph. Eur., Sigma Aldrich) solution from the pH = 5 (for the initial concentration of MO aqueous solution) to pH = 4, 3 and 2. A volume of 40 mL MO (reagent Ph Eur.) aqueous solution of 12 ppm initial concentration was used for all experiments. The catalyst (of 120 ppm concentration) was added to a glass container together with the solution and was magnetically stirred for 90 min. In order to determine MO degradation over time, the absorbance of the MO solution was monitored every 10 min using a UV-VIS spectrometer (portable Jaz spectrometer from Ocean Optics). For this, 2–3 mL of the mixed solution (catalyst with MO aqueous solution) was separated by centrifugation and the solution was then inserted into the UV-VIS cuvette. The experimental set-up used for the absorbance measurements was as described in a previous study [28]. The degradation rate of MO (%) was calculated with Equation (1):
D e g r a d a t i o n   r a t e ( % ) = ( A 0 A t ) A 0 · 100
where A0—the absorbance of the MO aqueous solution in the moment of catalyst addition (time t = 0), At—the absorbance of the MO solution at time t.
The catalytic experiments were performed at room temperature in dark conditions, whereas the photocatalytic experiments were performed under simulated solar radiation (provided by Sol2A 94042A, Oriel Instruments/Newport Corporation). The visible and respective UV irradiance at the surface of the mixed solution was 957 W/m2 and 1.02 mW/cm2, respectively. Moreover, a kinetic approach was employed for the undoped perovskite LaMnO3 materials in order to determine the reaction rate of catalytic MO degradation over LaMnO3 when the solution pH was changed from 2 to 5.

3. Results and Discussion

3.1. XRD Analysis

Analysis by X-ray diffraction was performed for the samples obtained as described in the previous section. The sol-gel method with thermal treatment at a low temperature obtains well-crystallized samples without secondary phases (Figure 1), as has been already reported in previous studies for undoped and doped/substituted LaMnO3 perovskite samples [15,16,17,24,25,26]. The diffractogram visualizations in Figure 1 indicate that some of the peaks are split for the undoped material compared to the 1% Pd2+, Ag+, or Y3+ doped LaMnO3 samples (insert Figure 1). The LaMnO3 compound obtained using the sol-gel method is indexed in the R-3 c space group [JCPDS card file no. 01-089-0678] and, for the doped materials, the Pm-3m space group is used [JCPDS card file no. 01-075-0440], as well as for the rare-earth-doped LaMnO3 materials in [14]. Therefore, upon doping in the LaMnO3 samples with Pd, Ag, or Y (this study), with rare earth Eu, Tb, and Ho [14], Pr or Eu [29], Na or Ca ion [30] changes are induced within the perovskite structure, which could have a great effect on the properties of the obtained materials.

3.2. FT-IR Analysis

Figure 2 shows the FT-IR spectra for pristine LaMnO3 and 1% Pd-, Ag-, or Y-doped LaMnO3 samples recorded in the 400–4000 cm−1 spectral range. It is observed that several absorption bands at 433, 619, 630, 853, 1373, 1477, 1637, 2924, and 3437 cm−1 are registered for these samples. The significant absorption bands for the LaMnO3 materials from this work are approximately located in the same spectral range as the perovskite LaMnO3 materials from the previous study obtained using the sol-gel synthesis method [14,17,31].
Two transmission bands around 620 and 430 cm−1 are observed for the materials; the band around 430 cm−1 directly corresponds to the bending mode (νb), owing to the change in the Mn–O–Mn bond angle. The frequency band around 620 cm−1 is related to the stretching mode (νs) of Mn–O or Mn–O–Mn bonds, which embraces the change in the bond length of Mn–O because of the internal motion [32]. These two bands are interrelated to the neighborhoods surrounding the MnO6 octahedral in the ABO3 perovskite and confirm the formation of a perovskite structure [33].
In the FT-IR spectra for the Pd-doped LaMnO3 and Y-doped LaMnO3 samples, some additional peaks at 853, 1373, 1477, 1637, and 2924 cm−1 appear. The low intensity peak located at 857 cm−1, which appears only in doped samples, could be attributed to the Me-O bond (Me = Pd, Y). The two bands located at 1477 and 1373 cm−1 can be attributed to the δOH and γOH vibrational modes of H2O molecules [34].
The 2500–3600 cm−1 region encloses stretching O–H modes (strong and wide depending on the environment), stretching C–H modes (generally strong), and stretching N–H modes typical in amine groups [35]. As in the synthesis process, no organic compounds were used, the 2924 cm−1 peak could be attributed to the O-H group from the water in the environment.
The broad bands of low intensity with the maximum around 3437 and 1637 cm−1 can be assigned to O-H stretching modes of surface-adsorbed water, due to the contact of the sample with the environment. These stretching vibrations of weakly-bound water interact with its environment via hydrogen bonding [36].

3.3. SEM-EDX/ Element Mapping

Figure 3 shows the SEM micrograph of undoped and doped LaMnO3 materials. In the case of the undoped and Pd-doped LaMnO3, the surface morphology shows that the particles are agglomerated into asymmetric formations, although these agglomerations are composed of extremely fine particles. The EDX mapping for the LaMnO3 compound has also been reported in [14], where the uniform distribution of elements was also confirmed.
The qualitative analysis of Ag- or Y-doped lanthanum manganite exhibit spherically aggregated particles free from agglomeration. The EDX spectra show, in addition to the purity of the obtained materials, the presence of the three dopants. Thus, quantification of the elements confirms that the expected level of doping is present in each sample, maintaining the stoichiometry of the compound. Moreover, elements mapping highlighted the uniform distribution of the component elements for each material (Figure 3).

3.4. Catalytic Activity of the LaMnO3 Nanomaterials

3.4.1. The Effect of the pH on the MO Absorption and on the Catalytic MO Degradation over Undoped LMO; Kinetic Parameters Determination

As a first step toward the evaluation of the catalytic properties of the LaMnO3 compound, the change in the visible absorption for the MO aqueous solution was investigated (Figure 4). The maximum absorption of the initially prepared solution of 12 ppm concentration (pH = 5) was registered at a wavelength of 467 nm. The increase in the solution acidity led to a shift of the absorbance band toward a higher wavelength (505 nm) and also to an increase of the band intensity from 0.92 (at pH = 5) to 1.29 (at pH = 4), 1.39 (at pH = 3), and 1.52 (at pH = 2). Visually, the bathochromic shift was observed by the color change from yellow-orange to orange-red nuances (as visualized in Figure 4 down). As reported in the literature, the redshift is obtained as a result of chromophoric group (NN bond) protonation of MO in the acidic medium, as the acidic structures absorbs light in the spectrum of the red-green domain [37,38]. Furthermore, the pH affects the surface charge of the LaMnO3 and the adsorption of the species on the material surface, determining the catalytic activity for MO degradation [9,39].
As represented in Figure 5, the removal of MO from the aqueous solution under dark conditions increased as the pH values decreased from 5 to 2. At pH = 2, the maximum removal of MO (97%) was achieved after 50 min of reaction, whereas at pH = 3, the maximum removal of MO (96%) was reached after 80 min. No degradation occurred at pH = 5, but, at pH = 4, 83% of MO from the aqueous was degraded after 90 min of reaction. The kinetics of the catalytic reactions carried out at pH = 2, 3, and 4 were studied based on the kinetics models of zero-, first-, and second-order reactions. The integrated forms of kinetics equations are:
Zero-order reaction kinetics
C t = C 0 k 0 t
First-order reaction kinetics
C t = C 0 e k 1 t
Second-order reaction kinetics
1 C t = 1 C 0 + k 2 t
where Ct—concentration of reactant at time t, C0—concentration of reactant at initial time, and k0, k1, and k2—rate constant for zero-, first-, and second-order reactions.
The constant rate of MO degradation by LaMnO3 materials carried out in the pH range of 2–4 was determined by plotting ln (Ct/C0), Ct − C0, and 1/Ct1/C0 versus time [40,41,42]. The concentration at time (t) was calculated considering the MO degradation rate and the known initial concentration of the solution. Therefore, the fitting parameters (regression coefficient—R2 and rate constant—k) results for MO degradation at different pH values are presented in Table 1. Based on the calculated regression coefficient, the degradation of MO catalyzed by the LaMnO3 can be attributed to the pseudo-first- and pseudo-second-order reaction for pH = 2, pH = 3, and pH = 4 with corresponding pseudo-constant rates of 0.07 min−1, 0.04 min−1, and 0.01 L mg−1 min−1, respectively [43].

3.4.2. The Dopant Effect on the Catalytic and Photocatalytic Properties of LaMnO3—Comparative Approach

Furthermore, the photocatalytic properties of undoped and 1% Pd-, Ag-, or Y-doped LaMnO3 were studied in order to investigate the impact of transition metals doping in the LaMnO3 perovskite structure (Figure 6). The reaction of MO degradation was carried out in acidic medium, and the solution pH was settled to 4 so the experimental conditions were less severe, involving minimal changes of the medium for the reaction to occur.
When compared to LaMnO3, the degradation rates obtained for Y-, Pd-, and Ag-doped LaMnO3 decreased by 54.63%, 60.31%, and 64.56% after 90 min of reaction. Under artificial solar irradiation, the degradation rate reached 94.7% for the undoped LaMnO3, with close values (84.5%, 88%, and 86%) for LaMnO3 doped with Y, Pd, and Ag, from which 4.65% of MO was degraded by photolysis. The improved efficiency of MO removal under solar irradiation indicates that Y-, Pd-, or Ag-doped LaMnO3 can be considered as solar light active photocatalysts, validating the multifunctional behavior of the perovskites as catalysts. Our findings indicate that the dopants did not improve, nor the catalytic either the photocatalytic properties extensively studied for LaMnO3, but, notably, the Ag doping led to a better performance for MO degradation under dark conditions when compared to Y and Pd, whereas Pd-, Ag-, or Y-doped LaMnO3 showed similar performances as photocatalysts. The MO degradation was induced by the well-known superior oxidation catalytic activity of the perovskites, as has already been reported in [44]. In the ABO3-type perovskite structure, the A cation is the larger in size and is 12-fold coordinated with oxygen anions, while the B cation is 6-fold coordinated with oxygen anions. From the catalytic activity perspective, the A cations are generally catalytically inactive, while the B cations are responsible for the catalytic response of the ABO3, so that B ions are situated at a relatively large distance (ca. 0.4 nm) away from each other, allowing, in this way, a reactant molecule to interact with a single site. Doping the A and B sites is a common practice by which to modify the catalytic response of the perovskites by altering the crystal structure and disturbing the oxidation states of the cations [3,5,12,34]. The structural changes followed the dopant incorporation into the perovskite lattice, depending on the ionic radius of the dopant; thus, in this study, the ionic radii are: Ag+ (1.15 Å), Pd2+ (0.86 Å), Y3+(0.9 Å).
Moreover, a blueshift of the MO absorbance appeared in the photocatalytic reaction (Figure 7), which can be explained by the agglomeration of MO molecules in the aqueous solution as the dye dilution increases, indicative of degradation instead of exclusive adsorption [10,11]. From the point of view of the mechanistic aspect of catalytic and photocatalytic reactions, the reaction occurs as a result of adsorption of the reactant on the catalyst surface, followed by the catalyzed reaction [7,45]. MO degradation is favored in acidic medium, as the pH changes leads to changes in MO adsorption on the catalyst surface and in the photodegradation mechanism. MO is an anionic dye and its adsorption is facilitated in acidic medium and, in addition, it has been reported to be less stable and prone to degradation [46,47]. In dark conditions, the degradation of MO can be explained by the interaction between MO and the catalyst surface, resulting in the degradation of azo bonds of MO and the generation of electrons. Following sequential reactions, the electrons lead to the formation of hydroxyl radicals. Under illumination, the pairs of electrons and holes are formed and lead to the formation of hydroxyl radicals. Therefore, hydroxyl radicals are the main oxidative species responsible for MO degradation, both in dark condition and under illumination [46,47,48,49,50]:
OH + MO   intermediates     SO 4 2 + NO 3 + NH 4 + CO 2 + H 2 O
Table 2 presents the results obtained in others studies related to catalytic MO degradation over LaMnO3 and other types of perovskites carried out in dark conditions or under irradiation (UV, light, or solar radiation). As an example, MO degradation is higher (83%) after 90 min of reaction at pH = 4 and in dark conditions; in this work, when compared with study [10], the degradation of MO is 73% at pH = 3.6. Similar MO degradation efficiencies are obtained in both this work and study [22], under illumination. By summarizing multiple studies regarding the catalytic and photocatalytic properties of perovskites for MO degradation, we consider that this work comes with relevant novelty in terms of the scientific literature by studying the effect of transitional dopants, and of the pH on the photodegradation of MO and on the assisted dark catalysis.

4. Conclusions

Undoped and 1% Pd-, Ag-, or Y-doped LaMnO3 perovskite materials obtained by the sol-gel method, followed by heat treatment at a low temperature (600 oC for 6 h), were studied in this research for the catalytic and photocatalytic degradation of methyl orange. The X-Ray diffraction patterns are characteristic of a perovskite structure with a different symmetry induced by doping (Pm-3m compared to R-3c for undoped LaMnO3). The catalytic activity of LaMnO3 was investigated for the removal of MO from the aqueous solution under dark conditions, and it was observed that it increased as the pH values decreased from 5 to 2; at pH = 2, the maximum removal of MO (97%) was achieved after 50 min of reaction, whereas at pH = 3, the maximum removal of MO (96%) was reached after 80 min; at pH = 4, 83% of MO from the aqueous was degraded after 90 min of reaction, while no degradation occurred at pH = 5. Moreover, the kinetics of catalytic reactions carried out for the different values of the pH were studied based on the kinetics model of zero-, first-, and second-order reactions.
When compared to LaMnO3, the degradation rates obtained for Y-, Pd-, or Ag-doped LaMnO3 decreased by 54.63%, 60.31%, and 64.56% after 90 min of reaction. Under artificial solar irradiation, the degradation rate reached 94.7% for the undoped LaMnO3 and close values (84.5%, 88%, and 86%) were obtained for Y-, Pd-, or Ag-doped LaMnO3 samples. On the other hand, the Ag-doped LaMnO3 led to better performance for MO degradation under dark conditions when compared to Y and Pd, which was probably related to some structural features facilitating this slight enhancement of properties.

Author Contributions

Conceptualization, P.S. and M.P.; methodology, P.S. and M.-G.I.; validation, P.S., M.P. and M.-G.I.; formal analysis, P.V., C.I., M.P., M.-G.I. and P.S.; investigation, P.V., M.P., M.-G.I. and P.S.; data curation, M.P. and M.-G.I.; writing—original draft preparation, P.V., M.P., M.-G.I. and P.S.; writing—review and editing, P.V., C.I., M.P., M.-G.I. and P.S.; supervision, P.S.; project administration, P.S.; funding acquisition, P.S., P.V. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Experimental Demonstrative Project 683PED/21/06/2022.

Data Availability Statement

The data that support the findings of this study are available from the authors, upon reasonable request.

Acknowledgments

Financial support for this work was provided by the Experimental Demonstrative Project 683PED/21/06/2022.

Conflicts of Interest

The authors have no relevant financial or non-financial interests to disclose.

References

  1. Rojas-Cervantes, M.L.; Castillejos, E. Perovskites as Catalysts in Advanced Oxidation Processes for Wastewater Treatment. Catalyst 2019, 9, 230. [Google Scholar] [CrossRef] [Green Version]
  2. Hu, J.; Men, J.; Ma, J.; Huang, H. Preparation of LaMnO3/graphene thin films and their photocatalytic activity. J. Rare Earths 2014, 32, 1126–1134. [Google Scholar] [CrossRef]
  3. Hu, J.; Liu, Y.; Men, J.; Zhang, L.; Huang, H. Ag modified LaMnO3 nanorods-reduced graphene oxide composite applied in the photocatalytic discoloration of direct green. Solid State Sci. 2016, 61, 239–245. [Google Scholar] [CrossRef]
  4. Gao, P.; Li, N.; Wang, A.; Wang, X.; Zhang, T. Perovskite LaMnO3 hollow nanospheres: The synthesis and the application in catalytic wet air oxidation of phenol. Mater. Lett. 2013, 92, 173–176. [Google Scholar] [CrossRef]
  5. Andrei, F.; Zavoianu, R.; Marcu, I.C. Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications. Materials 2020, 13, 5555. [Google Scholar] [CrossRef]
  6. Labhasetwar, N.; Saravanan, G.; Megarajan, S.K.; Manwar, N.; Khobragade, R.; Doggali, P.; Grasset, F. Perovskite-type catalytic materials for environmental applications. Sci. Technol. Adv. Mater. 2015, 16, 036002. [Google Scholar] [CrossRef]
  7. Dhinesh Kumar, R.; Thangappan, R.; Jayavel, R. Enhanced visible light photocatalytic activity of LaMnO3 nanostructures for water purification. Res. Chem. Intermed. 2018, 44, 4323–4337. [Google Scholar] [CrossRef]
  8. Priyatharshni, S.; Kumar, S.R.; Viswanathan, C.; Ponpandian, N. Morphologically tuned LaMnO3 as an efficient nanocatalyst for the removal of organic dye from aqueous solution under sunlight. J. Environ. Chem. Eng. 2020, 8, 104146. [Google Scholar] [CrossRef]
  9. Tran, T.H.; Phi, T.H.; Nguyen, H.N.; Pham, N.H.; Nguyen, C.V.; Ho, K.H.; Doan, Q.K.; Le, V.Q.; Nguyen, T.T.; Nguyen, V.T. Sr doped LaMnO3 nanoparticles prepared by microwave combustion method: A recyclable visible light photocatalyst. Results Phys. 2020, 19, 103417. [Google Scholar] [CrossRef]
  10. Rekavandi, N.; Malekzadeh, A.; Ghiasi, E. Methyl orange degradation using nano-LaMnO3 as a green catalyst under the mild conditions. Nanochem. Res. 2019, 4, 1–10. [Google Scholar]
  11. Dhiman, T.K.; Singh, S. Enhanced catalytic and photocatalytic degradation of organic pollutans Rhodamine -B by LaMnO3 nanoparticles synthesized by non -aqueous sol-gel route. Phys. Status Solidi 2019, 216, 1900012. [Google Scholar] [CrossRef]
  12. Hu, J.; Men, J.; Liu, Y.; Huang, H.; Jiao, T. One-pot synthesis of Ag-modified LaMnO3-graphene hybrid photocatalysts and application in the photocatalytic discoloration of an azo-dye. RSC Adv. 2015, 5, 54028–54036. [Google Scholar] [CrossRef]
  13. Jauhar, S.; Dhiman, M.; Bansal, S.; Singhal, S. Mn3+ ion in perovskite lattice: A potential Fenton’s reagent exhibiting remarkably enhanced degradation of cationic and anionic dyes. J. Sol-Gel Sci. Technol. 2015, 75, 124–133. [Google Scholar] [CrossRef]
  14. Šojić Merkulov, D.; Vlazan, P.; Poienar, M.; Bognár, S.; Ianasi, C.; Sfirloaga, P. Sustainable removal of 17α-ethynylestradiol from aqueous environment using rare earth doped lanthanum manganite nanomaterials. Catal. Today 2022. [Google Scholar] [CrossRef]
  15. Sfirloaga, P.; Poienar, M.; Malaescu, I.; Lungu, A.; Mihali, C.V.; Vlazan, P. Electrical conductivity of Ca-substituted lanthanum manganites. Ceram. Int. 2018, 44, 5823–5828. [Google Scholar] [CrossRef]
  16. Sfirloaga, P.; Sebarchievici, I.; Taranu, B.; Poienar, M.; Vlase, G.; Vlase, T.; Vlazan, P. Investigation of physico-chemical features of lanthanum manganite with nitrogen addition. J. Alloys Compd. 2020, 843, 155854. [Google Scholar] [CrossRef]
  17. Sfirloaga, P.; Poienar, M.; Malaescu, I.; Lungu, A.; Vlazan, P. Perovskite type lanthanum manganite: Morpho-structural analysis and electrical investigations. J. Rare Earths 2018, 36, 499–504. [Google Scholar] [CrossRef]
  18. El-Moez, A.; Mohamed, A.; Alvarez-Alonso, P.; Hernando, B. The intrinsic exchange bias effect in the LaMnO3 and LaFeO3 compounds. J. Alloys. Compd. 2021, 850, 156713. [Google Scholar] [CrossRef]
  19. Elsiddig, Z.A.; Xu, H.; Wang, D.; Zhang, W.; Guo, X.; Zhang, Y.; Sun, Z.; Chen, J. Modulating Mn4+ Ions and Oxygen Vacancies in Nonstoichiometric LaMnO3 Perovskite by a Facile Sol-Gel Method as High-Performance Supercapacitor Electrodes. Electrochim. Acta 2017, 253, 422–429. [Google Scholar] [CrossRef]
  20. Li, Y.; Xue, L.; Fan, L.; Yan, Y. The effect of citric acid to metal nitrates molar ratio on sol–gel combustion synthesis of nanocrystalline LaMnO3 powders. J. Alloys Compd. 2009, 478, 493–497. [Google Scholar] [CrossRef]
  21. Onrubia, J.A.; Pereda-Ayo, B.; De-La-Torre, U.; González-Velasco, J.R. Key factors in Sr-doped LaBO3 (B = Co or Mn) perovskites for NO oxidation in efficient diesel exhaust purification. Appl. Catal. B Environ. 2017, 213, 198–210. [Google Scholar] [CrossRef]
  22. Shaterian, M.; Enhessari, M.; h Rabbani, D.; Asghari, M.; Salavati-Niasari, M. Synthesis, Characterization and Photocatalytic Activity of LaMnO3 Nanoparticles. Appl. Surf. Sci. 2014, 318, 213–217. [Google Scholar] [CrossRef]
  23. Yin, X.; Wang, S.; Wang, B.; Shen, L. Perovskite-type LaMn1−xBxO3+δ (B = Fe, Co and Ni) as oxygen carriers for chemical looping steam methane reforming. Chem. Eng. J. 2021, 422, 128751. [Google Scholar] [CrossRef]
  24. Sfirloaga, P.; Malaescu, I.; Marin, C.N.; Poienar, M.; Vlazan, P. Effect of Fe-Doping on the Structural, Morphological and Electrical Properties of LaMnO3. AIP Conf. Proc. 2020, 2218, 040003. [Google Scholar]
  25. Sfirloaga, P.; Vlase, G.; Vlase, T.; Malaescu, I.; Marin, C.N.; Vlazan, P. Silver doping in lanthanum manganite materials: Structural and electrical properties. J. Therm. Anal. Calorim. 2020, 142, 1817–1823. [Google Scholar] [CrossRef]
  26. Sfirloaga, P.; Malaescu, I.; Marin, C.N.; Vlazan, P. The effect of partial substitution of Pd in LaMnO3 polycrystalline materials synthesized by sol–gel technique on the electrical performance. J. Sol-Gel Sci. Technol. 2019, 92, 537–545. [Google Scholar] [CrossRef]
  27. Navas, D.; Fuentes, S.; Castro-Alvarez, A.; Chavez-Angel, E. Review on Sol-Gel Synthesis of Perovskite and Oxide Nanomaterials. Gels 2021, 7, 275. [Google Scholar] [CrossRef]
  28. Taranu, B.O.; Ivanovici, M.-G.; Svera, P.; Vlazan, P.; Sfirloaga, P.; Poienar, M. Ni11(HPO3)8(OH)6 multifunctional materials: Electrodes for oxygen evolution reaction and potential visible-light active photocatalysts. J. Alloys Compd. 2020, 848, 156595. [Google Scholar] [CrossRef]
  29. Najjar, H.; Bâtis, H.; Lamonier, J.-F.; Mentré, O.; Giraudon, J.-M. Effect of praseodymium and europium doping in La1−xLnxMnO3+δ (Ln: Pr or Eu, 0 ≤ x ≤ 1) perovskite catalysts for total methane oxidation. Appl. Catal. A: Gen. 2014, 469, 98–107. [Google Scholar] [CrossRef]
  30. Malavasi, L.; Ritter, C.; Mozzati, M.C.; Tealdi, C.; Saiful Islam, M.; Azzoni, C.B.; Flor, G. Effects of cation vacancy distribution in doped LaMnO3+δ perovskites. J. Solid State Chem. 2005, 178, 2042–2049. [Google Scholar] [CrossRef] [Green Version]
  31. Taranu, B.-O.; Vlazan, P.; Svera, P.; Poienar, M.; Sfirloaga, P. New functional hybrid materials based on clay minerals for enhanced electrocatalytic activity. J. Alloys Compd. 2022, 892, 162239. [Google Scholar] [CrossRef]
  32. Gao, F.; Lewis, R.A.; Wang, X.L.; Dou, S.X. Infrared absorption of lanthanum manganites. Phys. C. Supercond. Appl. 2000, 341, 2235–2236. [Google Scholar] [CrossRef]
  33. Afify, M.S.; El Faham, M.M.; Eldemerdash, U.; El Rouby, W.M.A.; El-Dek, S.I. Room temperature ferromagnetism in Ag doped LaMnO3 nanoparticles. J. Alloys Compd. 2021, 861, 158570. [Google Scholar] [CrossRef]
  34. Ansari, A.A.; Ahmad, N.; Alam, M.; Adil, S.; Shahid, F.; Ramay, M.; Albadri, A.; Ahmad, A.; Al-Enizi, A.M.; Alrayes, B.F.; et al. Physico-chemical properties and catalytic activity of the sol-gel prepared Ce-ion doped LaMnO3 perovskites. Sci. Rep. 2019, 9, 7747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Silverstein, R.M.; Bassler, G.C.; Morrill, T.C. Spectrometric Identification of Organic Compounds, 4th ed.; John Wiley and Sons: New York, NY, USA, 1981. [Google Scholar]
  36. Aal, A.A.; Hammad, T.R.; Zawrah, M.; Battisha, I.K.; Abou Hammad, A.B. FTIR study of nanostructure perovskite BaTiO3 doped with both Fe3+ and Ni2+ ions prepared by sol-gel technique. Acta Phys. Pol. A 2014, 126, 1318–1322. [Google Scholar]
  37. Del Nero, J.; de Araujo, R.E.; Gomes, A.S.L.; de Melo, C.P. Theoretical and experimental investigation of the second hyperpolarizabilities of methyl orange. J. Chem. Phys. 2005, 122, 104506. [Google Scholar] [CrossRef]
  38. Elumalai, S.; Muthuraman, G. Recovery of Methyl Orange and Congo Red from aqueous solutions using tri-octyl amine (TOA) in benzene as carrier. Process Saf. Environ. Prot. 2015, 96, 177–183. [Google Scholar] [CrossRef]
  39. Rahman, M.; Pinky, T.A.; Mondal, D.C.; Abedin, M.; Hasan, M. The Study of the Photocatalytic Degradation of Methyl Orange in the Presence of Zinc Oxide (ZnO) Suspension. J. Mater. Sci. Res. Rev. 2020, 5, 1–14. [Google Scholar]
  40. Youssef, N.A.; Shaban, S.A.; Ibrahim, F.A.; Mahmoud, A.S. Degradation of methyl orange using Fenton catalytic reaction. Egypt. J. Pet. 2016, 25, 317–321. [Google Scholar] [CrossRef] [Green Version]
  41. Bala, S.S.; Alkhatib, A.J.; Bashir, S.S.; Abdulhadi, M. Photocatalytic Degradation of Indigo Carmine in Aqueous Solutions by the Antibacterial Agent Pefloxacin and UVA. Biomed. J. Sci. Tech. Res. 2018, 5, 4903–4909. [Google Scholar]
  42. Olajire, A.A.; Olajide, A.J. Kinetic Study of Decolorization of Methylene Blue with Sodium Sulphite in Aqueous Media: Influence of Transition Metal Ions. J. Phys. Chem. Biophys. 2014, 4, 2. [Google Scholar]
  43. Coates, E. Aggregation of Dyes in Aqueous Solutions. J. Soc. Dye. Colour. 1969, 85, 355–368. [Google Scholar] [CrossRef]
  44. Barresi, A.A.; Mazza, D.; Ronchetti, S.; Spinicci, R.; Vallino, M. Non-stoichiometry and catalytic activity in ABO3 perovskites: LaMnO3 and LaFeO3. Stud. Surf. Sci. Catal. 2000, 130, 1223–1228. [Google Scholar]
  45. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M. A Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative overview. RSC Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  46. Lin, X.; Huang, F.; Wang, W.; Shan, Z.; Shi, J. Methyl orange degradation over a novel Bi-based photocatalyst Bi 3SbO 7: Correlation of crystal structure to photocatalytic activity. Dye. Pigment. 2008, 78, 39–47. [Google Scholar] [CrossRef]
  47. Ghiasi, M.; Malekzadeh, A. Solar photocatalytic degradation of methyl orange over La0.7Sr0.3MnO3 nano-perovskite. Sep. Purif. Technol. 2014, 134, 12–19. [Google Scholar] [CrossRef]
  48. Guo, J.; Chen, X.; Shi, Y.; Lan, Y.; Qin, C. Rapid Photodegradation of Methyl Orange (MO) Assisted with Cu(II) and Tartaric Acid. PLoS ONE 2015, 10, e0134298. [Google Scholar] [CrossRef]
  49. Verduzco, L.E.; Garcia-Díaz, R.; Martinez, A.I.; Almanza Salgado, R.; Mendez-Arriaga, F.; Lozano-Morales, S.A.; Avendano-Alejo, M.; Padmasree, K.P. Degradation efficiency of methyl orange dye by La0.5Sr0.5CoO3 perovskite oxide under dark and UV irradiated conditions. Dye. Pigment. 2020, 183, 108743. [Google Scholar] [CrossRef]
  50. Naikwade, A.G.; Jagadale, M.B.; Kale, D.P.; Gophane, A.D.; Garadkar, K.M.; and Rashinkar, G.S. Photocatalytic Degradation of Methyl Orange by Magnetically Retrievable Supported Ionic Liquid Phase Photocatalyst. ACS Omega 2020, 5, 131–144. [Google Scholar]
  51. Huang, H.; Sun, G.; Hu, J.; Jiao, T. Single-Step Synthesis of LaMnO3/MWCNT Nanocomposites and Their Photocatalytic Activities. Nanomater. Nanotechnol. 2014, 4, 27. [Google Scholar] [CrossRef] [Green Version]
  52. Zhong, W.; Jiang, T.; Dang, Y.; He, J.; Chen, S.-Y.; Kuo, C.-H.; Kriz, D.; Meng, Y.; Meguerdichian, A.; Sui, S.L. Mechanism studies on methyl orange dye degradation by perovskite-type LaNiO3-δ under dark ambient conditions. Appl. Catal. A Gen. 2018, 549, 302–309. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for the undoped and 1% Pd-, Ag-, or Y-doped LaMnO3 perovskite materials. Insert: Zoom of the XRD patterns in the 2θ = 30.5–35.5° and 2θ = 62–80° regions.
Figure 1. XRD patterns for the undoped and 1% Pd-, Ag-, or Y-doped LaMnO3 perovskite materials. Insert: Zoom of the XRD patterns in the 2θ = 30.5–35.5° and 2θ = 62–80° regions.
Processes 10 02688 g001
Figure 2. FT-IR spectra of undoped LaMnO3 and 1% Pd-, Ag-, or Y-doped LaMnO3 materials.
Figure 2. FT-IR spectra of undoped LaMnO3 and 1% Pd-, Ag-, or Y-doped LaMnO3 materials.
Processes 10 02688 g002
Figure 3. SEM images, element mapping, EDX spectrum, and quantification of undoped, Ag-, Pd-, and Y-doped LaMnO3 materials.
Figure 3. SEM images, element mapping, EDX spectrum, and quantification of undoped, Ag-, Pd-, and Y-doped LaMnO3 materials.
Processes 10 02688 g003
Figure 4. Bathochromic shift of the MO solutions absorbance band caused by the increase of the aqueous solution acidity (up). Color of the MO aqueous solution before (left) and after (right) the addition of small volumes of HNO3 (0.1M) (down).
Figure 4. Bathochromic shift of the MO solutions absorbance band caused by the increase of the aqueous solution acidity (up). Color of the MO aqueous solution before (left) and after (right) the addition of small volumes of HNO3 (0.1M) (down).
Processes 10 02688 g004
Figure 5. Catalytic degradation of MO (40 mL of 12 ppm initial concentration) by LaMnO3 (120 ppm concentration) in the pH range values of 2–5.
Figure 5. Catalytic degradation of MO (40 mL of 12 ppm initial concentration) by LaMnO3 (120 ppm concentration) in the pH range values of 2–5.
Processes 10 02688 g005
Figure 6. Degradation of MO from aqueous solution (40 mL of 12 ppm initial concentration) by 1% Pd-, Ag-, or Y-doped and undoped LaMnO3 (120 ppm catalyst concentration) in acidic medium (pH = 4) under dark (left) and artificial solar light (right).
Figure 6. Degradation of MO from aqueous solution (40 mL of 12 ppm initial concentration) by 1% Pd-, Ag-, or Y-doped and undoped LaMnO3 (120 ppm catalyst concentration) in acidic medium (pH = 4) under dark (left) and artificial solar light (right).
Processes 10 02688 g006
Figure 7. Blueshift of MO absorbance peak in the photodegradation process by undoped and 1% Pd-, Ag-, and Y-doped LaMnO3.
Figure 7. Blueshift of MO absorbance peak in the photodegradation process by undoped and 1% Pd-, Ag-, and Y-doped LaMnO3.
Processes 10 02688 g007
Table 1. Fitting parameters calculated for catalytic degradation reaction of MO obtained at different pH values (2, 3, and 4).
Table 1. Fitting parameters calculated for catalytic degradation reaction of MO obtained at different pH values (2, 3, and 4).
Fitting Parameters (for Zero, First and Second Order Reaction)Reaction 1Reaction 2Reaction 3
pH = 2pH = 3pH = 4
zero order reactionk0 (mg L−1 min−1)0.200.100.08
R20.720.720.67
first order reactionk1 (min−1)0.070.040.02
R20.980.980.91
second order reactionk2 (L mg−1 min−1)0.050.030.01
R20.910.9140.99
Table 2. Scientific literature review of MO catalytic degradation over various types of perovskites.
Table 2. Scientific literature review of MO catalytic degradation over various types of perovskites.
Perovskite TypeCatalyst Concentration (ppm)Aqueous Solution Volume (mL)MO Concentration (ppm)pHMO Degradation Percent [%]Dark or IlluminationSource
LaMnO3200503-90% after 60 minVisible light[7]
LaMnO360406298% after 90 minVisible light[22]
LaMnO350010020-~57% after 90 minVisible
light
[51]
LaMnO3500206.52.6~90% in 90 mindark[10]
LaMnO3500206.53.3~87% in 90 mindark[10]
LaMnO3500206.53.6~73% in 90 mindark[10]
LaMnO3500206.53.1~100% in 60 minsolar[10]
LaNiO315001005-94.3% in 4 hdark[52]
La0.5Sr0.5CoO325100202.599.5% in 25 mindark[49]
La0.5Sr0.5CoO325100202.599.7% in 25 minUV radiation[49]
LaMnO3 1204012483% in 90 mindarkThis work
LaMnO31204012494.7% in 90 minSolar radiationThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sfirloaga, P.; Ivanovici, M.-G.; Poienar, M.; Ianasi, C.; Vlazan, P. Investigation of Catalytic and Photocatalytic Degradation of Methyl Orange Using Doped LaMnO3 Compounds. Processes 2022, 10, 2688. https://doi.org/10.3390/pr10122688

AMA Style

Sfirloaga P, Ivanovici M-G, Poienar M, Ianasi C, Vlazan P. Investigation of Catalytic and Photocatalytic Degradation of Methyl Orange Using Doped LaMnO3 Compounds. Processes. 2022; 10(12):2688. https://doi.org/10.3390/pr10122688

Chicago/Turabian Style

Sfirloaga, Paula, Madalina-Gabriela Ivanovici, Maria Poienar, Catalin Ianasi, and Paulina Vlazan. 2022. "Investigation of Catalytic and Photocatalytic Degradation of Methyl Orange Using Doped LaMnO3 Compounds" Processes 10, no. 12: 2688. https://doi.org/10.3390/pr10122688

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