Next Article in Journal
Mechanistic Insights into Photodegradation of Organic Dyes Using Heterostructure Photocatalysts
Next Article in Special Issue
Reactive Grinding Synthesis of LaBO3 (B: Mn, Fe) Perovskite; Properties for Toluene Total Oxidation
Previous Article in Journal
Selective Hydrogenation of 3-Nitrostyrene over a Co-promoted Pt Catalyst Supported on P-containing Activated Charcoal
Previous Article in Special Issue
BaTi0.8B0.2O3 (B = Mn, Fe, Co, Cu) LNT Catalysts: Effect of Partial Ti Substitution on NOx Storage Capacity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Oxidation of NO over LaCo1−xBxO3 (B = Mn, Ni) Perovskites for Nitric Acid Production

1
Department of Chemical Engineering, Norwegian University of Science and Technology (NTNU), Sem Sælands vei 4, NO-7491 Trondheim, Norway
2
SINTEF Industry, Kinetic and Catalysis Group, P.O. Box 4760 Torgarden, NO-7465 Trondheim, Norway
3
YARA Technology Center, Herøya Forskningspark, Bygg 92, Hydrovegen 67, NO-3936 Porsgrunn, Norway
*
Author to whom correspondence should be addressed.
Submission received: 31 March 2019 / Revised: 30 April 2019 / Accepted: 3 May 2019 / Published: 8 May 2019
(This article belongs to the Special Issue Catalysis by Metals on Perovskite-Type Oxides)

Abstract

:
Nitric acid (HNO3) is an important building block in the chemical industry. Industrial production takes place via the Ostwald process, where oxidation of NO to NO2 is one of the three chemical steps. The reaction is carried out as a homogeneous gas phase reaction. Introducing a catalyst for this reaction can lead to significant process intensification. A series of LaCo1−xMnxO3 (x = 0, 0.25, 0.5 and 1) and LaCo1−yNiyO3 (y = 0, 0.25, 0.50, 0.75 and 1) were synthesized by a sol-gel method and characterized using N2 adsorption, ex situ XRD, in situ XRD, SEM and TPR. All samples had low surface areas; between 8 and 12 m2/g. The formation of perovskites was confirmed by XRD. The crystallite size decreased linearly with the degree of substitution of Mn/Ni for partially doped samples. NO oxidation activity was tested using a feed (10% NO and 6% O2) that partly simulated nitric acid plant conditions. Amongst the undoped perovskites, LaCoO3 had the highest activity; with a conversion level of 24.9% at 350 °C; followed by LaNiO3 and LaMnO3. Substitution of LaCoO3 with 25% mol % Ni or Mn was found to be the optimum degree of substitution leading to an enhanced NO oxidation activity. The results showed that perovskites are promising catalysts for NO oxidation at industrial conditions.

Graphical Abstract

1. Introduction

Oxidation of nitric oxide (Equation (1)) is one of the few known third order reactions. The reaction is unusual, as the rate of reaction increases with a decrease in the temperature [1]
2NO + O2 ⇋ NO2     ΔrH298 = −113.8 kJ/mol
NO oxidation is a key reaction in lean NOx abatement technologies and in the Ostwald process for nitric acid production. In Ostwald’s process, NO oxidation is carried out as a non-catalytic process and the forward reaction is favored by the removal of heat and by providing sufficient residence time. Typical gas stream concentrations are 10% NO, 6% O2 and 15% H2O [2]. Using a catalyst for NO oxidation may lead to significant process intensification of the nitric acid plant. In addition to speeding up the oxidation process, it may reduce capital costs and increase heat recovery. Efforts have been made to find a catalyst effective under industrial conditions; but success so far has not been achieved [2]. To date, the process is carried out as a homogenous process in modern nitric acid plants. To the best of the authors knowledge, apart from an earlier patent [3], only two recent studies [4,5] report catalytic oxidation of NO at nitric acid plant conditions. However, catalytic oxidation of NO has been extensively studied with regards to lean NOx abatement technologies and reviewed by Russel and Epling [6] and Hong et al. [7]. In these studies, oxidation of NO is carried out at very lean concentrations of NO ranging from 100–1000 ppm of NO [8,9,10]. The huge difference in NO concentration between NOx abatement and nitric acid production makes the extrapolation of these findings to nitric acid plant conditions, questionable.
Although it has been demonstrated that platinum has significant catalytic activity for oxidation of NO to NO2 at nitric acid plant conditions [4,5] the high-cost and scarcity of platinum motivates the search for potential non-noble metal based catalysts.
Perovskites have gained particular interest as catalytic materials due to their high thermal stability, ease of synthesis and good catalytic activity [11]. Perovskites are represented by a general formula ABO3, where A represents a rare earth or alkaline earth cation and B represents a transition metal cation. The activity of perovskites can be tuned by partial substitution of A and/or B site cations with another element to obtain the desired properties [11]. The lanthanum-based perovskite LaBO3 (B = Co, Mn, Ni) have been found to be active for NO oxidation at lean NO conditions [12,13]. It has been demonstrated for LaCoO3, that partial doping of the A-site with strontium or cerium or doping of the B-site with manganese or nickel enhances NO oxidation activity of the perovskite [13,14,15,16].
In this work, lanthanum-based perovskites with cobalt, nickel and manganese (LaCoO3, LaNiO3 and LaMnO3) were synthesized by the sol-gel method using citric acid. Catalytic tests were performed using a dry feed (10% NO, 6% O2) at atmospheric pressure; partially simulating nitric acid plant conditions. The effect of B-site substitution was studied by preparing a series of LaCo1−xMnxO3 and LaCo1−yNiyO3 catalysts. The catalysts were characterized using N2 adsorption, X-ray diffraction (XRD), Scanning electron microscopy (SEM) and temperature programmed reduction (TPR). Catalyst structure during pretreatment and oxidation of NO was monitored using in situ XRD.

2. Results and Discussion

2.1. Catalyst Characterisation

Specific surface areas are summarized in Table 1. All samples have a relatively low surface area, in the range of 8–12 m2/g, which is a typical of perovskites prepared by the citrate method [14,17,18]. The surface area remains unaffected by the degree of substitution of Mn (x). However, an irregular change in surface area is observed with the degree of substitution of Ni (y). For y = 0.25 and 0.75, the surface area increased by 25% (from 8 to 10 m2/g) and 50% (from 8 to 12 m2/g) respectively. However, for y = 0.50 the surface area remains the same as for LaCoO3.
The XRD patterns shown in Figure 1 reveal the formation of phase-pure perovskite structure for all samples, and no peaks characteristic of the metallic oxides or carbonates were seen. The main characteristic peak at 2θ = 33° for LaCoO3 slightly shifts towards lower 2θ values with an increase in x and y, indicating an increase in the lattice parameters confirming previous findings [13,19,20]. This expansion can be explained by comparing the ionic radii of the species in the perovskite structure. The higher ionic radii of Ni3+ (0.56 Å) than Co3+ (0.52 Å) responsible for lattice expansion in LaCo1−yNiyO3 [20]. In case of LaCo1−xMnxO3, the average trivalent metal site is conserved by adjusting the ratio between Mn4+/Mn3+ and Co3+/Co2+ [21]. Therefore, the relative amount of Mn4+ (0.52 Å) present, Mn3+ (0.645 Å), Co2+ (0.82 Å), Co3+ (0.52 Å) dictates the overall lattice parameters.
Among undoped perovskites, LaCoO3 and LaMnO3 belong to the rhombohedral phase in agreement with the results reported in literature [22,23]. LaNiO3 belongs to the cubic phase, which is consistent with previous studies [24].
Crystallite sizes calculated using the Scherrer equation are summarized in Table 1. Figure 2 shows crystallite size as a function of the degree of substitution. The highest crystallite size of 28 nm was observed for LaCoO3 and a linear decrease was observed with an increase in x and y for partially substituted samples indicating that partial substitution effectively restrains the crystal growth. However, other contributions to XRD peak broadening such as strain cannot be ruled out. Partially substituted nickel perovskites had smaller crystallite size compared to their manganese counterparts.
Figure 3 shows the XRD pattern recorded in situ during pretreatment of LaCoO3 and LaMnO3. No change in structure is observed apart from lattice expansion with an increase in temperature. The structural stability of the perovskites is in accordance with the fact that they are calcined at higher temperatures in comparison with the pretreatment temperature of 500 °C. No change in structure was observed for LaCoO3 and LaMnO3 during steady state oxidation of NO at 350 °C indicating that the bulk structure of perovskite remains unaffected during the catalytic process. However, minor changes beyond the detectable range of XRD cannot be ruled out.
Figure 4 shows the SEM images of LaCoO3 with varying content of Mn and pure LaNiO3. The presence of agglomerated non-spherical particles is observed for all samples. A significant change in morphology is observed with the substitution of Co with Mn along with an increase in the extent of agglomeration (Figure 4b–d).
The H2-TPR profiles of LaCo1−xMnxO3 perovskites are given in Figure 5a. Three reduction peaks at 334, 373 and 526 °C are observed in the TPR profile of LaCoO3. The first two peaks are attributed to the reduction of Co3+ to Co2+, while the peak at 526 °C represents the reduction of Co2+ to Co0 leading to the destruction of the perovskite structure [25]. TPR of LaMnO3 shows two main reduction peaks at 383 and 818 °C. The first peak represents the reduction of Mn4+ to Mn3+, while the reduction of Mn3+ to Mn2+ occurs at elevated temperatures (above 700 °C), forming MnO and simultaneous collapse of the perovskite structure [26]. For x = 0.25 and 0.5, broad peaks overlapping reduction peaks of Co3+ to Co2+ and Mn4+ to Mn3+ are observed below 550 °C.
Figure 5b shows H2-TPR profiles of LaCo1−yNiyO3 perovskites. Reduction of LaNiO3 also proceeds via three peaks at 304 °C, one at 334 °C and one at 465 °C. A similar three-step reduction process for LaNiO3 has been reported [27] and in situ XRD revealed that reduction proceeds via formation of the La2NiO4 phase [27]. The highest peak is associated with the formation of Ni0 and La2O3 resulting in the destruction of the perovskite structure. Four reduction peaks were observed for y = 0.25. Comparison with LaCoO3 and LaNiO3 indicates that the first two peaks at 303 and 342 °C match with the reduction peaks of Ni3+ to Ni2+ and Co3+ to Co2+. Whereas the former two peaks at 459, 495 °C corresponds to further reduction to form the metallic phases (Ni0 and Co0). This reduction profile matches well with the previous findings [20]. In contrast to distinct reduction peaks for cobalt and nickel for LaCo0.75Ni0.25O3, only two broad peaks were observed for y = 0.50 and 0.75. The first peak overlaps with the reduction of Ni3+ to Ni2+ and Co3+ to Co2+ and increases from 313 °C for y = 0.50 to 346 °C for y = 0.75. Simultaneous reduction to metallic phases (Ni0 and Co0) was observed at 418 °C and 488 °C for y = 0.50 and 0.75, respectively.

2.2. NO Oxidation Activity

The oxidation of NO occurs as a homogeneous gas-phase reaction with a second order dependency in NO concentration. For this reason, contributions from gas phase conversion have been detected for studies performed at lean NO concentrations [28,29,30]. This necessitates the quantification of the gas phase contribution to NO oxidation in the current study, which involves such a high concentration of NO (10%). The blank run performed without catalyst is given in our previous publication [5]. The gas phase conversion gradually decreases with increasing temperature in the studied temperature range (150–450 °C); with a conversion of 6.1% at 350 °C.
The conversion of NO over LaCo1−xMnxO3 as a function of temperature is shown in Figure 6. The NO conversion and rate of reaction are summarized in Table 2. At low temperature, only gas phase conversion is observed for all perovskites. The catalytic activity for LaCoO3 starts at about 270 °C and increases gradually until it becomes limited by the thermodynamic equilibrium. This is in contrast to studies performed at lean NOx conditions, where the catalytic activity starts at significantly lower temperatures (150–200 °C); increases with a steeper slope, and becomes thermodynamically limited at ca 300 °C [14,15]. The catalytic conversion starts at a lower temperature (240 °C) for x = 0.25 in comparison to other perovskites and a significant increase in conversion was observed. A further increase in x to 0.5 leads to a substantial decrease in NO conversion to conversion levels even lower than what is observed for LaMnO3. Manganese is stable in valence states +III and +IV while cobalt is stable in valence states +II and +III. Ghiasi et al. [21] used X-ray absorption spectroscopy (XAS) to study the valence state of Mn and Co in a series of LaCo1−xMnxO3 perovskites and found that the average trivalent metal site is conserved by shifting the balance between Mn4+/Mn3+ in combination with Co3+/Co2+. The ratio between Mn4+/Mn3+ and Co3+/Co2+ decreases with a sequential increase in manganese content, the highest value being observed for LaCo0,75Mn0,25O3. The best activity of LaCo0.75Mn0.25O3 in the series of LaCo1−xMnxO3 perovskites may be attributed to the presence of highest content of Mn4+ as amongst different valence states of manganese, Mn4+ (MnO2) exhibits the highest activity for NO oxidation followed by Mn3+ (Mn2O3) and Mn2+ (Mn3O4) [31].
Figure 7 shows the conversion of NO over LaCo1−yNiyO3 as a function of temperature. Minor difference in gas phase conversion is observed at lower temperatures due to the differences in the packing of the catalyst bed. The substitution with 25 mol% Ni in LaCoO3 had a significant positive impact on the activity exhibiting a conversion of 30% at 350 °C. With a further increase in y, the activity gradually decreased until 75% nickel content. Similar conversion curves are exhibited by y = 0.5 and y = 1 perovskites. Although our results differ from Zhong et al. [13], who reported 70 mol% nickel substitution in LaCoO3 to yield the best results for NO oxidation, it can be argued that they used a co-precipitation method for preparation of perovskites and the activity was tested at substantially different feed concentration (400 ppm NO and 6% O2) compared to the present study. The increase in activity for LaCo0.75Ni0.25O3 can in part be attributed to the 25% increase in surface area. However, for LaCo0.25Ni0.75O3 the surface area increased with 50% compared to LaCoO3 and 33% compared to LaNiO3, while the catalytic activity decreased. Thus, it seems likely that the change in catalytic activity is more related to the chemical oxygen dynamics and to the redox properties than to changes in the surface area in this range. Ivanova et al. [32] reported a maximum in defect structures for x = 0.25 in a series of LaCo1−xNixO3, detected by EPR due to the presence of magnetic Ni clusters. However, this depends on the method of preparation and pre-treatment procedure. The highest activity of LaCo0.75Ni0.25O3 may thus be related to lattice defects not detectable by bulk XRD. The NO oxidation activity follows the order LaCoO3 > LaNiO3 > LaMnO3 for the undoped perovskites.
Figure 8 shows the rate of reaction as a function of crystallite size for LaCo1−xMnxO3 and LaCo1−yNixO3 perovskites. The rate of reaction increases linearly with crystallite size with x = 0.25 and y = 0.25 samples being the exception and not included in the linear fit. It should be kept in mind that the crystallite sizes were calculated using the Scherrer equation assuming spherical geometry. Though, the crystallites are not spherical as revealed through SEM images. The crystallite size estimates from the Scherrer equation might not be 100% accurate but they provide a fair comparison.
Partial substitution of LaCoO3 with either manganese or nickel leads to a modification in the redox properties, morphology, structure, crystallite size and valence state of the metal site. Therefore, it is difficult to pinpoint one governing factor determining the catalytic activity. In fact, it is a combination of several factors which contribute to dictating catalytic activity.

3. Materials and Methods

3.1. Catalyst Preparation

All catalysts were prepared using the citrate method [14]. Nitrate salts were used as starting materials: La(NO3)3·6H2O (Alfa Aesar, 99.9%, Kandel, Germany), Mn(NO3)2·4H2O (Alfa Aesar, 99.98%, Kandel, Germany), Co(NO3)3·6H2O (Acros Organics, 99%, Geel, Belgium) and Ni(NO3)2·6H2O (Acros Organics, 99%, Geel, Belgium). An appropriate amount of nitrate salts of the desired A and B sites were dissolved in deionized water with 10 wt.% excess citric acid (Sigma Aldrich, 99.5%, Munich, Germany). The solution was stirred for 1 h at room temperature and further stirred in an oil bath at 80 °C until a viscous gel was obtained. The gel was dried in static air overnight at 90 °C and then heated to 150 °C for 1 hr. The black, spongy material was crushed and calcined at 700 °C for 5 h in static air.

3.2. Catalyst Characterisation

N2 adsorption was used to measure the specific surface area. The sample (200 mg) was degassed at 200 °C overnight in VacPrep 061 Degasser (Norcross, GA, USA). Nitrogen adsorption was performed with a Miromeritics TriStar II 3020 Surface Area and Porosity Analyzer (Norcross, GA, USA) at −196 °C. The specific surface areas were calculated using the BET desorption branch of the isotherm [33].
Ex situ XRD patterns of as prepared samples were obtained with a Bruker D8 Advanced X-ray Diffractometer (Cambridge, United Kingdom) with a copper anode with an X-ray wavelength of 1.5418 Å. The measured angles (2θ) were scanned from 10° to 75° in 30 min; at a fixed divergence angle of 0.2°. The PDF database was used for phase identification. The average crystallite sizes were calculated according to the Scherrer equation [34] assuming spherical crystals (K = 0.9) using the diffraction peak between 2θ = 45–50°.
In situ powder XRD experiments were carried out at BM31 of the Swiss-Norwegian beam lines (SNBL) at the European Synchrotron Radiation Facility (ESRF). Catalyst sample (30mg, sieve fraction 53–90 µm) was fixed between two quartz wool plugs in a quartz capillary of 1 mm internal diameter (bed length: 10 mm). The capillary was then mounted in a custom cell [35] and exposed to X-rays for diffraction measurements. The temperature of the capillary reactor was controlled by a calibrated hot air blower. Powder X-ray diffraction patterns were collected with a 2D plate detector (Mar-345) using monochromatic radiation of wavelength 0.49324Å. Note that is different from the Cu Kα wavelength, which is used for acquiring ex situ XRD patterns. The instrumental peak broadening, wavelength calibration, and detector distance corrections were performed using a NIST 660a LaB6 standard. Mass flow controllers were used to feed NO, O2 and He in to produce the desired gas-feed composition (15 Ncm3/min; 1% NO, 6% O2, He balance). XRD patterns were recorded in situ during the pretreatment step (flowing 15 Ncm3/min of 6% O2/He by heating at 10 °C/min from ambient to 500 °C and holding for 1 hr) and steady state NO oxidation (1%NO, 6% O2 and balance He) at 350 °C for 2 h. A lower concentration of NO (1%) was used for the in situ studies due to experimental constraints at the beamline.
The morphological analysis of as-synthesized perovskites was performed by using an in-lens cold field emission electron microscope FE-S(T)EM, (Hitachi S-5500) in scanning electron microscopy (SEM) mode.
Temperature programmed reduction by H2 was performed with an Altamira BenchCAT Hybrid 1000 HP (Pittsburgh, PA, USA). The samples (100 mg) were pretreated in 50 Ncm3/min flow of Ar at 150 °C for 30 min, with a heating rate of 10 °C. TPR was conducted by heating at a rate of 5 °C/min from 50 to 850 °C with a 50 Ncm3/min flow of 10% H2/Ar.

3.3. Catalyst Activity Testing

Details of the reactor and experimental setup can be found in our previous publication [5]. Briefly, the activity of the catalysts was measured in a vertical stainless steel tubular reactor (internal diameter = 9.7 mm) operated at atmospheric pressure. The temperature was measured and controlled by a thermocouple, protected by a stainless steel jacket, inserted into the catalyst bed. A MKS MultiGas 2030-HS FTIR Gas Analyzer (path length 5.11 m, Cheshire, United Kingdom), calibrated at 1 bar and 191°C was used to analyze the composition of NO and NO2 in the product stream.
For activity measurements, 0.5 g of catalyst diluted with 2.75 g SiC was loaded into the reactor and held in place by quartz wool plugs. Prior to the activity tests, the catalysts were pretreated at 500 °C for 1 h in 200 Ncm3/min flow of 10% O2/Ar and subsequently cooled down in inert argon atmosphere. The activity of the catalysts was investigated by heating from 150 to 450 °C at a rate of 5 °C/min under a flow of 200 Ncm3/min of feed gas (10% NO, 6% O2 in balance Ar).
Conversion of NO to NO2 was calculated by the following equation:
NO conversion = α × [NO2]outlet/[NO]inlet
where [NO]inlet and [NO2]outlet are concentrations of NO at inlet and NO2 at the outlet of the reactor. Volume changes arising from the reaction is taken into account by the constant “α” [36] where α = 0.99. The closure of nitrogen balance across the reactor (99.5–100%) confirmed that all nitrogen is present as NO and NO2. Comparison of catalyst activity is performed at 350 °C where the reaction is in the kinetic regime, away from equilibrium. The reaction rate (rNO) calculations were performed by subtracting the homogeneous gas phase conversion of NO; hence, reflecting only the activity provided by the catalyst.

4. Conclusions

A series of lanthanum-based perovskites have been investigated for oxidation of NO using a feed containing 10% NO and 6% O2, thus partially simulating nitric acid plant conditions. Among the undoped perovskites, the NO oxidation activity follows the order LaCoO3 > LaNiO3 > LaMnO3. A significant increase in NO oxidation activity was achieved by partial substitution of cobalt in LaCoO3 with 25 mol% of either nickel or manganese. Further increase in the degree of Co substitution had a negative impact on activity.
From this work, perovskites are shown to be promising catalysts for oxidizing NO to NO2 at conditions representative of nitric acid plant operation. Low cost, ease of production and significant catalytic activity make perovskites attractive candidates as alternatives to noble metal catalysts.

Author Contributions

Conceptualization, A.u.R.S., D.W., B.C.E., R.L. and M.R.; methodology, A.S., S.M.H.; formal analysis, A.u.R.S., S.M.H., S.K.R. and M.Z.; investigation, A.u.R.S., S.M.H., S.K.R. and M.Z.; writing—original draft preparation, A.u.R.S.; writing—review and editing, S.M.H., S.K.R, M.Z., B.C.E., R.L, D.W. and M.R.; visualization, A.u.R.S.; supervision, B.C.E., R.L., D.W. and M.R.; project administration, M.R.; funding acquisition, B.C.E., R.L., D.W. and M.R.

Funding

This research was funded by iCSI (industrial Catalysis Science and Innovation) Centre for Research-based Innovation, which receives financial support from the Research Council of Norway, grant number 237922. The Research Council of Norway is also acknowledged for financial support to the Swiss-Norwegian Beamlines at ESRF, grant number 273608.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baulch, D.; Drysdale, D.; Horne, D. Homogeneous Gas Phase Reactions of the H2-N2-O2 System; CRC Press: Boca Raton, FL, USA, 1973; pp. 285–300. [Google Scholar]
  2. Honti, G. The Nitrogen Industry; Akademiai Kiado: Budapest, Hungary, 1976; pp. 400–413. [Google Scholar]
  3. Klingelhoefer, W.C. Nitric Oxide Oxidation. US2115173 A, 26 April 1938. [Google Scholar]
  4. Grande, C.A.; Andreassen, K.A.; Cavka, J.H.; Waller, D.; Lorentsen, O.-A.; Øien, H.; Zander, H.-J.; Poulston, S.; García, S.; Modeshia, D. Process intensification in nitric acid plants by catalytic oxidation of nitric oxide. Ind. Eng. Chem. Res. 2018, 57, 10180–10186. [Google Scholar] [CrossRef]
  5. Salman, A.u.R.; Enger, B.C.; Auvray, X.; Lødeng, R.; Menon, M.; Waller, D.; Rønning, M. Catalytic oxidation of NO to NO2 for nitric acid production over a Pt/Al2O3 catalyst. Appl. Catal. A Gen. 2018, 564, 142–146. [Google Scholar] [CrossRef]
  6. Russell, A.; Epling, W.S. Diesel oxidation catalysts. Catal. Rev. 2011, 53, 337–423. [Google Scholar] [CrossRef]
  7. Hong, Z.; Wang, Z.; Li, X. Catalytic oxidation of nitric oxide (NO) over different catalysts: An overview. Catal. Sci. Technol. 2017, 7, 3440–3452. [Google Scholar] [CrossRef]
  8. Auvray, X.; Olsson, L. Stability and activity of Pd-, Pt- and Pd–Pt catalysts supported on alumina for NO oxidation. Appl. Catal. B Environ. 2015, 168–169, 342–352. [Google Scholar] [CrossRef]
  9. Auvray, X.; Pingel, T.; Olsson, E.; Olsson, L. The effect gas composition during thermal aging on the dispersion and NO oxidation activity over Pt/Al2O3 catalysts. Appl. Catal. B Environ. 2013, 129, 517–527. [Google Scholar] [CrossRef]
  10. Després, J.; Elsener, M.; Koebel, M.; Kröcher, O.; Schnyder, B.; Wokaun, A. Catalytic oxidation of nitrogen monoxide over Pt/SiO2. Appl. Catal. B Environ. 2004, 50, 73–82. [Google Scholar] [CrossRef]
  11. Zhu, J.; Thomas, A. Perovskite-type mixed oxides as catalytic material for NO removal. Appl. Catal. B Environ. 2009, 92, 225–233. [Google Scholar] [CrossRef]
  12. 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]
  13. Zhong, S.; Sun, Y.; Xin, H.; Yang, C.; Chen, L.; Li, X. NO oxidation over Ni–Co perovskite catalysts. Chem. Eng. J. 2015, 275, 351–356. [Google Scholar] [CrossRef] [Green Version]
  14. Kim, C.H.; Qi, G.; Dahlberg, K.; Li, W. Strontium-doped perovskites rival platinum catalysts for treating NOx in simulated diesel exhaust. Science 2010, 327, 1624–1627. [Google Scholar] [CrossRef]
  15. Wen, Y.; Zhang, C.; He, H.; Yu, Y.; Teraoka, Y. Catalytic oxidation of nitrogen monoxide over La1−xCexCoO3 perovskites. Catal. Today 2007, 126, 400–405. [Google Scholar] [CrossRef]
  16. Wang, J.; Su, Y.; Wang, X.; Chen, J.; Zhao, Z.; Shen, M. The effect of partial substitution of Co in LaMnO3 synthesized by sol–gel methods for NO oxidation. Catal. Commun. 2012, 25, 106–109. [Google Scholar] [CrossRef]
  17. Taguchi, H.; Matsu-ura, S.-i.; Nagao, M.; Choso, T.; Tabata, K. Synthesis of LaMnO3+δ by firing gels using citric acid. J. Solid State Chem. 1997, 129, 60–65. [Google Scholar] [CrossRef]
  18. Taguchi, H.; Yamada, S.; Nagao, M.; Ichikawa, Y.; Tabata, K. Surface characterization of LaCoO3 synthesized using citric acid. Mater. Res. Bull. 2002, 37, 69–76. [Google Scholar] [CrossRef]
  19. Alifanti, M.; Auer, R.; Kirchnerova, J.; Thyrion, F.; Grange, P.; Delmon, B. Activity in methane combustion and sensitivity to sulfur poisoning of La1−xCexMn1−yCoyO3 perovskite oxides. Appl. Catal. B Environ. 2003, 41, 71–81. [Google Scholar] [CrossRef]
  20. Silva, C.R.B.; da Conceição, L.; Ribeiro, N.F.P.; Souza, M.M.V.M. Partial oxidation of methane over Ni–Co perovskite catalysts. Catal. Commun. 2011, 12, 665–668. [Google Scholar] [CrossRef]
  21. Ghiasi, M.; Delgado-Jaime, M.U.; Malekzadeh, A.; Wang, R.-P.; Miedema, P.S.; Beye, M.; de Groot, F.M.F. Mn and Co charge and spin evolutions in LaMn1–xCoxO3 nanoparticles. J. Phys. Chem. C 2016, 120, 8167–8174. [Google Scholar] [CrossRef]
  22. Chen, J.; Shen, M.; Wang, X.; Qi, G.; Wang, J.; Li, W. The influence of nonstoichiometry on LaMnO3 perovskite for catalytic NO oxidation. Appl. Catal. B Environ. 2013, 134–135, 251–257. [Google Scholar] [CrossRef]
  23. Pecchi, G.; Campos, C.; Peña, O.; Cadus, L.E. Structural, magnetic and catalytic properties of perovskite-type mixed oxides LaMn1−yCoyO3 (y = 0.0, 0.1, 0.3, 0.5, 0.7, 0.9, 1.0). J. Mol. Catal. A Chem. 2008, 282, 158–166. [Google Scholar] [CrossRef]
  24. Rakshit, S.; Gopalakrishnan, P.S. Oxygen nonstoichiometry and its effect on the structure of LaNiO3. J. Solid State Chem. 1994, 110, 28–31. [Google Scholar] [CrossRef]
  25. Zhou, C.; Feng, Z.; Zhang, Y.; Hu, L.; Chen, R.; Shan, B.; Yin, H.; Wang, W.G.; Huang, A. Enhanced catalytic activity for NO oxidation over Ba doped LaCoO3 catalyst. RSC Adv. 2015, 5, 28054–28059. [Google Scholar] [CrossRef]
  26. Qi, G.; Li, W. Pt-free, LaMnO3 based lean NOx trap catalysts. Catal. Today 2012, 184, 72–77. [Google Scholar] [CrossRef]
  27. Batiot-Dupeyrat, C.; Valderrama, G.; Meneses, A.; Martinez, F.; Barrault, J.; Tatibouët, J.M. Pulse study of Co2 reforming of methane over LaNiO3. Appl. Catal. A Gen. 2003, 248, 143–151. [Google Scholar] [CrossRef]
  28. Benard, S.; Retailleau, L.; Gaillard, F.; Vernoux, P.; Giroir-Fendler, A. Supported platinum catalysts for nitrogen oxide sensors. Appl. Catal. B Environ. 2005, 55, 11–21. [Google Scholar] [CrossRef]
  29. Mulla, S.S.; Chen, N.; Cumaranatunge, L.; Blau, G.E.; Zemlyanov, D.Y.; Delgass, W.N.; Epling, W.S.; Ribeiro, F.H. Reaction of NO and O2 to NO2 on Pt: Kinetics and catalyst deactivation. J. Catal. 2006, 241, 389–399. [Google Scholar] [CrossRef]
  30. Schmitz, P.J.; Kudla, R.J.; Drews, A.R.; Chen, A.E.; Lowe-Ma, C.K.; McCabe, R.W.; Schneider, W.F.; Goralski, C.T. NO oxidation over supported Pt: Impact of precursor, support, loading, and processing conditions evaluated via high throughput experimentation. Appl. Catal. B Environ. 2006, 67, 246–256. [Google Scholar] [CrossRef]
  31. Wang, H.; Chen, H.; Wang, Y.; Lyu, Y.-K. Performance and mechanism comparison of manganese oxides at different valence states for catalytic oxidation of NO. Chem. Eng. J. 2019, 361, 1161–1172. [Google Scholar] [CrossRef]
  32. Ivanova, S.; Senyshyn, A.; Zhecheva, E.; Tenchev, K.; Stoyanova, R.; Fuess, H. Crystal structure, microstructure and reducibility of LaNixCo1−xO3 and LaFexCo1−xO3 perovskites (0 <x ≤ 0.5). J. Solid State Chem. 2010, 183, 940–950. [Google Scholar]
  33. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. JACS 1938, 60, 309–319. [Google Scholar] [CrossRef]
  34. Scherrer, P. Bestimmung der größe und der inneren struktur von kolloidteilchen mittels röntgenstrahlen. Nachr. Ges. Wiss. Gött. Math. Phys. Kl. 1918, 1918, 98–100. [Google Scholar]
  35. Tsakoumis, N.E.; Voronov, A.; Rønning, M.; Beek, W.v.; Borg, Ø.; Rytter, E.; Holmen, A. Fischer–tropsch synthesis: An XAS/XRPD combined in situ study from catalyst activation to deactivation. J. Catal. 2012, 291, 138–148. [Google Scholar] [CrossRef]
  36. Fogler, H. Elements of Chemical Reaction Engineering; Prentice-Hall of India: New Delhi, India, 2006; p. 92. [Google Scholar]
Figure 1. XRD patterns of: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Figure 1. XRD patterns of: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Catalysts 09 00429 g001
Figure 2. Crystallite size as a function of the degree of substitution for LaCo1−xMnxO3 and LaCo1−yNiyO3 perovskites.
Figure 2. Crystallite size as a function of the degree of substitution for LaCo1−xMnxO3 and LaCo1−yNiyO3 perovskites.
Catalysts 09 00429 g002
Figure 3. XRD patterns (λ = 0.49324 Å) recorded in situ during pretreatment of: (a) LaCoO3; (b) LaMnO3 perovskites.
Figure 3. XRD patterns (λ = 0.49324 Å) recorded in situ during pretreatment of: (a) LaCoO3; (b) LaMnO3 perovskites.
Catalysts 09 00429 g003
Figure 4. SEM images: (a) LaCoO3; (b) LaCo0.75Mn0.25O3; (c) LaCo0.50Mn0.50O3; (d) LaMnO3; (e) LaNiO3.
Figure 4. SEM images: (a) LaCoO3; (b) LaCo0.75Mn0.25O3; (c) LaCo0.50Mn0.50O3; (d) LaMnO3; (e) LaNiO3.
Catalysts 09 00429 g004
Figure 5. TPR profiles of: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Figure 5. TPR profiles of: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Catalysts 09 00429 g005
Figure 6. Conversion of NO over LaCo1−xMnxO3 as a function of temperature.
Figure 6. Conversion of NO over LaCo1−xMnxO3 as a function of temperature.
Catalysts 09 00429 g006
Figure 7. Conversion of NO over LaCo1−yNiyO3 as a function of temperate.
Figure 7. Conversion of NO over LaCo1−yNiyO3 as a function of temperate.
Catalysts 09 00429 g007
Figure 8. Rate of reaction as a function of crystallite size: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Figure 8. Rate of reaction as a function of crystallite size: (a) LaCo1−xMnxO3; (b) LaCo1−yNiyO3 perovskites.
Catalysts 09 00429 g008
Table 1. Brunauer–Emmett–Teller (BET) surface area and crystallite size (d) calculated using the Scherrer equation.
Table 1. Brunauer–Emmett–Teller (BET) surface area and crystallite size (d) calculated using the Scherrer equation.
SampleSurface Area
(m2/g)
d
(nm)
LaCoO3828
LaCo0.75Mn0.25O3824
LaCo0.50Mn0.50O3718
LaMnO3921
LaCo0.75Ni0.25O31022
LaCo0.50Ni0.50O3817
LaCo0.25Ni0.75O3128
LaNiO3813
Table 2. NO oxidation activity of perovskites at 350 °C.
Table 2. NO oxidation activity of perovskites at 350 °C.
SampleXNO
(%)
rNO 1
(µmolgcat−1s−1)
LaCoO324.95.14
LaCo0.75Mn0.25O329.76.55
LaCo0.50Mn0.50O315.92.72
LaMnO318.03.27
LaCo0.75Ni0.25O330.36.70
LaCo0.50Ni0.50O323.24.74
LaCo0.25Ni0.75O316.42.86
LaNiO320.84.05
1 Catalytic activity obtained by subtracting the gas phase conversion of NO.

Share and Cite

MDPI and ACS Style

Salman, A.u.R.; Hyrve, S.M.; Regli, S.K.; Zubair, M.; Enger, B.C.; Lødeng, R.; Waller, D.; Rønning, M. Catalytic Oxidation of NO over LaCo1−xBxO3 (B = Mn, Ni) Perovskites for Nitric Acid Production. Catalysts 2019, 9, 429. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9050429

AMA Style

Salman AuR, Hyrve SM, Regli SK, Zubair M, Enger BC, Lødeng R, Waller D, Rønning M. Catalytic Oxidation of NO over LaCo1−xBxO3 (B = Mn, Ni) Perovskites for Nitric Acid Production. Catalysts. 2019; 9(5):429. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9050429

Chicago/Turabian Style

Salman, Ata ul Rauf, Signe Marit Hyrve, Samuel Konrad Regli, Muhammad Zubair, Bjørn Christian Enger, Rune Lødeng, David Waller, and Magnus Rønning. 2019. "Catalytic Oxidation of NO over LaCo1−xBxO3 (B = Mn, Ni) Perovskites for Nitric Acid Production" Catalysts 9, no. 5: 429. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9050429

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