Heterogeneously Catalyzed Aerobic Oxidation of ... - ACS Publications

Jun 27, 2018 - through the reductive activation of O2.39,40 Here, we focus ... catalyzed by BaRuO3 at 333 K was examined using O2 (0.1. MPa) as the so...
0 downloads 0 Views 3MB Size
Research Article www.acsami.org

Cite This: ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Heterogeneously Catalyzed Aerobic Oxidation of Sulfides with a BaRuO3 Nanoperovskite Keigo Kamata,†,‡ Kosei Sugahara,† Yuuki Kato,† Satoshi Muratsugu,‡,§ Yu Kumagai,‡,⊥ Fumiyasu Oba,† and Michikazu Hara†,#,*

Downloaded via TU MUENCHEN on July 9, 2018 at 07:08:55 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Laboratory for Materials and Structures, Institute of Innovative Research, Tokyo Institute of Technology, Nagatsuta-cho 4259, Midori-ku, Yokohama 226-8503, Japan Japan Science and Technology Agency (JST), ‡Precursory Research for Embryonic Science and Technology (PRESTO), and # Advanced Low Carbon Technology Research and Development Program (ALCA), 4-1-8 Honcho, Kawaguchi 332-0012, Japan § Department of Chemistry, Graduate School of Science, Nagoya University, Furo-cho, Chikusa, Nagoya, Aichi 464-8602, Japan ⊥ Materials Research Center for Element Strategy, Tokyo Institute of Technology, Yokohama 226-8503, Japan S Supporting Information *

ABSTRACT: A rhombohedral BaRuO3 nanoperovskite, which was synthesized by the sol−gel method using malic acid, could act as an efficient heterogeneous catalyst for the selective oxidation of various aromatic and aliphatic sulfides with molecular oxygen as the sole oxidant. BaRuO3 showed much higher catalytic activities than other catalysts, including ruthenium-based perovskite oxides, under mild reaction conditions. The catalyst could be recovered by simple filtration and reused several times without obvious loss of its high catalytic performance. The catalyst effect, 18O-labeling experiments, and kinetic and mechanistic studies showed that substrate oxidation proceeds with oxygen species caused by the solid. The crystal structure of ruthenium-based oxides is crucial to control the nature of the oxygen atoms and significantly affects their oxygen transfer reactivity. Density functional theory calculations revealed that the face-sharing octahedra in BaRuO3 likely are possible active sites in the present oxidation in sharp contrast to the corner-sharing octahedra in SrRuO3, CaRuO3, and RuO2. The superior oxygen transfer ability of BaRuO3 is also applicable to the quantitative conversion of dibenzothiophene into the corresponding sulfone and gram-scale oxidation of 4-methoxy thioanisole, in which 1.20 g (71% yield) of the analytically pure sulfoxide could be isolated. KEYWORDS: perovskite, ruthenium, selective oxidation, sulfide, molecular oxygen, oxygen transfer



INTRODUCTION Since the oxygenated products are useful as platform commodities and specialty chemicals, liquid-phase selective oxidation of organic substrates is an important reaction.1−4 Catalytic oxidation with molecular oxygen (O2) as the sole oxidant has received much attention because there is no requirement for stoichiometric amounts of traditional toxic organic and inorganic reagents, high contents of active oxygen, and no formation of toxic byproducts. Although various catalytic systems that are effective for aerobic selective oxidation have been reported, the scope of these catalysts still encounters many challenges.5−7 For example, the involvement of autoxidation reactions always suffers from low selectivity to oxygenated products, which leads to limited applicability (e.g., industrial production of terephthalic acid and ketone-alcohol (KA) oil (cyclohexanone and cyclohexanol)). Therefore, the design and development of new effective catalysts for selective oxidation using only O2 is required to overcome the drawbacks of conventional methods. © XXXX American Chemical Society

Sulfoxides and sulfones are useful organosulfur compounds as synthetic intermediates for natural products and biologically important molecules,8 ligands used for asymmetric catalysis,9,10 and oxygen atom donors.11 In addition, oxidative desulfurization is a promising method for the removal of some thiophenic sulfur compounds that are not effectively removed from fuels by hydrodesulfurization.12 There are many catalytic systems for the selective oxidation of sulfides using activated oxidants such as organic hydroperoxides and hydrogen peroxide (H2O2)13−19 or O2 in combination with reductants20−23 and photoirradiation.24−26 In particular, high activity (tens of thousands of turnover numbers (TONs) in some cases), excellent chemoand enantioselectivity, and broad substrate scope can be achieved for the sulfoxidation with tert-butyl hydroperoxide and H2O2 in the presence of transition metal catalysts such as Received: April 10, 2018 Accepted: June 27, 2018

A

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces Ti, V, Mn, Fe, W, and Re.13−19 On the other hand, examples of catalytic oxidation of inactive aryl sulfides with O2 as the sole oxidant without the need for additives are still limited.20−33 Ce- and Pd-based homogeneous systems promote the sulfoxidation under mild reaction conditions; however, the catalyst/product separation and reuse of the catalysts are very difficult. In addition, heterogeneous systems have several drawbacks, such as low applicability, activity, and selectivity, and requirements of high reaction temperatures, additives, or specific solvents to achieve high catalytic performance (see Table S1).29−33 In this context, the development of effective solid catalysts for aerobic oxidation of a variety of sulfides including inactive aryl sulfides under mild conditions remains a challenging subject of research. Perovskite oxides, having the general formula of ABO3, are one of the most promising piezoelectric, (multi)ferroelectric, magnetic, superconducting, and catalytic materials because of their structural simplicity and flexibility, good stability, and controllable physicochemical properties.34−38 The ideal perovskite has a cubic structure and can be described as the A cation occupying the 12-fold coordination site between cornersharing BO6 octahedra (Figure 1b). On the other hand, a

system has the following advantages: (1) high catalytic performance, even at low temperature, (2) wide substrate scope including application to the oxidative desulfurization of dibenzothiophene, and (3) recyclability. In this paper, we report the synthesis and aerobic sulfoxidation catalysis of BaRuO3 and investigate the kinetic and mechanistic aspects of the present sulfoxidation system. This study provides the first example of a perovskite catalyst applicable to selective oxidation of sulfides using O2 as the sole oxidant.



RESULTS AND DISCUSSION Catalytic Oxidation of Sulfides with O2 as the Sole Oxidant. BaRuO3 was synthesized by a sol−gel method using malic acid/metal acetate in combination with acid treatment (see details given in the Supporting Information).40 The characterization results using X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), and transmission electron microscopy (TEM) are shown in Figure 2. The XRD pattern of BaRuO3 indicates

Figure 2. (a) XRD patterns of BaRuO3 (upper) and rhombohedral 9R-BaRuO3 (lower, ICSD 162087). (b) XPS Ru 3p3/2 spectrum and (c) SEM and (d) TEM images of BaRuO3.

Figure 1. Structure of (a) tetragonal RuO2, (b) cubic SrRuO3, (c) orthorhombic CaRuO3, and (d) rhombohedral BaRuO3. Gray, green, blue, purple, and red spheres represent Ru, Sr, Ca, Ba, and O atoms, respectively.

that the crystal structure of the sample could be indexed to the rhombohedral BaRuO3 structure (ICSD 16208, space group R3m) and that impurity phases of RuO2 (Figure 1a) and BaCO3 are not observed (Figures 2a and S1).43 The grain size (dXRD) was calculated to be 26 nm using Scherrer’s equation with respect to the (110) diffraction lines. Figures 2c and 2d show SEM and TEM images of BaRuO3, respectively. The size of spherical nanoparticles was estimated to be 20−50 nm, and the distribution well agreed with the dXRD value. The BaRuO3 particles present the clear lattice fringes with d-spacings of 0.29 and 0.33 nm which would be assigned to the (110) and (104) planes of BaRuO3, respectively. The specific surface area of BaRuO3 was 25 m2·g−1, and the value was much larger than that of BaRuO3 (5 m2·g−1) synthesized by the coprecipitation method.44 The molar ratio of Ba to Ru was ca. 1:1 based on the elemental analysis of BaRuO3 using inductively coupled plasma-atomic emission spectroscopy (ICP-AES). The Raman spectrum of BaRuO3 showed bands at 616 (A1g), 457 (A1g), and 317 (Eg) cm−1, and the band positions are close to the previously reported values for 9R-BaRuO3 (Figure S2).45,46 Figure 2b shows the XPS Ru 3p spectrum of BaRuO3. The peak with a binding energy of 463.2 eV likely corresponds to ruthenium oxides.47 Orthorhombic CaRuO3 (Figure 1c) and

transformation from corner-sharing to face-sharing octahedra is induced when using large alkaline-earth cations (A2+) which results in the hexagonal or rhombohedral structures (e.g., Figure 1d).34 Despite such a structural diversity, catalysis over lanthanum-based and/or multicomponent perovskites with corner-sharing BO6 octahedra has been mainly investigated,35,37 and applications have been limited to electrochemical,38 photocatalytic,36 and high-temperature gas-phase reactions.35,37 In particular, the perovskite-catalyzed liquidphase organic reactions for application in green chemistry have been scarcely investigated.35 We have recently reported the synthesis and catalytic application of high-surface-area hexagonal SrMnO3 perovskite nanomaterials for the selective aerobic oxidation of organic substrates such as alcohols through the reductive activation of O2.39,40 Here, we focus on the possible oxidation ability of ruthenium-based perovskite catalysts.41,42 Rhombohedral BaRuO3, which consists of units of three RuO6 octahedra sharing faces in a partial chain (Figure 1d), could be an efficient solid catalyst for the liquid-phase oxidation of sulfides using O2 as the sole oxidant. The present B

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

well as the successive oxidation of 2a to 3a. The effectiveness of tertiary alcohols decreased with an increase in the alkyl chain length in the order of t-BuOH > tert-amyl alcohol >3methyl-3-pentanol, which indicates that the steric effect of coadsorped solvents probably affects the reaction of substrates with surface active oxygen atoms. In addition, t-BuOH would not act as a reductant but as a solvent according to the following results: (i) The oxidation of 1a in nonreducible odichlorobenzene efficiently proceeded without an induction period in a similar way to that in t-BuOH, while the successive oxidation of the corresponding sulfoxide was also promoted (Figure S6). (ii) No coproducts caused by the oxidation (or oxidative degradation) of t-BuOH as a reductant was observed. When the oxidation of 1a in t-BuOH was carried out with O2 (1.0 MPa), the reaction proceeded efficiently, even at 313 K, and the yield and selectivity for 2a were increased to 92% and 84%, respectively (entry 2). Although several effective Mo, Au, and Mn-based inorganic solid catalyst systems for the selective sulfoxidation of 1a have been reported,29−32 high reaction temperatures are typically needed to attain high yield of 2a. Therefore, this system promotes the oxidation under mild conditions; the temperature (313 K) was lower than those (373−423 K) for previously reported systems (Table S1). In addition, the successive conversion of 1a into 3a in tBuOH could be achieved by an increase in the reaction temperature to 373 K (entry 3). Thus, the product selectivity could be controlled by changing the reaction conditions. Next, the catalyst effect on the aerobic oxidation of 1a in tBuOH was investigated (Table 2). The reaction did not proceed without a catalyst (entry 14). The catalytic activity and specific surface area of BaRuO3 synthesized by the malicacid-aided method were higher than those of BaRuO3 synthesized by the polymerized complex method (entries 1 and 2).39 The catalytic activity of ARuO3 (A = Ca, Sr, and Ba) is strongly dependent on the type of A site atoms: Cubic SrRuO3 and orthorhombic CaRuO3 with corner-sharing octahedra exhibited much lower catalytic activity than BaRuO3 with face-sharing octahedra (entries 1, 3, and 4). Simple RuO2 and other Ru complexes (Ru(OAc)x, Ru(acac)3, and catalyst precursor before the calcination (see details given in the Supporting Information)) were almost inactive (entries 5−9). Hexagonal SrMnO3 perovskite was also inactive,49 and successful aerobic sulfoxidation required isobutyraldehyde as an additive.39 The catalytic activity of other Mn-based solid catalysts, such as activated MnO2 and Mo-doped α-MnO2 (Mo-OMS-2), which can act as efficient oxidation catalysts,30−32,50,51 was lower than that of BaRuO3, even at an O2 pressure of 1.0 MPa, despite the large surface areas of these Mn-based solid catalysts (42−325 m2/g) (entries 10−13). To examine whether the present oxidation catalysis is due to solid BaRuO3 or not, BaRuO3 was removed from the reaction mixture by hot filtration at ca. 30% conversion of 1a under the conditions (entry 1, Table 1). Although the filtrate was then heated again, further production of 2a was not observed (Figure 3a). Leaching of Ru or Ba species into the filtrate was not observed using ICP-AES (detection limits for Ru and Ba atoms of ca. 10 and 0.3 ppb, respectively). Therefore, the nature of the present oxidation catalysis was verified to be heterogeneous.52 The used catalyst could be easily recovered by filtration. The recovered BaRuO3 catalyst could then be reused three times. In this case, a significant decrease in the yield and selectivity of 2a was not observed (Figure 3b). The recycling experiments at 0.1 MPa of O2 were carried out to

cubic SrRuO3 (Figure 1b) with corner-sharing RuO6 octahedra were also successfully synthesized by the malic acid-aided sol− gel method and were characterized by XRD, N2-adsorption, XPS, SEM, and elemental analyses (Figures S3−S5).48 First, the oxidation of thioanisole (1a) in various solvents catalyzed by BaRuO3 at 333 K was examined using O2 (0.1 MPa) as the sole oxidant (Table 1). The main products were Table 1. Solvent Effect on the Oxidation of 1a Catalyzed by BaRuO3 with O2a

select. (%) entry

solvent

conv. of 1a (%)

yield (%)

2a

3a

1 2b 3c 4 5 6 7 8 9 10 11 12 13 14

t-BuOH t-BuOH t-BuOH chlorobenzene o-dichlorobenzene toluene tert-amyl alcohol 1,4-dioxane butylacetate n-octane diethylcarbonate DMA 3-methyl-3-pentanol 2-butanone

73 91 >99 64 56 44 44 28 20 18 8 8 7 8

73 92 94 63 55 42 42 27 19 17 7 7 6 6

79 84 99 >99 >99 >99

21 16 >99 48 54 44 21 15 10 19 SrRuO3 (−1.57 eV) > BaRuO3 (−0.99 eV), which suggests that the complete reduction of BaRuO3 is the most unfavorable reaction. The discrepancy between the trend of surface reduction rates (based on H2-TPR experiments) and that of energy change for bulk reduction (based on DFT calculations) would be caused by the significantly lower surface area of CaRuO3 than other Ru-based catalysts.69 These results support the idea that the oxygen transfer from BaRuO3 to a sulfide, as well as the reoxidation of the partially reduced BaRuO3−x species with O2, proceeds much more smoothly than with the other Ru-based oxides. In this case, μ2-face-sharing oxygen atoms and/or related adsorbed oxygen species (including the possibility of other representative active species of RuO, Ru−O radical, RuOO−, etc.)41,42 caused by the face-sharing octahedra in BaRuO3 likely play an important role in the present oxidation reaction. G

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

ACS Applied Materials & Interfaces





CONCLUSION In summary, BaRuO3 efficiently catalyzed the selective oxidation of various aryl and aliphatic sulfides into sulfoxides and/or sulfones under mild reaction conditions with O2 as the sole oxidant and without any additives. The present system is a reusable heterogeneous catalyst that has potential for application to the large-scale oxidation of thioanisoles and quantitative conversion of dibenzothiophene into the corresponding sulfone. On the basis of the catalyst effect, 18Olabeling experiments, kinetic, mechanistic, spectroscopic, and computational results, substrate oxidation most likely proceeds via oxygen atoms from the BaRuO3 oxide surface, and facesharing octahedra in BaRuO3 would play an important role in the sulfoxidation reaction. Such a difference of mechanism and specific structure in hexagonal perovskite catalyst can overcome the limitations including the applicability to aryl sulfides. This study suggests the importance of the structure−reactivity relationship for liquid-phase reactions over perovskite catalysts under mild reaction conditions. Perovskite and related materials have many polymorphs, which has a significant influence on their chemical and physical properties. This approach is a promising strategy for the design and development of highly efficient heterogeneous catalysts for functional group transformation reactions through the structure-dependent activation of substrates and/or reagents.



REFERENCES

(1) Punniyamurthy, T.; Velusamy, S.; Iqbal, J. Recent Advances in Transition Metal Catalyzed Oxidation of Organic Substrates with Molecular Oxygen. Chem. Rev. 2005, 105, 2329−2363. (2) Cavani, F.; Teles, J. H. Sustainability in Catalytic Oxidation: An Alternative Approach or a Structural Evolution? ChemSusChem 2009, 2, 508−534. (3) Mizuno, N.; Kamata, K. Catalytic Oxidation of Hydrocarbons with Hydrogen Peroxide by Vanadium-Based Polyoxometalates. Coord. Chem. Rev. 2011, 255, 2358−2370. (4) Engelmann, X.; Monte-Perez, I.; Ray, K. Oxidation Reactions with Bioinspired Mononuclear Non-Heme Metal−Oxo Complexes. Angew. Chem., Int. Ed. 2016, 55, 7632−7649. (5) Stahl, S. S. Palladium Oxidase Catalysis. Selective Oxidation of Organic Chemicals by Direct Dioxygen-Coupled Turnover. Angew. Chem., Int. Ed. 2004, 43, 3400−3420. (6) Neumann, R.; Khenkin, A. M. Molecular Oxygen and Oxidation Catalysis by Phosphovanadomolybdates. Chem. Commun. 2006, 2529−2538. (7) Ishikawa, S.; Ueda, W. Microporous Crystalline Mo-V Mixed Oxides for Selective Oxidations. Catal. Sci. Technol. 2016, 6, 617−629. (8) Carreño, M. C. Applications of Sulfoxides to Asymmetric Synthesis of Biologically Active Compounds. Chem. Rev. 1995, 95, 1717−1760. (9) Trost, B. M.; Rao, M. Development of Chiral Sulfoxide Ligands for Asymmetric Catalysis. Angew. Chem., Int. Ed. 2015, 54, 5026− 5043. (10) Otocka, S.; Kwiatkowska, M.; Madalinska, L.; Kielbasinski, P. Chiral Organosulfur Ligands/Catalysts with a Stereogenic Sulfur Atom: Applications in Asymmetric Synthesis. Chem. Rev. 2017, 117, 4147−4181. (11) Mancuso, A. J.; Swern, D. Synthesis 1981, 1981, 165−185. (12) Ito, E.; van Veen, J. A. R. On Novel Processes for Removing Sulphur from Refinery Streams. Catal. Today 2006, 116, 446−460. (13) Bäckvall, J.-E. In Modern Oxidation Methods; Bäckvall, J.-E., Ed.; Wiley-VCH: Weinheim, 2004; pp 193−222. (14) Legros, J.; Dehli, J. R.; Bolm, C. Applications of Catalytic Asymmetric Sulfide Oxidations to the Syntheses of Biologically Active Sulfoxides. Adv. Synth. Catal. 2005, 347, 19−31. (15) Baciocchi, E.; Gerini, M. F.; Lapi, A. Synthesis of Sulfoxides by the Hydrogen Peroxide Induced Oxidation of Sulfides Catalyzed by Iron Tetrakis(pentafluorophenyl)porphyrin: Scope and Chemoselectivity. J. Org. Chem. 2004, 69, 3586−3589. (16) Mba, M.; Prins, L. J.; Licini, G. C3-Symmetric Ti(IV) Triphenolate Amino Complexes as Sulfoxidation Catalysts with Aqueous Hydrogen Peroxide. Org. Lett. 2007, 9, 21−24. (17) Kamata, K.; Hirano, T.; Mizuno, N. Highly Efficient Oxidation of Sulfides with Hydrogen Peroxide Catalyzed by [SeO4{WO(O2)2}2]2−. Chem. Commun. 2009, 3958−3960. (18) Egami, H.; Katsuki, T. Fe(salan)-Catalyzed Asymmetric Oxidation of Sulfides with Hydrogen Peroxide in Water. J. Am. Chem. Soc. 2007, 129, 8940−8941. (19) Bolm, C.; Bienewald, F. Asymmetric Sulfide Oxidation with Vanadium Catalysts and H2O2. Angew. Chem., Int. Ed. Engl. 1996, 34, 2640−2642. (20) Iwahama, T.; Sakaguchi, S.; Ishii, Y. Selective Oxidation of Sulfides to Sulfoxides with Molecular Oxygen Catalyzed by NHydroxyphthalimide (NHPI) in the Presence of Alcohols. Tetrahedron Lett. 1998, 39, 9059−9062. (21) Okada, S.; Ikurumi, S.; Kamegawa, T.; Mori, K.; Yamashita, H. Structural Design of Pd/SiO2@Ti-Containing Mesoporous Silica Core-Shell Catalyst for Efficient One-Pot Oxidation Using in Situ Produced H2O2. J. Phys. Chem. C 2012, 116, 14360−14367. (22) Imada, Y.; Tonomura, I.; Komiya, N.; Naota, T. Aerobic Oxidation of Sulfides with a Vitamin B2-Derived Organocatalyst. Synlett 2013, 24, 1679−1682. (23) Arakawa, Y.; Yamanomoto, K.; Kita, H.; Minagawa, K.; Tanaka, M.; Haraguchi, N.; Itsuno, S.; Imada, Y. Design of Peptide-Containing

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.8b05343.



Research Article

Experimental details, three tables (comparison of previously reported aerobic sulfoxidation systems, isolated yields, and XPS results), and 12 figures (XRD patterns, Raman spectrum, SEM images, XPS spectra, reaction profiles, and NMR spectra) (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Keigo Kamata: 0000-0002-0624-8483 Satoshi Muratsugu: 0000-0002-3596-7380 Fumiyasu Oba: 0000-0001-7178-5333 Michikazu Hara: 0000-0003-3450-5704 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the PRESTO program (No. JPMJPR15S3) of the JST and a Grant-in-Aid (No. 15H04184) for Japan Society for the Promotion of Science (JSPS) Fellows and for Scientific Research from Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of Japan. The TEM analyses were performed through the Nagoya University Advanced characterization platform and the “Creation of Life Innovative Materials for Interdisciplinary and International Researcher Development” programs of MEXT, Japan. H

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces N5-Unmodified Neutral Flavins that Catalyze Aerobic Oxygenations. Chem. Sci. 2017, 8, 5468−5475. (24) Xu, H.-J.; Lin, Y.-C.; Wan, X.; Yang, C.-Y.; Feng, Y.-S. Selective Aerobic Oxidation of Sulfides to Sulfoxides Catalyzed by Coenzyme NAD+ Models. Tetrahedron 2010, 66, 8823−8827. (25) Suzuki, K.; Jeong, J.; Yamaguchi, K.; Mizuno, N. Photoredox Catalysis for Oxygenation/Deoxygenation Between Sulfides and Sulfoxides by Visible-Light-Responsive Polyoxometalates. New J. Chem. 2016, 40, 1014−1021. (26) Xu, Y.; Fu, Z.-C.; Cao, S.; Chen, Y.; Fu, W.-F. Highly Selective Oxidation of Sulfides on a CdS/C3N4 Catalyst with Dioxygen Under Visible-Light Irradiation. Catal. Sci. Technol. 2017, 7, 587−595. (27) Riley, D. P.; Smith, M. R.; Correa, P. E. Selective Molecular Oxygen Oxidation of Thio Ethers to Sulfoxides Catalyzed by Cerium(IV). J. Am. Chem. Soc. 1988, 110, 177−180. (28) Aldea, R.; Alper, H. Selective Aerobic Oxidation of Sulfides Using a Novel Palladium Complex as the Catalyst Precursor. J. Org. Chem. 1995, 60, 8365−8366. (29) Gao, Q.; Giordano, C.; Antonietti, M. Biomimetic Oxygen Activation by MoS2/Ta3N5 Nanocomposites for Selective Aerobic Oxidation. Angew. Chem., Int. Ed. 2012, 51, 11740−11744. (30) Taketoshi, A.; Concepción, P.; García, H.; Corma, A.; Haruta, M. Aerobic Oxidation of Sulfides to Sulfoxides Catalyzed by Gold/ Manganese Oxides. Bull. Chem. Soc. Jpn. 2013, 86, 1412−1418. (31) MnO2@wool, which was prepared from KMnO4 and natural white wool (sheep) support, can act as a recyclable heterogeneous catalyst for selective oxidation of sulfides to sulfoxide at ambient temperature: Shaabani, A.; Hezarkhani, Z.; Badali, E. Wool Supported Manganese Dioxide Nano-Scale Dispersion: A Biopolymer Based Catalyst for the Aerobic Oxidation of Organic Compounds. RSC Adv. 2015, 5, 61759−61767. (32) Uematsu, T.; Miyamoto, Y.; Ogasawara, Y.; Suzuki, K.; Yamaguchi, K.; Mizuno, N. Molybdenum-Doped α-MnO2 as an Efficient Reusable Heterogeneous Catalyst for Aerobic Sulfide Oxygenation. Catal. Sci. Technol. 2016, 6, 222−233. (33) It has been reported that the oxidation of alkyl sulfides at high pressure of molecular oxygen (7.2 MPa) proceeds even in the absence of catalysts. On the other hand, aryl sulfides are inactive under the same conditions. Thus, the oxidation of aryl sulfides requires the presence of catalysts: Correa, P. E.; Riley, D. P. Highly Selective Direct Oxidation of Thioethers to Sulfoxides Using Molecular Oxygen. J. Org. Chem. 1985, 50, 1787−1788. (34) Darriet, J.; Subramanian, M. A. Structural Relationships between Compounds Based on the Stacking of Mixed Layers Related to Hexagonal Perovskite-type Structures. J. Mater. Chem. 1995, 5, 543−552. (35) Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as Substitutes of Noble Metals for Heterogeneous Catalysis: Dream or Reality. Chem. Rev. 2014, 114, 10292−10368. (36) Wang, W.; Tade, M. O.; Shao, Z. Research Progress of Perovskite Materials in Photocatalysis- and Photovoltaics-related Energy Conversion and Environmental Treatment. Chem. Soc. Rev. 2015, 44, 5371−5408. (37) Zhu, H.; Zhang, P.; Dai, S. Recent Advances of LanthanumBased Perovskite Oxides for Catalysis. ACS Catal. 2015, 5, 6370− 6385. (38) Zhu, Y.; Zhou, W.; Shao, Z. Perovskite/Carbon Composites: Applications in Oxygen Electrocatalysis. Small 2017, 13, 1603793. (39) Kawasaki, S.; Kamata, K.; Hara, M. Dioxygen Activation by a Hexagonal SrMnO3 Perovskite Catalyst for Aerobic Liquid-Phase Oxidation. ChemCatChem 2016, 8, 3247−3253. (40) Sugahara, K.; Kamata, K.; Muratsugu, S.; Hara, M. Amino AcidAided Synthesis of a Hexagonal SrMnO3 Nanoperovskite Catalyst for Aerobic Oxidation. ACS Omega 2017, 2, 1608−1616. (41) Naota, T.; Takaya, H.; Murahashi, S.-I. Ruthenium-Catalyzed Reactions for Organic Synthesis. Chem. Rev. 1998, 98, 2599−2660.

(42) Murahashi, S.; Zhang, D. Ruthenium Catalyzed Biomimetic Oxidation in Organic Synthesis Inspired by Cytochrome P-450. Chem. Soc. Rev. 2008, 37, 1490−1501. (43) Donohue, P. C.; Katz, L.; Ward, R. The Crystal Structure of Barium Ruthenium Oxide and Related Compounds. Inorg. Chem. 1965, 4, 306−310. (44) Doggali, P.; Grasset, F.; Cador, O.; Rayalu, S.; Teraoka, Y.; Labhsetwar, N. Studies on Catalytic and Structural Properties of BaRuO3 Type Perovskite Material for Diesel Soot Oxidation. J. Environ. Chem. Eng. 2014, 2, 340−343. (45) Quilty, J.; Trodahl, H. J.; Edgar, A. Raman Spectroscopy of Barium Ruthenium Oxide BaRuO3. Solid State Commun. 1993, 86, 369−371. (46) Lee, Y. S.; Noh, T. W.; Park, J. H.; Lee, K. B.; Cao, G.; Crow, J. E.; Lee, M. K.; Eom, C. B.; Oh, E. J.; Yang, I.-S. TemperatureDependent Raman Spectroscopy in BaRuO3 Systems. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 65, 235113. (47) Morgan, D. J. Resolving Ruthenium: XPS Studies of Common Ruthenium Materials. Surf. Interface Anal. 2015, 47, 1072−1079. (48) Bensch, W. Structure and Thermochemical Reactivity of CaRuO3 and SrRuO3. Solid State Ionics 1990, 43, 171−177. (49) Other perovskite-type oxide catalysts containing alkaline-earth metals ABO3−δ (AB = CaMn, BaMn, CaFe, SrFe, BaFe, Sr1.2Co, and BaCo) synthesized by our method39 were inactive for the sulfoxidation under the present reaction conditions. (50) Fatiadi, A. J. Active Manganese Dioxide Oxidation in Organic Chemistry - Part II. Synthesis 1976, 1976, 133−167. (51) Hayashi, E.; Komanoya, T.; Kamata, K.; Hara, M. Heterogeneously-Catalyzed Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid with MnO2. ChemSusChem 2017, 10, 654−658. (52) Sheldon, R. A.; Wallau, M.; Arends, I. W. C. E.; Schuchardt, U. Heterogeneous Catalysts for Liquid-Phase Oxidations: Philosophers’ Stones or Trojan Horses? Acc. Chem. Res. 1998, 31, 485−493. (53) Davies, D. M.; Deary, M. E.; Quill, K.; Smith, R. A. BorateCatalyzed Reactions of Hydrogen Peroxide: Kinetics and Mechanism of the Oxidation of Organic Sulfides by Peroxoborates. Chem. - Eur. J. 2005, 11, 3552−3558. (54) Imada, Y.; Iida, H.; Ono, S.; Murahashi, S. Flavin Catalyzed Oxidations of Sulfides and Amines with Molecular Oxygen. J. Am. Chem. Soc. 2003, 125, 2868−2869. (55) Bonesi, S. M.; Fagnoni, M.; Albini, A. Hammett Correlations in the Photosensitized Oxidation of 4-Substituted Thioanisoles. J. Org. Chem. 2004, 69, 928−935. (56) Chatterjee, S.; Paine, T. K. Hydroxylation versus Halogenation of Aliphatic C−H Bonds by a Dioxygen-Derived Iron−Oxygen Oxidant: Functional Mimicking of Iron Halogenases. Angew. Chem., Int. Ed. 2016, 55, 7717−7722. (57) Kamata, K.; Hirano, T.; Ishimoto, R.; Mizuno, N. Sulfoxidation with Hydrogen Peroxide Catalyzed by [SeO4{WO(O2)2}2]2−. Dalton Trans. 2010, 39, 5509−5518. (58) Tang, N. F.; Jiang, Z. X.; Li, C. Oxidation of Refractory SulfurContaining Compounds with Molecular Oxygen Catalyzed by Vanadoperiodate. Green Chem. 2015, 17, 817−820. (59) Lu, H.; Gao, J.; Jiang, Z.; Yang, Y.; Song, B.; Li, C. Oxidative Desulfurization of Dibenzothiophene with Molecular Oxygen Using Emulsion Catalysis. Chem. Commun. 2007, 150−152. (60) Che, Y.; Ma, W.; Ren, Y.; Chen, C.; Zhang, X.; Zhao, J.; Zang, L. Photooxidation of Dibenzothiophene and 4,6-Dimethyldibenzothiophene Sensitized by N-Methylquinolinium Tetrafluoborate: Mechanism and Intermediates Investigation. J. Phys. Chem. B 2005, 109, 8270−8276. (61) Sampanthar, J. T.; Xiao, H.; Dou, H.; Nah, T. Y.; Rong, X.; Kwan, W. P. A Novel Oxidative Desulfurization Process to Remove Refractory Sulfur Compounds from Diesel Fuel. Appl. Catal., B 2006, 63, 85−93. (62) Zhang, W.; Zhang, H.; Xiao, J.; Zhao, Z. X.; Yu, M. X.; Li, Z. Carbon Nanotube Catalysts for Oxidative Desulfurization of A Model I

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces Diesel Fuel Using Molecular Oxygen. Green Chem. 2014, 16, 211− 220. (63) Gomez-Paricio, A.; Santiago-Portillo, A.; Navalon, S.; Concepcion, P.; Alvaro, M.; Garcia, H. MIL-101 Promotes the Efficient Aerobic Oxidative Desulfurization of Dibenzothiophenes. Green Chem. 2016, 18, 508−515. (64) Vannice, M. A. An Analysis of the Mars−van Krevelen Rate Expression. Catal. Today 2007, 123, 18−22. (65) The 16O2/18O2 exchange reaction did not occur under the reaction conditions without 1a. In addition, the aerobic oxidation of 2,3-dimethyl-2-butene, which is easily trapped by singlet oxygen than sulfides, did not proceeded with BaRuO3. These results also support suggest that the origin of oxygen atoms in oxygenated products (sulfoxides and sulfones) is not from the external gaseous dioxygen including singlet oxygen but from the oxygen atoms in BaRuO3: Bonesi, S. M.; Fagnoni, M.; Monti, S.; Albini, A. Reaction of Singlet Oxygen with Some Benzylic Sulfides. Tetrahedron 2006, 62, 10716− 10723. (66) Riley, D. P.; Shumate, R. F. Mechanistic Studies of the Selective Oxygen Oxidation of Sulfides to Sulfoxides Catalyzed by Dihaloruthenium(II) Complexes. J. Am. Chem. Soc. 1984, 106, 3179−3184. (67) Al-Ajlouni, A. M.; Daiafla, T. M.; El-Khateeb, M. New Nitrophenyl-Substituted Polyperoxotungstate Catalyst: A More Active and Selective for the Oxidation of Sulfides by Hydrogen Peroxide. J. Mol. Catal. A: Chem. 2007, 275, 139−147. (68) Yu, J.; Zhao, D.; Xu, X.; Wang, X.; Zhang, N. Study on RuO2/ SnO2: Novel and Active Catalysts for CO and CH4 Oxidation. ChemCatChem 2012, 4, 1122−1132. (69) CaRuO3 also showed multistep reduction peaks;: a relatively weak peak at 500 K and an intense peak at ca. 740 K which are higher than those of BaRuO3 and SrRuO3. It has been reported that the decrease in the surface area causes the shift of reduction peaks toward higher temperatures due to the increased diffusional resistance. Irusta, S.; Pina, M. P.; Menéndez, M.; Santamaría, J. Catalytic Combustion of Volatile Organic Compounds over La-Based Perovskites. J. Catal. 1998, 179, 400−412. (70) McFarland, E. W.; Metiu, H. Catalysis by Doped Oxides. Chem. Rev. 2013, 113, 4391−4427. (71) Agarwal, V.; Metiu, H. Energy of Oxygen-Vacancy Formation on Oxide Surfaces: Role of the Spatial Distribution. J. Phys. Chem. C 2016, 120, 2320−2323. (72) Natile, M. M.; Ugel, E.; Maccato, C.; Glisenti, A. LaCoO3: Effect of Synthesis Conditions on Properties and Reactivity. Appl. Catal., B 2007, 72, 351−362.

J

DOI: 10.1021/acsami.8b05343 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX