Next Article in Journal
Application of Chelating Agents to Enhance Fenton Process in Soil Remediation: A Review
Previous Article in Journal
Octahedral Growth of PtPd Nanocrystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Oxidation of Benzyl Alcohol to Benzaldehyde on Au8 and Au6Pd2 Clusters: A DFT Study on the Reaction Mechanism

by
Jitlada Kodchasee
1,
Chanon Chanloi
1,
Pongtanawat Khemthong
2,
Boontida Uapipatanakul
1,
Masahiro Ehara
3,4,* and
Karan Bobuatong
1,*
1
Department of Chemistry, Faculty of Science and Technology, Rajamangala University of Technology Thanyaburi, Pathum Thani 12110, Thailand
2
National Nanotechnology Center (NANOTEC), National Science and Technology, Development Agency, 111 Thailand Science Park, Paholyothin Rd., Pathum Thani 12120, Thailand
3
Institute for Molecular Science and Research Center for Computational Science, 38 Nishigo-Naka, Myodaiji, Okazaki 444-8585, Japan
4
Elements Strategy Initiative for Catalysts and Batteries (ESICB), Kyoto University, Kyoto Daigaku-Katsura, Kyoto 615-8510, Japan
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(6), 720; https://doi.org/10.3390/catal11060720
Submission received: 17 May 2021 / Revised: 6 June 2021 / Accepted: 7 June 2021 / Published: 9 June 2021
(This article belongs to the Section Computational Catalysis)

Abstract

:
Density functional theory calculations were performed to investigate the reaction mechanism of the aerobic oxidation of benzyl alcohol to benzaldehyde catalyzed by Au and Au–Pd clusters. Two consecutive reaction mechanisms were examined with Au8 and Au6Pd2 clusters: (1) the oxidation of benzyl alcohol with dissociated O atoms on metal clusters generating benzaldehyde and H2O; and (2) oxidation with adsorbed oxygen molecules generating benzaldehyde and H2O2. The calculations show that the aerobic oxidation of benzyl alcohol energetically prefers to proceed in the former mechanism, which agrees with the experimental observation. We demonstrate that the role of Au centers around the activation of molecular oxygen to peroxide-like species, which are capable of the H–abstraction of benzyl alcohol. The roles of Pd in the Au6Pd2 cluster are: (1) increasing the electron distribution to neighboring Au atoms, which facilitates the activation of O2; and (2) stabilizing the adsorption complex and transition states by the interaction between positively charged Pd atoms and the π-bond of benzyl alcohol, both of which are the origin of the lower energy barriers than those of Au8.

Graphical Abstract

1. Introduction

Bimetallic systems have gained much attention in both academic and industrial research fields because they often exhibit advantageous properties such as catalytic, electronic, and optical properties, compared with those of their pure constituent metals [1,2,3,4]. Among the numerous metals that have been investigated in combination, Au–Pd bimetallic nanoclusters were found to be efficient catalysts for a wide variety of chemical reactions because they enable one to perform reactions under mild conditions, providing high selectivity for the desired products [1,2,3,4,5,6,7]. For instance, their excellent catalytic performance was shown for the direct synthesis of H2O2 [7,8,9,10,11], the synthesis of vinyl acetate [12,13], the epoxidation of alkenes [14,15], the hydrogenation of unsaturated hydrocarbons [16,17], the oxidation of greenhouse gases [18,19], coupling reactions [20,21,22], and the oxidation of alcohols to carbonylic compounds [23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39].
The oxidation of alcohols, in particular the selective oxidation of primary alcohols to aldehydes, is an important process for the synthesis of fine chemicals and highly valuable intermediates. The oxidation of alcohols is commercially conducted by stoichiometric oxygen donors, such as chromate or permanganate producing aldehydes. However, this produces a large amount of toxic waste, which is environmentally unacceptable [40,41]. In this respect, a series of experimental studies have been devoted to developing catalysts to eliminate this drawback over the past two decades. Gold nanoclusters were reported to be effective catalysts for the selective oxidation of alcohols when used alone; however, significant stability and selectivity improvements were observed with the incorporation of a second metal promoter, such as Pd [5]. It was demonstrated that Au–Pd bimetallic clusters have remarkable catalytic activity for the selective oxidation of benzyl alcohol, which is one of the most frequently employed alcohols for the catalytic activity of a Au–Pd bimetallic catalyst under solvent-free conditions, and is comparable to, or even better than, Au catalysts [23,24,25,26,27,28,29]. Regarding mechanistic studies on AuPd bimetallic-catalyzed selective oxidation of benzyl alcohol, two different mechanisms of monometallic Au- and Pd-supported nanoparticles for alcohol oxidation have been proposed experimentally [42,43,44]: in the case of Au, the rate-determining step involves H-atom abstraction by a superoxo-like oxygen species adsorbed on Au, whereas, for Au–Pd, the rate-determining step is the Hβ elimination of the adsorbed alkoxide forming the aldehyde [23]. Nevertheless, the precise reaction mechanisms of the Au–Pd bimetallic catalyst, in particular the mechanistic role of Au and Pd atoms in accelerating such reaction, remain the subject of considerable debate and are yet to be fully understood.
In the past decade, quantum chemical calculations have been widely applied to understand the structure and catalytic activity of various nanocluster catalysts, including gold and gold-based bimetallic systems [45,46,47,48,49,50,51,52,53,54,55]. Density functional theory (DFT) calculations with the generalized gradient approximation (GGA) functional have been frequently used because of their low computational cost, but some issues still remain with respect to their validity [54,55]. For example, GGA tends to overestimate the binding energy of O2 on gold clusters compared with that calculated by CCSD(T) [54] and the reaction energies of the oxidation of methanol by CR-CC(2,3) [56]. However, it was demonstrated that the DFT calculations with M06-L [57] and M06 [58] functionals are practical choices for the theoretical study of nanoscale gold and gold-palladium catalysis [59,60,61,62,63,64]. Unlike with GGA approximations, the use of these functionals shows how to avoid overestimating the bond energies, leading to an accurate description of the 2D-to-3D structural transition of gold clusters. Karanjit et al. demonstrated that the use of M06 functionals and anionic Au8 clusters is a suitable model for simulating the aerobic oxidation (oxidation with molecular oxygen) of methanol on PVP (N-vinyl-2-pyrrolidone) supported gold clusters [64]. However, the catalytic activity of Au–Pd clusters for the oxidation of benzyl alcohol has yet to be investigated using suitable modeling.
In this work, we performed a theoretical study to better understand how gold and palladium affect the energetics and the reaction mechanism of the aerobic oxidation of benzyl alcohol over Au8 and Au6Pd2 using DFT calculations with the M06 functional. We examined the low-energy structures of Au8 isomers and then constructed the low-energy structures of Au6Pd2 isomers as a simple model for Au–Pd bimetallic clusters enriched in gold atoms. Next, we investigated the possible reaction pathways as well as the role of Au and Pd in the oxidation reaction of benzyl alcohol to benzaldehyde on both Au8 and Au6Pd2 clusters.

2. Results and Discussion

2.1. Structures of Au8 and Au6Pd2

The Au8 cluster was chosen as a model catalyst because, firstly, Au8 is a stable gold cluster that can be detected in low-temperature experiments [65,66,67,68,69], and, secondly, it is one of the most commonly employed Au nanocatalysts in theoretical studies [69,70]. Subsequently, we performed a systematic search for the energetically minimum structures of Au8 and Au6Pd2. AuPd is gold-rich and shows the optimum catalytic activity for the aerobic oxidation of alcohol, such as benzyl alcohol and cinnamyl alcohol without the formation of acid derivatives [27,71]. The lowest-energy structures of Au8 and Au6Pd2 are shown in Figure 1 while other structures and the relative energies are provided in the Supplementary Material (Figures S1 and S2 and Tables S1 and S2). The most stable of the Au8 and Au6Pd2 structures have D4h and nonplanar Cs symmetry, respectively, in which two Pd atoms of Au6Pd2 are located in the adjacent positions. It was found that the introduction of two Pd atoms in the Au8 cluster slightly increases the partial charge at the neighboring Au atoms due to the lower electron affinity of the Pd atoms. This characteristic of Au6Pd2 clusters facilitates orbital interaction between O2 and apex gold atoms, which induces the binding of O2 to the Au6Pd2 cluster, which is described in the next section.
The Mulliken charges presented in Figure 1 indicate that the charge distribution is not uniform over all the Au8 atoms. High electron densities appear at the apex sites where the gold atoms are less coordinated; the highly coordinated gold atoms have a lower electron density. In the present case, the Mulliken charges provided a more reasonable picture than those obtained by natural bond orbital analysis (NBO). For instance, NBO charges at the apex sites of the metal nanoclusters appear to be positive, which is not reasonable considering the charge distribution and reactivity. The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of the Au8 cluster, which are depicted in Figure 2a, reveal that HOMO is localized at low-coordinated gold atoms. This agrees well with the report of Metiu et al. [72], which demonstrates that the valence electrons of the gold cluster are not delocalized over the entire cluster, but localized at the apex, corner, and step. Localization of the HOMO on the apex sites of metal nanoclusters is correlated with the activation of a small molecule, i.e., O2, because it facilitates this molecule adsorbing effectively on apex sites. Although it was reported that the complete basis set limit (CBS) beyond the MP2 and CCSD(T) level of theory is required for predicting the dimensionality of the Au8 cluster [73], the practical choice of DFT calculations with moderate basis sets which reproduce the results of wavefunction theories is useful for simulating the catalytic reactions of gold clusters. Accordingly, the hybrid meta-GGA functional and M06 functional employed in this study provide a similar trend in energetics and electronic properties when compared with previous theoretical studies. Based on the present results together with those of previous studies [59,60,61,62,63,64], it was reasonable to adopt the present computational method for further calculations of the aerobic oxidation of benzyl alcohol.

2.2. Pathways and Energetics: The Formation of Benzaldehyde and H2O

The catalytic cycle of the oxidation of benzyl alcohol examined in this work is schematically displayed in Figure 3. The adsorption of O2 on the cluster is considered a first step for the aerobic oxidation of alcohols. First, we studied the adsorption behavior of O2 on Au8 and Au6Pd2 clusters. The optimized adsorption complexes are shown in Figure 4, and the structural parameters are provided in Table S3 of the Supplementary Material. In the case of Au8, the O2 molecule is most preferably adsorbed at the apex Au atom, which possesses the highest negative charge, followed by coadsorption with a benzyl alcohol molecule. The adsorption of O2 has been reported as the initial step in the oxidation of methanol on gold clusters [59,60], because it results in a pronounced cooperative effect in the activation of molecular oxygen and catalytic activity for the oxidation of methanol. The O2 molecule acts as the charge acceptor from Au8 and Au6Pd2. The partial charges at apex Au atoms attached to O2 decrease from δq = −0.094 to −0.081 and δq = −0.145 to −0.068 for Au8 and Au6Pd2 clusters, respectively (Table S3). The partial charge on the O2 molecule is δq = −0.181 and δq = −0.213 for Au8 and Au6Pd2 clusters, respectively; the negative charge is enhanced in the Au6Pd2 cluster. Considering isolated species, O2 and metal clusters (Figure 2a) illustrate the adsorption complexes, whereas the adsorbed O2 illustrates the bent-triatomic bonding with the apex gold atom (Figure 2b). If the adsorption occurred by charge transfer from the Au cluster into a π*-type orbital (LUMO) of O2, we would expect that electrons would be gained in a spatial region resembling this orbital. This turned out to occur in the presented adsorption complexes we examined. The example is depicted in Figure 2b. The frontier orbitals of the adsorption complexes suggest that the overlapping of LUMOs between O2 and apex gold atom readily induce the adsorption process. As a result, the O–O bond of the adsorbed O2 molecules is increased to 1.29 and 1.30 Å for Au8 and Au6Pd2 clusters, respectively. These O–O bonds are elongated in between the reported experimental values of O2 (1.207 Å) [74] and O2 (1.347 Å) [75]. These results suggest that charge transfer occurs from the cluster to the adsorbed molecular oxygen, leading to the formation of a superoxo-like species, which was experimentally observed as an active species in the selective oxidation of alcohol [76].
For the next step, various orientations between and benzyl alcohol on the Au8 and Au6Pd2 clusters were taken into account for initially designed geometries. The most stable geometries optimized at the chosen level of theory are shown in Figure 4, Figure 5, Figure 6 and Figure 7 and employed for the further step of the transition state search; meanwhile, other adsorption sites and configurations considered in this work are depicted in Figures S3 and S4 of the Supplementary Material. The energy profile from the adsorption of substrates to the generation of benzaldehyde and superoxo species is depicted in Figure 5. The calculated adsorption energy of O2 and benzyl alcohol on the Au6Pd2 cluster is −26.8 kcal/mol, which is much larger than that on the Au8 cluster of 11.4 kcal/mol. This result can be understood by analyzing the frontier molecular orbitals and the charges on the atoms of the adsorption complexes. The HOMO of the adsorption complexes is delocalized over Pd atoms and the benzene moiety of benzyl alcohol, whereas this characteristic is not observed over Au8 (Figure 4). Moreover, the Mulliken charge analysis confirms a charge transfer from the benzene unit of benzyl alcohol to Pd atoms, whereby the charges on the Pd atoms decrease from δq = 0.147 to 0.101. These results suggest that, once the adsorption complex between benzyl alcohol and O2 on Au–Pd bimetallic surface is formed, the interaction between positively charged Pd atoms and benzyl alcohol via π-bond interaction and the charge transfer from catalyst cluster to O2 is significant for stabilizing this adsorption complex.
The Pd atoms not only enhance the electron density on the neighboring Au atoms and are responsible for the stability of the adsorption complex, but also significantly stabilize the transition state structures in the oxidation reaction. As shown in Figure 5, in the presence of two Pd atoms of Au6Pd2, the transition state of H–abstraction by a superoxo-like species (TS_1B) is much more stable than that of Au8 (TS_1A); the energy barrier of Au6Pd2 is lower by 2.4 times than Au8 (5.5 vs. 13.4 kcal/mol). The 2D to 3D evolution of the Au cluster at TS_1A and thereafter is expected, since the fully relaxed geometry optimization can change the cluster geometry considerably and/or its charge distribution [58] Consequently, the transition state structure reorients to form the interaction between the benzene fragment of benzyl alcohol and Au cluster. The transition state leads to the formation of a hydroperoxyl-like species, AuOOH (Int_1A and Int_1B), which is considered the key intermediate in the oxidation of alcohols [76] The relative energies of these intermediates were estimated to be −8.1 and −22.6 kcal/mol for Au8 and Au6Pd2, respectively. The difference in the stability of the intermediates is due to the benzylic alkoxide in the latter case occupying a bridge position between two positively charged Pd atoms, resulting in the stronger interaction, which is also indicated by the shorter bond distance of O–Pd compared with O–Au (2.17 vs. 2.25 Å). Subsequently, these key intermediates are transformed to either benzaldehyde and H2O via Hβ–abstraction (Figure 5 and Figure 6) or benzaldehyde and H2O2 (Figure 7).
For the next step, the benzaldehyde formation via Hβ–abstraction (TS_2A and TS_2B) readily occurs on the metal clusters and passes the Hβ to the neighboring hydroperoxyl-like species. Figure 5 shows that this step can be achieved with the energy barriers of 17.9 and 7.2 kcal/mol for Au8 and Au6Pd2, respectively. The calculated activation energies are lower than those obtained from the competitive pathway (21.8 and 12.4 kcal/mol, respectively), which are discussed in the next section. The present result agrees well with the experimental observation in which the use of molecular oxygen as an oxidant in the oxidation of alcohols generates water as a byproduct [43,76]. After overcoming the low-energy barriers, TS_2A and TS_2B, benzaldehyde, and H2O can be desorbed, leaving chemisorbed atomic oxygen in the bridge positions on Au8 and Au6Pd2.
The resultant chemisorbed atomic oxygen on the cluster (Au8-O and Au6Pd2-O species) can be removed by the reaction with another benzyl alcohol. The energy profile of this route with the intermediate and transition-state structures is shown in Figure 5. This reaction pathway is initiated by the adsorption of benzyl alcohol on the cluster, forming a hydrogen bond between the hydroxyl group and the atomic oxygen with calculated energies of −26.3 and −36.2 kcal/mol for Au8 and Au6Pd2, respectively. The adsorption energies are substantially affected by the charge distribution and the relaxation near the bridge positions of the atomic oxygen. Table S5 in the Supplementary Material shows that the charge distribution of chemisorbed atomic oxygen on Au6Pd2 is more pronounced than on Au8 (δq = −0.663 vs. −0.612), leading to strong hydrogen bonding and yielding a more stable adsorption complex. In the transition state, the O–H bond of benzyl alcohol is activated by chemisorbed atomic oxygen. In this case, an explanation for the high adsorption energies of this elementary step is the difference in the charge distribution for the atomic oxygen on catalyst clusters. For example, the negative charge on the chemisorbed atomic oxygen of the Int_4A and Int_4B is larger than that on the O1 atom of superoxo-like species in the Ads_A and Ads_B (−0.612 vs. −0.578, and −0.663 vs. −0.580, respectively). In other words, the atomic oxygen on Au6Pd2 possesses more basicity than on Au8. This results in strong hydrogen bonding between the adsorbed atomic oxygen (proton acceptor) and the hydroxyl group of benzyl alcohol (proton donor) in Int_4. Subsequently, the dissociation of the O1–H1 bond and the formation of the O3–H1 bond occur. In this step, the atomic oxygen moves closer to the apex gold atom; meanwhile, the phenyl group of benzyl alcohol is only slightly affected in the course of the reaction. The calculated energy barriers relative to the adsorption complex were 9.5 and 3.4 kcal/mol for Au8 and Au6Pd2, respectively. The resulting benzylic alkoxide occupies a bridge position on Au8 and Au6Pd2, as was found in Int_1A and Int_1B. Subsequently, the Hβ atom is transferred to a surface hydroxyl group with the optimized C–Hβ distances of 1.28 and 1.19 Å for Au8 and Au6Pd2, respectively; meanwhile, the C–O bond length shortens to 1.34 Å, indicating the characteristic of the carbonyl group produces benzaldehyde and H2O. Consequently, the products, i.e., benzaldehyde and H2O, are adsorbed at catalyst clusters with the calculated relative energies of −35.1 and −52.7 kcal/mol for Au8 and Au6Pd2, respectively. These predicted reactions are exothermic; benzaldehyde and H2O require desorption energies of 20.3 and 22.8 kcal/mol for Au8 and Au6Pd2, respectively.

2.3. The Formation of Benzaldehyde and H2O2

As noted above, another possibility exists for the formation of benzaldehyde and H2O2. The Hβ–abstractions by oxygen atoms attached to the apex gold atoms occur while the interaction between the benzyl fragment of benzyl alcohol and the metal clusters is formed. Figure 7 shows that the activation energy of this step is 21.8 and 12.4 kcal/mol for Au8 and Au6Pd2, respectively. These values are slightly higher than those obtained from the benzaldehyde formation with H2O in the previous section. Thus, we finally conclude that the O–H bond cleavage activated by chemisorbed atomic oxygen is kinetically preferable to the superoxo-like species.
Comparing the energy barriers between the proposed reaction pathways, we found that the mechanism of Hβ–abstraction requires higher activation energies than that of the O–H bond dissociation for both catalysts. Similar trends have also been reported in experimental [43,77,78,79] and theoretical [80] studies; the presence of a chemisorbed oxygen atom on the gold nanoparticle has a strong influence on the first step of the reaction, decreasing the barrier energy for the deprotonation of the alcohol hydroxyl group, but failing to enhance the dissociation of the C–Hβ bond. Therefore, the latter step can be considered as the rate-determining step for alcohol oxidation on gold nanoclusters.
Overall, the energy profiles of reactions presented in this study show that the presence of Pd in the Au6Pd2 bimetallic cluster can significantly enhance the catalytic activity compared with the Au8 monometallic cluster. This result agrees well with the reported experimental study whereby enhancements in the selectivity and conversion towards benzaldehyde were observed for gold-rich Au–Pd catalysts, which were maximized for Au80–Pd20 compared with palladium-rich ones [71]. Finally, we believe that the present insights into the reaction mechanism and energetics for the aerobic oxidation of benzyl alcohol to benzaldehyde might be useful for developing bimetallic catalysts for the aerobic oxidation of alcohols.

3. Computational Details and Models

All DFT calculations were carried out with the Gaussian09 suit of program [81]. We used hybrid meta-exchange-correlation functionals, the M06 functional, which has been shown to be a reasonably useful functional for studying the structures and catalytic activities of gold clusters, such as the dimensionality of gold clusters and various reactions [61,62,63,64]. All-electron calculations for the noble metals in the lower half of the periodic table are difficult; therefore, the scalar relativistic effective core potentials (RECP) for the Au (Xe core) and Pd (Kr core) atoms were applied to describe the core electrons. The LanL2DZ basis sets [82,83,84] were introduced to describe the outer shell (5s5p5d6s) of the Au atoms and the outer shell (4s4p4d) of the Pd atoms, where pure d and pure f functions were employed for all calculations. The 6-31G(d,p) basis sets were applied for H, C, and O atoms. Since it was reported that the Minnesota functionals are sensitive to the choice of the grid for numerical integrations, the ultrafine integration grid of Gaussian09 was used. In the optimization process, all metal clusters and substrates were allowed to fully relax without any symmetry constraints. The convergence thresholds of the optimization were set to the default values. All the optimized ground state structures were confirmed as local minima, vibrational analysis was performed on the optimized structures, and none of the imaginary frequencies were found. The transition state search was achieved using the QST3 version of the transit-guided quasi-Newton (STQN) method, with 30 images for both the forward and backward direction corresponding to the direction of the transition vector from the TS. All the optimized activated complexes and transition states consisted of only one imaginary frequency. The Mulliken and NBO charge analysis for the most stable structure of Au8 and Au6Pd2 was evaluated using LANL2TZ. The final energy profiles were presented with the Gibbs energy change (∆G°298.15) obtained by frequency calculations with triple-zeta-quality basis sets; the 6-311++G(d,p) basis sets were employed for C, H and O atoms, whereas the LanL2TZ basis set was used for Au and Pd.

4. Conclusions

In this work, we showed how the presence of Pd in Au–Pd clusters can significantly affect the reaction energetics of the aerobic oxidation of benzyl alcohol to benzaldehyde. The results can be summarized as follows:
(1)
The reaction is initiated by the activation of the O2 molecule on a low-coordinated Au atom forming a peroxide-like species, followed by hydrogen abstraction from the hydroxyl group of benzyl alcohol. Subsequently, C–Hβ bond dissociation occurs to produce either benzaldehyde and H2O2 or benzaldehyde and H2O. The Au cluster is an effective catalyst, marking a significant improvement in the stability of the adsorbed substrates, as observed in the Au–Pd cluster.
(2)
The promotional roles of Pd in the Au-Pd cluster involve the enhancement in the electron distribution to the next-nearest-neighbor Au atom, which facilitates the activation of O2 and substantially affects the stability of the adsorption complex and transition states, yielding low-energy barriers compared with the Au cluster.
(3)
The reaction preferentially proceeds via deprotonation of the O–H bond of benzyl alcohol followed by C–Hβ bond dissociation, producing benzaldehyde and H2O. Another possible reaction pathway for H2O2 formation can also be responsible for the reasonable catalytic activity of this reaction because of the low energy barrier.
(4)
The calculated energy profiles show that the activation of the O–H bond of benzyl alcohol is significantly facilitated by the presence of active oxygen species bound to the catalyst clusters (superoxide-like and chemisorbed atomic oxygen species), whereas the effect is reduced by the barrier for C–Hβ bond dissociation that occurs on both Au and AuPd clusters. Thus, the latter step is considered to be the rate-determining step, which is also in agreement with experimental observations.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11060720/s1, Figure S1: Optimized structures of Au8, Figure S2: Optimized structures of Au6Pd2, Figure S3: Initial positions of benzyl alcohol, intermediates and products on Au8 and Au6Pd2, Figure S4–S7: Optimized structures for the complexes between benzyl alcohol and O2 on Au8 and Au6Pd2 clusters, Table S1: The relative energy of the Au8 isomers, Table S2: The relative energy of the Au6Pd2 isomers, Table S3–S7: Tables of structural parameters and Cartesian coordinates of all the species in this work.

Author Contributions

Conceptualization, K.B.; Formal analysis, B.U.; Funding acquisition, K.B.; Investigation, J.K. and C.C.; Methodology, M.E. and K.B.; Project administration, K.B.; Resources, M.E.; Software, M.E.; Supervision, M.E.; Validation, P.K., M.E. and K.B.; Writing—original draft, K.B.; Writing—review & editing, M.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a Research and Researcher for industry (RRi) grant of the Thailand Research Fund (TRF; grant number MSD60I0097 to J.K.). K.B. gratefully acknowledges support from The Thailand Research Fund (grant number MRG5980142) and Commission on Higher Education, Ministry of Education, THAILAND. The authors acknowledge the financial support from the Nanotechnology Platform Program (Molecule and Material Synthesis) of the Ministry of Education, Culture, Sports, Science, and Technology of Japan (grant number S-21-MS-0022). The APC was funded by Institute for Molecular Science, Okazaki, Japan.

Acknowledgments

The computations were partially performed in the Research Center for Computational Science, Okazaki, Japan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, M.; Li, C.; Fang, J. Noble-Metal Based Random Alloy and Intermetallic Nanocrystals: Syntheses and Applications. Chem. Rev. 2021, 121, 736–795. [Google Scholar] [CrossRef]
  2. Zhang, L.; Xie, Z.; Gong, J. Shape-controlled synthesis of Au–Pd bimetallic nanocrystals for catalytic applications. Chem. Soc. Rev. 2016, 45, 3916–3934. [Google Scholar] [CrossRef] [PubMed]
  3. Ma, T.; Liang, F. Au−Pd Nanostars with Low Pd Content: Controllable Preparation and Remarkable Performance in Catalysis. J. Phys. Chem. C 2020, 124, 7812–7822. [Google Scholar] [CrossRef]
  4. Sankar, M.; Dimitratos, N.; Miedziak, P.J.; Wells, P.P.; Kiely, C.J.; Hutchings, G.J. Designing bimetallic catalysts for a green and sustainable future. Chem. Soc. Rev. 2012, 41, 8099–8139. [Google Scholar] [CrossRef]
  5. Hutchings, G.J. Nanocrystalline gold and gold palladium alloy catalysts for chemical synthesis. Chem. Commun. 2008, 10, 1148–1164. [Google Scholar] [CrossRef] [PubMed]
  6. Kaiser, J.; Leppert, L.; Welz, H.; Polzer, F.; Wunder, S.; Wanderka, N.; Albrecht, M.; Lunkenbein, T.; Breu, J.; Kummel, S.; et al. Catalytic activity of nanoalloys from gold and palladium. Phys. Chem. Chem. Phys. 2012, 14, 6487–6495. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, J.; Jin, R. Heterogeneous catalysis by gold and gold-based bimetal nanoclusters. Nano Today 2018, 18, 86–102. [Google Scholar] [CrossRef]
  8. Edwards, J.K.; Freakley, S.J.; Carley, A.F.; Kiely, C.J.; Hutchings, G.J. Strategies for Designing Supported Gold–Palladium Bimetallic Catalysts for the Direct Synthesis of Hydrogen Peroxide. Acc. Chem. Res. 2014, 47, 845–854. [Google Scholar] [CrossRef] [PubMed]
  9. Freakley, S.J.; Piccinini, M.; Edwards, J.K.; Ntainjua, E.N.; Moulijn, J.A.; Hutchings, G.J. Effect of Reaction Conditions on the Direct Synthesis of Hydrogen Peroxide with a AuPd/TiO2 Catalyst in a Flow Reactor. ACS Catal. 2013, 3, 487–501. [Google Scholar] [CrossRef]
  10. Edwards, J.K.; Pritchard, J.; Lu, L.; Piccinini, M.; Shaw, G.; Carley, A.F.; Morgan, D.J.; Kiely, C.J.; Hutchings, G.J. The Direct Synthesis of Hydrogen Peroxide Using Platinum-Promoted Gold–Palladium Catalysts. Angew. Chem. Int. Ed. 2014, 53, 2381–2384. [Google Scholar] [CrossRef] [PubMed]
  11. Ricciardulli, T.; Gorthy, S.; Adams, J.S.; Thompson, C.; Karim, A.M.; Neurock, M.; Flaherty, D.W. Effect of Pd Coordination and Isolation on the Catalytic Reduction of O2 to H2O2 over PdAu Bimetallic Nanoparticles. J. Am. Chem. Soc. 2021, 143, 5445–5464. [Google Scholar] [CrossRef]
  12. Neurock, M.; Tysoe, W.T. Mechanistic Insights in the Catalytic Synthesis of Vinyl Acetate on Palladium and Gold/Palladium Alloy Surfaces. Top. Catal. 2013, 56, 1314–1332. [Google Scholar] [CrossRef]
  13. Calaza, F.; Li, Z.; Garvey, M.; Neurock, M.; Tysoe, W.T. Reactivity and Selectivity in the Au/Pd(111) Alloy-Catalyzed Vinyl Acetate Synthesis. Catal. Lett. 2013, 143, 756–762. [Google Scholar] [CrossRef]
  14. Alshammari, H. Solvent-free aerobic oxidation of 1-octene using graphite-supported gold–palladium nanoparticle catalysts. Reac. Kinet. Mech. Cat. 2016, 119, 149–163. [Google Scholar] [CrossRef]
  15. Bawaked, S.; He, Q.; Dummer, N.F.; Carley, A.F.; Knight, D.W.; Bethell, D.; Kiely, C.J.; Hutchings, G.J. Selective oxidation of alkenes using graphite-supported gold-palladiumcatalysts. Catal. Sci. Technol. 2011, 1, 747–759. [Google Scholar] [CrossRef]
  16. Cattaneo, S.; Freakley, S.J.; Morgan, D.J.; Sankar, M.; Dimitratos, N.; Hutchings, G.J. Cinnamaldehyde hydrogenation using Au–Pd catalysts prepared by sol immobilisatio. Catal. Sci. Technol. 2018, 8, 1677–1685. [Google Scholar] [CrossRef]
  17. Liu, J.; Shan, J.; Lucci, F.R.; Cao, S.; Sykes, E.C.H.; Flytzani-Stephanopoulos, M. Palladium–gold single atom alloy catalysts for liquid phase selective hydrogenation of 1-hexyne. Catal. Sci. Technol. 2017, 7, 4276–4284. [Google Scholar] [CrossRef]
  18. Luneau, M.; Shirman, T.; Filie, A.; Timoshenko, J.; Chen, W.; Trimpalis, A.; Flytzani-Stephanopoulos, M.; Kaxiras, E.; Frenkel, A.I.; Aizenberg, J.; et al. Dilute Pd/Au Alloy Nanoparticles Embedded in Colloid-Templated Porous SiO2: Stable Au-Based Oxidation Catalysts. Chem. Mater. 2019, 31, 5759–5768. [Google Scholar] [CrossRef]
  19. Han, S.; Mullins, C.B. Catalytic Reactions on Pd–Au Bimetallic Model Catalysts. Acc. Chem. Res. 2021, 54, 379–387. [Google Scholar] [CrossRef] [PubMed]
  20. Dhital, R.N.; Kamonsatikul, C.; Somsook, E.; Bobuatong, K.; Ehara, M.; Karanjit, S.; Sakurai, H. Low-Temperature Carbon-Chlorine Bond Activation by Bimetallic Gold/Palladium Alloy Nanoclusters: An Application to Ullmann Coupling. J. Am. Chem. Soc. 2012, 134, 20250–20253. [Google Scholar] [CrossRef] [PubMed]
  21. Xiao, Q.; Sarina, S.; Jaatinen, E.; Jia, J.; Arnold, D.P.; Liu, H.; Zhu, H. Efficient photocatalytic Suzuki cross-coupling reactions on Au–Pd alloy nanoparticles under visible light irradiation. Green Chem. 2014, 16, 4272–4285. [Google Scholar] [CrossRef]
  22. Kadu, B.S. Suzuki–Miyaura cross coupling reaction: Recent advancements in catalysis and organic synthesis. Catal. Sci. Technol. 2021, 11, 1186–1221. [Google Scholar] [CrossRef]
  23. Meenakshisundaram, S.; Nowicka, E.; Miedziak, P.J.; Brett, G.L.; Jenkins, R.L.; Dimitratos, N.; Taylor, S.H.; Knight, D.W.; Bethell, D.; Hutchings, G.J. Oxidation of alcohols using supported gold and gold–palladium nanoparticles. Faraday Discuss. 2010, 145, 341–356. [Google Scholar] [CrossRef]
  24. Zhu, X.; Guo, Q.; Sun, Y.; Chen, S.; Wang, J.-Q.; Wu, M.; Fu, W.; Tang, Y.; Duan, X.; Chen, D.; et al. Optimising surface d charge of AuPd nanoalloy catalysts for enhanced catalytic activity. Nat. Commun. 2019, 10, 1428–1438. [Google Scholar] [CrossRef] [Green Version]
  25. McClure, J.P.; Boltersdorf, J.; Baker, D.R.; Farinha, T.G.; Dzuricky, N.; Villegas, C.E.P.; Rocha, A.R.; Leite, M.S. Structure–Property-Performance Relationship of Ultrathin Pd–Au Alloy Catalyst Layers for Low-Temperature Ethanol Oxidation in Alkaline Media. ACS Appl. Mater. Interfaces 2019, 11, 24919–24932. [Google Scholar] [CrossRef]
  26. Enache, D.I.; Edwards, J.K.; Landon, P.; Solsona-Espriu, B.; Carley, A.F.; Herzing, A.A.; Watanabe, M.; Kiely, C.J.; Knight, D.W.; Hutchings, G.J. Solvent-Free Oxidation of Primary Alcohols to Aldehydes Using Au-Pd/TiO2 Catalysts. Science 2006, 311, 362–365. [Google Scholar] [CrossRef]
  27. Evangelisti, C.; Schiavi, E.; Aronica, L.A.; Caporusso, A.M.; Vitulli, G.; Bertinetti, L.; Martra, G.; Balerna, A.; Mobilio, S. Bimetallic Gold–Palladium vapour derived catalysts: The role of structural features on their catalytic activity. J. Catal. 2012, 286, 224–236. [Google Scholar] [CrossRef]
  28. Li, X.; Feng, J.; Perdjon, M.; Oh, R.; Zhao, W.; Huang, X.; Liu, S. Investigations of supported Au-Pd nanoparticles on synthesized CeO2 with different morphologies and application in solvent-free benzyl alcohol oxidation. Appl. Surf. Sci. 2020, 505, 144473. [Google Scholar] [CrossRef]
  29. Tareq, S.; Yap, Y.H.T.; Saleh, T.A.; Abdullah, A.H.; Rashid, U.; Saiman, M.I. Synthesis of bimetallic gold-pallidum loaded on carbon as efficient catalysts for the oxidation of benzyl alcohol into benzaldehyde. J. Mol. Liq. 2018, 271, 885–891. [Google Scholar] [CrossRef]
  30. Wang, Z.; Song, Y.; Zou, J.; Li, L.; Yu, Y.; Wu, L. The cooperation effect in the Au-Pd/LDH for promoting photocatalytic selective oxidation of benzyl alcohol. Catal. Sci. Technol. 2018, 8, 268–275. [Google Scholar] [CrossRef]
  31. Cui, W.; Xiao, Q.; Sarina, S.; Ao, W.; Xie, M.; Zhu, H.; Bao, Z. Au–Pd alloy nanoparticle catalyzed selective oxidation of benzyl alcohol and tandem synthesis of imines at ambient conditions. Catal. Today 2014, 235, 152–159. [Google Scholar] [CrossRef]
  32. Zhang, W.; Xiao, Z.; Wang, J.; Fu, W.; Tan, R.; Yin, D. Selective Aerobic Oxidation of Alcohols over Gold-Palladium Alloy Catalysts Using Air at Atmospheric Pressure in Water. ChemCatChem 2019, 11, 1779–1788. [Google Scholar] [CrossRef]
  33. Lee, A.F.; Ellis, C.V.; Wilson, K.; Hondow, N.S. In situ studies of titania-supported Au shell–Pd core nanoparticles for the selective aerobic oxidation of crotyl alcohol. Catal. Today 2010, 157, 243–249. [Google Scholar] [CrossRef]
  34. Ksar, F.A.; Ramos, L.; Keita, B.; Nadjo, L.; Beaunier, P.; Remita, H. Bimetallic Palladium−Gold Nanostructures: Application in Ethanol Oxidation. Chem. Mater. 2009, 21, 3677–3683. [Google Scholar] [CrossRef]
  35. Brett, G.L.; He, Q.; Hammond, C.; Miedziak, P.J.; Dimitratos, N.; Sankar, M.; Herzing, A.A.; Conte, M.; Lopez-Sanchez, J.A.; Kiely, C.J.; et al. Selective Oxidation of Glycerol by Highly Active Bimetallic Catalysts at Ambient Temperature under Base-Free Conditions. Angew. Chem. Int. Ed. 2011, 50, 10136–10139. [Google Scholar] [CrossRef]
  36. Jiang, T.; Jia, C.; Zhang, L.; He, S.; Sang, Y.; Li, H.; Li, Y.; Xu, X.; Liu, H. Gold and gold–palladium alloy nanoparticles on heterostructured TiO2 nanobelts as plasmonic photocatalysts for benzyl alcohol oxidation. Nanoscale 2015, 7, 209–217. [Google Scholar] [CrossRef]
  37. Dhital, R.N.; Sakurai, H. Gold– and gold–palladium/poly(1-vinylpyrrolidin-2-one) nanoclusters as quasi-homogeneous catalyst for aerobic oxidation of glycerol. Tetrahedron Lett. 2011, 52, 2633–2637. [Google Scholar]
  38. Alshammari, H.M.; Alshammari, A.S.; Humaidi, J.R.; Alzahrani, S.A.; Alhumaimess, M.S.; Aldosari, O.F.; Hassan, H.M.A. Au-Pd Bimetallic Nanocatalysts Incorporated into Carbon Nanotubes (CNTs) for Selective Oxidation of Alkenes and Alcohol. Processes 2020, 8, 1380–1390. [Google Scholar] [CrossRef]
  39. Tsai, C.-H.; Xu, M.; Kunal, P.; Trewyn, B.G. Aerobic oxidative esterification of primary alcohols over Pd-Au bimetallic catalysts supported on mesoporous silica nanoparticles. Catal. Today 2018, 306, 81–88. [Google Scholar] [CrossRef]
  40. Lee, D.G.; Spitzer, U.A. Aqueous dichromate oxidation of primary alcohols. J. Org. Chem. 1970, 35, 3589–3590. [Google Scholar] [CrossRef]
  41. Menger, F.M.; Lee, C. Oxidations with solid potassium permanganate. Tetrahedron Lett. 1981, 22, 1655–1656. [Google Scholar] [CrossRef]
  42. Keresszegi, C.; Bürgi, T.; Mallat, T.; Baiker, A. On the Role of Oxygen in the Liquid-Phase Aerobic Oxidation of Alcohols on Palladium. J. Catal. 2002, 211, 244–251. [Google Scholar] [CrossRef]
  43. Tsunoyama, H.; Sakurai, H.; Negishi, Y.; Tsukuda, T. Size-Specific Catalytic Activity of Polymer-Stabilized Gold Nanoclusters for Aerobic Alcohol Oxidation in Water. J. Am. Chem. Soc. 2005, 127, 9374–9375. [Google Scholar] [CrossRef]
  44. Corma, A.; Garcia, H. Supported gold nanoparticles as catalysts for organic reactions. Chem. Soc. Rev. 2008, 37, 2096–2126. [Google Scholar] [CrossRef]
  45. Pyykkö, P. Theoretical Chemistry of Gold. Angew. Chem. Int. Ed. 2004, 43, 4412–4456. [Google Scholar] [CrossRef]
  46. Pyykko, P. Theoretical chemistry of gold. III. Chem. Soc. Rev. 2008, 37, 1967–1997. [Google Scholar] [CrossRef]
  47. Coquet, R.; Howard, K.L.; Willock, D.J. Theory and simulation in heterogeneous gold catalysis. Chem. Soc. Rev. 2008, 37, 2046–2076. [Google Scholar] [CrossRef]
  48. Hammer, B.; Nørskov, J.K. Theoretical surface science and catalysis—Calculations and concepts. In Advances in Catalysis; Bruce, C., Gates, H.K., Eds.; Academic Press: Cambridge, MA, USA, 2000; Volume 45, pp. 71–129. [Google Scholar]
  49. Norskov, J.K.; Bligaard, T.; Rossmeisl, J.; Christensen, C.H. Towards the computational design of solid catalysts. Nat. Chem. 2009, 1, 37–46. [Google Scholar] [CrossRef]
  50. Li, J.; Ishihara, T.; Yoshizawa, K. Theoretical Revisit of the Direct Synthesis of H2O2 on Pd and Au@Pd Surfaces: A Comprehensive Mechanistic Study. J. Phys. Chem. C 2011, 115, 25359–25367. [Google Scholar] [CrossRef]
  51. Zhang, J.; Jin, H.; Sullivan, M.B.; Lim, F.C.H.; Wu, P. Study of Pd–Au bimetallic catalysts for CO oxidation reaction by DFT calculations. Phys. Chem. Chem. Phys. 2009, 11, 1441–1446. [Google Scholar] [CrossRef]
  52. Atoa, Y.; Hayashi, A.; Koga, H.; Kawakami, T.; Yamanaka, S.; Okumura, M. Exploring a near-Hartree–Fock–Kohn–Sham approach to study electronic properties of azobenzene in interaction with gold: From clusters to the Au(111) surface. Chem. Phys. Lett. 2019, 724, 15–121. [Google Scholar]
  53. Miyazaki, R.; Jin, X.; Yoshii, D.; Yatabe, T.; Yabe, T.; Mizuno, N.; Yamaguchi, K.; Hasegawa, J.Y. Mechanistic study of C–H bond activation by O2 on negatively charged Au clusters: α,β-dehydrogenation of 1-methyl-4-piperidone by supported Au catalysts. Catal. Sci. Technol. 2021, 11, 3333–3346. [Google Scholar] [CrossRef]
  54. Varganov, S.A.; Olson, R.M.; Gordon, M.S.; Metiu, H. The interaction of oxygen with small gold clusters. J. Chem. Phys. 2003, 119, 2531–2537. [Google Scholar] [CrossRef] [Green Version]
  55. Madsen, G.K.H.; Hammer, B. Effect of subsurface Ti-interstitials on the bonding of small gold clusters on rutile TiO2(110). J. Chem. Phys. 2009, 130, 044704. [Google Scholar] [CrossRef] [PubMed]
  56. Hansen, J.; Ehara, M.; Piecuch, P. Aerobic oxidation of methanol to formic acid on Au8: Benchmark analysis based on completely renormalized coupled-cluster and density functional theory calculations. J. Phys. Chem. A 2013, 117, 10416–10427. [Google Scholar] [CrossRef]
  57. Zhao, Y.; Truhlar, D.G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101. [Google Scholar] [CrossRef] [Green Version]
  58. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  59. Mantina, M.; Valero, R.; Truhlar, D.G. Validation study of the ability of density functionals to predict the planar-to-three-dimensional structural transition in anionic gold clusters. J. Chem. Phys. 2009, 131, 064705–064706. [Google Scholar] [CrossRef] [PubMed]
  60. Ferrighi, L.; Hammer, B.; Madsen, G.K.H. 2D−3D transition for cationic and anionic gold clusters: A kinetic energy density functional study. J. Am. Chem. Soc. 2009, 131, 10605–10609. [Google Scholar] [CrossRef]
  61. Bobuatong, K.; Karanjit, S.; Fukuda, R.; Ehara, M.; Sakurai, H. Aerobic oxidation of methanol to formic acid on Au20: A theoretical study on the reaction mechanism. Phys. Chem. Chem. Phys. 2012, 14, 3103–3111. [Google Scholar] [CrossRef]
  62. Bobuatong, K.; Sakurai, H.; Ehara, M. Intramolecular hydroamination by a primary amine of an unactivated alkene on gold nanoclusters: A DFT study. ChemCatChem 2017, 9, 4490–4500. [Google Scholar]
  63. Ehara, M.; Priyakumar, U.D. Gold-Palladium nanocluster catalysts for homocoupling: Electronic structure and interface dynamics. Chem. Rec. 2019, 19, 947–959. [Google Scholar] [CrossRef]
  64. Karanjit, S.; Bobuatong, K.; Fukuda, R.; Ehara, M.; Sakurai, H. Mechanism of the aerobic oxidation of methanol to formic acid on Au8: A DFT study. Int. J. Quantum Chem. 2012, 113, 428–436. [Google Scholar] [CrossRef]
  65. Oliver-Meseguer, J.; Cabrero-Antonino, J.R.; Irene Domínguez, I.; Leyva-Pérez, A.; Corma, A. Small Gold Clusters Formed in Solution Give Reaction Turnover Numbers of 107 at Room Temperature. Science 2012, 338, 1452–1455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Xu, J.; Xu, S.; Chen, M.; Zhu1, Y. Unlocking the catalytic activity of an eight-atom gold cluster with a Pd atom. Nanoscale 2020, 12, 6020–6028. [Google Scholar] [CrossRef]
  67. Cordón, J.; Jiménez-Osés, G.; López-de-Luzuriaga, J.M.; Monge, M. The key role of Au-substrate interactions in catalytic gold subnanoclusters. Nat. Commun. 2017, 8, 1657–1664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Oliver-Meseguer, J.; Leyva-Pérez, A.; Corma, A. Very Small (3–6 Atoms) Gold Cluster Catalyzed Carbon–Carbon and Carbon–Heteroatom Bond-Forming Reactions in Solution. ChemCatChem 2013, 5, 3509–3515. [Google Scholar] [CrossRef]
  69. Landman, U.; Yoon, B.; Zhang, C.; Heiz, U.; Arenz, M. Factors in gold nanocatalysis: Oxidation of CO in the non-scalable size regime. Top. Catal. 2007, 44, 145–158. [Google Scholar] [CrossRef] [Green Version]
  70. Yang, M.; Zhou, M.; Zhang, A.; Zhang, C. Graphene Oxide: An Ideal Support for Gold Nanocatalysts. J. Phys. Chem. C 2012, 116, 22336–22340. [Google Scholar] [CrossRef]
  71. Villa, A.; Janjic, N.; Spontoni, P.; Wang, D.; Su, D.S.; Prati, L. Au–Pd/AC as catalysts for alcohol oxidation: Effect of reaction parameters on catalytic activity and selectivity. Appl. Catal. A Gen. 2009, 364, 221–228. [Google Scholar] [CrossRef]
  72. Vijay, A.; Mills, G.; Metiu, H. Adsorption of gold on stoichiometric and reduced rutile TiO2 (110) surfaces. J. Chem. Phys. 2003, 118, 6536–6551. [Google Scholar] [CrossRef]
  73. Diefenbach, M.; Kim, K.S. Spatial structure of Au8: Importance of basis set completeness and geometry relaxation. J. Phys. Chem. B 2006, 110, 21639–21642. [Google Scholar] [CrossRef] [PubMed]
  74. Huber, K.P.; Herzberg, G. Molecular Spectra and Molecular Structure; Van Nostrand Reinhold: New York, NY, USA, 1979. [Google Scholar]
  75. Travers, M.J.; Cowles, D.C.; Ellison, G.B. Reinvestigation of the electron affinities of O2 and NO. Chem. Phys. Lett. 1989, 164, 449–455. [Google Scholar] [CrossRef]
  76. Tsunoyama, H.; Ichikuni, N.; Sakurai, H.; Tsukuda, T. Effect of electronic structures of Au clusters stabilized by Poly(N-vinyl-2-pyrrolidone) on aerobic oxidation catalysis. J. Am. Chem. Soc. 2009, 131, 7086–7093. [Google Scholar] [CrossRef]
  77. Tsukuda, T.; Tsunoyama, H.; Sakurai, H. Aerobic oxidations catalyzed by colloidal nanogold. Chem. Asian J. 2011, 6, 736–748. [Google Scholar] [CrossRef] [PubMed]
  78. Pina, C.D.; Falletta, E.; Prati, L.; Rossi, M. Selective oxidation using gold. Chem. Soc. Rev. 2008, 37, 2077–2095. [Google Scholar] [CrossRef]
  79. Zope, B.N.; Hibbitts, D.D.; Neurock, M.; Davis, R.J. Reactivity of the Gold/Water Interface During Selective Oxidation Catalysis. Science 2010, 330, 74–78. [Google Scholar] [CrossRef] [Green Version]
  80. Boronat, M.; Corma, A. Molecular approaches to catalysis Naked gold nanoparticles as quasi-molecular catalysts for green processes. J. Catal. 2011, 284, 138–147. [Google Scholar] [CrossRef]
  81. Frisch, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision A. 02; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  82. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  83. Feller, D. The role of databases in support of computational chemistry calculations. J. Comp. Chem. 1996, 17, 1571–1586. [Google Scholar] [CrossRef]
  84. Schuchardt, K.L.; Didier, B.T.; Elsethagen, T.; Sun, L.; Gurumoorthi, V.; Chase, J.; Li, J.; Windus, T.L. Basis Set Exchange:  A Community database for computational sciences. J. Chem. Inf. Model 2007, 47, 1045–1052. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The lowest-energy structure of Au8 and Au6Pd2 with Mulliken atomic charges, and NBO analysis.
Figure 1. The lowest-energy structure of Au8 and Au6Pd2 with Mulliken atomic charges, and NBO analysis.
Catalysts 11 00720 g001
Figure 2. (a) HOMO and LUMO of O2, Au8 and Au6Pd2 (b) Frontier orbitals of the adsorption complex of Au8 and Au6Pd2 with O2 (the contour value is 0.02 electron/Å3).
Figure 2. (a) HOMO and LUMO of O2, Au8 and Au6Pd2 (b) Frontier orbitals of the adsorption complex of Au8 and Au6Pd2 with O2 (the contour value is 0.02 electron/Å3).
Catalysts 11 00720 g002
Figure 3. Catalytic cycle of the aerobic oxidation of benzyl alcohol to benzaldehyde over Au8 and Au6Pd2.
Figure 3. Catalytic cycle of the aerobic oxidation of benzyl alcohol to benzaldehyde over Au8 and Au6Pd2.
Catalysts 11 00720 g003
Figure 4. Frontier orbitals of the adsorption complex of Au8 and Au6Pd2 with benzyl alcohol and O2.
Figure 4. Frontier orbitals of the adsorption complex of Au8 and Au6Pd2 with benzyl alcohol and O2.
Catalysts 11 00720 g004
Figure 5. Calculated energy profile and optimized structures for the aerobic oxidation of benzyl alcohol over Au8 and Au6Pd2 generating benzaldehyde and H2O. All energy values are shown in kcal/mol and the distances are given in Å.
Figure 5. Calculated energy profile and optimized structures for the aerobic oxidation of benzyl alcohol over Au8 and Au6Pd2 generating benzaldehyde and H2O. All energy values are shown in kcal/mol and the distances are given in Å.
Catalysts 11 00720 g005
Figure 6. Calculated energy profile and optimized structures for the aerobic oxidation of benzyl alcohol over Au8 and Au6Pd2 generating benzaldehyde and H2O (cont’d). All energy values are shown in kcal/mol and the distances are given in Å.
Figure 6. Calculated energy profile and optimized structures for the aerobic oxidation of benzyl alcohol over Au8 and Au6Pd2 generating benzaldehyde and H2O (cont’d). All energy values are shown in kcal/mol and the distances are given in Å.
Catalysts 11 00720 g006
Figure 7. Calculated energy profile and optimized structures for the aerobic oxidation of second benzyl alcohol over Au8-O and Au6Pd2-O species generating benzaldehyde and H2O2. All energy values are shown in kcal/mol and the distances are given in Å.
Figure 7. Calculated energy profile and optimized structures for the aerobic oxidation of second benzyl alcohol over Au8-O and Au6Pd2-O species generating benzaldehyde and H2O2. All energy values are shown in kcal/mol and the distances are given in Å.
Catalysts 11 00720 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kodchasee, J.; Chanloi, C.; Khemthong, P.; Uapipatanakul, B.; Ehara, M.; Bobuatong, K. Catalytic Oxidation of Benzyl Alcohol to Benzaldehyde on Au8 and Au6Pd2 Clusters: A DFT Study on the Reaction Mechanism. Catalysts 2021, 11, 720. https://doi.org/10.3390/catal11060720

AMA Style

Kodchasee J, Chanloi C, Khemthong P, Uapipatanakul B, Ehara M, Bobuatong K. Catalytic Oxidation of Benzyl Alcohol to Benzaldehyde on Au8 and Au6Pd2 Clusters: A DFT Study on the Reaction Mechanism. Catalysts. 2021; 11(6):720. https://doi.org/10.3390/catal11060720

Chicago/Turabian Style

Kodchasee, Jitlada, Chanon Chanloi, Pongtanawat Khemthong, Boontida Uapipatanakul, Masahiro Ehara, and Karan Bobuatong. 2021. "Catalytic Oxidation of Benzyl Alcohol to Benzaldehyde on Au8 and Au6Pd2 Clusters: A DFT Study on the Reaction Mechanism" Catalysts 11, no. 6: 720. https://doi.org/10.3390/catal11060720

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