Next Article in Journal
Straightforward Design for Phenoxy-Imine Catalytic Activity in Ethylene Polymerization: Theoretical Prediction
Previous Article in Journal
Titanate Nanotube-Supported Au–Rh Bimetallic Catalysts: Characterization and Their Catalytic Performances in Hydroformylation of Vinyl Acetate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Theoretical Studies on the Direct Propylene Epoxidation Using Gold-Based Catalysts: A Mini-Review

1
Department of Chemical and Biological Engineering, The University of Alabama, Tuscaloosa, AL 35487, USA
2
Department of Chemical and Materials Engineering, The University of Alabama in Huntsville, Huntsville, AL 35899, USA
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(10), 421; https://doi.org/10.3390/catal8100421
Submission received: 3 September 2018 / Revised: 22 September 2018 / Accepted: 25 September 2018 / Published: 27 September 2018
(This article belongs to the Section Computational Catalysis)

Abstract

:
Direct propylene epoxidation using Au-based catalysts is an important gas-phase reaction and is clearly a promising route for the future industrial production of propylene oxide (PO). For instance, gold nanoparticles or clusters that consist of a small number of atoms demonstrate unique and even unexpected properties, since the high ratio of surface to bulk atoms can provide new reaction pathways with lower activation barriers. Support materials can have a remarkable effect on Au nanoparticles or clusters due to charge transfer. Moreover, Au (or Au-based alloy, such as Au–Pd) can be loaded on supports to form active interfacial sites (or multiple interfaces). Model studies are needed to help probe the underlying mechanistic aspects and identify key factors controlling the activity and selectivity. The current theoretical/computational progress on this system is reviewed with respect to the molecular- and catalyst-level aspects (e.g., first-principles calculations and kinetic modeling) of propylene epoxidation over Au-based catalysts. This includes an analysis of H2 and O2 adsorption, H2O2 (OOH) species formation, epoxidation of propylene into PO, as well as possible byproduct formation. These studies have provided a better understanding of the nature of the active centers and the dominant reaction mechanisms, and thus, could potentially be used to design novel catalysts with improved efficiency.

1. Introduction

Due to the growing environmental issues in chemical synthesis and processing, there has been increasing interest in the development of new processes for minimizing pollution and reducing energy consumption. Heterogeneous catalysis plays a crucial role in this respect, since it is widely used to reduce emissions from automobiles, reduce byproducts, and improve the selectivity of important chemical products in the petrochemical as well as other chemical industries [1,2,3,4,5]. Computational and experimental investigators have been working together to better understand fundamental characteristics of a wide range of catalytic reactions. In this paper, we focus on the selective epoxidation of propylene to yield propylene oxide (PO) using heterogeneous gold-based catalysts. The catalytic activity of gold and gold alloy nanoparticles (a few nm in size) has been ascribed to various mechanisms, involving reactions occurring at neutral gold atoms that differ from bulk gold atoms (due to different degrees of coordination), quantum size effects that change the electronic structures of nanoparticles, or charge modification of gold atoms through the interaction with oxide supports [6,7,8].
Computational studies have been enormously valuable in describing the structural and electronic character of size-selected gold nanoparticles and determining the catalytic reaction mechanisms for the propylene epoxidation process, and have already yielded important outcomes for the development of catalytic materials. Motivated by these growing computational investigations and the need to incorporate this theoretical knowledge into catalyst design, we review the theoretical studies relevant to direct propylene epoxidation in the gas phase with Au-based catalysts. In the following sections, we provide a brief background of propylene epoxidation, and we then highlight different aspects of the propylene epoxidation process, including H2/O2 adsorption, H2O2 synthesis, and possible epoxidation mechanisms. Finally, conclusions and perspectives are given, along with current challenges and opportunities in the field.

2. Background

Propylene oxide (PO) is a key intermediate in the chemical industry. For instance, PO is mainly used to produce polyether polyols (65%), as well as propene glycol (30%) and propene glycol ethers (4%) (the second and third largest applications, respectively), which are mainly applied to manufacture commercial products such as adhesives, solvents, and foams [9,10,11]. The annual worldwide production of PO in 2013 amounted to 8.06 million tons and will likely go beyond 9.56 million tons in 2018, and this market is annually growing by ~4–5% [11]. The currently available processes are not optimal, since the chlorohydrin process to produce PO suffers from environmental concerns, and the hydroperoxide process (the halcon process) is challenged by poor economics [9,11]. For instance, in the chlorohydrin process, a large amount of byproduct CaCl2 (2.2 tons per 1.0 ton of PO) is coproduced, along with toxic chlorinated organic compounds (several hundred grams per ton of PO). One problem of the hydroperoxide process is the production of a coproduct (either styrene or tert-butyl alcohol, depending on which variant of the hydroperoxide process is applied) in a fixed amount, leading to a mismatch of market demand between PO and co-products.
Compared to the two methods aforementioned, using hydrogen peroxide (H2O2) as the oxidant allows for a high selectivity (95%) for PO production, and it is more environmentally friendly since water is the only byproduct. The major disadvantage for the commercialization of this process is that hydrogen peroxide and PO have comparable market values on a molar basis, which makes it impossible to run the process profitably at this time. Moreover, it requires multiple reaction steps in the liquid phase for purification, which consumes huge amounts of energy. The in situ production of the hydrogen peroxide is currently under construction by Dow-BASF, to solve the high cost of hydrogen peroxide for propylene epoxidation [12].
Another route toward PO production is the direct propylene (C3H6) epoxidation using gold/titania catalyst in the presence of a mixture of hydrogen and oxygen. For a long time, coinage metal gold was regarded as a catalytically inert material. This common assumption was dramatically shifted by the discovery of low-temperature CO oxidation with metal oxide supported gold nanoparticles [13]. Approximately 20 years ago, Haruta and co-workers discovered that the direct vapor-phase propylene epoxidation with molecular H2 and O2 catalyzed by nano-sized gold deposited on TiO2 (anatase) was a viable route [14]. The largest advantage of gold-on-titania catalysts is that they are able to epoxidize propylene very selectively (>90%) even under mild conditions (typically 323 K and 1 bar), but the activity is relatively low.
Since that time, considerable efforts have been dedicated to enhancing Au-based catalysts, as well as understanding the reaction features of this novel system. For instance, gold particles which are hemispherical in shape are highly active, while spherical nanoparticles possess fairly weak activity [15]. Estimates of the optimal gold nanoparticle size have reduced from 2~5 nm in diameter to sizes smaller than 2 nm [11,14,16,17,18]. Haruta’s group initially started working with Au nanoparticles supported on TiO2, but they gradually transitioned to mesoporous titanosilicates (Ti-SiO2) as the supports [19,20]. Delgass’s group has focused on microporous titanosilicate (e.g., TS-1) supports, which demonstrate better results, even with low Au loadings [21,22]. The fact that the PO formation requires the presence of H2 and O2, and the fact that C3H6 can be epoxidized by H2O2 over titania, generates the common speculation that a peroxide species formed on gold is involved in the propylene epoxidation reaction. In other words, it is proposed that the active oxygen species, hydrogen peroxide, is readily synthesized on nano-sized Au via H2/O2 mixtures, and it subsequently transfers to a neighboring Ti site to epoxidize propylene into PO. During this process, the typically low propylene conversion (<2%) and low hydrogen utilization efficiency (<30%) still need to be improved, despite the fact that the selectivity toward PO is high.
Attempts to improve propylene conversion by raising the reaction temperature inevitably lead to decreasing PO selectivity due to consecutive PO reactions (e.g., isomerization, oligomerization, and oxidative cracking) [17,23,24,25,26]. For instance, by elevating the reaction temperature from 90 °C to 160 °C, there is an appreciable increase in C3H6 conversion from 2.5 to 9.8% over Au/mesoporous Ti-SiO2, while the PO selectivity decreases from 95 to 90%. The decrease in selectivity is likely due to the formation of byproducts (e.g., propanal, acetone, ethanal, CO2, dioxane, and acids) generated by secondary reactions [19,27]. The byproduct generation is proposed to be related to the dispersity of Ti sites. For example, PO combustion is significantly suppressed if Ti sites are isolated, as in TS-1 and in mesoporous Ti–SiO2 [28,29]. It is worth noting that acrolein is another significant byproduct [25]. The allylic hydrogen of propylene is acidic and labile, and easily suffers from the nucleophilic attack of the oxygen radicals bound to Au nanoparticles, since the oxygen species behave as a Brønsted base [10,30,31,32]. Thus, silver-based catalysts, used for commercial production of ethylene oxide (EO) via ethylene epoxidation, lead to propylene combustion versus epoxidation due to the activation of allylic hydrogen by surface O. On the other hand, investigations with regard to gold catalysts indicate that they may be commercially feasible for propylene oxide production [33].
To date, there have been several reviews about the experimental synthesis of PO in the gas phase, including discussions about the development and improvement of Au-based catalysts, propylene conversion, hydrogen efficiency, and PO selectivity [9,11,16,17,34]. At the same time, many theoretical studies have focused on propylene epoxidation with H2 and O2 to assist in understanding various aspects of this catalytic reaction. For instance, first-principles calculations can be employed to investigate the elementary events and underlying mechanisms, but it is usually difficult to make statements about the performance of the catalysts under different reaction conditions. In this respect, kinetic modeling studies are needed to evaluate the catalytic activity and product selectivity at larger time and length scales, particularly with several competing reaction pathways. For instance, first-principles based kinetic Monte Carlo (KMC) models can help unravel the reaction mechanism and catalyst complexities in heterogeneous catalytic systems [35,36,37,38]. These computation and simulation studies provide detailed insight on how to screen the size and composition of Au nanoparticles, as well as identify catalytic sites and reaction features for the development of commercially viable catalyst for direct PO production.

3. The Adsorption of O2 and H2 on Au

Bulk gold surfaces are usually regarded as poor catalysts, at variance with other transition metal surfaces. However, as compared to Au surfaces, small Au particles are known to be active catalysts for a variety of reactions. The adsorption of molecular O2 and H2 on small Au clusters in the gas phase is the first step of the overall propylene epoxidation reaction. Thus, this is a key step, since the adsorption energy (Eads) can be used as an indicator of activity, and it is helpful for developing a comprehensive understanding of Au catalysis.
Salisbury and co-workers investigated O2 adsorption on Au n clusters (n = 2 to 22). They found that Au n clusters with even n strongly adsorb O2 molecules, while Au n clusters with odd n cannot. Salisbury et al. supposed that the adsorbed oxygen molecule exists in the form of O 2 species due to charge transfer from the gold [39].
Okumura et al. and Yoon et al., using density functional theory (DFT), studied the interaction between O2 and Au clusters [40,41]. They found that the anionic Au atoms in gold clusters interact strongly with O2 (−12.1 kcal/mol), in contrast with neutral Au atoms (−3~8 kcal/mol), suggesting that negatively charged Au atoms are the active sites for oxygenation. As an extension, Mills et al. suggested that neutral Aun clusters with an odd n adsorb O2 more strongly in contrast with those with even n [42], which agrees well with the recent findings of Chen’s group [43]. At the same time, they mentioned that O2 does not adsorb on flat Au surfaces, indicating that the interaction between O2 and an Au surface is weak. While in the DFT studies of Mavrikakis and coworkers, they found that, although molecular O2 does not adsorb on Au(111), it does bind to a stretched Au(111) surface, as well as to both unstretched and stretched Au(211) surfaces with binding energies of −1.84, −3.46, and −6.0 kcal/mol, respectively, implying that steps and tensile strain facilitate the adsorption of O2 on Au surfaces [7,44].
There is a general agreement regarding the relationship between the increasing activity of gold nanoparticles and a decrease in particle size (leading to a higher concentration of low coordinated gold atoms), as well as particle morphology [6,45]. For instance, Wang and Gong found that an icosahedral Au32 cluster can dissociate molecular O2, but the detailed dissociation pathways and the activation energies are not described [46]. Taketsugu et al. studied O2 adsorption on small neutral gold clusters (Au3–Au12), and found that O2 adsorption induced structural transformations in gold clusters [47]. Pacchioni et al. also studied O2 adsorption and dissociation on neutral, positively, and negatively charged Aun clusters varying from 5 to 79 atoms [48,49]. It was found that the charging effect of gold clusters is prominent for very small sizes up to about 20–25 gold atoms, in which negatively charged gold clusters are beneficial for O2 activation, while positively charged clusters are not. The effect of charge disappears with the increasing gold cluster size, and Au38 seems to be the most reactive. When the cluster size increases beyond about 40 atoms, the effect of size and shape becomes a more significant factor governing the reactivity. The electronic properties beyond Au79 linearly scale to the bulk gold properties, and the catalytic activity thus gradually dies out [50].
Corma’s group, using DFT calculations, investigated the adsorption and dissociation of O2 on extended Au(111) and Au(100) surfaces, isolated Au nanoparticles with different sizes (Au13 and Au38) and shapes, and Au nanoparticles supported on TiO2 [51]. The adsorption energy of O2 was found to be correlated with gold particle size, and the particle morphology was proved to be a crucial factor for activating O2 dissociation. They found a linear relationship between the activation energy of O2 and the net charge on adsorbed O2, indicating that the degree of charge transfer determines the molecular O2 activation. Moreover, they discovered that most of the sites for O2 adsorption and dissociation are located at the Au/support interface (when gold particles are loaded on a TiO2 support). This is in good agreement with other investigations, with regard to the high catalytic activity of gold particles loaded on metal oxide supports such as CeO2, TiO2, and Fe2O3 [6,52,53,54,55,56,57], as well as metal carbides [58]. Nevertheless, there is still a debate about whether the supports are involved in the formation of charged (positively or negatively, depending on oxidized or reduced support surfaces) Au sites due to a transfer of electron density, if they are involved in the activation of reactants (e.g., O2), or if they are limited to merely stabilizing the gold particles [59,60,61].
The adsorption of O2 on alloy clusters of gold has also been investigated [62,63]. Joshi et al. analyzed the adsorption of O2 on AunMm (m, n = 0 to 3; M = Cu, Ag, Pd, Pt, and Na) clusters, in which they demonstrated that the alloy trimers containing only one Au atom are most reactive toward O2, while those with two Au atoms are least reactive [63]. The O2 binding energy (BE) follows the trend: BE (MAu2) < BE (M3) ≤ BE (M2Au). A Natural Bond Orbital (NBO) analysis indicated that all of the clusters donate electron density to the adsorbed O2. Polynskaya et al. calculated the adsorption energies of O2 on clusters (following the order: Au20 < Au19Ag < Ag19Au ≈ Ag20), and they found dissociative O2 adsorption is possible on Au19Ag, Ag19Au, and Ag20 clusters [64]. Recently, Feng et al. compared the adsorption of O2 on the surfaces of pure Au nanoparticles and Au–Ag bimetallic nanoparticles, and found that O2 adsorption on the Au–Ag alloy is stronger [65]. Their Bader charge calculations showed that the Au atoms on the surface of Au–Ag alloy nanoparticles are negatively charged (prior to O2 adsorption), implying that there is electron transfer from Ag to Au. When O2 adsorbs on the pure Au or Au–Ag alloy nanoparticles, the adsorbed O2 can accept electron charge from both nanoparticles, but the O2 withdraws more charge from the Au–Ag nanoparticle surface than from the pure Au nanoparticle surface. This suggests that the bimetallic Au–Ag nanoparticles tend to donate more electrons to the adsorbed O2, thereby facilitating O2 activation.
In contrast with relatively weak O2 adsorption on Au, H2 adsorption on Au usually follows a dissociative pathway. Gordon’s group investigated the reactions of molecular hydrogen with small gold clusters and showed that molecular H2 easily binds to neutral Au2 and Au3 clusters, and then dissociates into two hydrogen atoms [66]. Through DFT calculations, Barrio et al. studied the interaction between H2 and Au(111), Au(100), and Au clusters (Au14, Au25, and Au29). They found that the H–H bond was spontaneously elongated up to cleavage, without an apparent activation barrier, when H2 approached the Au clusters. H2 dissociation was also found to be accompanied by the deformation of clusters in order to stabilize the dissociated H atoms. This is different from H2 dissociation on the more rigid icosahedral Au13 cluster, with an activation energy of 6.93 kcal/mol [67]. Recently, studies of H2 adsorption and dissociation on gold clusters demonstrated not only that the effect of low coordinated Au atoms is important, but also that the fluxionality (the flexibility of the structure) and ensemble effects play key roles in the bonding and dissociation of H2 [68]. Simultaneously, Barrio et al. found that the flat (111) and (100) surfaces of bulk gold are not active towards H2 dissociation [69,70], which is in agreement with previous studies [71]. Flores et al. also obtained a similar conclusion that stretched gold nanowires are more catalytically active for H2 dissociation in comparison to the Au(111) surface (H2 dissociation barriers: 9.22 vs. 23.06 kcal/mol) [72].
In addition to Au surfaces and nanoparticles, H2 dissociation at the perimeter sites of Au/TiO2 has been explored by several groups. For instance, Hu et al. investigated the dissociation of molecular H2 at an Au/TiO2 interface, as well as atomic H spillover from Au nanoparticles to a TiO2 support [73]. In their studies, H2 dissociation is possible via two different modes, namely homolytic and heterolytic routes. In the homolytic pathway, H2 dissociates on the Au atoms only, while in the heterolytic one, H2 dissociation occurs on one Au atom and a nearby surface O (from TiO2). It was found that the heterolytic dissociation of molecular H2 is more favored than the homolytic dissociation (with activation barriers of 8.53 vs. 14.76 kcal/mol) at the perimeter sites of Au/TiO2. Using DFT calculations, Takeda’s group verified the transition state structure O2−–H+–H–Au of H2 dissociation at the perimeter of TiO2, as Hückel’s theory initially suggested [74]. Simultaneously, they demonstrated that H2 dissociation at the perimeter sites has the lowest activation barrier (6.23 kcal/mol) in contrast with those on Au(100), Au(321), and the TiO2 support (17.53, 14.76, and 18.68 kcal/mol, respectively). The calculated activation barrier of 6.23 kcal/mol is consistent with the theoretical result (6.69 kcal/mol) from Hu’s group aforementioned but is higher than the reported 3.69 kcal/mol from Neurock’s group [75]. Their results are lower than the experimental estimate of 8.76 kcal/mol [76].

4. The Synthesis of H2O2 from O2 and H2

Hydrogen peroxide (H2O2) is a green oxidant, and it has been listed as one of the 100 most important chemicals in the world [77]. The annual worldwide production of H2O2 in 2010 was about 3 million tons [78], in which the auto-oxidation (AO) method accounts for more than 95% of the synthesis. However, the AO process is accompanied by problems of exhaust gas emission, solid waste, and waste water [77]. Currently, most commercial H2O2 production is directed towards applications in wastewater treatment, hydroquinone production, bleaching of textile and pulp, removal of organic pollutants, etc. [79,80,81]. In recent years, commercial H2O2 has been used in chemical synthesis, such as the epoxidation of alkenes [82]. Unlike the AO process, the direct synthesis of H2O2 with H2 and O2 emerges as an atomically economic and green chemical reaction, catalyzed by noble metals, such as Pd, Pt, Au, and Ru. Since this review focuses on propylene epoxidation with Au-based catalysts, we highlight the relevant theoretical studies of H2O2 production from H2 and O2 involving gold catalysts.
Thomson’s group proposed a reaction path for H2O2 formation on a neutral Au3 cluster as shown in Figure 1 [83]. The first step is nondissociative O2 adsorption on an Au trimer with an end-on geometry (3–6 kcal/mol binding energy). As molecular H2 approaches the adsorbed O2, the H–H bond gradually elongates to breakage and one H directly bonds to one Au atom, and the other H bonds to an adsorbed O2 in the form of an –OOH (Intermediate A in Figure 1). The second H2 then inserts between the nearest Au atom and the oxygen end of the hydroperoxy group (–OOH), leading to H2O2 formation (Intermediate B in Figure 1). The generated H2O2 desorbs, followed by the exothermic addition of molecular O2, to form Intermediate D, which undergoes a structural rearrangement to regenerate Intermediate A, allowing this catalytic cycle to repeat. The H2O2 desorption has the highest activation barrier (ΔEact = 8.6 kcal/mol) during the formation of hydrogen peroxide. Likewise, Kacprzak et al. have demonstrated that small neutral gold clusters (e.g., Au2 and Au4) can catalyze H2O2 formation, and have identified possible reaction pathways for both H2O2 and H2O synthesis [84]. In their studies, both metadynamics and constrained molecular dynamics (MD) simulations highlight important fluctuations of the Au cluster during the reaction process, which is consistent with the geometric fluxionality of Au particles aforementioned [69]. Later, Thomson et al. compared the catalytic activity of Au3, Au 4 + , Au5, and Au 5 clusters in the gas-phase reaction of H2 and O2 to form H2O2 [85]. It was found that both neutral and charged Au clusters are active in the peroxide formation reaction, but Au 4 + is the most active one for H2O2 formation based on the Gibbs free energy of activation.
As mentioned above, the adsorption of O2 on Au particles is weak. However, it has been shown that the reactivity of Au(111) and Au(100) surfaces and clusters toward O2 is noticeably enhanced with the presence of preadsorbed hydrogen. In this case, the interaction between Au and O2 becomes stronger due to the formation of O–H bonds, leading to easier formation of the active oxidant species OOH [70]. The hydrogenation of OOH was found to be highly exothermic, with a low activation barrier, thereby allowing the easy and fast formation of H2O2. Moreover, the partial and complete reduction of O2 by hydrogen on the (111) surfaces of transition metals (e.g., Rh, Ir, Ni, Pd, Pt, Cu, Ag, and Au) have also been systematically studied, and the relevant elementary reactions are listed in Equations (1)–(14) [86]. Several hydrogenated forms of oxygen are investigated as possible reaction intermediates, including peroxyl (OOH), aquoxyl (OOHH), hydrogen peroxide (HOOH), trihydrogen peroxide (HOOHH), hydroxyl (OH), and water (H2O). It is clearly seen that the elementary reactions concerning O–O bond breakage (e.g., O–O, O–OH, HO–OH, and O–OHH) directly or indirectly bring about H2O formation. The results demonstrate that (1) the dissociative adsorption of O2 prevails on Rh, Ir, Ni, and Cu facets, and the reduction of the dissociated O species is kinetically hindered due to the strong O binding to the metals; (2) the complete reduction to H2O via species OOH, HOOH, and OOHH prevails on Pd, Pt, and Ag, since they are capable of catalyzing O–O bond cleavage and O–H bond formation; (3) Au cannot easily break the O–O bond due to the weak binding of O2. Therefore, Au is more selective towards the partial reduction of O2 to generate H2O2.
Adsorption
O2 (g) → O2
Hydrogenation
O2 + H → OOH
OOH + H → HOOH
OOH + H → OOHH
HOOH + H → HOOHH
O + H → OH
OH + H → H2O
2OH → H2O + O
O–O bond scission
O2 → 2O
OOH → OH + O
HOOH → 2OH
OOHH → H2O + O
Desorption
HOOH → HOOH (g)
H2O → H2O (g)
Since Hutchings and coworkers discovered the synergistic effect of bimetallic PdAu on the direct synthesis H2O2 in 2002 [87], several analyses of H2O2 formation from H2 and O2 on alloy clusters have been reported. Through first principles calculations, Thomson’s group predicted the catalytic activity of Ag–Au, Cu–Au, and Pd–Au dimers and trimers for H2O2 formation, including several elementary steps: (1) molecular adsorption of O2, (2) the formation of OOH species due to the first H addition, and (3) the formation of H2O2 based on the second H addition [88]. For each step, the activation energy (Δ E act ) and the change in the Gibbs free energy (Δ G act ) were computed. It was shown that the H2O2 formation on small Ag–Au and Cu–Au clusters is not feasible, mainly because of the unfavorable thermodynamics of H2 addition. On the other hand, the formation of the OOH and H2O2 species is both thermodynamically and kinetically favorable on the PdAu dimer, while it is thermodynamically unfavorable on PdAu2 and Pd2Au trimers. Consequently, the PdAu dimer is proposed as a potentially active cluster for both OOH and H2O2 formation (in contrast with other clusters), implying that H2O2 formation is very sensitive to the size as well as the composition of the alloying clusters. This result is consistent with Ham’s studies [89,90], in which they demonstrated that the surface reactivity of bimetallic alloys is mostly influenced by the creation of unique mixed-metal surface sites (the so-called ensemble (geometry) effect [91]). Three different Pd monomer systems in the slab and cluster geometries involving AuPdM/Pd(111), AuPdM/Au(111) (M represents a Pd monomer), and Au41Pd@Pd13 are used to examine how surface electronic structure impacts the ensemble effect for H2O2 formation. In contrast with Au41Pd@Pd13 and AuPdM/Au(111) cases, AuPdM/Pd(111) has a higher H2O2 selectivity. The AuPdM/Pd(111) shows lower activation barriers for OOH and H2O2 formation (12.22 and 8.53 kcal/mol, respectively) compared to that on Au41Pd@Pd13 (16.60 and 14.30 kcal/mol, respectively) and AuPdM/Au(111) (11.76 and 16.37 kcal/mol, respectively). However, the activation barriers for OOH and H2O2 decomposition are higher (18.68 and 7.15 kcal/mol, respectively), as compared to that on Au41Pd@Pd13 (10.38 and 3.92 kcal/mol, respectively) and AuPdM/Au(111) (7.38 and 5.30 kcal/mol, respectively). The reactivity enhancement of the Pd and Au surface atoms is a consequence of the reduced coordination number, as well as the inherent mechanical strain. As a result, the enhanced activity of Pd and its adjacent Au atoms makes the bond breakage of O–O, O–OH, and HO–OH easier, leading to decreased H2O2 selectivity. In the formation of H2O2 on Au20 and Au19Pd clusters, Beletskaya et al. also concluded that the substitution of gold atoms by palladium in Au20 results in an increase in the activity of catalyst for H2O2 formation, while low coordinated palladium atoms are also responsible for decreasing the selectivity due to water formation [92].
Yoshizawa et al. examined the formation of H2O2 from H2 and O2 on Pd(111) and Au@Pd(111) and 10 possible elementary steps were considered [93,94,95]. All of the side reactions are involved in O–O bond cleavage (e.g., O–O, O–OH, and HO–OH), leading to the nonselective formation of H2O. It is proposed that the competition of H2O2 and H2O formation is actually the competition between the O–O bond and O–M bond (M is Pd in the case of the Pd(111) and Au in the case of the Au@Pd(111)). The O–Pd bond is usually stronger than the O–O bonds in the OOH and H2O2 species, making the O–O bond breakage facile, while the O–Au bond is generally weaker than the O–O bonds, suppressing the O–O bond dissociation and thus facilitating the release of H2O2. Similar conclusions are also obtained in Todorovic’s findings [96], in which they demonstrate that the H2O formation mainly originates from O2 dissociation on Pd(111), while it is formed from H2O2 decomposition on Au(111). One intuitive explanation is that each additional H atom weakens the bond of the molecular backbone and increases the O–O bond length. Nørskov et al. have shown that the O–O bond length increases as O2 (1.34 Å), OOH (1.44 Å), and H2O2 (1.50 Å) on an Au12 cluster [97]. Moreover, using well-established scaling and Brønsted-Evans-Polanyi (BEP) relations in combination with the concept of the degree of catalyst control, they predict a search direction for promising H2O2 synthesis catalysts, where binary Au–Pd and Au–Ag alloys are most promising (taking stability considerations into account).

5. PO Reaction Mechanism

Haruta and co-workers first demonstrated the direct propylene epoxidation using gas-phase H2 and O2 over Au/TiO2 with high selectivity [14]. Since then, Au nanoparticles supported on several other Ti-containing supports (such as TS-1 [10,29], Ti-MCM-41 [98], and Ti-MCM-48 [26]) have been found to be active for propylene epoxidation in the gas phase. In particular, Au nanoparticles loaded on a TS-1 support were first synthesized at Enichem (Italy) [99] and were identified as a possible candidate for the future commercial production of PO. Because of the adequate activity, high selectivity, and extraordinary stability of TS-1, liquid-phase epoxidation using H2O2 over TS-1 is feasible for PO production [100,101]. However, in view of the fact that H2O2 and PO have comparable market values on a molar basis, currently, it is difficult to run this process profitably. An inelastic neutron scattering (INS) study [102] and in situ UV spectroscopy study [103] have offered direct spectroscopic evidence for the formation of H2O2 and the hydroperoxy radical (OOH) species, originating from H2–O2 related reactions over Au/TiO2. Thus, a similar propylene epoxidation mechanism in the gas phase using in situ formed H2O2 over Au/TS-1 catalyst is expected.
In many cases, current theoretical studies are on par with classical experimental analyses, so theory and experiment can drive each other towards a better understanding of the catalytic reaction mechanism of propylene epoxidation and catalyst design. In the following sections, we summarize the application of computational catalysis and kinetic modeling for the propylene epoxidation mechanism (as well as possible side reactions), involving (1) epoxidation using H2O2 over TS-1; (2) H2O2 formation and propylene epoxidation on Au; and (3) the synergistic effect of an Au/Ti-containing support.

5.1. Reaction Mechanism on Titanosilicate

In comparison to Au/TiO2, gold particles supported on titanosilicate (TS-1) present much higher catalytic activity for the direct propylene epoxidation reaction, suggesting that the microstructure of the Ti sites is crucial. Through DFT calculations, Thomson’s group demonstrated that the defects (Ti sites located near Si vacancies in Ti-substituted silicate lattices, shown in Figure 2) of TS-1 catalysts promote the reactivity of propylene epoxidation with H2O2 [32]. The results showed that, in the defect-free model, Ti–OOH formation is endothermic (about 6 kcal/mol) after the addition of H2O2 to the hydroxyl group on Ti, and the interaction between propylene and hydroperoxy is highly repulsive. In the defect model, Ti–OOH formation is exothermic (−9 kcal/mol), and the activation barrier (ΔEact) for propylene epoxidation is about 9.3 kcal/mol as propylene approaches the intermediate Ti–OOH (the transition state is exhibited in Figure 3), which is highly exothermic (−41.1 kcal/mol). With regard to the effect of the Ti defect site, they proposed that the absence of one Si atom reduces steric repulsions and the generated defect Ti site is more flexible, thereby stabilizing the transition-state formation through structural relaxation.
In their later studies, Thomson et al. compared five possible mechanisms for propylene epoxidation using H2O2/TS-1 at the same level of theory [104]. The first one is the Sinclair and Catlow mechanism (ethylene in the original work) [105,106], including two main steps: (1) the formation of the reactive hydroperoxy intermediate Ti–OOH on a tripodal site (representing the exterior termination surface of TS-1 crystallites) and (2) propylene attack on the reactive Ti–OOH to form PO. The two activation energies calculated are 7.9 and 8.5 kcal/mol for the hydroperoxy formation and epoxidation step, respectively. The second case is the Vayssilov and van Santen mechanism (ethylene in the original work) [107], in which the preadsorbed H2O2 on a fully tetrahedral Ti site is reacted with an incoming propylene molecule, overcoming an activation barrier of 19.0 kcal/mol. The third case is the Munakata et al. mechanism [108], where the intermediate Si–O–O–Ti forms first with a two-step reaction via one oxygen atom from H2O2 located in the Ti coordination sphere. The activation barriers are 21.2 and 4.3 kcal/mol (difference in the Gibbs free energy at 298 K), corresponding to transition-state geometries (2) and (4), respectively, as shown in Figure 4a. The oxygen from the Si–O–O–Ti species then inserts into the C=C double bond of the adsorbed propylene molecule (ΔEact = 1.2 kcal/mol), as shown in Figure 4b.
The last two mechanisms involve defect sites, based on a partial silanol nest model and a full silanol nest model, respectively. As Figure 2 shows, a Ti site with one defect neighbor means a total of three neighboring hydroxyl groups (or silanol groups, Si–OH). In the partial silanol nest model, only one Si–OH group is modeled instead of three, and the epoxidation mechanism mentioned above is studied, with the transition-state geometry shown in Figure 3 [32]. In the full silanol nest model, the formation of a hydroperoxy intermediate (Ti–OOH) overcomes an activation barrier of 8.9 kcal/mol (Gibbs free energy at 298 K), with the transition-state structure shown in Figure 5a. Once the generated H2O is removed from the active site, propylene can adsorb and react with the active Ti–OOH species. The activation energy for the transition-state structure (Figure 5b) is only 4.6 kcal/mol (Gibbs free energy at 298 K). In comparison to a partial silanol nest model, the Ti/defect site pathway with a full silanol nest is energetically more favorable with regard to the propylene epoxidation in the H2O2/TS-1 or H2/O2/Au/TS-1 catalytic systems.
Recently, Song’s group performed DFT calculations to investigate the propylene epoxidation mechanism (stepwise and concerted mechanisms) with H2O2 over tripodal and Ti/defect (a full silanol nest model) sites, respectively (both tripodal and Ti/defect represent two types of internal Ti active sites) [109]. It was found that, on tripodal sites, both stepwise (through a three-membered ring (3MR) Ti–(η2–OOH)) and concerted (through a five-membered ring (5MR) Ti–(η1–H2O2)) mechanisms are possible because of comparable barriers. On the other hand, over the Ti/defect site, both stepwise mechanisms are feasible either via a 3MR Ti–(η2–OOH) or a 5MR Ti–(η1–OOH) intermediate, and the concerted mechanism proceeds through a 5MR Ti–(η1–H2O2) intermediate. Among them, the stepwise mechanism via the 5MR Ti–(η1–OOH) intermediate over the Ti/defect site is kinetically most favorable by comparison of the energy barriers of various pathways. Their studies indicate that both Ti–OOH and Ti–H2O2 species are probably responsible for propylene epoxidation to produce PO.

5.2. Reaction Mechanism on Gold

In addition to the aforementioned reaction mechanisms on TS-1 in which propylene attacks active Ti–OOH (or Ti–H2O2) species, recent DFT studies in combination with Fourier transform infrared (FTIR) spectroscopy have identified the adsorption of propylene and PO on the Au/TiO2 catalyst interface [110,111]. Propylene was found to bind to a single atomic Au atom site via a (πσ)-interaction and bind to the TiO2 surface via a π-interaction, and this interaction is also distinguishable through the C=C bond stretching frequency of propylene (Figure 6). The results reveal that the propylene–Au (particularly low coordinated Au atoms, e.g., corner and edge sites) interaction has a stronger binding energy in comparison to the propylene–TiO2 interaction, suggesting a possible epoxidation route to form PO on Au particles. Friend et al. also investigated the adsorption of propylene on an Au(111) surface, and the adsorption strength was found to follow the trend: defect-free < vacancy < step < adatom [30], indicating that the low coordinated Au sites (e.g., step and adatom) can activate propylene to some extent.
When referring to the olefin epoxidation process, the most commonly assumed mechanism involves the formation of an oxametallacycle intermediate. For example, ethylene epoxidation on Ag(111) to produce ethylene oxide (EO) is considered to undergo an oxametallacycle intermediate, namely, OMME (oxygen-metal-metal-ethylene, a five-member ring) or OME (oxygen-metal-ethylene, a four-member ring) [112,113,114]. These oxametallacycle complexes have been experimentally identified on the silver surface [115,116,117]. The OME complex with a single Ag atom in its ring is suggested to be much more selective toward EO formation. In contrast, the OMME intermediate with two Ag atoms incorporated preferably supports acetaldehyde production [118].
Using DFT calculations, Thomson et al. explored propylene epoxidation over a neutral Au3 cluster in the presence of H2 and O2 [119]. A side-on O2 adsorption on Au3 is followed by dissociative addition of H2 to generate an OOH species and a lone H located on the gold trimer. The more electrophilic O atom (proximal to the Au) of the Au–OOH group attacks the C=C of an approaching propylene to generate PO (Figure 7, ΔEact = 19.6 kcal/mol). The residual OH reacts with H to form water. The propylene epoxidation step is identified to be the rate-determining step (RDS).
Roldan et al. investigated the partial oxidation of propylene using atomic oxygen adsorbed on Au(111) by means of DFT calculations, and they observed different intermediates with respect to different reaction pathways [120]. Surface atomic oxygen acts as a Brönsted base, stripping the allylic hydrogen (acidic hydrogen) of propylene to yield allyl and OH radicals. On the other hand, oxygen works as a Lewis acid, receiving electronic density from the C=C bond of propylene to lead to the formation of an oxametallacycle intermediate OMMP (oxygen-metal-metal-propylene, a five-member ring) incorporating two Au atoms. The OMMP can either be OMMP1 in which atomic O is bound to the end C of the C=C and the middle C of propylene is linked to the Au surface, or OMMP2 in which the atomic O is bound to the middle C of propylene and the end C of C=C is linked to gold surface, as demonstrated in Figure 8. Both OMMP1 and OMMP2 can evolve to the final product PO, and OMMP1 is also able to evolve to propanal. However, results show that the formation of OMMP on this extended surface is thermodynamically favored but kinetically inhibited (ΔEact = 14.07 kcal/mol). In contrast, the abstraction of the allylic hydrogen leading to an allyl radical is much more kinetically favored (ΔEact = 6.23 kcal/mol).
The results from Roldan et al. are consistent with Moskaleva’s studies on Au(321) [24], in which the stripping of the methyl hydrogen of propylene by atomic O to yield allyl was found to be facile, leading to the final product allyl alcohol or acrolein. Allyl alcohol is known to convert to acrolein on O-covered Au(111), and this is why acrolein is the main partial oxidation product captured on gold-based catalysts [121,122,123,124,125]. Moreover, OMMP (from C3H6 + O) will preferentially react to form propanal (from OMMP1) or acetone (from OMMP2) via a 1,2-H shift reaction instead of PO. At the same time, Moskaleva took into account an OMP (oxygen-metal-propylene, a four-member ring) species as a possible oxametallacycle intermediate incorporating one Au atom, and they found that the ring closure of OMP to form PO (epoxide) is less favored, as compared to its isomerization into propanal or acetone. Thus, the formation of acrolein, propanal, and acetone explains the low selectivity of PO on the bulk gold surface. Nonetheless, if H2 is added to O2 as a co-reactant, high PO selectivity can be achieved, in part due to the fact that H2 can help reduce the concentration of atomic O. On the other hand, Moskaleva suggested that OOH species (formed from H2 + O2) serve as the oxidant, and the “hydroperoxametallacycle” HO–OMMP cannot as easily isomerize to acetone or propanal, in contrast with OMMP. Additionally, the allylic H abstraction of propylene by OOH has a much higher energy barrier than that by O, leading to less acrolein.
In addition to PO production from H2/O2 sources, PO formation has also been detected experimentally from reactions with H2O/O2 mixtures (with a C3H6 conversion of 0.88% and a PO selectivity of 52%), indicating the formation of an OOH species from the reaction O2 + H2O ⇋ OOH + OH. Although the current performance is not sufficient for commercial application, it does highlight an interesting eco-friendly route for PO production [126,127,128]. Landman et al. first reported the formation of OOH from coadsorbed H2O and O2 on Au8/MgO(100) for a CO oxidation reaction [129]. The role of H2O in propylene epoxidation on gold particles has also been computationally explored. For instance, Chang et al. proposed a catalytic reaction mechanism of propylene epoxidation with an H2O/O2 mixture on Au38 and Au10 clusters, using DFT calculations [130,131]. In their studies, O2 reacts with H2O to form OOH and OH species after their coadsorption on Au38 clusters, and the generated OOH species then attack the C=C bond of the incoming propylene molecule to form a four-member ring (OMP) intermediate (ΔEact = 19.13 kcal/mol). Finally, PO is produced via a ring closure of OMP (ΔEact = 13.15 kcal/mol) and desorbs from the Au38 cluster (Figure 9a). The other PO production pathway on the Au38 cluster is not via the OMP intermediate but directly from C3H6 + OOH ⇋ PO + OH (ΔEact = 22.65 kcal/mol, Figure 9a), which also occurs on the Au10 cluster (ΔEact = 19.03 kcal/mol, Figure 9b). In the later studies of Chang et al., they reported the reaction mechanism of propylene epoxidation with an H2O/O2 mixture over Au7/α-Al2O3(0001), by means of DFT calculations and ab initio molecular dynamics (AIMD) [132]. H2O was found to be easily dissociated on coordinatively unsaturated Al surface sites to form a hydroxylated Al2O3 surface. The activation (hydrogenation) of molecular O2 via an OOH species (adsorbed O2 on a Au7 cluster reacts with H2O adsorbed at the gold/oxide interface site: O2 + H2O ⇋ OOH + OH) is identified to be a feasible pathway from both DFT calculations and AIMD simulations. The resulting OOH species epoxidizes the propylene molecule to first form a “hydroperoxametallacycle” intermediate. OMMP is then generated along with OH stripping, to finally produce PO via the ring closure of the OMMP intermediate.
Although many studies have indicated that the epoxidation of propylene using molecular oxygen is unlikely [11,30], Hu et al. demonstrated its possibility with DFT calculations on Au(111), in which OMMP is derived from O–O bond cleavage of an oxametallyacycle intermediate OOMMP [31]. Two key competing reactions, OMMP formation (leading to PO, acrolein, and propanal) and allylic hydrogen stripping of propylene (leading to acrolein, CO2, etc.), were compared via both atomic and molecular O2 mechanisms. In the atomic O mechanism, allylic hydrogen abstraction is much easier to proceed with lower activation barriers, in contrast with OMMP formation, indicating a poor PO selectivity when using atomic O as an oxidant. While in the molecular O2 mechanism, PO selectivity is enhanced since the activation barrier for OMMP formation is much lower than that of the stripping of methyl hydrogen.

5.3. Reaction Mechanism on Au/Ti Interface Sites

With regard to different mechanisms mentioned in Section 5.1, H2O2 species adsorb on the nondefect and defect Ti site in the form of Ti–H2O2 or further dissociate to form Ti–OOH species. Propylene then adsorbs around the Ti–OOH (or Ti–H2O2) species, followed by the epoxidation step, suggesting a “sequential” mechanism (H2O2 formation on Au nanoparticles first and then diffusion to Ti sites). Nonetheless, it was shown that propylene adsorption in TS-1 pores without any Au particles or clusters (Eads = −10.0 kcal/mol) is significantly weaker than that on Au/Ti interface sites (Eads = −20.0 kcal/mol) [133], which is consistent with the experimentally observed adsorption behavior [134,135]. Thus, the “simultaneous” mechanism (2-site) that does not require Ti–OOH and Ti–H2O2 species is proposed to proceed at the Au/Ti interfacial sites (it is actually OOH species on Au nanoparticles or H–Au–OOH species) [133]. In this mechanism, propylene molecules adsorbed on interfacial Ti sites will attack OOH or H–Au–OOH species formed on Au particles; thus, Ti plays an indirect role.
Although there are different opinions about the active sites for the epoxidation reaction, a reasonable proximity between Au and Ti sites is generally considered to be indispensable for the PO reaction, suggesting the existence of a well-defined reaction zone [110,111,133,135,136]. Previous experiments have indicated that the perimeters between Au particles and Ti sites are active for the epoxidation reaction [16,137], as well as several side reactions (isomerization and oxidative cracking, etc.) [25,138].
At a larger modeling scale, Turner’s group used kinetic Monte Carlo (KMC) simulations to investigate propylene epoxidation over an Au/TiO2/SiO2 catalyst. In their model, H2O2 is first generated from coadsorbed H2 and O2 on Au nanoparticles, and H2O2 can then either degrade into water or diffuse to TiO2 sites supported on SiO2 to epoxidize propylene molecules [139]. The PO formation rates predicted by KMC simulations are consistent with experimental reports corresponding to different temperatures and feed concentrations of O2 [140]. In their later kinetic modeling studies [141], several key side reactions encountered during PO formation are taken into account (acrolein formation, propanal and acetone formation from OMP isomerization, oxidative cracking into CO2, etc.), affecting the overall PO selectivity. According to Figure 10, it is the OOH species formed on Au nanoparticles that directly transfers to the neighboring Ti sites (interfacial Ti sites at the dual Au/Ti sites) to form active Ti–OOH species and epoxidize propylene to form an oxametallacycle OMP (OMC’ or OMC”, depending on which carbon of propylene is connected with O). Their KMC simulations closely reproduce the experimental findings [140]. For instance, it was found that the O2 feed concentration has a slight effect on PO selectivity (since O2 adsorption on Au particles is very weak), which is in good agreement with the kinetic tests of Taylor and Chen [142,143]. At the same time, both KMC simulations and experiments demonstrate that PO selectivity increases along with decreasing reaction temperature and increasing H2/C3H6 feed concentration ratio, since low temperature and high H2 feed concentration can help suppress side reactions that contribute to acrolein, propanal, acetone, CO2, and ethanal. It seems that the synergy of the Au/Ti dual interface sites plays a central role in the reaction network of propylene epoxidation.
It is worth noting that, although low temperatures yield high PO selectivity, the propylene conversion is typically very low (<2%) [138,144]. Efforts to enhance propylene conversion by raising the reaction temperature inevitably lead to decreasing PO selectivity because of consecutive PO reactions (oligomerization, oxidative cracking). In particular, the oxidative cracking into acids and further into CO2 is correlated with decarbonylation and decarboxylation of carboxylic acids [17,23,24,25,26]. However, only a few studies have investigated the role of carboxylic acid oxidation on Au-based catalysts [23,145,146,147]. The active sites were found to reside at the Au/TiO2 interface, again highlighting the synergistic effect of interface sites in heterogeneous catalysis.
Matthew’s group studied the partial oxidation of acetic acid by O2 at the dual perimeter sites of an Au/TiO2 catalyst using first-principle DFT calculations. They identified that a novel gold ketenylidene (Au2=C=C=O) intermediate is formed from the deoxygenation of acetic acid [145,146]. The ketenylidene species is also identified by its measured characteristic stretching frequency ν(CO) = 2040 cm−1 using in situ infrared spectroscopy. As a comparison, no ketenylidene formation was observed on Au/SiO2 or a TiO2 blank sample, suggesting the involvement of dual catalytic Ti4+ and Au perimeter sites. The production of CO2 and H2O is also observed when raising the reaction temperature to 473 K, due to the total oxidation of the ketenylidene species. The decarboxylation and decarbonylation of longer chain carboxylic acids at the Au/TiO2 interface has also been investigated [23,147]. For instance, O2 activation and the oxidation of propionic and butyric acids to form propionate and butyrate first proceed at the dual Au/TiO2 interface. Dehydrogenation of propionate and butyrate to form acrylate and crotonate then occurs, followed by the further oxidation and the subsequent C–C and C–O cleavage to generate the Au2=C=C=O intermediate. The formed ketenylidene species can be hydrogenated to produce H2=C=C=O (g), which either readily desorbs from the catalyst surface or undergoes a full oxidation to form CO2 and H2O.

6. Conclusions and Perspectives

Direct gaseous-phase epoxidation of propylene over gold-based catalysts in the presence of H2 and O2 has been extensively studied using theoretical techniques. These studies have provided fundamental insights into H2 and O2 adsorption, H2O2 (OOH) species formation, and propylene epoxidation with gold-based catalysts. For realistic evaluation of propylene epoxidation, the large number of reaction species as well as the structural complexity of supported catalysts require detailed investigations of different reaction pathways on the various active sites, such as metal (gold), support (TiO2, TS-1, etc.), and the metal/support interface. As the reaction proceeds, changes in the composition of surface species usually take place and may influence catalytic activity. Capturing the dynamic behavior of catalytic surface is challenging and herein KMC simulations aid to detect possible reaction events “on the fly” and prove significant to explicitly capture the temporal evolution of the catalytic system. On the other hand, it demands that the validation of the KMC results in a detailed comparison of experimental findings, and underscores the need for methods able to quantify the effect of the uncertainty of certain parameters toward simulation results. Additionally, in complex systems, sensitivity analysis is essential to identify the main parameters influencing the observed behaviors and characterize such influences.
Until now the reaction mechanism toward propylene epoxidation over Au/titania/silica is still debated, and the most convincing one is based on the synergy between Au particles and isolated tetrahedral titanium (or titania), where hydroperoxyl species (OOH or H2O2) is formed on Au nanoparticles first, and then transfers to the adjacent Ti sites to form an active Ti–OOH intermediate, followed by the epoxidation of propylene to produce PO. In this process, some byproducts are coproduced at the Au/support interface and lead to catalyst deactivation. Additionally, a higher propylene conversion level (>10%) and higher hydrogen efficiency (>50%) will be essential for a commercial process to be profitable.
It is known that H2O2 is a weak acid, and as suggested in our previous studies [141], Brønsted acid-functionalized supports may promote OOH spillover, thereby enhancing the probability of OOH transferring to the neighboring Ti sites. This would potentially reduce OOH decomposition to O on the Au surface, leading to more PO production. Additionally, it is proposed that bimetallic alloys (e.g., Au–Pd and Au–Ru) or trimetallic alloys (e.g., Au–Pd–Ru) supported on TiO2 can be highly active for H2O2 production, since this leads to a significant synergistic enhancement in activity (from Pd and Ru) and selectivity (from Au) [148,149,150,151]. Thus, such Au–Pd, Au–Ru, and Au–Pd–Ru alloy nanoparticles loaded on TS-1 support may lead to improved propylene conversion, as well as improved hydrogen utilization efficiency. In addition to Ti-based supports, other reducible support candidates (such as CeO2, Fe2O3, and Co3O4) need to be theoretically explored, since they can help anchor Au nanoparticles to some extent and promote O2 adsorption and activation at the interfacial sites [16,53,152,153,154]. Improved computational approaches will enable the modeling of multifunctional catalysts, created by multicomponent nanoparticles deposited on supports. These theoretical models can lead to more efficient prediction of advanced catalysis by building a fundamental understanding on active sites of multicomponent nanoparticles, multiple interfaces, and intermediate transfer.

Funding

This research was funded by the National Science Foundation (CBET-1510485 and CBET-1511820) and a UA System Collaboration Grant.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shelef, M.; Graham, G.W. Why rhodium in automotive three-way catalysts? Catal. Rev. 1994, 36, 433–457. [Google Scholar] [CrossRef]
  2. Hoebink, J.H.B.J.; van Gemert, R.A.; van den Tillaart, J.A.A.; Marin, G.B. Competing reactions in three-way catalytic converters: Modelling of the NOx conversion maximum in the light-off curves under net oxidising conditions. Chem. Eng. Sci. 2000, 55, 1573–1581. [Google Scholar] [CrossRef]
  3. Angelici, R.J. Heterogeneous catalysis of the hydrodesulfurization of thiophenes in petroleum: An organometallic perspective of the mechanism. Acc. Chem. Res. 1988, 21, 387–394. [Google Scholar] [CrossRef]
  4. Barteau, M.A. New perspectives on direct heterogeneous olefin epoxidation. Top. Catal. 2003, 22, 3–8. [Google Scholar] [CrossRef]
  5. Hutchings, G.J. Promotion in heterogeneous catalysis: A topic requiring a new approach? Catal. Lett. 2001, 75, 1–12. [Google Scholar] [CrossRef]
  6. Valden, M.; Lai, X.; Goodman, D.W. Onset of catalytic activity of gold clusters on titania with the appearance of nonmetallic properties. Science 1998, 281, 1647–1650. [Google Scholar] [CrossRef] [PubMed]
  7. Mavrikakis, M.; Stoltze, P.; Nørskov, J.K. Making gold less noble. Catal. Lett. 2000, 64, 101–106. [Google Scholar] [CrossRef]
  8. Yoon, B.; Häkkinen, H.; Landman, U.; Wörz, A.S.; Antonietti, J.-M.; Abbet, S.; Judai, K.; Heiz, U. Charging effects on bonding and catalyzed oxidation of CO on Au8 clusters on MgO. Science 2005, 307, 403–407. [Google Scholar] [CrossRef] [PubMed]
  9. Nijhuis, T.A.; Makkee, M.; Moulijn, J.A.; Weckhuysen, B.M. The production of propene oxide: Catalytic processes and recent developments. Ind. Eng. Chem. Res. 2006, 45, 3447–3459. [Google Scholar] [CrossRef]
  10. Yap, N.; Andres, R.P.; Delgass, W.N. Reactivity and stability of Au in and on TS-1 for epoxidation of propylene with H2 and O2. J. Catal. 2004, 226, 156–170. [Google Scholar] [CrossRef]
  11. Huang, J.; Haruta, M. Gas-phase propene epoxidation over coinage metal catalysts. Res. Chem. Intermed. 2012, 38, 1–24. [Google Scholar] [CrossRef]
  12. Tullo, A. Dow, BASF to build propylene oxide. Chem. Eng. News 2004, 82, 15. [Google Scholar]
  13. Masatake, H.; Tetsuhiko, K.; Hiroshi, S.; Nobumasa, Y. Novel gold catalysts for the oxidation of carbon monoxide at a temperature far below 0 °C. Chem. Lett. 1987, 16, 405–408. [Google Scholar]
  14. Hayashi, T.; Tanaka, K.; Haruta, M. Selective vapor-phase epoxidation of propylene over Au/TiO2 catalysts in the presence of oxygen and hydrogen. J. Catal. 1998, 178, 566–575. [Google Scholar] [CrossRef]
  15. Haruta, M. Spiers memorial lecture role of perimeter interfaces in catalysis by gold nanoparticles. Faraday Discuss. 2011, 152, 11–32. [Google Scholar] [CrossRef] [PubMed]
  16. Haruta, M.; Daté, M. Advances in the catalysis of Au nanoparticles. Appl. Catal. A 2001, 222, 427–437. [Google Scholar] [CrossRef]
  17. Okumura, M.; Fujitani, T.; Huang, J.; Ishida, T. A career in catalysis: Masatake haruta. ACS Catal. 2015, 5, 4699–4707. [Google Scholar] [CrossRef]
  18. Huang, J.; Takei, T.; Akita, T.; Ohashi, H.; Haruta, M. Gold clusters supported on alkaline treated TS-1 for highly efficient propene epoxidation with O2 and H2. Appl. Catal. B 2010, 95, 430–438. [Google Scholar] [CrossRef]
  19. Sinha, A.K.; Sindhu, S.; Susumu, T.; Masatake, H. A three-dimensional mesoporous titanosilicate support for gold nanoparticles: Vapor-phase epoxidation of propene with high conversion. Angew. Chem. Int. Ed. 2004, 43, 1546–1548. [Google Scholar] [CrossRef] [PubMed]
  20. Chowdhury, B.; Bravo-Suárez, J.J.; Daté, M.; Tsubota, S.; Haruta, M. Trimethylamine as a gas-phase promoter: Highly efficient epoxidation of propylene over supported gold catalysts. Angew. Chem. Int. Ed. 2006, 45, 412–415. [Google Scholar] [CrossRef] [PubMed]
  21. Taylor, B.; Lauterbach, J.; Delgass, W.N. Gas-phase epoxidation of propylene over small gold ensembles on TS-1. Appl. Catal. A 2005, 291, 188–198. [Google Scholar] [CrossRef]
  22. Taylor, B.; Lauterbach, J.; Delgass, W.N. The effect of mesoporous scale defects on the activity of Au/TS-1 for the epoxidation of propylene. Catal. Today 2007, 123, 50–58. [Google Scholar] [CrossRef]
  23. McEntee, M.; Tang, W.; Neurock, M.; Yates, J.T. Mechanistic insights into the catalytic oxidation of carboxylic acids on Au/TiO2: Partial oxidation of propionic and butyric acid to gold ketenylidene through unsaturated acids. ACS Catal. 2015, 5, 744–753. [Google Scholar] [CrossRef]
  24. Moskaleva, L.V. Theoretical mechanistic insights into propylene epoxidation on Au-based catalysts: Surface O versus OOH as oxidizing agents. Catal. Today 2016, 278, 45–55. [Google Scholar] [CrossRef]
  25. Stangland, E.E.; Stavens, K.B.; Andres, R.P.; Delgass, W.N. Characterization of gold–titania catalysts via oxidation of propylene to propylene oxide. J. Catal. 2000, 191, 332–347. [Google Scholar] [CrossRef]
  26. Uphade, B.S.; Akita, T.; Nakamura, T.; Haruta, M. Vapor-phase epoxidation of propene using H2 and O2 over Au/Ti–MCM-48. J. Catal. 2002, 209, 331–340. [Google Scholar] [CrossRef]
  27. Kapoor, M.P.; Sinha, A.K.; Seelan, S.; Inagaki, S.; Tsubota, S.; Yoshida, H.; Haruta, M. Hydrophobicity induced vapor-phase oxidation of propene over gold supported on titanium incorporated hybrid mesoporous silsesquioxane. Chem. Commun. 2002, 2902–2903. [Google Scholar] [CrossRef]
  28. Stangland, E.E.; Taylor, B.; Andres, R.P.; Delgass, W.N. Direct vapor phase propylene epoxidation over deposition-precipitation gold-titania catalysts in the presence of H2/O2: Effects of support, neutralizing agent, and pretreatment. J. Phys. Chem. B 2005, 109, 2321–2330. [Google Scholar] [CrossRef] [PubMed]
  29. Nijhuis, T.A.; Huizinga, B.J.; Makkee, M.; Moulijn, J.A. Direct epoxidation of propene using gold dispersed on TS-1 and other titanium-containing supports. Ind. Eng. Chem. Res. 1999, 38, 884–891. [Google Scholar] [CrossRef]
  30. Baker, T.A.; Xu, B.; Jensen, S.C.; Friend, C.M.; Kaxiras, E. Role of defects in propene adsorption and reaction on a partially O-covered Au(111) surface. Catal. Sci. Technol. 2011, 1, 1166–1174. [Google Scholar] [CrossRef]
  31. Dai, Y.; Chen, Z.; Guo, Y.; Lu, G.; Zhao, Y.; Wang, H.; Hu, P. Significant enhancement of the selectivity of propylene epoxidation for propylene oxide: A molecular oxygen mechanism. Phys. Chem. Chem. Phys. 2017, 19, 25129–25139. [Google Scholar] [CrossRef] [PubMed]
  32. Wells, D.H.; Delgass, W.N.; Thomson, K.T. Evidence of defect-promoted reactivity for epoxidation of propylene in titanosilicate (TS-1) catalysts: A dft study. J. Am. Chem. Soc. 2004, 126, 2956–2962. [Google Scholar] [CrossRef] [PubMed]
  33. Corti, C.W.; Holliday, R.J.; Thompson, D.T. Developing new industrial applications for gold: Gold nanotechnology. Gold Bull. 2002, 35, 111–117. [Google Scholar] [CrossRef]
  34. Min, B.K.; Friend, C.M. Heterogeneous gold-based catalysis for green chemistry: Low-temperature CO oxidation and propene oxidation. Chem. Rev. 2007, 107, 2709–2724. [Google Scholar] [CrossRef] [PubMed]
  35. Mei, D.; Neurock, M.; Smith, C.M. Hydrogenation of acetylene–ethylene mixtures over Pd and Pd–Ag alloys: First-principles-based kinetic Monte Carlo simulations. J. Catal. 2009, 268, 181–195. [Google Scholar] [CrossRef]
  36. Mei, D.; Du, J.; Neurock, M. First-principles-based kinetic Monte Carlo simulation of nitric oxide reduction over platinum nanoparticles under lean-burn conditions. Ind. Eng. Chem. Res. 2010, 49, 10364–10373. [Google Scholar] [CrossRef]
  37. Hess, F.; Over, H. Rate-determining step or rate-determining configuration? The deacon reaction over RuO2(110) studied by DFT-based KMC simulations. ACS Catal. 2017, 7, 128–138. [Google Scholar] [CrossRef]
  38. Hess, F.; Sack, C.; Langsdorf, D.; Over, H. Probing the activity of different oxygen species in the CO oxidation over RuO2(110) by combining transient reflection–absorption infrared spectroscopy with kinetic Monte Carlo simulations. ACS Catal. 2017, 7, 8420–8428. [Google Scholar] [CrossRef]
  39. Salisbury, B.E.; Wallace, W.T.; Whetten, R.L. Low-temperature activation of molecular oxygen by gold clusters: A stoichiometric process correlated to electron affinity. Chem. Phys. 2000, 262, 131–141. [Google Scholar] [CrossRef]
  40. Okumura, M.; Kitagawa, Y.; Haruta, M.; Yamaguchi, K. DFT studies of interaction between O2 and Au clusters. The role of anionic surface Au atoms on Au clusters for catalyzed oxygenation. Chem. Phys. Lett. 2001, 346, 163–168. [Google Scholar] [CrossRef]
  41. Yoon, B.; Häkkinen, H.; Landman, U. Interaction of O2 with gold clusters: Molecular and dissociative adsorption. J. Phys. Chem. A 2003, 107, 4066–4071. [Google Scholar] [CrossRef]
  42. Mills, G.; Gordon, M.S.; Metiu, H. The adsorption of molecular oxygen on neutral and negative Aun clusters (n = 2–5). Chem. Phys. Lett. 2002, 359, 493–499. [Google Scholar] [CrossRef]
  43. Shi, H.X.; Sun, W.G.; Kuang, X.Y.; Lu, C.; Xia, X.X.; Chen, B.L.; Hermann, A. Probing the interactions of O2 with small gold cluster Au n Q (n = 2–10, Q = 0, −1): A neutral chemisorbed complex Au5O2 cluster predicted. J. Phys. Chem. C 2017, 121, 24886–24893. [Google Scholar] [CrossRef]
  44. Xu, Y.; Mavrikakis, M. Adsorption and dissociation of O2 on gold surfaces: Effect of steps and strain. J. Phys. Chem. B 2003, 107, 9298–9307. [Google Scholar] [CrossRef]
  45. Lopez, N.; Nørskov, J.K. Catalytic CO oxidation by a gold nanoparticle: A density functional study. J. Am. Chem. Soc. 2002, 124, 11262–11263. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, Y.; Gong, X.G. First-principles study of interaction of cluster Au32 with CO, H2, and O2. J. Chem. Phys. 2006, 125, 124703. [Google Scholar] [CrossRef] [PubMed]
  47. Gao, M.; Horita, D.; Ono, Y.; Lyalin, A.; Maeda, S.; Taketsugu, T. Isomerization in gold clusters upon O2 adsorption. J. Phys. Chem. C 2017, 121, 2661–2668. [Google Scholar] [CrossRef]
  48. Roldan, A.; Ricart, J.M.; Illas, F.; Pacchioni, G. O2 adsorption and dissociation on neutral, positively and negatively charged Aun (n = 5–79) clusters. Phys. Chem. Chem. Phys. 2010, 12, 10723–10729. [Google Scholar] [CrossRef] [PubMed]
  49. Alberto, R.; Silvia, G.; Manel, R.J.; Francesc, I. Critical size for O2 dissociation by Au nanoparticles. ChemPhysChem 2009, 10, 348–351. [Google Scholar]
  50. Roldán, A.; Viñes, F.; Illas, F.; Ricart, J.M.; Neyman, K.M. Density functional studies of coinage metal nanoparticles: Scalability of their properties to bulk. Theor. Chem. Accounts 2008, 120, 565–573. [Google Scholar] [CrossRef]
  51. Boronat, M.; Corma, A. Oxygen activation on gold nanoparticles: Separating the influence of particle size, particle shape and support interaction. Dalton Trans. 2010, 39, 8538–8546. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, M.S.; Goodman, D.W. The structure of catalytically active gold on titania. Science 2004, 306, 252–255. [Google Scholar] [CrossRef] [PubMed]
  53. Bond, G.C.; Thompson, D.T. Catalysis by gold. Catal. Rev. 1999, 41, 319–388. [Google Scholar] [CrossRef]
  54. Carrettin, S.; Concepción, P.; Corma, A.; Lopez Nieto, J.M.; Puntes, V.F. Nanocrystalline CeO2 increases the activity of Au for CO oxidation by two orders of magnitude. Angew. Chem. Int. Ed. 2004, 43, 2538–2540. [Google Scholar] [CrossRef] [PubMed]
  55. Guzman, J.; Carrettin, S.; Corma, A. Spectroscopic evidence for the supply of reactive oxygen during CO oxidation catalyzed by gold supported on nanocrystalline CeO2. J. Am. Chem. Soc. 2005, 127, 3286–3287. [Google Scholar] [CrossRef] [PubMed]
  56. Hernández, N.C.; Sanz, J.F.; Rodriguez, J.A. Unravelling the origin of the high-catalytic activity of supported Au: A density-functional theory-based interpretation. J. Am. Chem. Soc. 2006, 128, 15600–15601. [Google Scholar] [CrossRef] [PubMed]
  57. Kotobuki, M.; Leppelt, R.; Hansgen, D.A.; Widmann, D.; Behm, R.J. Reactive oxygen on a Au/TiO2 supported catalyst. J. Catal. 2009, 264, 67–76. [Google Scholar] [CrossRef]
  58. Florez, E.; Feria, L.; Viñes, F.; Rodriguez, J.A.; Illas, F. Effect of the support on the electronic structure of Au nanoparticles supported on transition metal carbides: Choice of the best substrate for Au activation. J. Phys. Chem. C 2009, 113, 19994–20001. [Google Scholar] [CrossRef]
  59. Schubert, M.M.; Hackenberg, S.; van Veen, A.C.; Muhler, M.; Plzak, V.; Behm, R.J. CO oxidation over supported gold catalysts—“Inert” and “active” support materials and their role for the oxygen supply during reaction. J. Catal. 2001, 197, 113–122. [Google Scholar] [CrossRef]
  60. Wang, J.G.; Hammer, B. Oxidation state of oxide supported nanometric gold. Top. Catal. 2007, 44, 49–56. [Google Scholar] [CrossRef]
  61. Laursen, S.; Linic, S. Strong chemical interactions between au and off-stoichiometric defects on TiO2 as a possible source of chemical activity of nanosized Au supported on the oxide. J. Phys. Chem. C 2009, 113, 6689–6693. [Google Scholar] [CrossRef]
  62. Greeley, J.; Nørskov, J.K. A general scheme for the estimation of oxygen binding energies on binary transition metal surface alloys. Surf. Sci. 2005, 592, 104–111. [Google Scholar] [CrossRef]
  63. Joshi, A.M.; Delgass, W.N.; Thomson, K.T. Analysis of O2 adsorption on binary−alloy clusters of gold: Energetics and correlations. J. Phys. Chem. B 2006, 110, 23373–23387. [Google Scholar] [CrossRef] [PubMed]
  64. Polynskaya, Y.G.; Pichugina, D.A.; Kuz’menko, N.E. Correlation between electronic properties and reactivity toward oxygen of tetrahedral gold–silver clusters. Comput. Theor. Chem. 2015, 1055, 61–67. [Google Scholar] [CrossRef]
  65. Feng, X.; Yang, J.; Duan, X.; Cao, Y.; Chen, B.; Chen, W.; Lin, D.; Qian, G.; Chen, D.; Yang, C.; et al. Enhanced catalytic performance for propene epoxidation with H2 and O2 over bimetallic Au–Ag/uncalcined titanium silicate-1 catalysts. ACS Catal. 2018, 7799–7808. [Google Scholar] [CrossRef]
  66. Varganov, S.A.; Olson, R.M.; Gordon, M.S.; Mills, G.; Metiu, H. A study of the reactions of molecular hydrogen with small gold clusters. J. Chem. Phys. 2004, 120, 5169–5175. [Google Scholar] [CrossRef] [PubMed]
  67. Okumura, M.; Kitagawa, Y.; Haruta, M.; Yamaguchi, K. The interaction of neutral and charged Au clusters with O2, CO and H2. Appl. Catal. A 2005, 291, 37–44. [Google Scholar] [CrossRef]
  68. Gao, M.; Lyalin, A.; Takagi, M.; Maeda, S.; Taketsugu, T. Reactivity of gold clusters in the regime of structural fluxionality. J. Phys. Chem. C 2015, 119, 11120–11130. [Google Scholar] [CrossRef]
  69. Barrio, L.; Liu, P.; Rodríguez, J.A.; Campos-Martín, J.M.; Fierro, J.L.G. A density functional theory study of the dissociation of H2 on gold clusters: Importance of fluxionality and ensemble effects. J. Chem. Phys. 2006, 125, 164715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Barrio, L.; Liu, P.; Rodriguez, J.A.; Campos-Martin, J.M.; Fierro, J.L.G. Effects of hydrogen on the reactivity of O2 toward gold nanoparticles and surfaces. J. Phys. Chem. C 2007, 111, 19001–19008. [Google Scholar] [CrossRef]
  71. Hammer, B.; Norskov, J.K. Why gold is the noblest of all the metals. Nature 1995, 376, 238–240. [Google Scholar] [CrossRef]
  72. Jelínek, P.; Pérez, R.; Ortega, J.; Flores, F. Hydrogen dissociation over Au nanowires and the fractional conductance quantum. Phys. Rev. Lett. 2006, 96, 046803. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, B.; Cao, X.-M.; Gong, X.-Q.; Hu, P. A density functional theory study of hydrogen dissociation and diffusion at the perimeter sites of Au/TiO2. Phys. Chem. Chem. Phys. 2012, 14, 3741–3745. [Google Scholar] [CrossRef] [PubMed]
  74. Sun, K.; Kohyama, M.; Tanaka, S.; Takeda, S. A study on the mechanism for H2 dissociation on Au/TiO2 catalysts. J. Phys. Chem. C 2014, 118, 1611–1617. [Google Scholar] [CrossRef]
  75. Green, I.X.; Tang, W.; Neurock, M.; Yates, J.T. Low-temperature catalytic H2 oxidation over Au nanoparticle/TiO2 dual perimeter sites. Angew. Chem. Int. Ed. 2011, 50, 10186–10189. [Google Scholar] [CrossRef] [PubMed]
  76. Fujitani, T.; Nakamura, I.; Akita, T.; Okumura, M.; Haruta, M. Hydrogen dissociation by gold clusters. Angew. Chem. Int. Ed. 2009, 48, 9515–9518. [Google Scholar] [CrossRef] [PubMed]
  77. Yi, Y.; Wang, L.; Li, G.; Guo, H. A review on research progress in the direct synthesis of hydrogen peroxide from hydrogen and oxygen: Noble-metal catalytic method, fuel-cell method and plasma method. Catal. Sci. Technol. 2016, 6, 1593–1610. [Google Scholar] [CrossRef]
  78. Solvay, Solvay’s Position and Strategy in Hydrogen Peroxide. London Investors Morning. Available online: http://www.solvay.com/EN/Investors/FinancialCalendarAndPresentations/Financial-Calendar/Calendar%20Documents/20100930_London-Morning.pdf (accessed on 30 September 2010).
  79. Ronald, H.; Achim, L. Applications of transition-metal catalysts to textile and wood-pulp bleaching. Angew. Chem. Int. Ed. 2006, 45, 206–222. [Google Scholar]
  80. Hess, H.T. Kirk-Othmer Encyclopedia of Chemical Engineering; Wiley: New York, NY, USA, 1995; p. 961. [Google Scholar]
  81. Pugge, J.W.a.L. Catalysis and Zeolites: Fundamentals and Applications; Springer: Berlin, Germany, 1996. [Google Scholar]
  82. Lane, B.S.; Burgess, K. Metal-catalyzed epoxidations of alkenes with hydrogen peroxide. Chem. Rev. 2003, 103, 2457–2474. [Google Scholar] [CrossRef] [PubMed]
  83. Wells, D.H.; Delgass, W.N.; Thomson, K.T. Formation of hydrogen peroxide from H2 and O2 over a neutral gold trimer: A DFT study. J. Catal. 2004, 225, 69–77. [Google Scholar] [CrossRef]
  84. Kacprzak, K.A.; Akola, J.; Hakkinen, H. First-principles simulations of hydrogen peroxide formation catalyzed by small neutral gold clusters. Phys. Chem. Chem. Phys. 2009, 11, 6359–6364. [Google Scholar] [CrossRef] [PubMed]
  85. Joshi, A.M.; Delgass, W.N.; Thomson, K.T. Comparison of the catalytic activity of Au3, Au 4 + , Au5, and Au 5 in the gas-phase reaction of H2 and O2 to form hydrogen peroxide: A density functional theory investigation. J. Phys. Chem. B 2005, 109, 22392–22406. [Google Scholar] [CrossRef] [PubMed]
  86. Ford, D.C.; Nilekar, A.U.; Xu, Y.; Mavrikakis, M. Partial and complete reduction of O2 by hydrogen on transition metal surfaces. Surf. Sci. 2010, 604, 1565–1575. [Google Scholar] [CrossRef]
  87. Landon, P.; Collier, P.J.; Papworth, A.J.; Kiely, C.J.; Hutchings, G.J. Direct formation of hydrogen peroxide from H2/O2 using a gold catalyst. Chem. Commun. 2002, 2058–2059. [Google Scholar] [CrossRef]
  88. Joshi, A.M.; Delgass, W.N.; Thomson, K.T. Investigation of gold–silver, gold–copper, and gold–palladium dimers and trimers for hydrogen peroxide formation from H2 and O2. J. Phys. Chem. C 2007, 111, 7384–7395. [Google Scholar] [CrossRef]
  89. Ham, H.C.; Hwang, G.S.; Han, J.; Nam, S.W.; Lim, T.H. On the role of Pd ensembles in selective H2O2 formation on PdAu alloys. J. Phys. Chem. C 2009, 113, 12943–12945. [Google Scholar] [CrossRef]
  90. Ham, H.C.; Hwang, G.S.; Han, J.; Nam, S.W.; Lim, T.H. Geometric parameter effects on ensemble contributions to catalysis: H2O2 formation from H2 and O2 on AuPd alloys. A first principles study. J. Phys. Chem. C 2010, 114, 14922–14928. [Google Scholar] [CrossRef]
  91. Liu, P.; Norskov, J.K. Ligand and ensemble effects in adsorption on alloy surfaces. Phys. Chem. Chem. Phys. 2001, 3, 3814–3818. [Google Scholar] [CrossRef]
  92. Beletskaya, A.V.; Pichugina, D.A.; Shestakov, A.F.; Kuz’menko, N.E. Formation of H2O2 on Au20 and Au19Pd clusters: Understanding the structure effect on the atomic level. J. Phys. Chem. A 2013, 117, 6817–6826. [Google Scholar] [CrossRef] [PubMed]
  93. Staykov, A.; Kamachi, T.; Ishihara, T.; Yoshizawa, K. Theoretical study of the direct synthesis of H2O2 on Pd and Pd/Au surfaces. J. Phys. Chem. C 2008, 112, 19501–19505. [Google Scholar] [CrossRef]
  94. Li, J.; Staykov, A.; Ishihara, T.; Yoshizawa, K. Theoretical study of the decomposition and hydrogenation of H2O2 on Pd and Au@Pd surfaces: Understanding toward high selectivity of H2O2 synthesis. J. Phys. Chem. C 2011, 115, 7392–7398. [Google Scholar] [CrossRef]
  95. 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]
  96. Todorovic, R.; Meyer, R.J. A comparative density functional theory study of the direct synthesis of H2O2 on Pd, Pt and Au surfaces. Catal. Today 2011, 160, 242–248. [Google Scholar] [CrossRef]
  97. Grabow, L.C.; Hvolbæk, B.; Falsig, H.; Nørskov, J.K. Search directions for direct H2O2 synthesis catalysts starting from Au12 nanoclusters. Top. Catal. 2012, 55, 336–344. [Google Scholar] [CrossRef]
  98. Uphade, B.S.; Yamada, Y.; Akita, T.; Nakamura, T.; Haruta, M. Synthesis and characterization of Ti-MCM-41 and vapor-phase epoxidation of propylene using H2 and O2 over Au/Ti-MCM-41. Appl. Catal. A 2001, 215, 137–148. [Google Scholar] [CrossRef]
  99. Taramasso, M.; Perego, G.; Notari, B. Preparation of Porous Crystalline Synthetic Material Comprised of Silicon and Titanium Oxides. U.S. Patent 4,410,501, 18 October 1983. [Google Scholar]
  100. Wenping, F.; Yaquan, W.; Guoqiang, W.; Yi, L.; Juan, X.; Hainan, S.; Teng, Z.; Shuhai, W.; Xiaoxue, W.; Pengxu, Y. Liquid phase propylene epoxidation with H2O2 on TS-1/SiO2 catalyst in a fixed-bed reactor: Experiments and deactivation kinetics. J. Chem. Technol. Biotechnol. 2015, 90, 1489–1496. [Google Scholar]
  101. Clerici, M.G.; Bellussi, G.; Romano, U. Synthesis of propylene oxide from propylene and hydrogen peroxide catalyzed by titanium silicalite. J. Catal. 1991, 129, 159–167. [Google Scholar] [CrossRef]
  102. Sivadinarayana, C.; Choudhary, T.V.; Daemen, L.L.; Eckert, J.; Goodman, D.W. The nature of the surface species formed on Au/TiO2 during the reaction of H2 and O2: An inelastic neutron scattering study. J. Am. Chem. Soc. 2004, 126, 38–39. [Google Scholar] [CrossRef] [PubMed]
  103. Chowdhury, B.; Bravo-Suárez, J.J.; Mimura, N.; Lu, J.; Bando, K.K.; Tsubota, S.; Haruta, M. In situ UV−vis and EPR study on the formation of hydroperoxide species during direct gas phase propylene epoxidation over Au/Ti-SiO2 catalyst. J. Phys. Chem. B 2006, 110, 22995–22999. [Google Scholar] [CrossRef] [PubMed]
  104. Wells, D.H.; Joshi, A.M.; Delgass, W.N.; Thomson, K.T. A quantum chemical study of comparison of various propylene epoxidation mechanisms using H2O2 and TS-1 catalyst. J. Phys. Chem. B 2006, 110, 14627–14639. [Google Scholar] [CrossRef] [PubMed]
  105. Bellussi, G.; Carati, A.; Clerici, M.G.; Maddinelli, G.; Millini, R. Reactions of titanium silicalite with protic molecules and hydrogen peroxide. J. Catal. 1992, 133, 220–230. [Google Scholar] [CrossRef]
  106. Sinclair, P.E.; Catlow, C.R.A. Quantum chemical study of the mechanism of partial oxidation reactivity in titanosilicate catalysts: Active site formation, oxygen transfer, and catalyst deactivation. J. Phys. Chem. B 1999, 103, 1084–1095. [Google Scholar] [CrossRef]
  107. Vayssilov, G.N.; van Santen, R.A. Catalytic activity of titanium silicalites—A DFT study. J. Catal. 1998, 175, 170–174. [Google Scholar] [CrossRef]
  108. Munakata, H.; Oumi, Y.; Miyamoto, A. A DFT study on peroxo-complex in titanosilicate catalyst: Hydrogen peroxide activation on titanosilicalite-1 catalyst and reaction mechanisms for catalytic olefin epoxidation and for hydroxylamine formation from ammonia. J. Phys. Chem. B 2001, 105, 3493–3501. [Google Scholar] [CrossRef]
  109. Nie, X.; Ji, X.; Chen, Y.; Guo, X.; Song, C. Mechanistic investigation of propylene epoxidation with H2O2 over TS-1: Active site formation, intermediate identification, and oxygen transfer pathway. Mol. Catal. 2017, 441, 150–167. [Google Scholar] [CrossRef]
  110. Panayotov, D.; McEntee, M.; Burrows, S.; Driscoll, D.; Tang, W.; Neurock, M.; Morris, J. Infrared studies of propene and propene oxide adsorption on nanoparticulate Au/TiO2. Surf. Sci. 2016, 652, 172–182. [Google Scholar] [CrossRef]
  111. Driscoll, D.M.; Tang, W.; Burrows, S.P.; Panayotov, D.A.; Neurock, M.; McEntee, M.; Morris, J.R. Binding sites, geometry, and energetics of propene at nanoparticulate Au/TiO2. J. Phys. Chem. C 2017, 121, 1683–1689. [Google Scholar] [CrossRef]
  112. Bocquet, M.-L.; Loffreda, D. Ethene epoxidation selectivity inhibited by twisted oxametallacycle: A DFT study on Ag surface-oxide. J. Am. Chem. Soc. 2005, 127, 17207–17215. [Google Scholar] [CrossRef] [PubMed]
  113. Mavrikakis, M.; Doren, D.J.; Barteau, M.A. Density functional theory calculations for simple oxametallacycles: Trends across the periodic table. J. Phys. Chem. B 1998, 102, 394–399. [Google Scholar] [CrossRef]
  114. Torres, D.; Lopez, N.; Illas, F.; Lambert, R.M. Why copper is intrinsically more selective than silver in alkene epoxidation: Ethylene oxidation on Cu(111) versus Ag(111). J. Am. Chem. Soc. 2005, 127, 10774–10775. [Google Scholar] [CrossRef] [PubMed]
  115. Jones, G.S.; Mavrikakis, M.; Barteau, M.A.; Vohs, J.M. First synthesis, experimental and theoretical vibrational spectra of an oxametallacycle on a metal surface. J. Am. Chem. Soc. 1998, 120, 3196–3204. [Google Scholar] [CrossRef]
  116. Linic, S.; Barteau, M.A. Formation of a stable surface oxametallacycle that produces ethylene oxide. J. Am. Chem. Soc. 2002, 124, 310–317. [Google Scholar] [CrossRef] [PubMed]
  117. Linic, S.; Medlin, J.W.; Barteau, M.A. Synthesis of oxametallacycles from 2-iodoethanol on Ag(111) and the structure dependence of their reactivity. Langmuir 2002, 18, 5197–5204. [Google Scholar] [CrossRef]
  118. Özbek, M.O.; van Santen, R.A. The mechanism of ethylene epoxidation catalysis. Catal. Lett. 2013, 143, 131–141. [Google Scholar] [CrossRef]
  119. Joshi, A.M.; Delgass, W.N.; Thomson, K.T. Partial oxidation of propylene to propylene oxide over a neutral gold trimer in the gas phase: A density functional theory study. J. Phys. Chem. B 2006, 110, 2572–2581. [Google Scholar] [CrossRef] [PubMed]
  120. Roldan, A.; Torres, D.; Ricart, J.M.; Illas, F. On the effectiveness of partial oxidation of propylene by gold: A density functional theory study. J. Mol. Catal. A Chem. 2009, 306, 6–10. [Google Scholar] [CrossRef]
  121. Deng, X.; Min, B.K.; Liu, X.; Friend, C.M. Partial oxidation of propene on oxygen-covered Au(111). J. Phys. Chem. B 2006, 110, 15982–15987. [Google Scholar] [CrossRef] [PubMed]
  122. Cant, N.W.; Hall, W.K. Catalytic oxidation. Iv. Ethylene and propylene oxidation over gold. J. Phys. Chem. 1971, 75, 2914–2921. [Google Scholar] [CrossRef]
  123. Geenen, P.V.; Boss, H.J.; Pott, G.T. A study of the vapor-phase epoxidation of propylene and ethylene on silver and silver-gold alloy catalysts. J. Catal. 1982, 77, 499–510. [Google Scholar] [CrossRef]
  124. Mullen, G.M.; Zhang, L.; Evans, E.J.; Yan, T.; Henkelman, G.; Mullins, C.B. Oxygen and hydroxyl species induce multiple reaction pathways for the partial oxidation of allyl alcohol on gold. J. Am. Chem. Soc. 2014, 136, 6489–6498. [Google Scholar] [CrossRef] [PubMed]
  125. Mullen, G.M.; Zhang, L.; Evans, E.J.; Yan, T.; Henkelman, G.; Mullins, C.B. Control of selectivity in allylic alcohol oxidation on gold surfaces: The role of oxygen adatoms and hydroxyl species. Phys. Chem. Chem. Phys. 2015, 17, 4730–4738. [Google Scholar] [CrossRef] [PubMed]
  126. Ojeda, M.; Iglesia, E. Catalytic epoxidation of propene with H2O-O2 reactants on Au/TiO2. Chem. Commun. 2009, 352–354. [Google Scholar] [CrossRef] [PubMed]
  127. Lee, S.; Molina, L.M.; Lopez, M.J.; Alonso, J.A.; Hammer, B.; Lee, B.; Seifert, S.; Winans, R.E.; Elam, J.W.; Pellin, M.J.; Vajda, S. Selective propene epoxidation on immobilized Au6–10 clusters: The effect of hydrogen and water on activity and selectivity. Angew. Chem. Int. Ed. 2009, 48, 1467–1471. [Google Scholar] [CrossRef] [PubMed]
  128. Jiahui, H.; Tomoki, A.; Jérémy, F.; Tadahiro, F.; Takashi, T.; Masatake, H. Propene epoxidation with dioxygen catalyzed by gold clusters. Angew. Chem. Int. Ed. 2009, 48, 7862–7866. [Google Scholar]
  129. Bongiorno, A.; Landman, U. Water-enhanced catalysis of CO oxidation on free and supported gold nanoclusters. Phys. Rev. Lett. 2005, 95, 106102. [Google Scholar] [CrossRef] [PubMed]
  130. Chang, C.-R.; Wang, Y.-G.; Li, J. Theoretical investigations of the catalytic role of water in propene epoxidation on gold nanoclusters: A hydroperoxyl-mediated pathway. Nano Res. 2011, 4, 131–142. [Google Scholar] [CrossRef]
  131. Chang, C.-R.; Huang, Z.-Q.; Li, J. Hydrogenation of molecular oxygen to hydroperoxyl: An alternative pathway for O2 activation on nanogold catalysts. Nano Res. 2015, 8, 3737–3748. [Google Scholar] [CrossRef]
  132. Liu, J.C.; Tang, Y.; Chang, C.R.; Wang, Y.G.; Li, J. Mechanistic insights into propene epoxidation with O2–H2O mixture on Au7/α-Al2O3: A hydroproxyl pathway from ab initio molecular dynamics simulations. ACS Catal. 2016, 6, 2525–2535. [Google Scholar] [CrossRef]
  133. Joshi, A.M.; Delgass, W.N.; Thomson, K.T. Mechanistic implications of Aun/Ti-lattice proximity for propylene epoxidation. J. Phys. Chem. C 2007, 111, 7841–7844. [Google Scholar] [CrossRef]
  134. Ajo, H.M.; Bondzie, V.A.; Campbell, C.T. Propene adsorption on gold particles on TiO2(110). Catal. Lett. 2002, 78, 359–368. [Google Scholar] [CrossRef]
  135. Nijhuis, T.A.; Gardner, T.Q.; Weckhuysen, B.M. Modeling of kinetics and deactivation in the direct epoxidation of propene over gold–titania catalysts. J. Catal. 2005, 236, 153–163. [Google Scholar] [CrossRef]
  136. Lee, W.S.; Lai, L.C.; Cem Akatay, M.; Stach, E.A.; Ribeiro, F.H.; Delgass, W.N. Probing the gold active sites in Au/TS-1 for gas-phase epoxidation of propylene in the presence of hydrogen and oxygen. J. Catal. 2012, 296, 31–42. [Google Scholar] [CrossRef]
  137. Haruta, M. Size- and support-dependency in the catalysis of gold. Catal. Today 1997, 36, 153–166. [Google Scholar] [CrossRef]
  138. Nijhuis, T.A.; Visser, T.; Weckhuysen, B.M. The role of gold in gold–titania epoxidation catalysts. Angew. Chem. Int. Ed. 2005, 44, 1115–1118. [Google Scholar] [CrossRef] [PubMed]
  139. Turner, C.H.; Ji, J.; Lu, Z.; Lei, Y. Analysis of the propylene epoxidation mechanism on supported gold nanoparticles. Chem. Eng. Sci. 2017, 174, 229–237. [Google Scholar] [CrossRef]
  140. Lu, Z.; Piernavieja-Hermida, M.; Turner, C.H.; Wu, Z.; Lei, Y. Effects of TiO2 in low temperature propylene epoxidation using gold catalysts. J. Phys. Chem. C 2018, 122, 1688–1698. [Google Scholar] [CrossRef]
  141. Ji, J.; Lu, Z.; Lei, Y.; Turner, C.H. Mechanistic insights into the direct propylene epoxidation using Au nanoparticles dispersed on TiO2/SiO2. Chem. Eng. Sci. 2018, 191, 169–182. [Google Scholar] [CrossRef]
  142. Chen, J.; Halin, S.J.A.; Schouten, J.C.; Nijhuis, T.A. Kinetic study of propylene epoxidation with H2 and O2 over Au/Ti-SiO2 in the explosive regime. Faraday Discuss. 2011, 152, 321–336. [Google Scholar] [CrossRef] [PubMed]
  143. Taylor, B.; Lauterbach, J.; Blau, G.E.; Delgass, W.N. Reaction kinetic analysis of the gas-phase epoxidation of propylene over Au/TS-1. J. Catal. 2006, 242, 142–152. [Google Scholar] [CrossRef]
  144. Zwijnenburg, A.; Makkee, M.; Moulijn, J.A. Increasing the low propene epoxidation product yield of gold/titania-based catalysts. Appl. Catal. A 2004, 270, 49–56. [Google Scholar] [CrossRef]
  145. Green, I.X.; Tang, W.; Neurock, M.; Yates, J.T. Localized partial oxidation of acetic acid at the dual perimeter sites of the Au/TiO2 catalyst—Formation of gold ketenylidene. J. Am. Chem. Soc. 2012, 134, 13569–13572. [Google Scholar] [CrossRef] [PubMed]
  146. Green, I.X.; Tang, W.; Neurock, M.; Yates, J.T. Mechanistic insights into the partial oxidation of acetic acid by O2 at the dual perimeter sites of a Au/TiO2 catalyst. Faraday Discuss. 2013, 162, 247–265. [Google Scholar] [CrossRef] [PubMed]
  147. McEntee, M.; Tang, W.; Neurock, M.; Yates, J.T. Selective catalytic oxidative-dehydrogenation of carboxylic acids—Acrylate and crotonate formation at the Au/TiO2 interface. J. Am. Chem. Soc. 2014, 136, 5116–5120. [Google Scholar] [CrossRef] [PubMed]
  148. Ntainjua, E.N.; Freakley, S.J.; Hutchings, G.J. Direct synthesis of hydrogen peroxide using ruthenium catalysts. Top. Catal. 2012, 55, 718–722. [Google Scholar] [CrossRef]
  149. 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]
  150. Gudarzi, D.; Ratchananusorn, W.; Turunen, I.; Heinonen, M.; Salmi, T. Promotional effects of au in Pd–Au bimetallic catalysts supported on activated carbon cloth (ACC) for direct synthesis of H2O2 from H2 and O2. Catal. Today 2015, 248, 58–68. [Google Scholar] [CrossRef]
  151. Yohei, N.; Tatsumi, I.; Yuiko, H.; Kotaro, K.; Kenji, K.; Hiroshige, M. Nanocolloidal Pd-Au as catalyst for the direct synthesis of hydrogen peroxide from H2 and O2. ChemSusChem 2008, 1, 619–621. [Google Scholar]
  152. Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. Gold catalysts prepared by coprecipitation for low-temperature oxidation of hydrogen and of carbon monoxide. J. Catal. 1989, 115, 301–309. [Google Scholar] [CrossRef]
  153. Haruta, M. Catalysis of gold nanoparticles deposited on metal oxides. CATTECH 2002, 6, 102–115. [Google Scholar] [CrossRef]
  154. Hashmi, A.S.; Hutchings, G.J. Gold catalysis. Angew. Chem. Int. Ed. 2006, 45, 7896–7936. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Catalytic cycle for the formation of H2O2 over Au3 (Au3 cluster is represented by the triangle in each intermediate geometry) with two transition state geometries (TS A and TS D), identified with square brackets, and intermediates (AD). Energy differences are shown in kcal/mol under each reaction arrow [83]. Copyright © 2004 Elsevier Inc.
Figure 1. Catalytic cycle for the formation of H2O2 over Au3 (Au3 cluster is represented by the triangle in each intermediate geometry) with two transition state geometries (TS A and TS D), identified with square brackets, and intermediates (AD). Energy differences are shown in kcal/mol under each reaction arrow [83]. Copyright © 2004 Elsevier Inc.
Catalysts 08 00421 g001
Figure 2. Schematic view of (A) a fully tetrahedral Ti site bonded with four O–Si groups, and (B) the same Ti site located near a silicon vacancy forming a silanol nest [32]. Copyright © 2004 American Chemical Society.
Figure 2. Schematic view of (A) a fully tetrahedral Ti site bonded with four O–Si groups, and (B) the same Ti site located near a silicon vacancy forming a silanol nest [32]. Copyright © 2004 American Chemical Society.
Catalysts 08 00421 g002
Figure 3. Transition state for the reaction of the hydroperoxy intermediate on a defect Ti site with propylene [32]. (Ti = large white, O = red, Si = blue, H = small white spheres). Copyright © 2004 American Chemical Society.
Figure 3. Transition state for the reaction of the hydroperoxy intermediate on a defect Ti site with propylene [32]. (Ti = large white, O = red, Si = blue, H = small white spheres). Copyright © 2004 American Chemical Society.
Catalysts 08 00421 g003
Figure 4. Munakata et al. mechanism: (a) the formation of Si–O–O–Ti species, where (1) is the preadsorbed complex of H2O2 on TS-1, (2) is the transition-state geometry for the first step in oxygen insertion to the peroxo intermediate on the closed Ti site, (3) is the stable intermediate geometry prior to the second step in oxygen insertion to the peroxo intermediate, and (4) is the transition-state geometry for the second step in oxygen insertion to the peroxo intermediate; (b) propylene epoxidation, where (1) is the preadsorbed complex of propylene with the peroxo intermediate on the closed Ti site in TS-1, and (2) is the transition-state geometry during the propylene epoxidation [104]. Distances shown are in units of angstroms. Copyright © 2006 American Chemical Society.
Figure 4. Munakata et al. mechanism: (a) the formation of Si–O–O–Ti species, where (1) is the preadsorbed complex of H2O2 on TS-1, (2) is the transition-state geometry for the first step in oxygen insertion to the peroxo intermediate on the closed Ti site, (3) is the stable intermediate geometry prior to the second step in oxygen insertion to the peroxo intermediate, and (4) is the transition-state geometry for the second step in oxygen insertion to the peroxo intermediate; (b) propylene epoxidation, where (1) is the preadsorbed complex of propylene with the peroxo intermediate on the closed Ti site in TS-1, and (2) is the transition-state geometry during the propylene epoxidation [104]. Distances shown are in units of angstroms. Copyright © 2006 American Chemical Society.
Catalysts 08 00421 g004
Figure 5. Ti/defect mechanism for the full silanol system: (a) transition-state geometry for the formation of the hydroperoxy intermediate on the Ti site of TS-1, and (b) transition-state geometry for propylene epoxidation [104]. Distances shown are in units of angstroms. Copyright © 2006 American Chemical Society.
Figure 5. Ti/defect mechanism for the full silanol system: (a) transition-state geometry for the formation of the hydroperoxy intermediate on the Ti site of TS-1, and (b) transition-state geometry for propylene epoxidation [104]. Distances shown are in units of angstroms. Copyright © 2006 American Chemical Society.
Catalysts 08 00421 g005
Figure 6. Adsorption structure and stretching frequencies of the C=C bond of propylene on Au/TiO2 catalyst [111]. Copyright © 2017 American Chemical Society.
Figure 6. Adsorption structure and stretching frequencies of the C=C bond of propylene on Au/TiO2 catalyst [111]. Copyright © 2017 American Chemical Society.
Catalysts 08 00421 g006
Figure 7. Structures of the transition state and final product during the epoxidation step on Au3. Distances are in angstroms [119]. Copyright © 2006 American Chemical Society.
Figure 7. Structures of the transition state and final product during the epoxidation step on Au3. Distances are in angstroms [119]. Copyright © 2006 American Chemical Society.
Catalysts 08 00421 g007
Figure 8. Structures of oxametallacycle intermediates: OMMP1 (left) and OMMP2 (right) on Au(111) [120]. Copyright © 2009 Elsevier B.V.
Figure 8. Structures of oxametallacycle intermediates: OMMP1 (left) and OMMP2 (right) on Au(111) [120]. Copyright © 2009 Elsevier B.V.
Catalysts 08 00421 g008
Figure 9. PO formation (a) on Au38 cluster where two different pathways are shown with their corresponding transition-state geometries and activation energies; (b) on an Au10 cluster where only one path is possible, along with the corresponding transition-state geometry and activation energy [130]. Copyright © 2011 Elsevier B.V.
Figure 9. PO formation (a) on Au38 cluster where two different pathways are shown with their corresponding transition-state geometries and activation energies; (b) on an Au10 cluster where only one path is possible, along with the corresponding transition-state geometry and activation energy [130]. Copyright © 2011 Elsevier B.V.
Catalysts 08 00421 g009
Figure 10. Reaction network at the Au/Ti interface. The * symbol represents a bare Au site, and represents a bare interfacial Ti site that is in contact with Au nanoparticles (for clarity, Au nanoparticle is not shown here). The species attached to * or indicates one Au surface-bound species or one Ti surface-bound species, respectively [141]. Copyright © 2018 Elsevier Ltd.
Figure 10. Reaction network at the Au/Ti interface. The * symbol represents a bare Au site, and represents a bare interfacial Ti site that is in contact with Au nanoparticles (for clarity, Au nanoparticle is not shown here). The species attached to * or indicates one Au surface-bound species or one Ti surface-bound species, respectively [141]. Copyright © 2018 Elsevier Ltd.
Catalysts 08 00421 g010

Share and Cite

MDPI and ACS Style

Ji, J.; Lu, Z.; Lei, Y.; Turner, C.H. Theoretical Studies on the Direct Propylene Epoxidation Using Gold-Based Catalysts: A Mini-Review. Catalysts 2018, 8, 421. https://doi.org/10.3390/catal8100421

AMA Style

Ji J, Lu Z, Lei Y, Turner CH. Theoretical Studies on the Direct Propylene Epoxidation Using Gold-Based Catalysts: A Mini-Review. Catalysts. 2018; 8(10):421. https://doi.org/10.3390/catal8100421

Chicago/Turabian Style

Ji, Jingjing, Zheng Lu, Yu Lei, and C. Heath Turner. 2018. "Theoretical Studies on the Direct Propylene Epoxidation Using Gold-Based Catalysts: A Mini-Review" Catalysts 8, no. 10: 421. https://doi.org/10.3390/catal8100421

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