Next Article in Journal
AFX Zeolite for Use as a Support of NH3-SCR Catalyst Mining through AICE Joint Research Project of Industries–Academia–Academia
Next Article in Special Issue
Influence of Co-Precipitation Agent on the Structure, Texture and Catalytic Activity of Au-CeO2 Catalysts in Low-Temperature Oxidation of Benzyl Alcohol
Previous Article in Journal
Surface Modifications of 2D-Ti3C2O2 by Nonmetal Doping for Obtaining High Hydrogen Evolution Reaction Activity: A Computational Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Synthesis of Methyl Methacrylate by One Step Oxidative Esterification over Zn-Al-Mixed Oxides Supported Gold Nanocatalysts

1
Hangzhou Institute of Advanced studies, Zhejiang Normal University, Hangzhou 311231, China
2
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
3
The State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
*
Authors to whom correspondence should be addressed.
Submission received: 18 December 2020 / Revised: 12 January 2021 / Accepted: 21 January 2021 / Published: 25 January 2021
(This article belongs to the Special Issue Advance in Selective Alcohol and Polyol Oxidation Catalysis)

Abstract

:
Methyl methacrylate (MMA) is an important monomer in fine chemicals. The synthesis of MMA by one-step oxidative esterification from methacrolein with methanol over a heterogeneous catalyst with high activity, selectivity and stability is highly desirable. Herein, Zn-Al-hydrotalcites (HTs)-supported atomically precise Au25 nanoclusters with different molar ratios of Zn2+/Al3+ were prepared and used as the precursors for this reaction. They exhibited good performances in comparison with the gold catalysts prepared by the deposition precipitation method. The structural and electronic properties were evaluated by various characterization technologies, including X-ray diffraction (XRD), high-resolution transmission electron microscopy (HRTEM), in situ diffuse reflectance infrared Fourier-transform spectroscopy (DRIFTS) of CO adsorption, X-ray photoelectron spectroscopy (XPS), and CO2 temperature-programmed desorption (TPD). The combined characterization results suggested that the adsorption property of gold and the basicity of the catalyst contributes to their high activities. Substrates extended experiments and stability tests implied the potential application of Zn-Al-mixed oxides supported gold catalysts, which paves a new way for supported gold catalyst in the one-step oxidation esterification reaction.

Graphical Abstract

1. Introduction

Methyl methacrylate (MMA) is an important monomer for the synthesis of poly-methyl methacrylate (PMMA), which is widely used for producing acrylic plastic [1]. A recent analysis originated from the global MMA market indicated that the demand for MMA may reach $9.8 billion by 2027 [2]. Thus, a noteworthy growth in MMA monomers will appear in the next several years. Traditionally, the dominant industrial process for the synthesis of MMA has been the acetone cyanohydrin (ACH) route [3], in which acetone and hydrocyanic acid are used as the raw materials. At the same time, the ACH process involves environmentally unfriendly issues that employ strong sulfuric acid as a catalyst and inevitably generate ammonium bisulfate waste [3,4].
Currently, isobutene (IB) oxidation routes that stem from isobutene and C4 ingredients for producing MMA are attracting great interest due to the economic and environmental preponderance [5,6]. In these processes, isobutene is first oxidized to methacrolein (MAL) and then oxidized to methacrylic acid (MAA), followed by esterification of MAA to MMA [5]. Instead, MAL could also react with methanol to form MMA by one-step oxidative esterification (Scheme 1) [6]. In view of energy efficiency and environmental cost, the direct oxidative esterification process is more attractive in comparison with the two-step method. However, it is still very challenging for the development of efficient and selective catalytic systems under moderate conditions by a one-step method [7,8].
Supported gold catalysts are generally believed to exhibit good performances in oxidation reactions under moderate conditions [9,10,11,12,13,14]. The special property enables it to be a potential candidate to replace traditional Pd-based catalysts in one-step oxidative esterification [15,16,17,18,19,20,21,22]. Previously, K. Suzuki at Asahi Kasei developed a SiO2–Al2O3–MgO multiple oxides-supported AuNiOx catalyst with core–shell structure [15]. The catalyst showed good performance for MAL to MMA, with the conversion of ∼60% and selectivity of 96–97%. After this, Wang et al. observed good activity and selectivity over the Au/MgO catalyst, which showed conversion of 98% and selectivity of 99%, respectively [16]. Moreover, gold catalysts supported on different transition metal oxides and hydroxyapatite have been prepared in this reaction to evaluate the catalytic performances [17,18,19,20,21,22]. Their comparing results are shown in Table S1. Based on these results, the property of the support and/or the particle size of gold has been demonstrated to be crucial for one-step oxidative esterification [16,17,21]. The support with strong basic sites and the gold with a small average particle size are said to benefit the transformation of MAL to MMA [16,21]. However, although the activity and selectivity of supported gold catalysts exhibited great advantages to traditional Pd-based catalysts, the stability was among one of the big challenges.
Recently, layered double hydroxides (HTs) or hydrotalcite-type compounds, which have a general formula of [M1−x2+Mx3+(OH)2]x+·[(An−)x/n]·yH2O, have attracted great interest due to their high surface area, homogeneously dispersed metal atoms and acid–base properties [23,24,25]. Up to now, they have been utilized in a wide variety of oxidative reactions, such as the oxidation of benzyl alcohol [26], the oxidative steam reforming of methanol [27], the styrene epoxidation [28] and, etc. Previously, Zn-Al-mixed oxide-supported gold catalysts were demonstrated by us to display special catalytic performances in selective hydrogenation reactions [29,30,31]. The interaction between the gold and the support made it a great candidate for the synthesis of gold catalysts with controllable sizes [29,31]. Moreover, the inert activation for the vinyl group enables it a potential nominee for selective hydrogenation or oxidation reaction.
Herein, to synthesize the gold catalyst with homogeneous size, atomically precisely Au25 nanoclusters were preliminarily prepared to use as a precursor; Zn-Al hydrotalcites with different Zn2+/Al3+ molar ratios were used as the precursors of supports. The catalysts were applied for one-step oxidative esterification of methacrolein with methanol to the synthesis of MMA. In comparison with the catalytic performances of gold catalysts prepared by the deposition precipitation method, the Au25/ZnxAl-400 (x = 3,2,1) catalysts exhibited comparable better activity. N2-physical adsorption–desorption, X-ray diffraction (XRD), transmission electron microscopy (TEM), in situ diffuse reflectance infrared Fourier-transform spectroscopy (DRIFTS) of CO adsorption, X-ray photoelectron spectroscopy (XPS) and CO2 temperature-programmed desorption (TPD) were utilized to characterize the structural and electronic properties of the catalysts, from which the relationship about the structure and catalytic performances were reasonably revealed.

2. Results and Discussion

2.1. Catalytic Performances

Zn-Al-HTs with different molar ratios supported gold nanoclusters were used as the precursors of the catalysts. Their catalytic performances were successively investigated in one-step oxidative esterification of MAL with methanol (MeOH) for the synthesis of MMA (Scheme 2).
From the results, the supported gold catalyst with Zn2+/Al3+ molar ratio of 3 showed conversion of 84.9% and selectivity of 95.1% as the reaction was conducted at 353 K and 3 atm O2 for 2 h. (Table 1, entry 1). When the Zn2+/Al3+ ratio decreased from 3 to 2, the conversion of MAL increased to 93.1%, with the selectivity of MMA maintained at 94.8% (Table 1, entry 2). Further decreased the Zn2+/Al3+ ratio to 1, both the activity and selectivity reduced, with the value being 90% and 92%, respectively (Table 1, entry 3). To be noted, although the final conversion of MAL was the highest over the Au25/Zn2Al-400 catalyst, the turnover frequency (TOF) based on total gold metals was slightly lower than that of the Au25/Zn3Al-400 and Au25/Zn1Al-400 catalysts. The reason may be originated from the influence of the Zn2+/Al3+ molar ratio on the adsorption of reactant, which leads to the difference of initial reaction rate to final reaction rate.
Hence, to study the evolution of the product distribution during the reaction process of MAL with MeOH, the dynamic experiments were carried out on the Au25/ZnxAl-400 catalysts with various Zn2+/Al3+ molar ratios. The yields of MAL and products with reaction time were plotted in Figure 1a–c. From the results, the product distributions over the three catalysts were similar within the reaction progress. After prolonging the reaction time, MMA was formed as the sacrifice of MAL. In the meantime, few acetals and byproducts appeared via the reaction of MAL and impurity aldehyde with MeOH (Figure S1). The reaction rates over the three catalysts decreased gradually from beginning to end, suggesting the dependence of activity on the concentration of the reactant. However, the selectivity of MMA could be well maintained at above 90% despite the various Zn2+/Al3+ ratios.
Under the same conditions, the supported gold catalysts prepared by the traditional deposition–precipitation (DP) method were synthesized for this reaction to make a comparison. The results are shown in Table 1, entry 4, with Zn2+/Al3+ molar ratios of 2. As seen, the conversion of MAL over the Au/Zn2Al-400 catalyst was 57.1%, which fell much lower than that of the Au25/ZnxAl-400. Simultaneously, the TOF value was several times lower than that of the Au25/ZnxAl-400. It suggested that the utilization of Au25 nanoclusters as the precursor of gold greatly improves the activity. Following characterizations about the geometric and electronic structure of the gold catalysts were provided to reveal the structure–activity relationship.
Moreover, to disclose the effect of support for this reaction, the heat-treated ZnxAl-HTs at 400 °C were also applied in this reaction under the same conditions (Table 1, entry 5–7). As seen, although 6.4%, 13.2% and 16.9% of the MAL conversions were observed over the Zn3Al-400, Zn2Al-400 and Zn1Al-400 samples, no target product of MMA appeared in the system. It indicated the significant role of gold species towards one-step oxidative esterification of MAL to MMA. Moreover, to eliminate the effect of device deviation, the blank experiment with no catalysts or supports was conducted at the same condition (Table 1, entry 8). Almost no transformation of MAL to MMA could be observed, indicating the reaction between MAL and MeOH indeed occurred on our catalysts. To be noted, when compared the catalytic performance of Au25/Zn2Al-400 with the other catalysts in previous work, the activity and selectivity of our catalysts were among one the best-performing catalysts under similar conditions. Such results are shown in Table S1.

2.2. Structure of the Catalysts

The chemical analysis of Zn-Al-mixed oxides supported gold catalysts were measured by inductively coupled plasma atomic emission spectroscopy (ICP-AES). The loadings of gold on the Au25/Zn3Al-400, Au25/Zn2Al-400, Au25/Zn1Al-400 and Au/Zn2Al-400 catalysts after heat treatment were 1.3, 1.2, 1.3 and 1.3 wt%, respectively, which agreed well with the nominal values. The actual molar ratios of Zn2+/Al3+ were 3.3, 2.4 and 1.2, respectively, with the content of Zn2+ and Al3+ of 21–32 wt% and 3.3–7.4 wt%, respectively (Table S2). The pore structure parameters such as the specific surface area, the pore volume and pore width were evaluated by the N2-physical adsorption–desorption test (Table 2). The isotherms were exhibited in Figure S2. Based on the classification of IUPAC [32,33], the above catalysts exhibited typical IV type isotherms with obviously H3-type hysteresis loop, suggesting that the mesoporous structures existed in the above catalysts, which may result from the collapse of hydrotalcites structure after heat treatment.
The crystallographic phases of the catalysts before and after calcination were analyzed by powder XRD. The patterns are shown in Figure 2. Before the thermal treatment, all of the catalysts exhibited the characteristic reflections of the hydrotalcite lamellar structure with intense diffraction peaks at 11.7°, 23.6°, 34.6°, 39.7°, 46.8°, 60.4° and 61.8° (Figure 2a). These peaks could be assigned to (003), (006), (012), (015), (018), (110) and (113) planes of standard zinc-aluminum hydrotalcite (JCPDS NO. 00-048-1023). After calcination, the lamellar structure of hydrotalcites went into collapse, with the formation of Zn-Al-mixed oxides. The XRD patterns of the samples after heat treatment at 400 °C are shown in Figure 2b, in which obvious sharp peaks corresponding to reflections of ZnO phase (JCPDS NO. 01-089-0510) were observed, with the peaks located at around 32.1°, 34.6°, 36.5°, 47.6°, 56.8° and 62.7°. It demonstrated that Zn and Al atoms were homogeneously distributed in the framework of the catalysts. Noteworthy, no diffraction peaks ascribed as face-centered cubic gold were observed in all of the catalysts, perhaps due to the small particle sizes or low concentration of Au.
In order to examine the particle sizes and lattice parameters of gold, TEM and HRTEM were performed on the Au25/ZnxAl-400 (x = 3, 2, 1) and Au/Zn2Al-400 (DP) catalysts (Figure 3). More than 200 nanoparticles in representative regions were included to count the average particle sizes. As shown from TEM images, all of the gold particles after thermal treatment were uniformly dispersed on the catalysts, with similar particle sizes of 2.8 ± 0.7 nm (Figure 3a1), 2.5 ± 0.8 nm (Figure 3b1), 2.5 ± 0.7 nm (Figure 3c1) and 2.8 ± 0.9 nm (Figure 3d1) over the Au25/Zn3Al-400, Au25/Zn2Al-400, Au25/Zn1Al-400 and Au/Zn2Al-400 catalysts, respectively. Moreover, the gold particles with the shape of hexagon, sphere or ellipsoid could be observed, with clear boundaries between gold and supports. The interplanar spacing of 2.42 Å, 2.60 Å and 2.80 Å corresponded to (101), (100) and (002) lattice fringes of the supports that were consistent with the XRD results.

2.3. Electronic Property of the Catalyst

The in situ DRIFTS of CO adsorption was conducted to characterize the state of gold and active species on the surface of the Au25/ZnxAl-400 and Au/Zn2Al-400 catalysts (Figure 4). The double peaks at 2119 and 2168 cm−1 could be ascribed to the adsorption of the gas phase CO, which disappeared gradually with the purging of helium. An obvious band at around 2102–2105 cm−1 was presented in the Au25/ZnxAl-400 catalysts (Figure 4a–c. It could be ascribed to the CO molecule adsorbed on metallic gold [34]. Whereas on the Au/Zn2Al-400 catalyst, no adsorption peak of CO was observed in the spectra except for the gaseous CO, indicating the adsorption of CO was very weak on the Au/Zn2Al-400 catalyst (Figure 4d). That may result in the poor activity of Au/Zn2Al-400 in oxidation esterification than that of the Au25/ZnxAl-400 (x = 3, 2, 1) catalysts.
The valence state of gold was also provided by X-ray photoelectron spectra (XPS). Figure 5 displayed a wide spectrum of the core levels of Au, Zn, Al, C, O (Figure 5a) and the high-resolution XPS spectra of Au 4f, C 1s, Zn 2p, Al 2p and O 1s on the Au25/ZnxAl-400 catalysts. Although part of the peak of Au 4f5/2 was overlapped by the signal of Zn 3p (Figure 5b), the branches of Au 4f7/2 could be deconvoluted from the peak. Obviously, the metallic gold (Au0) dominates the surface of gold species, with the binding energy at around 83.5 eV, which agreed well with the in situ CO DRIFT spectra. Moreover, three deconvoluted peaks located at about 288.3, 285.6 and 284.0 eV were presented in C 1s XPS spectra (Figure 5c), which could be assigned to the C=O, C–O and C–C bonds [35]. It indicated there were still carbonate on the catalysts even after high–temperature calcination at 400 °C. Additionally, O 1s XPS fitting results displayed the existence of O2−, O–H and C=O species [36,37], which are named Oα, Oβ and Oγ, respectively (Figure 5f). The lattice oxygen of Oα bounded to the metal cation of Zn2+ and Al3+, which was supported by Zn 2p (Figure 5d) and Al 2p (Figure 5e) XPS spectra. To be noted, with decreasing of Zn2+/Al3+ molar ratios, the valence state of Zn, Al, O species varied correspondingly, indicating the slight variation of the structure over the Zn-Al mixed oxides. Moreover, the oxygen in the form of a hydroxyl group (Oβ) provided some basic sites of the catalysts that may promote the reaction since the abundance of previous works highlights the effect of basicity for catalytical performances [16,19,38].

2.4. The Basic Property of the Catalyst

Previously, several groups have presented the significance of basicity towards oxidative esterification of MAL with methanol [16,19,38]. It was said that the catalysts with a large density of basic sites could accelerate the formation of hemiacetal intermediate, thus, contribute to the formation of MMA [16,38]. To analyze the basicity of the supported gold catalyst, CO2 temperature-programmed desorption (TPD) was subsequently employed on the Au25/ZnxAl-400 and Au/Zn2Al-400 catalysts. The total basicity of the catalysts ranged from 430–640 µmol/g, which is shown in Table 2. The fitted desorption curves according to CO2 desorption temperature were then displayed in Figure 6. From the results, all of the catalysts exhibited obvious bands below 250 °C and big bands from 350 to 900 °C. The peak below 250 °C could be assigned to weak basic sites of the catalysts [17,21]. However, the desorption peak of CO2 at the higher temperature would have resulted from the medium-strength basic sites (250–470 °C), strong basic sites (470–650 °C) and super strong basic sites (above 650 °C) [21,39]. Through quantitative analysis of the basic sites, the gold catalysts supported on Zn2Al-400 possess more strong basic sites (470–650 °C) than that on the Zn3Al-400 and Zn1Al-400. While using Au25 clusters as the precursor of gold, more super-strong basic sites (above 650 °C) were observed in Au25/ZnxAl-400 catalysts, which might be connected with their good catalytic performances. Moreover, to exclude the influence of decomposition of the residual hydrotalcites on the results of CO2–TPD, the comparison experiments were conducted on Au25/ZnxAl-400 and Au/ZnAl-400 catalysts without adsorption of CO2, which showed a clear absence of CO2 desorption peak by acquiring the TCD and MS signal simultaneously (Figure S3).

2.5. The Active Sites and Reaction Mechanism

According to the catalytic performances on supported gold catalysts and calcined supports (Table 1), the active species were supposed to locate at the interface of gold particles and the supports, since the support itself was very hard to activate oxygen and transform the MAL to MMA (Table 1, entry 5–7). Moreover, the small gold particles were generally supposed to benefit the catalytic activity due to the abundant uncoordinated gold species over small gold particles [40,41]. On account of the XRD and TEM results, the average particle size of gold on the Au/Zn2Al-400 catalyst (~2.8 nm, Figure 3d) was close to that of the Au25/Zn2Al-400 (~2.5 nm, Figure 3b), while the activity of it decreased dramatically from 93.1% (Table 1, entry 2) to 57.1% (Table 1, entry 4), and the TOF value reduced from 1626 h−1 to 684 h−1. Thus, the decreased activity might not be attributed merely to the size of gold.
Further characterization of the valence state of gold and the adsorption property of the catalysts were conducted by XPS and in situ CO DRIFT spectra. As shown in Figure 4, a strong adsorption peak of CO was observed on metallic gold over the Au25/ZnxAl-400 catalyst (Figure 4a–c), while no adsorption band of CO could be found on the DP-method-prepared Au/Zn2Al-400 catalyst (Figure 4d). That is, the poor adsorption behavior of the Au/Zn2Al-400 catalyst is very different from that of the Au25/ZnxAl-400, thus, lead to its poor activity. Moreover, CO2–TPD results indicated more super-strong basic sites existed in Au25/ZnxAl-400 catalysts than that in Au/Zn2Al-400 (Figure 6), which may facilitate the transformation of MAL to hemiacetal intermediate. Further characterization about the intrinsic origin of the performance was still needed to make it clear. In contrast, the thiolate-protected gold clusters were demonstrated useful as the precursor of gold in preparation of the good-performing gold catalysts with controllable sizes.
On the supported gold catalysts, the reaction mechanism for the oxidative esterification of MAL to MMA in the presence of molecular oxygen was generally involved with a two-step mechanism in the earlier study [17,38,42]. Based on this, herein, a postulated pathway over the Au25/ZnxAl-400 catalyst is proposed in detail in the combination of the characterization results and catalytic performances (Scheme 3). In the pathway, oxygen and methanol molecule was first approaching the interface and then interacted with gold particles and basic sites to form the adsorbed atomic oxygen and methoxy. After this, the nucleophilic attack of methoxy to the carbon of methyl acrolein forms the surface hemiacetal intermediate. Moreover, finally, with the assistant of atomic oxygen, the hemiacetal intermediate transforms to the target ester via β-H elimination. The acetal appeared as the byproduct, indicating the competing reaction of hemiacetal intermediate attacked by methoxy and adsorbed oxygen. That revealed the selectivity of MMA was probably dependent on the dissociation of oxygen on well-dispersed gold particles and/or the basic sites derived from the support.

2.6. Substrate Universality of the Catalyst

To examine the effect of the substrates on reactivity, a range of aldehydes and alcohols was extrapolated for one-step oxidative esterification, including isobutyl aldehyde, butyraldehyde, benzaldehyde and, etc. The results are shown in Table 3. As seen, different aldehydes could react with MeOH or ethanol on the Au25/Zn2Al-400 catalysts to form the corresponding esters, indicating the catalyst developed in this study had excellent substrate universality. Thereinto, better activity and selectivity were preferentially achieved over the saturated aldehyde, whereas substrates with a vinyl group or phenyl ring had comparably low activity. Based on the earlier study, the substrates with electron-donating substituents promoted the reaction. In contrast, substrates with electron-withdrawing substituents impeded the formation of ester [18].

2.7. Stability of Catalysts in Fixed Bed Reactor

Stability was one of the most important targets for industrial catalysts. Thus, the developed Au25/Zn2Al-400 catalyst was then evaluated in a fixed-bed reactor for synthesis of MMA through oxidative esterification. Preliminary, the optimization of reaction conditions was conducted over the Au25/Zn2Al-400 catalyst as a function of reaction temperature, gas hourly space velocity (GHSV) and liquid hourly space velocity (LHSV). Their catalytic performances as a consequence of the above factors are shown in Figure 7. Obviously, the catalytic activity and selectivity were largely affected by GHSV (Figure 7b) and LHSV (Figure 7c). In contrast, the temperature seems to make no difference in selectivity (Figure 7a). Based on such results, the reaction conditions at 353 K, 2.1 atm O2, GHSV: 3000 h−1 and LHSV: 10 h−1 were designated for fixed-bed reactor from overall consideration. As seen from the stability results in Figure 7d, the selectivity of MMA over the Au25/Zn2Al-400 catalyst maintained at above 90% despite the initial drop. Though the activities decreased from ~90% to ~70% after 60 h, it tends to get a balance. Thus, the potential application of Au25/Zn2Al-400 catalyst for this reaction could be revealed. The same trends were also presented in Au25/Zn3Al-400 and Au25/Zn1Al-400 catalysts, in which both of the catalysts showed good stabilities in one-step oxidative esterification (Figure S4).

3. Materials and Methods

3.1. Chemicals

All chemicals were used as-received and without any purification. All glassware was washed with Aqua Regia and rinsed with ethanol and ultrapure water. Ultrapure water (18.2 MΩ) was used throughout this work. Hydrogen tetrachloroaurate hydrate (HAuCl4·3H2O), sodium carbonate (Na2CO3, AR), sodium hydroxide (NaOH, AR), aluminum nitrate hydrate ((Al(NO3)3·9H2O), 99 wt%) and methanol were purchased from Sinopharm Company (Beijing, China). Sodium borohydride (NaBH4, 97 wt%) was purchased from Shanghai Lingfeng Chemical Reagent Company (Shanghai, China). Methacrolein (MAL, 95 wt%) was purchased from Shanghai Macklin Biochemical Technology Company (Shanghai, China). Methyl methacrylate (MMA, 99.5 wt%), cysteine, zinc nitrate hydrate ((Zn(NO3)2·6H2O), 99 wt%) and ortho-xylene (CP) were purchased from Aladdin Industrial Corporation (Shanghai, China).

3.2. Preparation of the Supported Gold Clusters

The Zn-Al-HTs supported Au25 nanoclusters catalysts were prepared as follows: 30 mg of the Au25 clusters were dispersed into 20 mL of ultrapure water. Then, 2.00 g of the Zn-Al-HTs supports were added into the above suspension under vigorous stirring. About one hour later, the products were washed with ultrapure water and collected by centrifugal separation. Within the process, Zn-Al-HTs and the atomically precisely Au25 nanoclusters were homemade according to our previous work [26,27,28].
The Zn-Al-HTs were prepared through the coprecipitation method. Typically, Zn(NO3)2·6H2O (0.21, 0.14 and 0.07 mol) and Al(NO3)3·9H2O (0.07 mol) with Zn-Al atomic ratios of 3, 2, 1 were mixed by adding 200 mL of ultrapure water to obtain the solution A. NaOH (0.438 mol) and Na2CO3 (0.113 mol) were placed into a 1000 mL beaker with 200 mL of ultrapure water to obtain the solution B. The solution A was slowly pumped into the solution B in a water bath at 70 °C. Moreover, then, the gel aged at the same temperature for another 24 h at constant stirring. Afterward, the solid was filtered, washed, dried and ground to obtain the ZnxAl-HT (x = 3, 2, 1).
Au25 clusters were prepared by a NaOH-mediated NaBH4 reduction method [26,27,28]. First, 5.0 mL of HAuCl4 solution (110 mM) and 150 mL of cysteine solution (5.5 mM) were successively added to 200 mL of ultrapure water under stirring. Then, 30 mL of 1 M NaOH solution was introduced into the above mixture at one time. About 5 min later, fresh sodium borohydride solution was quickly added into the above solution, followed by vigorous stirring for 3 h. Finally, the products were collected and washed with ethanol–water (V/V = 3:1) and dried by lyophilization. The obtained dried Zn-Al-HTs supported Au25 nanoclusters samples were defined as Au25/ZnxAl-HT. Before the catalytic test, the samples were calcined at 673 K for 2 h, with a heating rate of 5 °C/min. The obtained catalysts were denoted as Au25/ZnxAl-400.

3.3. Preparation of Supported Gold Catalysts with DP Method

The supported gold catalyst with the same Zn-Al-mixed oxide was prepared by the traditional deposition–precipitation (DP) method. In a typical synthesis, an aqueous solution of HAuCl4 (Au content: 25 g/L, 1 mL) was added to a 50 mL beaker. Then, certain amounts of Zn2Al-HT powders were introduced into the solution under vigorous stirring. After this, 1 M NaOH solution was added to adjust the pH value of the suspension liquid to ~10. Then, the reaction was allowed to proceed at room temperature for 2 h. The obtained solids were washed, dried and calcined at 673 K for 2 h, with the heating rate of 5 °C/min. The catalysts produced by this method were denoted as Au/Zn2Al-400.

3.4. Catalytic Test

Catalytic testing of synthesis of MMA by one-step oxidative esterification of MAL with methanol was carried out in a stainless-steel autoclave equipped with a pressure gauge and magnetic stirring. Before reaction, 0.396 mL of MAL (0.96 M), 0.121 mL of internal standard (ortho-xylene) and 4.5 mL of methanol were put into the vessel. Then, certain amounts of catalysts were introduced into the autoclave. After sealing, the autoclave was flushed with oxygen six times and then pressurized at 3 atm. To initiate the reaction, the reactor was heated to 353 K in a water bath without stirring until the temperature reached the specified value. After the reaction, the autoclave was put into cold water to terminate the reaction. Different substrates of aldehydes and alcohols for the synthesis of esters were conducted in similar conditions. The products and reactants were analyzed by gas chromatography/mass spectrometry (GC, Agilent 7890 A, Wilmington, NC, USA), which was equipped with an HP-5 (30 m, 0.25 mm inner diameter) capillary column.

3.5. Stability Test

The lifetime of the catalyst was performed in a stainless-steel fixed bed reactor (inner diameter, 10.0 mm; length, 660 mm) equipped with a K-type thermocouple for control of the reaction temperature and a mass flow controller for regulating the gas flow rate. The feed gas was a mixture of oxygen and nitrogen, with a volume ratio of 7/93. Prior to reaction, 0.5 g of the Au25/ZnxAl-400 catalysts were introduced in the middle of the reactor, with quartz sand placed above and below the catalyst bed to hold it in place. In a typical experiment, a reaction mixture of MAL and methanol (molar ratio: 1/40) was introduced into the system using a double plunger liquid pump. The reaction was carried out under the following conditions: 353 K, 3 MPa, liquid hourly space velocity (LHSV) = 10 h−1 and gas hourly space velocity (GHSV) = 3000 h−1. Before introducing of reaction mixture into the system, the temperature of the reactor was raised to a specific value. The total pressure of 3 MPa (oxygen partial pressure of 0.21 MPa) was achieved by regulating the back-pressure valve. The liquid products were analyzed by gas chromatography with a flame ionization detector (FID), which was equipped with an HP-5 capillary column.

3.6. Characterization

The actual loadings of gold, zinc and aluminum were measured with an inductively coupled plasma atomic emission spectroscopy (ICP-AES) on an IRIS Intrepid II XSP instrument (Thermo Electron Corporation, Madison, WI, USA). The X-ray powder diffraction (XRD) patterns were determined on a PW3040/60 X’Pert PRO (PANalytical, Almelo, The Netherlands) diffractometer equipped with a Cu Kα radiation source (λ = 0.15432 nm) operating at 40 kV and 40 mA. The N2-physical adsorption–desorption tests were measured at 77 K using an AutoSorb-1 instrument (Boynton Beach, FL, USA). Prior to the measurements, the catalysts were treated at vacuum for 2 h at 120 °C. The transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) were recorded on a JEOL JEM-2100 F (Tokyo, Japan) microscope at 200 kV. The in situ diffuse reflectance infrared Fourier-transform spectroscopy of CO adsorption (CO-DRIFTS) were acquired with a Bruker INVENIO Fourier-transform infrared spectrometer (Karlsruhe, Germany), equipped with an MCT detector in the range of 600–4000 cm−1. The X-ray photoelectron spectra (XPS) were conducted on an ESCALAB 250 X-ray photoelectron spectrometer (Madison, WI, USA) equipped with a monochromated Al Kα anode. The binding energies were calibrated for surface charging by referencing them to the energy of the C 1s peak at 285 eV. The temperature-programmed desorption of carbon dioxide (CO2–TPD) experiments were performed on Autochem II 2920 Instruments (Norcross, GA, USA) with a thermal conductivity detector and mass spectrometry. Prior to the analysis, about 50 mg of the catalysts were added into a U-type quartz tube reactor. Then the samples were heated in a flow of helium at 400 °C for 0.5 h. After the temperature decreased to 50 °C, the catalysts were saturated with CO2 at 80 °C for 120 min. Finally, the samples were purged in helium at 50 °C for 95 min. The CO2–TPD signal was recorded from 50 °C to 1000 °C at a rate of 10 °C/min with a cold trap.

4. Conclusions

In summary, Zn-Al-HTs supported Au25 nanoclusters were designed and used as the precursors for direct oxidative esterification of MAL with methanol to form MMA. The Au25/ZnxAl-400 catalysts showed comparable better performances with higher activity and selectivity than the gold catalysts prepared by the DP method. XRD and TEM characterizations suggested that the Au25/ZnxAl-400 catalysts have good dispersion and small average particle sizes of gold. In situ CO DRIFT spectra and Au 4f XPS spectra indicated the metallic gold with strong CO adsorption property might attribute to this high activity. Moreover, the strong basic sites revealed by CO2–TPD were said to facilitate the formation of hemiacetal intermediate, which contributes to the formation of the final product. Combined with the characterization results with catalytic performances, the active sites were supposed to locate at the interface between small gold particles and the support. Further substrate scope experiments and stability test in a fixed-bed reactor suggested the Au25/ZnxAl-400 catalysts have potential in these kinds of reactions. This work also provides a good reference for designing the gold catalyst with high activity and selectivity.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/11/2/162/s1, Table S1: Summarization of catalytic performances of supported gold catalysts; Table S2: Element analysis of supported gold catalysts with various Zn/Al molar ratios; Figure S1: N2 adsorption–desorption isotherms and pore size distributions of the supported gold catalysts; Figure S2: Chromatogram analysis of the liquid product before and after oxidative esterification between MAL and CH3OH; Figure S3: TPD profiles of supported gold catalysts without adsorption of CO2; Figure S4: Stabilities of supported gold catalysts with various Zn/Al ratios for the synthesis of MMA from MAL and MeOH.

Author Contributions

Y.T. and Y.D. conceived the study; H.L. performed the experiments; X.C., W.Y. contributed with the materials; J.L. contributed with characterizations; C.H. contributed with the useful advice; All the authors contributed to the writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China under Grant No. 21803056 and 21973084. Zhejiang Provincial Natural Science Foundation of China under Grant No. LQ20B030005 and LQ20B030008.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors would like to thank the public testing platform of Zhejiang Normal University and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Thanks to the technical support with ICP, XRD, TEM, HRTEM and XPS measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nagai, K. New developments in the production of methyl methacrylate. Appl. Catal. A Gen. 2001, 221, 367–377. [Google Scholar] [CrossRef]
  2. Global Industry Analysts Inc, Methyl Methacrylate (MMA)—Global Market Trajectory Analytics. Available online: https://www.researchandmarkets.com/categories/chemicals-materials (accessed on 5 December 2020).
  3. Wang, B.; Li, H.; Zhu, J.; Sun, W.; Chen, S. Preparation and characterization of mono-/multi-metallic hydrophobic catalysts for the oxidative esterification of methacrolein to methyl methacrylate. J. Mol. Catal. A Chem. 2013, 379, 322–326. [Google Scholar] [CrossRef]
  4. Tian, Y.; Li, Y.; Zuo, C.; Yin, D.; Wang, L.; Zheng, Y.; Huang, H.; Fu, Z.; Wang, M. Ionic-Liquid-Modified Porous Au/CeMnOx Nanorods for Methyl Methacrylate (MMA) Synthesis via Direct Oxidative Esterification. Chemnanomat 2019, 5, 1361–1366. [Google Scholar] [CrossRef]
  5. Diao, Y.; He, H.; Yang, P.; Wang, L.; Zhang, S. Optimizing the structure of supported Pd catalyst for direct oxidative esterification of methacrolein with methanol. Chem. Eng. Sci. 2015, 135, 128–136. [Google Scholar] [CrossRef]
  6. Diao, Y.; Yang, P.; Yan, R.; Jiang, L.; Wang, L.; Zhang, H.; Li, C.; Li, Z.; Zhang, S. Deactivation and regeneration of the supported bimetallic Pd–Pb catalyst in direct oxidative esterification of methacrolein with methanol. Appl. Catal. B Environ. 2013, 142, 329–336. [Google Scholar] [CrossRef]
  7. Mahboub, M.J.D.; Dubois, J.-L.; Cavani, F.; Rostamizadeh, M.; Patience, G.S. Catalysis for the synthesis of methacrylic acid and methyl methacrylate. Chem. Soc. Rev. 2018, 47, 7703–7738. [Google Scholar] [CrossRef] [PubMed]
  8. Yamamatsu, S.; Yamaguchi, T.; Yokota, K.; Nagano, O.; Chono, M.; Aoshima, A. Development of Catalyst Technology for Producing Methyl Methacrylate (MMA) by Direct Methyl Esterification. Catal. Surv. Asia 2010, 14, 124–131. [Google Scholar] [CrossRef]
  9. 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]
  10. Liu, S.Y.; Yang, S.M. Complete oxidation of 2-propanol over gold-based catalysts supported on metal oxides. Appl. Catal. A Gen. 2008, 334, 92–99. [Google Scholar] [CrossRef]
  11. Turner, M.; Golovko, V.B.; Vaughan, O.P.H.; Abdulkin, P.; Berenguer-Murcia, A.; Tikhov, M.S.; Johnson, B.F.G.; Lambert, R.M. Selective oxidation with dioxygen by gold nanoparticle catalysts derived from 55-atom clusters. Nat. Cell Biol. 2008, 454, 981–983. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Cui, X.; Shi, F.; Deng, Y. Nano-Gold Catalysis in Fine Chemical Synthesis. Chem. Rev. 2011, 112, 2467–2505. [Google Scholar] [CrossRef] [PubMed]
  13. Tan, Y.; Liu, H.; Liu, X.Y.; Wang, A.; Liu, C.; Zhang, T. Effective removal of the protective ligands from Au nanoclusters by ambient pressure nonthermal plasma treatment for CO oxidation. Chin. J. Catal. 2018, 39, 929–936. [Google Scholar] [CrossRef]
  14. Liu, L.; Li, H.; Tan, Y.; Chen, X.; Lin, R.; Yang, W.; Huang, C.; Wang, S.; Wang, X.; Liu, X.Y.; et al. Metal-Support Synergy of Supported Gold Nanoclusters in Selective Oxidation of Alcohols. Catalysts 2020, 10, 107. [Google Scholar] [CrossRef] [Green Version]
  15. Suzuki, K.; Yamaguchi, T.; Matsushita, K.; Iitsuka, C.; Miura, J.; Akaogi, T.; Ishida, H. Aerobic Oxidative Esterification of Aldehydes with Alcohols by Gold–Nickel Oxide Nanoparticle Catalysts with a Core–Shell Structure. ACS Catal. 2013, 3, 1845–1849. [Google Scholar] [CrossRef]
  16. Wan, X.; Deng, W.; Zhang, Q.; Wang, Y. Magnesia-supported gold nanoparticles as efficient catalysts for oxidative esterification of aldehydes or alcohols with methanol to methyl esters. Catal. Today 2014, 233, 147–154. [Google Scholar] [CrossRef]
  17. Li, Y.; Wang, L.; Yan, R.; Han, J.; Zhang, S. Gold nanoparticles supported on Ce–Zr oxides for the oxidative esterification of aldehydes to esters. Catal. Sci. Technol. 2015, 5, 3682–3692. [Google Scholar] [CrossRef]
  18. Zuo, C.; Tian, Y.; Zheng, Y.; Wang, L.; Fu, Z.; Jiao, T.; Wang, M.; Huang, H.; Li, Y. One step oxidative esterification of methacrolein with methanol over Au-CeO2/γ-Al2O3 catalysts. Catal. Commun. 2019, 124, 51–55. [Google Scholar] [CrossRef]
  19. Tian, Y.; Li, Y.; Zheng, Y.; Wang, M.; Zuo, C.; Huang, H.; Yin, D.; Fu, Z.; Tan, J.; Zhou, Z. Nano-Au/MCeOx catalysts for the direct oxidative esterification of methylacrolein to methyl esters. Ind. Eng. Chem. Res. 2019, 58, 19397–19405. [Google Scholar] [CrossRef]
  20. Paul, B.; Khatun, R.; Sharma, S.K.; Adak, S.; Singh, G.; Das, D.; Siddiqui, N.; Bhandari, S.; Joshi, V.; Sasaki, T.; et al. Fabrication of Au Nanoparticles Supported on One-Dimensional La2O3 Nanorods for Selective Esterification of Methacrolein to Methyl Methacrylate with Molecular Oxygen. ACS Sustain. Chem. Eng. 2019, 7, 3982–3994. [Google Scholar] [CrossRef]
  21. Gao, J.; Fan, G.; Yang, L.; Cao, X.; Zhang, P.; Li, F. Oxidative Esterification of Methacrolein to Methyl Methacrylate over Gold Nanoparticles on Hydroxyapatite. ChemCatChem 2017, 9, 1230–1241. [Google Scholar] [CrossRef]
  22. Li, H.; Li, J.; Tan, Y.; Ding, Y. Research progress on the production technology and noble metal catalyst for synthesis of methyl methacrylate. Chem. Ind. Eng. Prog. 2021, 40, 183–194. [Google Scholar]
  23. Seftel, E.; Popovici, E.; Mertens, M.; De Witte, K.; Van Tendeloo, G.; Cool, P.; VanSant, E.F. Zn–Al layered double hydroxides: Synthesis, characterization and photocatalytic application. Microporous Mesoporous Mater. 2008, 113, 296–304. [Google Scholar] [CrossRef]
  24. Hernández, W.Y.; Aliç, F.; Verberckmoes, A.; Van Der Voort, P. Tuning the acidic–basic properties by Zn-substitution in Mg–Al hydrotalcites as optimal catalysts for the aldol condensation reaction. J. Mater. Sci. 2016, 52, 628–642. [Google Scholar] [CrossRef] [Green Version]
  25. Fang, L.; Luo, W.; Meng, Y.; Zhou, X.; Pan, G.; Ni, Z.; Xia, S. Visible-Light-Promoted Selective Hydrogenation of Crotonaldehyde by Au Supported ZnAl-Layered Double Hydroxides: Catalytic Property, Kinetics, and Mechanism Investigation. J. Phys. Chem. C 2018, 122, 17358–17369. [Google Scholar] [CrossRef]
  26. Fang, W.; Chen, J.; Zhang, Q.; Deng, W.; Wang, Y. Hydrotalcite-Supported Gold Catalyst for the Oxidant-Free Dehydrogenation of Benzyl Alcohol: Studies on Support and Gold Size Effects. Chem. A Eur. J. 2010, 17, 1247–1256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Turco, M.; Bagnasco, G.; Costantino, U.; Marmottini, F.; Montanari, T.; Ramis, G.; Busca, G. Production of hydrogen from oxidative steam reforming of methanolII. Catalytic activity and reaction mechanism on Cu/ZnO/Al2O3 hydrotalcite-derived catalysts. J. Catal. 2004, 228, 56–65. [Google Scholar] [CrossRef]
  28. Kirm, I.; Medina, F.; Rodríguez, X.; Cesteros, Y.; Salagre, P.; Sueiras, J. Epoxidation of styrene with hydrogen peroxide using hydrotalcites as heterogeneous catalysts. Appl. Catal. A Gen. 2004, 272, 175–185. [Google Scholar] [CrossRef]
  29. Tan, Y.; Liu, X.; Zhang, L.; Wang, A.; Li, L.; Pan, X.; Miao, S.; Haruta, M.; Wei, H.; Wang, H.; et al. ZnAl-Hydrotalcite-Supported Au25 Nanoclusters as Precatalysts for Chemoselective Hydrogenation of 3-Nitrostyrene. Angew. Chem. Int. Ed. 2017, 56, 2709–2713. [Google Scholar] [CrossRef]
  30. Tan, Y.; Liu, X.Y.; Li, L.; Kang, L.; Wang, A.; Zhang, T. Effects of divalent metal ions of hydrotalcites on catalytic behavior of supported gold nanocatalysts for chemoselective hydrogenation of 3-nitrostyrene. J. Catal. 2018, 364, 174–182. [Google Scholar] [CrossRef]
  31. Tan, Y.; Liu, X.Y.; Zhang, L.; Liu, F.; Wang, A.; Zhang, T. Producing of cinnamyl alcohol from cinnamaldehyde over supported gold nanocatalyst. Chin. J. Catal. 2021, 42, 470–481. [Google Scholar] [CrossRef]
  32. Sing, K.S.; Williams, R.T. Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials. Adsorpt. Sci. Technol. 2004, 22, 773–782. [Google Scholar] [CrossRef]
  33. Włodarczyk-Stasiak, M.; Jamroz, J. Specific surface area and porosity of starch extrudates determined from nitrogen adsorption data. J. Food Eng. 2009, 93, 379–385. [Google Scholar] [CrossRef]
  34. Mihaylov, M.Y.; Knözinger, H.; Hadjiivanov, K.; Gates, B.C. Characterization of the Oxidation States of Supported Gold Species by IR Spectroscopy of Adsorbed CO. Chem. Ing. Tech. 2007, 79, 795–806. [Google Scholar] [CrossRef]
  35. George, G.; Saravanakumar, M. Correction to: Facile synthesis of carbon-coated layered double hydroxide and its comparative characterisation with Zn–Al LDH: Application on crystal violet and malachite green dye adsorption—Isotherm, kinetics and Box-Behnken design. Environ. Sci. Pollut. Res. 2018, 25, 30255–30256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Gil, J.J.; Aguilar-Martínez, O.; Piña-Pérez, Y.; Pérez-Hernández, R.; Santolalla-Vargas, C.E.; Gómez, R.; Tzompantzi, F. Efficient ZnS–ZnO/ZnAl-LDH composite for H2 production by photocatalysis. Renew. Energy 2020, 145, 124–132. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Hua, Z.; Lang, J.; Song, Y.; Zhang, Q.; Han, Q.; Fan, H.; Gao, M.; Li, X.; Fan, H. Eco-friendly nanostructured Zn–Al layered double hydroxide photocatalysts with enhanced photocatalytic activity. CrystEngComm 2019, 21, 4607–4619. [Google Scholar] [CrossRef]
  38. Trimpalis, A.; Giannakakis, G.; Cao, S.; Flytzani-Stephanopoulos, M. NiAu single atom alloys for the selective oxidation of methacrolein with methanol to methyl methacrylate. Catal. Today 2020, 355, 804–814. [Google Scholar] [CrossRef]
  39. Li, J.; Chao, H.; Kai, T.; Hao, X.; Zhu, Z.; Hu, Z. CO2 atmosphere-enhanced methanol aromatization over the NiO-HZSM-5 catalyst. Rsc. Adv. 2014, 4, 44377–44385. [Google Scholar] [CrossRef]
  40. Shi, Q.; Qin, Z.; Xu, H.; Li, G. Heterogeneous Cross-Coupling over Gold Nanoclusters. Nanomaterials 2019, 9, 838. [Google Scholar] [CrossRef] [Green Version]
  41. Taketoshi, A.; Haruta, M. ChemInform Abstract: Size- and Structure-Specificity in Catalysis by Gold Clusters. Chem. Lett. 2014, 45, 380–387. [Google Scholar] [CrossRef] [Green Version]
  42. Xu, B.; Liu, X.; Haubrich, J.; Friend, C.M. Vapour-phase gold-surface-mediated coupling of aldehydes with methanol. Nat. Chem. 2009, 2, 61–65. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Isobutene oxidation routes to produce methyl methacrylate (MMA).
Scheme 1. Isobutene oxidation routes to produce methyl methacrylate (MMA).
Catalysts 11 00162 sch001
Scheme 2. Reaction pathway of methacrolein (MAL) with MeOH to form MMA.
Scheme 2. Reaction pathway of methacrolein (MAL) with MeOH to form MMA.
Catalysts 11 00162 sch002
Figure 1. The products distributions of reaction with time courses over the (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400 and (c) Au25/Zn1Al-400 catalysts. Reaction conditions: MAL/MeOH = 1/23 (molar ratio), catalyst: 0.05 g, temperature: 353 K, O2 pressure: 3 atm.
Figure 1. The products distributions of reaction with time courses over the (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400 and (c) Au25/Zn1Al-400 catalysts. Reaction conditions: MAL/MeOH = 1/23 (molar ratio), catalyst: 0.05 g, temperature: 353 K, O2 pressure: 3 atm.
Catalysts 11 00162 g001
Figure 2. XRD patterns of Zn-Al-HTs supported gold nanoclusters before (a) and after (b) heat treatment.
Figure 2. XRD patterns of Zn-Al-HTs supported gold nanoclusters before (a) and after (b) heat treatment.
Catalysts 11 00162 g002
Figure 3. TEM (a1–d1) and HRTEM (a2–d2) images of supported gold catalysts: (a12) Au25/Zn3Al-400; (b12) Au25/Zn2Al-400; (c12) Au25/Zn1Al-400; (d12) Au/Zn2Al-400 (deposition–precipitation (DP)).
Figure 3. TEM (a1–d1) and HRTEM (a2–d2) images of supported gold catalysts: (a12) Au25/Zn3Al-400; (b12) Au25/Zn2Al-400; (c12) Au25/Zn1Al-400; (d12) Au/Zn2Al-400 (deposition–precipitation (DP)).
Catalysts 11 00162 g003
Figure 4. In situ diffuse reflectance infrared Fourier–transform spectroscopy of CO adsorption (CO–DRIFT) spectra of supported gold catalysts: (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400; (c) Au25/Zn1Al-400; (d) Au/Zn2Al-400 (DP).
Figure 4. In situ diffuse reflectance infrared Fourier–transform spectroscopy of CO adsorption (CO–DRIFT) spectra of supported gold catalysts: (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400; (c) Au25/Zn1Al-400; (d) Au/Zn2Al-400 (DP).
Catalysts 11 00162 g004
Figure 5. (a) XPS wide spectra and (b) Au 4f, (c) C 1s, (d) Zn 2p, (e) Al 2p (f) O 1s of XPS spectra over the Au25/ZnxAl-400 catalyst: (1) Zn2+/Al3+ = 1; (2) Zn2+/Al3+ = 2; (3) Zn2+/Al3+ = 3.
Figure 5. (a) XPS wide spectra and (b) Au 4f, (c) C 1s, (d) Zn 2p, (e) Al 2p (f) O 1s of XPS spectra over the Au25/ZnxAl-400 catalyst: (1) Zn2+/Al3+ = 1; (2) Zn2+/Al3+ = 2; (3) Zn2+/Al3+ = 3.
Catalysts 11 00162 g005
Figure 6. CO2–TPD of supported gold catalysts: (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400; (c) Au25/Zn1Al-400; (d) Au/Zn2Al-400.
Figure 6. CO2–TPD of supported gold catalysts: (a) Au25/Zn3Al-400; (b) Au25/Zn2Al-400; (c) Au25/Zn1Al-400; (d) Au/Zn2Al-400.
Catalysts 11 00162 g006
Scheme 3. Reaction pathway of MAL with MeOH to form MMA.
Scheme 3. Reaction pathway of MAL with MeOH to form MMA.
Catalysts 11 00162 sch003
Figure 7. The oxidative esterification performances from MAL and methanol over Au25/ZnAl-400 catalyst as a function of (a) Reaction temperature at 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1, LHSV: 10 h−1; (b) gas hourly space velocity (GHSV) at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, LHSV: 10 h−1, and (c) liquid hourly space velocity (LHSV) at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1; (d) reaction time at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1, LHSV: 10 h−1.
Figure 7. The oxidative esterification performances from MAL and methanol over Au25/ZnAl-400 catalyst as a function of (a) Reaction temperature at 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1, LHSV: 10 h−1; (b) gas hourly space velocity (GHSV) at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, LHSV: 10 h−1, and (c) liquid hourly space velocity (LHSV) at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1; (d) reaction time at 353 K, 2.1 atm O2, MAL/MeOH = 1/40 (molar ratio), catalyst: 0.5 g, GHSV: 3000 h−1, LHSV: 10 h−1.
Catalysts 11 00162 g007
Table 1. Catalytic activities of MAL with MeOH to form MMA over supported gold catalysts.
Table 1. Catalytic activities of MAL with MeOH to form MMA over supported gold catalysts.
EntryCatalystsConversion (%) aSelectivity (%) aTOF (h−1) b
MMAACEOthers
1Au25/Zn3Al-40084.995.10.74.21734
2Au25/Zn2Al-40093.194.80.84.41626
3Au25/Zn1Al-40090.192.00.57.51675
4Au/Zn2Al-40057.194.13.42.5684
5Zn3Al-4006.4099.0--
6Zn2Al-40013.2094.8--
7Zn1Al-40016.9091.0--
8Blank0----
Reaction conditions: MAL/MeOH = 1/23(molar ratio), V(MeOH) = 4.5 mL, Catalyst: 0.1 g (Au: 0.1 mol%), T = 353 K, PO2 = 3 atm, t = 2 h. a Conversion and selectivity were analyzed by gas chromatography; b turnover frequency (TOF) of different catalysts was measured by moles of converted MAL per mole of gold per hour, with the conversion of MAL below 20%.
Table 2. Textural properties of Zn-Al-mixed oxides supported gold nanoclusters.
Table 2. Textural properties of Zn-Al-mixed oxides supported gold nanoclusters.
EntryCatalystLoadings of Gold (wt %) aSurface Area (m2/g) bPore Volume (cm3/g) cHalf Pore Width (nm) dParticle Size of Gold
(nm) e
Total Basicity (µmol CO2/gcat) f
1Au25/Zn3Al-4001.2749.10.241.712.8640
2Au25/Zn2Al-4001.1860.50.141.532.5500
3Au25/Zn1Al-4001.3280.90.271.712.5430
4Au/Zn2Al-4001.3271.00.081.902.8540
a The actual loadings of gold were evaluated by inductively coupled plasma atomic emission spectroscopy (ICP-AES) measurement. b Surface area was calculated from the Brunauer-Emmet-Teller (BET) equation at P/P0 range of 0.05–0.3. c Total pore volume was calculated at P/P0 = 0.98. d Pore width was calculated from the adsorption branch of the isotherm—BJH. e The mean particle sizes of gold were calculated from TEM measurements. f Total basicity was calculated from CO2–TPD.
Table 3. Catalytic performances of various aldehydes and alcohols to form the esters by oxidative esterification over the Au25/Zn2Al-400 catalyst.
Table 3. Catalytic performances of various aldehydes and alcohols to form the esters by oxidative esterification over the Au25/Zn2Al-400 catalyst.
EntryAldehydeAlcoholProductConversion (%)Selectivity (%)
1 Catalysts 11 00162 i001Methanol Catalysts 11 00162 i00269.492.0
2 Catalysts 11 00162 i003Ethanol Catalysts 11 00162 i00453.964.4
3 Catalysts 11 00162 i005Methanol Catalysts 11 00162 i00687.898.9
4 Catalysts 11 00162 i007Ethanol Catalysts 11 00162 i00891.297.9
5 Catalysts 11 00162 i009Methanol Catalysts 11 00162 i01087.895.2
6 Catalysts 11 00162 i011Ethanol Catalysts 11 00162 i01293.691.9
7 Catalysts 11 00162 i013Methanol Catalysts 11 00162 i01445.275.4
8 Catalysts 11 00162 i015Ethanol Catalysts 11 00162 i01632.490.3
Reaction conditions: MAL/MeOH = 1/23 (molar ratio), V(MeOH) = 4.5 mL, catalyst: 0.05 g, T = 353 K, PO2 = 3 atm, t = 2 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, H.; Tan, Y.; Chen, X.; Yang, W.; Huang, C.; Li, J.; Ding, Y. Efficient Synthesis of Methyl Methacrylate by One Step Oxidative Esterification over Zn-Al-Mixed Oxides Supported Gold Nanocatalysts. Catalysts 2021, 11, 162. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020162

AMA Style

Li H, Tan Y, Chen X, Yang W, Huang C, Li J, Ding Y. Efficient Synthesis of Methyl Methacrylate by One Step Oxidative Esterification over Zn-Al-Mixed Oxides Supported Gold Nanocatalysts. Catalysts. 2021; 11(2):162. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020162

Chicago/Turabian Style

Li, Huayin, Yuan Tan, Xingkun Chen, Wenshao Yang, Chuanqi Huang, Jie Li, and Yunjie Ding. 2021. "Efficient Synthesis of Methyl Methacrylate by One Step Oxidative Esterification over Zn-Al-Mixed Oxides Supported Gold Nanocatalysts" Catalysts 11, no. 2: 162. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11020162

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