Вы находитесь на странице: 1из 10

Journal of Catalysis 302 (2013) 1019

Contents lists available at SciVerse ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Palladium-doped silicaalumina catalysts obtained from double-ame FSP for chemoselective hydrogenation of the model aromatic ketone acetophenone
Zichun Wang a,1, Suman Pokhrel b,1, Mengmeng Chen a, Michael Hunger c, Lutz Mdler b,, Jun Huang a,
a

Laboratory for Catalysis Engineering, School of Chemical and Biomolecular Engineering, The University of Sydney, NSW 2006, Australia Foundation Institute of Materials Science (IWT), Department of Production Engineering, University of Bremen, Germany c Institute of Chemical Technology, University of Stuttgart, D-70550 Stuttgart, Germany
b

a r t i c l e

i n f o

a b s t r a c t
Chemoselective hydrogenation of aromatic ketones is an important reaction in the production of ne chemicals and pharmaceuticals. A typical example of this class of reactions is the hydrogenation of acetophenone (Aph) over the supported noble metal catalysts. In this research, Pd/silicaalumina (Pd/ SA) catalysts have been prepared for Aph hydrogenation with the emerging double-ame spray pyrolysis system (decouples the two pathways of support formation and metal loading). The Pd particles offered identical electronic properties of Pd surface, which contributed the similar chemoselectivity for the hydrogenation of carbonyl groups on all double-ame-derived catalysts. This revealed a striking difference in Pd surface activity between double-ame and other 5%Pd/SA catalysts. While for supported-Pd catalysts synthesized by other methods, a signicant ionic effect of support acidity on the surface Pd particles is reported, and this kind of surface electronic change was not observed for double-ame catalysts. The reaction rate of Aph hydrogenation could be strongly enhanced (TOF from 1.3 102 s1 to 4.5 102 s1) through tuning the density of surface Brnsted acid sites on supports of 5%Pd/SA via various Si/Al ratios. These advantages of double-ame-derived catalysts clearly demonstrate that double-ame spray pyrolysis can efciently tune nano-catalysts and their bifunctional activities for specic surface reactions. 2013 Elsevier Inc. All rights reserved.

Article history: Received 23 October 2012 Revised 18 February 2013 Accepted 20 February 2013 Available online 29 March 2013 Keywords: Palladium nano-catalysts Silicaalumina Double-ame spray pyrolysis Solid-state NMR Chemoselective hydrogenation

1. Introduction Supported metal catalysts are widely used in chemical industry, such as hydrogenation, oxidation, CC coupling reaction, catalytic reforming, as well as biomass gasication, and FischerTropsch synthesis [13]. The key properties of these materials are related to their unique physicochemical structure, which is essential in catalytic processes [4]. These properties have also been inuenced by the nature of the support material, such as high specic surface area and uniform dispersion of noble metals, resulting in an enhanced catalytic performance. Aiming at synthesizing ideal catalysts with such properties for a number of chemical reactions, many efforts have been devoted to the development of different synthetic methods for active supported-metal catalysts over several decades [519]. A variety of methods, such as depositionprecipitation, coprecipitation, and impregnation/ion-exchange, have been utilized for the synthesis of supported metal catalysts [2022]. These methods
Corresponding authors. Fax: +49 421 218 51211 (L. Mdler), fax: +61 2 9351 2854 (J. Huang). E-mail addresses: lmaedler@iwt.uni-bremen.de (L. Mdler), jun.huang@sydney.edu.au (J. Huang). 1 Equally contribution to this work.
0021-9517/$ - see front matter 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.jcat.2013.02.017

require time-consuming tedious multiple chemical processes for the preparation of the desired catalysts. Recently, fast aerosol synthetic technique has been developed and applied in preparing nano-catalysts [2326]. The synthesis of support and metal particles can be carried out in a single step using ame spray pyrolysis allowing the production of catalysts in micro-seconds from their liquid precursors. These ame-derived catalysts are ultrane nanoparticles with high surface area. Their compositions, structure, and supported metal properties can be easily adjusted through ame conditions and precursor compositions [2628]. Normally, the properties of the supported catalyst are strongly inuenced by the composition and morphology (size and shape) of the support material [1]. Changing the particle size of active metal with different support compositions is widely reported for catalysts prepared by both the classical methods [2932] and the emerging ame spray pyrolysis [6,7,11,12]. As reviewed earlier [33,34], the size of the supported metals has strong inuence on their catalytic performance, especially for hydrogenations [29], oxidations [35], cross-coupling reactions [36], and electron transfer reactions [37]. Therefore, many nano-porous supports such as zeolites, MCM-41, SBA-15 have been used to prepare the size-conned nano-catalysts [3840]. However, the poor diffusion of big molecules and easily blocked channels by metal nanoparticles

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

11

inside nano-pores of these catalysts limit their applications. Obtaining the narrow size distribution of the metal nanoparticles on the supports during the variation of support compositions is still a challenge. Aiming at producing these ideal catalysts, the double-ame spray pyrolysis (double FSP) has been exploited in the present investigation. The double FSP system is able to deliver support precursor and metal precursor through two independent nozzles [19]. This system can efciently reduce the interaction between growing support and metal particle formation. It also allows the control over the formation of support and metal loading during multi-component FSP synthesis. Here, we report a series of Pd/SA catalysts with Pd nanoparticles on the silicaalumina supports with different Si/Al ratios. The noble metal and support materials were sprayed simultaneously from two independent nozzles where the two aerosol streams mixed in situ to form nal catalytic component in a single step. The morphology and surface properties of the catalysts were characterized using microscopic, XPS, and solid-state NMR spectroscopic techniques. Chemoselective hydrogenation of acetophenone was used as a test reaction, since hydrogenation or hydrogenolysis reactions are sensitive to the change of electronic properties of supported metal nanoparticles induced through different support acidity [6,13]. In addition, the selective hydrogenation of acetophenone has practical relevance for the generation of phenylethanol for perfumery industry and ethylbenzene for polymer and fuel industry [5,41]. 2. Experimental 2.1. Catalyst preparation by double FSP Metallorganic-based precursors such as palladium acetylacetonate, aluminum secondary butoxide, tetraethyl orthosilicate (TEOS) (99.9% purity, SigmaAdrich) were used for the production of the catalysts. Aluminum secondary butoxide was rstly dissolved in xylene and added in the TEOS prepared in acetic acid and methanol (1:1 by volume). At Si/Al ratio of 30/70, gel was formed due to higher content of aluminum butoxide in TEOS. FSP requires a low viscous liquid for combustion which was not possible for the Si/Al ratio of 30/70. To overcome this problem, the gel obtained from the higher content of Al butoxideTEOS was further dissolved using 510 mL of methanol. Both AlSi precursor solutions (0.5 M concentration by metal) and Pd(acac)2 dissolved in acetic acid and methanol (1:1 ratio by volume, 2.43 mM by metal) were fed into two independent nozzles placed at 20 to the base and 11 cm apart (see Fig. 1) such that the aerosol streams were homogeneously mixed at a distance of 16 cm perpendicular to the base. Two syringe pumps with a ow rate of 5 mL/min supplied the liquid precursor separately through two independent nozzles for combustion. The precursor solutions (owing at the rate of 5 mL/min to the nozzle) were atomized with O2 (at a ow rate of 3.2 mL/min) maintaining a pressure drop of 1.5 bar at the nozzle tip. The nomenclature of the 5%Pd/silicaalumina powders is dened as Pd/SA-X, where X is 0, 10, 30, and 70, assigned for the weight% of Al in the support mixture. A photograph of the two merging ame sprays is shown in Fig. 1. For example, to obtain Pd/SA-0, 50 mL of 2.4 mM Pd2+ solution was sprayed in one nozzle, and 50 mL of 0.5 M Si4+ precursor solution was sprayed in the other. Combustion of the dispersed droplets was initiated through a premixed CH4 and O2 (1.5 L/min, 3.2 L/min) ame surrounding each nozzle. 2.2. Physical characterization 2.2.1. BET measurements Nitrogen adsorptiondesorption measurements (BET) were performed at 77 K using a 4000e NOVA system to determine the

Fig. 1. The double-ame spray pyrolysis (double FSP) system used for production of the Pd doped silicaalumina nanoparticles.

specic surface areas. Before measurements, the samples were degassed at 200 C under vacuum to remove adsorbants from the surface. The primary particle size and specic surface areas were determined by BET measurement and summarized in columns 1 and 2 of Table 1. BET surface area measurement is related to the average equivalent primary particle size by the equation [27] dBET = 6000/(qpSA) (in nm), where dBET is the average diameter of a spherical particle, SA represents the measured surface area of the powder in m2/g, and qp is the theoretical density in g/cm3. 2.2.2. X-ray characterization The 5% Pd doped SA powders (0, 10, 30, 70) were pressed on to a single crystalline Si substrates and were loaded in a Philips PW 1800 diffraction system. The instrument was equipped with Ni-ltered Cu Ka (k = 0.154 nm) radiation, 1=4 xed divergence, primary and secondary Soller slit with 0.04 rad aperture, circular sample holder with 16 mm diameter, and XCelerator detector (position sensitive in a range of 2.122 2h with 127 channels, yielding a channel width of 0.01671 2h), applying a continuous scan in the range of 15140 2h and an integration step width of 0.0334 2h. The measuring time step was nominally 170 s, to be divided by 127 channels to yield effective measuring time of 1.34 s per 0.0334 2h. 2.2.3. TEM characterization For preparing the TEM samples, a small amount of the sample material was dispersed in 5 mL of ethanol (AR grade, Strem) in an ultrasonic bath and sonicated for 60 min. A drop of the colloidal suspension was placed on a copper grid coated with carbon lm. The samples were dried in ambient air. Large regions of the sample were scanned before the investigation of the particle morphology. The low magnication TEM, the corresponding selected area electron diffractions (SAEDs), and high-resolution transmission electron microscopy (HRTEM) studies were carried out using FEI Titan 80/300 ST unit. The instrument was equipped with a Cs corrector for spherical aberration of the objective lens. For imaging, a high-angle annular dark eld (HAADF) detector, a GATAN post-column imaging lter was used. In all investigations, the microscope was operated at an acceleration voltage of 300 kV. 2.2.4. XPS characterization Pd surface analysis was performed by XPS (ESCALAB250Xi spectrometer, Thermo Scientic, UK) employing a mono-chromated Al Ka (1486.68 eV) X-ray source operated at 15.2 kV and 164 W

12

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

Table 1 Properties of double FSP Pd/SA catalysts with different Si/Al ratios. BET surface areas, primary particle size, concentrations of Brnsted acid sites, and fractions (atom %) of aluminum species with different oxygen coordination. Pd/SAs Pd/SA-0 Pd/SA-10 Pd/SA-30 Pd/SA-70
a

Surface area (m2/g) 329 264 193 151

BET primary particle size (nm) 6.9 8.2 10.2 11.2

Brnsted acidic OH (mmol/g) 5.1 102 11 102 9.5 102

Brnsted acidic OH (mol%)a 3.6 11.9 11.4

AlIV (at.%) 87.3 36.7 25.3

AlV (at.%) 9.4 28.7 47.2

AlVI (at.%) 3.3 34.7 27.5

Brnsted acidic OH (mol%) = (number of Brnsted acidic OH groups/total number of OH groups on supports) 100%.

(10.8 mA). The photoelectron takeoff angle was 90 measured with respect to the surface of the sample. Typical operating pressure in the vacuum chamber was 2 109 mbar. 2.3. Solid-state NMR studies For 27Al and 29Si MAS NMR investigation, all samples were exposed to the saturated vapor of Ca(NO3)2 solution at ambient temperature overnight in a desiccators for fully hydration. Before 1 H and 13C MAS NMR experiments, the samples were dehydrated at 723 K in vacuum pressure less than 102 bar for 12 h in glass tubes. The dehydrated samples were sealed or utilized for in situ loading with ammonia or acetone-2-13C (99.5% 13C-enriched, SigmaAldrich) on a vacuum line, followed by evacuation at 393 K for 1 h (for ammonia) or at room temperature for 2 h (for acetone) to remove weakly hydrogen-bonded or physisorbed molecules. Subsequently, the samples were transferred into the MAS NMR rotors under dry nitrogen gas inside a glove box. 1 H, 27Al, and 13C MAS NMR investigations were carried out on a Bruker Avance III 400 WB spectrometer at resonance frequencies of 400.1, 104.3, and 100.6 MHz with the sample spinning rate of 8 kHz using 4 mm MAS rotors. Spectra were recorded after single-pulse p/2 and p/6 excitation with repetition times of 20 s and 0.5 s for studying 1H and 27Al nuclei, respectively. 27Al MQMAS NMR investigations were performed on the Bruker Avance III 400 WB spectrometer at the resonance frequency of 104.3 MHz, using the three-pulse z-lter pulse sequence with pulse lengths of 5.5, 1.6, and 20 ls, a repetition time of 200 ms and 64 t1 increments. Quantitative 1H MAS NMR measurements were performed using zeolite H,NaY (35% ion-exchanged) as an external intensity standard, which contains 58.5 mg zeolite H,NaY with 1.776 mmol OH/ g. 13C cross-polarization (CP) MAS NMR spectra were recorded with the contact time of 4 ms and the repetition time of 4 s. 29Si MAS NMR experiments were performed on the same spectrometer at the resonance frequency of 79.5 MHz and with the sample spinning rate of 4 kHz using a 7 mm MAS rotor [6]. For 29Si MAS NMR, single-pulse p/2 excitation and high-power proton decoupling with a recycle delay of 20 s was applied. To separate the different signals and for the quantitative evaluation of spectra, the data were processed using the Bruker software WINNMR and WINFIT. 2.4. Catalytic hydrogenation Prior to the hydrogenation, the catalyst materials were reduced in a quartz reactor under a hydrogen ow rate of 50 mL/min for 1 h at 573 K. After cooling down under a nitrogen ow rate of 50 mL/ min, the reduced catalysts were immediately transferred to a 50 mL stainless steel autoclave, which was purged with nitrogen followed by hydrogen for three times. Acetophenone (60 mg) was dissolved in hexane (6 mL) and added to the autoclave. The molar ratio of Pd on the catalyst to reactant was 1:101. The reaction was carried out at a hydrogen pressure of 3 bar at room temperature under magnetic stirring for 3 h. The product was collected at intervals of 10, 20, 30, 60, 90, 120, and 180 min. The conversion and composition of the product mixtures were analyzed using a

Shimadzu GCMS-QP2010 Ultra with a RTX-5MS capillary column (30 m 0.25 mm 0.25 lm) and quantied by Shimadzu GCFID equipped with a RTX-5 capillary column (30 m 0.32 mm 3 lm). Bicyclohexyl (0.1 mmol) was used as an internal standard. The selectivity to specic product(s) i (Si) was calculated as Si (%) = (ci)/[(cAPh)0 (cAPh)], where (ci) is the molar concentration of the products, such as 1-phenylethanol (PhE), and (cAPh)0 and (cAPh) correspond to the molar concentration of acetophenone before and after the reaction, respectively.

3. Results and discussion 3.1. Catalyst morphology The surface areas of Pd/SA-X [X = 0, 10, 30, and 70] are summarized in Table 1. The BET measurements show that the specic surface areas of the samples decreased from 329 to 151 m2/g with increasing aluminum content. This range is similar to the surface areas of Pd/SAs (366174 m2/g) prepared with the single FSP reported earlier [6]. The considerable decrease in the specic surface area and the increase in the particle size (Table 1) with increasing aluminum content clearly indicate the homogeneous introduction of aluminum into the silicon oxide. SEM microscopy investigation of the samples with different Si/Al ratio is presented in Supplementary Fig. S1. No large aggregated particles were observed, and the particles were uniformly mixed. The particles are highly amorphous in nature as indicated from the XRD patterns shown in Fig. 2. For all samples, a broad peak at 2h = 25 is corresponding to amorphous silica [42]. No crystalline alumina were detected on Pd/SA-0, Pd/SA-10, and Pd/SA-30 catalysts. Only a small amount of silicon aluminate (48 and 68 2h in Fig. 2) was detected for Pd/SA-70 having the highest Al content in this study. It should be noted that no Pd crystal was detected in the XRD patterns, and Pd particles were well dispersed in nanosize on the supports as shown in TEM and STEM images in Fig. 3. STEM (Fig. 3a, d, g, and j) and TEM (Fig. 3b, e, h, and k) images showed that all Pd/SAs catalysts contained Pd nanoparticles with the mean size of 24 nm and silicaalumina supports have the spherical morphology. HRTEM images (Fig. 3c, f, g and i) showed highly crystalline single Pd particle dispersed on various supports (see Fig. 4ad). SEM images of double FSP PdSA (see Supporting information Fig. S1) showed that the silicon and aluminum content were mixed homogeneously during double-ame synthesis, and the mean particle size between 2 and 4 nm was observed for each catalyst. For the single-ame-made 5%Pd/SA catalyst, however, the average Pd nanoparticle size was 12 nm for Pd/SA-0 (no Al) and 36 nm for Pd/SAs with Al inside the supports [6]. In general, the size of metal particles differed with support compositions prepared by either single FSP system or other methods [6,7,11,12,39]. This is caused by the strong interaction between the multi-components during synthesis leading to difculty in controlling the growth of metal nanoparticles. However, the uniform metal nanoparticles size is essential for their catalytic performance for contributing high selectivity in many reactions [33,34]. Using

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

13

Fig. 2. XRD patterns of Pd doped silicaalumina nanoparticles prepared using the double-ame spray pyrolysis. The low intensity peaks at around 48 and 68 2h represent a very small amounts of silicon aluminate formed during the synthesis of Pd/SA-70.

double FSP system with individual nozzles for the support and the metal precursors, the interaction between support nucleation/ growth and metal nucleation/deposition was minimized compared to the single FSP system. Therefore, the composition of the support precursor had less effect on the formation of metal particles. Also, formation of the support material was not obviously inuenced by the composition or formation of metal particles. This indicates that the metal loading and support composition could be exibly tuned with the double FSP system in order to obtain the desired metal and support properties for enhanced heterogeneously catalyzed reactions. 3.2. Surface acidity and local structure of supports The supported nano-catalysts always show bifunctional catalytic activities through both the support and the active metal during reactions. As one popular example of solid acid, the surface acidity of silicaalumina supports associated with active metal nanoparticles show signicant inuence in many catalytic reactions, especially in hydrogenation [6,12,43]. The local structure of supports not only shows the formation of acid sites, but also relates to the dispersion of metal nanoparticles on the supports. Therefore, the acidity and local structure of the Pd/SAs were investigated by multinuclear MAS NMR spectroscopy. 1 H MAS NMR spectroscopy was employed to investigate the surface hydroxyl protons. As shown in Fig. 5a, the 1H MAS NMR spectrum of double FSP Pd/SA-0 showed a strong peak at d1H = 1.8 ppm which is assigned to silanol groups. When there are neighboring silanol groups, hydrogen-bonds are formed and cause an additional signal at 2.6 ppm, which was clearly detected in 1H MAS NMR spectra of the single-ame-made silica and 5%Pd/SA-0 catalyst [6,44]. However, this signal was difcult to be observed in the spectra for double FSP 5%Pd/SA-0 (Fig. 5a). Both single and double FSP Pd/SA-0 have similar surface areas and the same amount of Pd loading, but the size and distribution of the metal particles on the support was different. In present work, the absence of a signal at 2.6 ppm indicates that the Pd nanoparticles obtained from double FSP Pd/SA-0 are mainly located in the vicinity of SiOH groups and block interactions with neighboring silanols.

When introducing aluminum into the support precursor, the aluminum atoms are incorporated into the silica network, as indicated by the low-eld shift of the 29Si MAS NMR signals from 110 ppm to 90 ppm going from the double FSP Pd/SA-10 to the Pd/SA-70 catalyst (Supporting information, Fig. S2). The 27Al MAS NMR spectra of double FSP Pd/SA-10 catalyst in Fig. 6ac consist of peaks at ca. 3, 30, and 58 ppm. For larger aluminum contents, a signicant increase in the signal at 30 ppm occurs. The above-mentioned signals are assigned to tetrahedrally coordinated (AlIV), penta-coordinated (AlV), and octahedrally coordinated (AlVI) aluminum species, respectively, involved in weak quadrupolar interactions as shown in Fig. 6d. A contribution of strained tetrahedrally coordinated aluminum at 30 ppm can be excluded. While AlIV signals at ca. 58 ppm are generally explained by framework of aluminum species at Al(OSi)4 sites, the AlV signal at 30 ppm could be a hint for the presence of pseudo-bridging with nearby silanol [44,45]. Aluminum atoms in the vicinity of SiOH groups enhance the acid strength of silanol groups. Adsorption of basic molecules, such as NH3, is a suitable method for studying the number of surface OH groups with enhanced acid strength [46]. Upon adsorption of NH3 at room temperature and subsequent evacuation at 393 K for 1 h, ammonium ions causing 1H MAS NMR signals at d1H = 6.57.0 ppm are formed on acidic surface OH groups. The absence of a peak at 6.57.0 ppm in the 1H MAS NMR spectrum of NH3-loaded FSP Pd/SA-0, shown in Fig. 5a top, indicates that the silanol groups of this material are not able to protonate absorbed ammonia. Loading NH3 on the aluminum-containing double FSP Pd/SAs catalysts, however, led to 1H NMR signals of ammonium ions at 6.57.0 ppm (Fig. 5bd, top). This nding conrmed that the presence of SiOH groups with enhanced Brnsted acidity is caused by the incorporation of aluminum into the support material. The NMR intensities of the ammonium signals were utilized to quantify the number of Brnsted acid sites. The concentration and molar fraction of these Brnsted acid sites are summarized in Table 1 (3 and 4 column). As shown in Fig. 5, the intensity of the ammonium signal at 6.57.0 ppm increases (from 5.1 102 to 11 102 mmol/g in Table 1) with increasing aluminum content from 0% to 30%. Further increase in aluminum content to 70%, a slight decrease in the 1H NMR signals of NH 4 was observed as shown in Fig. 5c and d (from 11 102 to 9.5 102 mmol/g in Table 1). The concentration and molar fraction of Brnsted acid sites increases with the aluminum content up to 30%, and slightly decreases when the aluminum content is increased to 70%. The concentration of Brnsted acid sites of the single FSP Pd/SAs materials studied in our earlier work [6], however, increased with the aluminum content up to 70%. This discrepancy between the number of Brnsted acid site observed through single and double FSP Pd/SAs catalysts with high aluminum contents may be due to the property of the small Pd particles. In the sample obtained by double ame, the Pd particles are mainly located in the vicinity of SiOH groups blocking the interactions with neighboring aluminum species and reducing the surface Brnsted acid sites. Like single FSP silicaalumina [44] and Pd/SA materials [6], the intensity of the 27Al MAS NMR signal of the penta-coordinated aluminum (AlV) species at ca. 30 ppm increases with the aluminum content from 10% to 70% (see Fig. 6, Table 1). Due to simultaneous increase in the number of silanol groups with enhanced acid strength (vide supra), an electron withdrawing effect of AlV species on neighboring silanols is assumed to take place via pseudo-bridging silanol [44,45]. Using 13C-enriched acetone (CH313COCH3) as a probe molecule, the acidic strength of surface sites formed on the Pd/SA supports can be investigated by the chemical shift of the carbonyl signal [46]. For Pd/SA-10 and Pd/SA-30, the strong signal at 214 ppm in 13C MAS NMR spectra was assigned to acetone-2-13C adsorbed on SiOH groups with enhanced acid strength by

14

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

Fig. 3. Transmission electron microscopy (STEM, TEM and HRTEM modes) of Pd/SA-X [X = 0, 10, 30 and 70, where X is the Al-content] catalysts. The bright white spots observed in the STEM images and the black dots in the TEM images are Pd particles on the surface. The lattice distance of 0.2210.223 nm derived from HRTEM matches reasonably with 100% peak at 0.224 nm from XRD (Pdf le no. 00-046-1043).

neighboring aluminum species. This signal was shifted to 217 ppm in the spectrum of acetone-loaded Pd/SA-70. This indicates an enhanced strength of acid sites of Pd/SA-70 compared to low aluminum contents in the present investigation. A strong 13C NMR peak at 225 ppm was observed after adsorption of acetone-2-13C on double FSP Pd/SA-70 (Fig. 7), but it was not detected on the single FSP Pd/SA-70 [6]. This type of low-eld shift in the acetone signal has been previously observed for the single-ame-made silicaalumina (without Pd) particles with higher aluminum content of 70% [44]. Clearly, the double FSP synthesis method can minimize the inuence of Pd nucleation/deposition on the support formation. It is well known that the formation of surface acid sites on amorphous silicaalumina is related to the local structure of silanol groups. Mixing Pd and silicaalumina precursor solutions followed by spraying in a single nozzle system inuences the formation of the surface silicaalumina network. The change of the local structure of the SiOH groups during the incorporation of Pd atoms should directly change the surface acidity of the support material. During double FSP synthesis, the amorphous silicaalumina support is formed as in a single FSP system with minimum effects through Pd loading. When a single aluminum is incorporated in the vicinity of a silanol, the acid strength of this surface OH group is

slightly enhanced. For silicaalumina support material with high aluminum contents, a larger number of aluminum atoms are incorporated into the silica network, and an increasing number of penta-coordinated aluminum (AlV) species occurs as presented in the 27 Al MAS NMR spectra. Assuming AlV species are close to silanol groups, an increasing number of the electron withdrawing centers would lead to a further enhancement of the acids strength of neighboring silanol groups [47]. Lewis acid sites occur with increasing aluminum content as shown in a low-eld 13C MAS NMR signal at d13C = 235245 ppm observed upon adsorption of acetone-2-13C (Fig. 7) [48]. The signals at d13C = 29 and 75 ppm appearing in the spectra shown in Fig. 7 are caused by the conversion of some acetone molecules to diacetone alcohol observed for acetone-2-13C on c-Al2O3 [49]. 3.3. Surface electronic property of Pd particles As reported in many cases [6,7,11,13], the acidity could strongly inuence the catalytic performance of the supported metal nanoparticles. This is attributed to the interaction between the acidic support and the active metal, inducing ionic effects required for the catalytic reactions [6,7,11,13]. XPS has been widely used to

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

15

Fig. 4. Pd particle size distribution (histogram curve) of (a) Pd/SA-0, (b) Pd/SA-10, (c) Pd/SA-30 and (d) Pd/SA-70 (0, 10, 30 and 70 are the Al content in Si).

investigate the surface electronic properties of supported metal nanoparticles [50], and this technique is also employed in the present investigation to study the surface properties of these metal particles distributed in the support. The XPS spectra of Pd3d5/2 and Pd3d3/2 on supports with various acidities are shown in Fig. 8. The binding energies of all double FSPderived samples occur at 335.4 and 340.8 eV, respectively. This nding indicates that for the supported metal catalysts synthesized with double FSP, the electronic properties of metal nanoparticles are independent of the support acidity. This is different from the metal supported on zeolites or mixed oxides where the ionic effects of noble metal particles is known to decrease with increasing the support acidity [13,50]. This metalsupport interaction has also been demonstrated with IR investigation of CO adsorption on noble metal with different support acidity [6,51,52], where a decrease in electron density on the supported metal nanoparticles has been observed. Pd nanoparticles on various silicaalumina supports prepared by double FSP have similar electronic properties and should show similar catalytic performance during hydrogenation reaction. Combining the similar Pd electronic properties with tunable support acidity as described previously, Pd/SAs would show their unique bifunctional catalytic performance in chemoselective hydrogenation. 3.4. Chemoselective hydrogenation of acetophenone The chemoselective hydrogenation of acetophenone (Aph) was carried out over Pd/SAs. The reaction mechanism has been published elsewhere [6,9]. As shown in Scheme 1, the selective hydrogenation of carbonyl group could yield 1-phenylethanol (PhE) and ethylbenzene (EB), and the hydrogenation of phenyl group would produce cyclohexylmethylketone (CMK), ethylcyclohexane (EC), and 1-cyclohexylethanol (CE) [12]. The hydrogenation of Aph started immediately at room temperature under the pressure of 3 bar and PhE was the only product from all the catalysts in 90 min. With time, more and more PhE

was produced, where small amounts of PhE were converted to EB. As shown in Table 2, the selectivity for PhE (SPhE) and EB (SEB) was 9698% and 24% at the Aph conversion of ca. 50% on all supported Pd nano-catalysts. There were no CMK, CE, and EC detected. All catalysts had similar chemoselectivity in the hydrogenation, and only the carbonyl group rather than the aromatic ring was involved in the reactions. This was different with previous observation, in which increasing supports acidity strongly enhanced the selectivity of the phenyl group during the hydrogenation of aromatic ketones at the similar reaction conditions [12]. The reason for the unique chemoselectivity on the double FSP catalysts in the present work could be explained through the similar surface electronic properties of Pd nanoparticles on various silicaalumina supports as detected by XPS. The selectivity to carbonyl groups in the whole reaction range is 100%. However, the yields of PhE were changing from 17% to 59% related to the alumina compositions of catalyst supports. It was proposed that the reaction carried out at the low temperature results in the lower conversion of reactants. And the lower reactant conversion sometimes contributes to the high selectivity of products. When the reactions carried out at the same temperatures on single-ame Pd/SAs (at 100170 C, solvent free [6]) or Pt/SAs (at R.T., Aph hydrogenation with solvent [12]), the catalysts with higher support acidity (related to higher catalytic activity) contributed higher Aph conversion and lower PhE selectivity during the same reaction time. However, all Pd nano-catalysts with various silicaalumina supports in this research have the same chemoselectivity during Aph hydrogenation, even the activity of catalysts increased ca. 4 times with increasing the support acidity. Obviously, tuning support acidity did not change the chemoselectivity of double FSP Pd catalyst, but strongly enhanced their reaction rates in Aph hydrogenation. As shown in Fig. 9, Pd/SA30 has the highest conversion in the reaction and Pd/SA-0 is the most inactive. According to the Pd dispersion data obtained upon CO chemisorption (see Supporting information), the sequence of TOFs for the catalysts in reactions correlates very well with the

16

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

(d) Pd/SA-70

(c) Pd/SA-70

(b) Pd/SA-30

(a) Pd/SA-10

Fig. 5. 1H MAS NMR spectra of double FSP Pd/SA-0 (a), Pd/SA-10 (b), Pd/SA-30 (c) and Pd/SA-70 (d) recorded after dehydrated at 723 K before and after loading with NH3 and subsequent evacuation at 393 K for 1 h. Fig. 6. 27Al MAS NMR spectra of double FSP Pd/SA-0 (a), Pd/SA-30 (b), and Pd/SA-70 (c), obtained with fully hydrated samples, the 27Al MQMAS NMR spectrum of double FSP Pd/SA-70 (d).

density of Brnsted acid sites on their supports: TOF of 1.3 102 s1 for Pd/SA-0 with no Brnsted acid sites, TOF of 2.0 102 s1 for Pd/SA-10 with 5.1 102 mmol/g Brnsted acid sites, TOF of 2.8 102 s1 for Pd/SA-70 with 9.5 102 mmol/g Brnsted acid sites, and TOF of 4.5 102 s1 for Pd/SA-30 with 11 102 mmol/g Brnsted acid sites. Among them, both Pd/SA-

10 and Pd/SA-30 have similar acid strength (dominant peaks at d13C = 214 ppm in Fig. 7 after adsorption of acetone-2-13C). Pd/ SA-30 has ca. 2 times higher density of Brnsted acid sites compared with Pd/SA-10, which contributed ca. 2 times higher TOFs

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

17

(c) Pd/SA-70
217 225 240 75 29

Acetophenone (Aph) OH O

(b) Pd/SA-30
1-Phenylethanol (PhE)
228 240

OH

Cyclohexylmethylketone (CMK)

1-Cyclohexylethanol (CE)

(a) Pd/SA-10

214

Ethylbenzene (EB)
235
Scheme 1.

Ethylcyclohexane (EC)

240

180

120

60

13C / ppm
Fig. 7. 13C CP/MAS NMR spectra of double FSP Pd/SA-0 (a), Pd/SA-10 (b), Pd/SA-30 (c) and Pd/SA-70 (d) recorded after loading of 13C-2-acetone and subsequent evacuation at room temperature.

Table 2 Chemoselective hydrogenation of acetophenone over double FSP Pd/SA catalysts with different Si/Al ratios. Pd/SAs Pd/SA-0 Pd/SA-10 Pd/SA-30 Pd/SA-70
a b

YPhEa% 16.8 31.8 58.7 40.0

SPhEb% 98 97 96.3 96.5

SEBb% 2 3 3.7 3.5

TOFs (s1)c 1.3 102 2.0 102 4.5 102 2.8 102

YPhE is the yield to PhE at the reaction time of 3 h. SPhE and SEB are the selectivity to PhE and EB, respectively, obtained at the Aph conversion of 50%. c TOFs for Pd/SA catalysts were calculated based on Pd dispersion obtained by CO chemisorptions in Supporting information.

Fig. 8. Binding energy of Pd3d5/2 and Pd3d3/2 lines of FSP Pd/SAs samples.

during the reaction. Pd/SA-70 has strongest acid sites (two strong signals at d13C = 217 and 225 ppm in Fig. 7) and just slightly less amount of Brnsted acid sites than Pd/SA-30. However, the TOF for Pd/SA-70 was much lower than Pd/SA-30. It indicates that enhancing the acidic strength of supports cannot accelerate the Aph hydrogenation process. As described by the recent investigation using in situ ATR-IR, Aph can be adsorbed on both the support and the metal nanoparticles, while the hydrogenation was performed on metal surface between adsorbed g1 (O) Aph and dissociated hydrogen [53]. For double FSP Pd/SAs, the similar surface electronic properties should cause similar hydrogen dissociation energy and the similar Aph chem-adsorption energy during the hydrogenation on Pd surface. Therefore, when Aph was adsorbed on Pd surface, the reaction rate should be the same. However, the experimental TOFs were totally different for all catalysts during the reaction. One possible reason is

Fig. 9. The conversion of acetophenone (Aph) over double FSP Pd/SAs materials as function of reaction times at 3 bar hydrogen pressure and 293 K.

the different diffusion rates of Aph on the surface of the catalysts. It was observed that Aph could be desorbed from support surface and diffuses to the metal particle surface for hydrogenation [53]. The high density of acid sites on the support surface could increase the Aph adsorption on the catalysts and also enhance the concentration of Aph around the local environment of Pd-sites due to adsorption-desorption balance promoting diffusion of Aph to the Pd surface for the hydrogenation. However, if the acid sites are

18

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019

too strong, Aph is difcult desorption from support and quickly diffused to Pd surface for the hydrogenation. Another possible reason is that the hydrogenation of Aph performed on the surface acid sites of support with the H-spillover species from metal nanoparticles during the reaction. According to the H-spillover mechanism, H-spilling species created on the metal surface can migrate on the support surface and react with the reactant at a one surface acid site only [5459]. The higher density of acid sites on support could help to adsorb more Aph molecules and offer more surface sites for the reaction with H-spillover species from Pd surface, which may contribute higher reaction rates during chemoselective hydrogenation. 4. Conclusions In the present work, the emerging double-ame spray pyrolysis has been used for the synthesis of Pd nanoparticles on silicaalumina supports. The Pd precursor and mixed silicaalumina precursor were sprayed simultaneously from two different nozzles where the two aerosol streams mixed in situ to form the nal Pd/silica alumina catalysts in a single step. The obtained catalysts have a uniform spherical morphology and their surface areas are decreased from 329 to 151 m2/g with increasing aluminum content. These results are similar to those of the single-ame-derived Pd/ silicaalumina catalysts. Unlike Pd size from 1 to 6 nm on the single FSP Pd catalysts, however, double FSP provided highly crystalline Pd nanoparticles independent of the support composition. Decoupling the two path ways of support formation and metal loading by the double FSP system could minimize the interaction between support nucleation/growth and metal nucleation/ deposition. The uniform Pd nanoparticles were highly dispersed on the supports and mainly located in the vicinity of SiOH groups, which could well combine their bifunctional activities from both Pd active sites and support acid sites in the reaction. The local structure and the surface acidity of supports were changed with various Si/Al ratios. The Pd particles had mean size at 24 nm and the same surface electronic property. This unique property of double FSP catalysts contributed the uniform chemoselectivity to carbonyl group during the hydrogenation of Aph (SPhE 9698% and SEB 24%). It is different from many previous reports that the size of metal particles and the ionic effect on metal particles were changed simultaneously with support acidity. Increasing supports acidity of those catalysts strongly enhanced the selectivity to the phenyl groups during the hydrogenation of aromatic ketones. However, increasing the density of support acid sites of double FSP catalysts extremely enhanced the Aph conversion during hydrogenation. TOF was increased from 1.3 102 s1 to 4.5 102 s1, when the density of Brnsted acid sites was increased up to 11 102 mmol/g. The present research offered an alternative route for designing and producing high-performance supported metal catalysts. The double FSP system is a promising technique to exibly tune the structure and catalytic properties of support and metal nanoparticles for target products and desired catalytic reactions, particularly in chemoselective/enantioselective hydrogenation. Acknowledgments This work was supported by the Early Career Research Scheme and the Major Equipment Scheme from the University of Sydney. S.P. and L.M. would like to thank National Science Foundation, Environmental Protection Agency (Cooperative Agreement No. DBI-0830117), US Public Health Service Grants U19 ES019528 (UCLA Center for NanoBiology and Predictive Toxicology), RO1 ES016746, and RC2 ES018766. Authors would also like

to thank Prof. A. Rosenauer and Dr. M. Schowalter, Department of Physics for microscopic measurements. Appendix A. Supplementary material Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.jcat.2013.02.017. References
[1] R.J. White, R. Luque, V.L. Budarin, J.H. Clark, D.J. Macquarrie, Chem. Soc. Rev. 38 (2009) 481. [2] G.W. Huber, S. Iborra, A. Corma, Chem. Rev. 106 (2006) 4044. [3] M. Minnermann, S. Pokhrel, K. Thiel, R. Henkel, J. Birkenstock, T. Laurus, A. Zargham, J.I. Flege, V. Zielasek, E. Piskorska-Hommel, J. Falta, L. Mdler, M.J. Bumer, Phys. Chem. C 115 (2011) 1302. [4] A.M. Doyle, S.K. Shaikhutdinov, S.D. Jackson, H. Freund, Angew. Chem., Int. Ed. 42 (2003) 5240. [5] A.A. Dabbawala, N. Sudheesh, H.C. Bajaj, Dalton Trans. 41 (2012) 2910. [6] J. Huang, Y. Jiang, N. van Vegten, M. Hunger, A. Baiker, J. Catal. 281 (2011) 352. [7] F. Hoxha, E. Schmidt, T. Mallat, B. Schimmoeller, S.E. Pratsinis, A. Baiker, J. Catal. 278 (2011) 94. [8] C. Chaisuk, P. Boonpitak, J. Panpranot, O. Mekasuwandumrong, Catal. Commun. 12 (2011) 917. [9] N.M. Bertero, A.F. Trasarti, C.R. Apesteguia, A. Marchi, Appl. Catal. A Gen. 394 (2011) 228. [10] N.M. Bertero, C.R. Apesteguia, A.J. Marchi, Catal. Today 172 (2011) 171. [11] E. Schmidt, F. Hoxha, T. Mallat, A. Baiker, J. Catal. 274 (2010) 117. [12] B. Schimmoeller, F. Hoxha, T. Mallat, F. Krumeich, S.E. Pratsinis, A. Baiker, Appl. Catal. A Gen. 374 (2010) 48. [13] F. Hoxha, B. Schimmoeller, Z. Cakl, A. Urakawa, T. Mallat, S.E. Pratsinis, A. Baiker, J. Catal. 271 (2010) 115. [14] N. van Vegten, M. Maciejewski, F. Krumeich, A. Baiker, Appl. Catal. B Environ. 93 (2009) 38. [15] P. Haider, A. Baiker, J. Catal. 248 (2007) 175. [16] R. Strobel, F. Krumeich, W.J. Stark, S.E. Pratsinis, A. Baiker, J. Catal. 222 (2004) 307. [17] R. Mueller, R. Jossen, H.K. Kammler, S.E. Pratsinis, Aiche J. 50 (2004) 3085. [18] D. Dumitriu, A.R. Bally, C. Ballif, P. Hones, P.E. Schmid, R. Sanjins, F. Lvy, V.I. Prvulescu, Appl. Catal. B Environ. 25 (2000) 83. [19] R. Strobel, M. Piacentini, L. Mdler, M. Maciejewski, A. Baiker, S.E. Pratsinis, Chem. Mater. 18 (2006) 2532. [20] K.P. de Jong, in: G. Poncelet, P.A. Jacob, P. Grange, B. Delmon (Eds.), Studies in Surf. Sci. Catal., vol. 63, Elsevier, 1991, p. 19 [21] T. Lopez, M. Asomoza, P. Bosch, E. Garcia-Figueroa, R. Gomez, J. Catal. 138 (1992) 463. [22] C. Sivaraj, C. Contescu, J.A. Schwarz, J. Catal. 132 (1991) 422. [23] B. Schimmoeller, S.E. Pratsinis, A. Baiker, ChemCatChem 3 (2011) 1234. [24] R. Strobel, A. Baiker, S.E. Pratsinis, Adv. Powder Technol. 17 (2006) 457. [25] W.Y. Teoh, R. Amal, L. Mdler, Nanoscale 2 (2010) 1324. [26] H.K. Kammler, L. Mdler, S.E. Pratsinis, Chem. Eng. Technol. 24 (2001) 583. [27] S. Pokhrel, J. Birkenstock, M. Schowalter, A. Rosenauer, L. Mdler, Cryst. Growth Des. 10 (2010) 632. [28] L. Mdler, H.K. Kammler, R. Mueller, S.E. Pratsinis, J. Aerosol. Sci. 33 (2002) 369. [29] K. Shimizu, Y. Miyamoto, T. Kawasaki, T. Tanji, Y. Tai, A. Satsuma, J. Phys. Chem. C 113 (2009) 17803. [30] V. Dal Santo, A. Gallo, A. Naldoni, L. Sordelli, Inorg. Chim. Acta 380 (2012) 216. [31] T. Kimura, T. Miyazawa, J. Nishikawa, S. Kado, K. Okumura, T. Miyao, S. Naito, K. Kunimori, K. Tomishige, Appl. Catal. B Environ. 68 (2006) 160. [32] S. Handjani, E. Marceau, J. Blanchard, J.M. Krafft, M. Che, P. Mki-Arvela, N. Kumar, J. Wrn, D.Y. Murzin, J. Catal. 282 (2011) 228. [33] C.J. Jia, F. Schth, Phys. Chem. Chem. Phys. 13 (2011) 2457. [34] M. Che, C.O. Bennett, Adv. Catal. 36 (1989) 55. [35] M. Arenz, K.J.J. Mayrhofer, V. Stamenkovic, B.B. Blizanac, T. Tomoyuki, P.N. Ross, N.M. Markovic, J. Am. Chem. Soc. 127 (2005) 6819. [36] J. Han, Y. Liu, R. Guo, J. Am. Chem. Soc. 131 (2009) 2060. [37] S. Pokhrel, S. George, Z.X. Ji, B.L. Henderson, T. Xia, L.J. Li, J.I. Zink, A.E. Nel, L. Mdler, J. Am. Chem. Soc. 133 (2011) 11270. [38] A.B. Laursen, K.T. Hjholt, L.F. Lundegaard, S.B. Simonsen, S. Helveg, F. Schth, M. Paul, J.D. Grunwaldt, S. Kegnoes, C.H. Christensen, K. Egeblad, Angew. Chem., Int. Ed. 49 (2010) 3504. [39] C. Wu, L. Wang, P.T. Williams, J. Shi, J. Huang, Appl. Catal. B Environ. 108 (2011) 6. [40] A. Ungureanu, B. Dragoi, A. Chirieac, S. Royer, D. Duprez, E. Dumitriu, J. Mater. Chem. 21 (2011) 12529. [41] G.W. Huber, S. Iborra, A. Corma, Chem. Rev. 106 (2006) 4044. [42] C. Gerardin, S. Sundaresan, J. Benziger, A. Navrotsky, Chem. Mater. 6 (1994) 160. [43] C.E. Volckmar, M. Bron, U. Bentrup, A. Martin, P. Claus, J. Catal. 261 (2009) 1. [44] J. Huang, N. van Vegten, Y. Jiang, M. Hunger, A. Baiker, Angew. Chem., Int. Ed. 49 (2010) 7776. [45] C. Chizallet, P. Raybaud, Angew. Chem., Int. Ed. 48 (2009) 2891.

Z. Wang et al. / Journal of Catalysis 302 (2013) 1019 [46] Y. Jiang, J. Huang, W. Dai, M. Hunger, Solid State Nucl. Magn. Reson. 39 (2011) 116. [47] S.H. Li, A.M. Zheng, Y.C. Su, H.L. Zhang, L. Chen, J. Yang, C.H. Ye, F. Deng, J. Am. Chem. Soc. 129 (2007) 11161. [48] J. Yang, M.J. Janik, D. Ma, A.M. Zheng, M.J. Zhang, M. Neurock, R.J. Davis, C.H. Ye, F. Deng, J. Am. Chem. Soc. 127 (2005) 18274. [49] A.I. Biaglow, R.J. Gorte, D. White, J. Catal. 150 (1994) 221. [50] B.L. Mojet, J.T. Miller, D.E. Ramaker, D.C. Koningsberger, J. Catal. 186 (1999) 373. [51] A.Y. Stakheev, Y. Zhang, A.V. Ivanov, G.N. Baeva, D.E. Ramaker, D.C. Koningsberger, J. Phys. Chem. C 111 (2007) 3938.

19

[52] D.C. Koningsberger, D.E. Ramaker, J.T. Miller, J. de Graaf, B.L. Mojet, Top. Catal. 15 (2001) 35. [53] M. Chen, N. Maeda, A. Baiker, J. Huang, ACS Catal. (2012) 2007. [54] G.M. Pajonk, Appl. Catal. A Gen. 202 (2000) 157. [55] F. Roessner, U. Roland, T.J. Braunschweig, Chem. Soc. Faraday Trans. 91 (1995) 1539. [56] U. Roland, T. Braunschweig, F. Roessner, J. Mol. Catal. A Chem. 127 (1997) 61. [57] F. Gao, A.D. Allian, H.J. Zhang, S.Y. Cheng, M. Garland, J. Catal. 241 (2006) 189. [58] V.S. Ranade, R. Prins, Chem. Eur. J. 6 (2000) 313. [59] Y. Li, G.H. Lai, R.X. Zhou, Appl. Surf. Sci. 253 (2007) 4978.

Вам также может понравиться