ARTICLE IN PRESS
JID: JECHEM
[m5G;August 30, 2018;4:50]
Journal of Energy Chemistry xxx (2018) xxx–xxx
Contents lists available at ScienceDirect
Journal of Energy Chemistry journal homepage: www.elsevier.com/locate/jechem
http://www.journals.elsevier.com/ journal-of-energy-chemistry/
Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene Fei Huang a,b, Zhimin Jia a,b, Jiangyong Diao a, Hua Yuan c, Dangsheng Su a, Hongyang Liu a,∗
Q1
a b c
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, Liaoning, China School of Materials Science and Engineering, University of Science and Technology of China, Hefei 230026, Anhui, China College of Chemistry, Chongqing Normal University, Chongqing 401331, China
a r t i c l e
i n f o
Article history: Received 10 July 2018 Revised 17 August 2018 Accepted 21 August 2018 Available online xxx Keywords: Selective hydrogenation Palladium nanoclusters Nanocarbon support Liquid-phase hydrogenation
Q2
a b s t r a c t We report a nanocarbon material with nanodiamond (ND) core and graphene shell (ND@G) as a support for Pd nanocatalysts. The designed catalyst performed good selectivity of styrene (85.2%) at full conversion of phenylacetylene and superior stability under mild conditions. Supported Pd catalysts are characterized by means of high resolution transmission electron microscopy (HRTEM), Raman, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS) and H2 temperature-programmed reduction (H2 -TPR). The results clearly show that formation of the strong metal-support interaction (SMSI) between Pd nanoclusters and the defective graphene shell helpfully modifies the selectivity and stability of the Pd-based catalysts. © 2018 Published by Elsevier B.V. and Science Press on behalf of Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences
1
1. Introduction
2 3
Selective hydrogenation of alkynes to alkenes, without further reduction to alkanes, is of fundamental importance for the production of fine chemicals as well as polymers throughout the petrochemical and agrochemical industries [1–3]. For instance, the preliminary removal of trace phenylacetylene from styrene feeds is a crucial purification process in olefin industry as phenylacetylene is regarded as not only a harmful component in styrene feedstocks, but also a poison for subsequent olefin polymerization catalysts [3–5]. However, this process requires some precise regulation because the process commonly accompanies with two undesired reactions: full hydrogenation of pheylacetylene to ethylbenzene and excessive hydrogenation of styrene to ethylbenzene. Therefore, the choice of an efficient catalyst which can still maintain a high selectivity of styrene at phenylacetylene conversion near 100 % is of paramount importance. Supported palladium-based catalysts have been widely utilized and discussed for decades in semihydrogenation process owing to their excellent catalytic activity [6,7]. Commercial supported Pd catalysts for this reaction are modified with harmful surface modifiers, such as lead acetate or quinoline (Lindlar catalyst) [8,9]. These catalysts are toxic and large amounts of wastes and environ-
4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22
∗
Corresponding author. E-mail address:
[email protected] (H. Liu).
mental pollutants (Pb or sulfur compounds) are produced together with the desired products. With the environmental concerns, it is proposed to develop non-toxic catalysts to realize green catalytic process. Several significant efforts have been devoted to create Pdbased catalysts that can selectively produce olefins from alkynes. For example, selectively depositing metals or metal oxides on the surface of Pd nanoparticles (NPs) facilitate the alkenes selectivity via geometric and/or electronic efforts [10–12]. Fabrication of intermetallic compounds, such as Pd-Ga, Pd-Zn, Pd-Ag, Pd-Cu, promotes the catalytic selectivity by isolating the active Pd sites in the crystallographic structure [13–17]. Formation single atom Pd catalysts enhances the catalytic selectivity to alkenes through modulating the adsorption properties [18]. Besides, it must be noted that supporting material holds great significance to catalytic performance, since its interaction with the active species may greatly influence the properties of the resulting catalysts [19–21]. Various studies have shown experimentally and theoretically that hydrogen dissolved on active metal Pd surface can diffuse to subsurface sites of Pd to form subsurface-hydrogen species (β palladiumhydride phase, β -H) [22–24]. β -H coordinated with Pd strongly in crystal lattices was found to disfavor selective partial hydrogenation by enhancing undesired total hydrogenation [25]. Furthermore, some studies have also been reported that Pd carbide (Pd-Cx ), formed by diffusing carbonaceous deposits into Pd lattices favors the regenerative of the involved hydrogen species [26,27]. Thus, formation of the strong interaction between Pd and carbon
https://doi.org/10.1016/j.jechem.2018.08.006 2095-4956/© 2018 Published by Elsevier B.V. and Science Press on behalf of Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48
JID: JECHEM 2
ARTICLE IN PRESS
72
supports could helpfully modify the selectivity of the Pd-based catalysts. For an idea green hydrogenation process, the hydrogenation process should be performed under mild conditions using molecular hydrogen as hydrogen source. In the present, we will report the semihydrogenation of phenylacetylene over Pd/ND@G under mild conditions. In our previous studies, ND@G composing a nanodiamond core and a defective, ultrathin graphene nanoshell was used to support noble metals and the resulting catalysts exhibited superior catalytic activity in CO oxidation and dehydrogenation of alkanes [21,29,30]. In this work, the ND@G material was firstly used in liquid-phase hydrogenation reaction systems to modify the selectivity of the incorporated Pd nanoclusters through a strong metal-support interaction (SMSI) and to promote stability of the Pd clusters. After immobilization Pd NPs, the resulting catalysts (1%Pd/ND@G) were evaluated in the selective hydrogenation of phenylacetylene under mild conditions and compared with onionlike carbon (OLC) supported Pd catalysts (1%Pd/OLC) and traditional carbon nanotubes (CNTs) supported Pd catalysts (1%Pd/CNTs) and commercial carbon supported Pd catalysts (commercial Pd/C). 1% Pd/ND@G exhibits an outstanding selectivity of styrene (85.2%) at total phenylacetylene conversion. We demonstrated that the defects on ND@G account for the SMSI and the SMSI facilitates the catalytic selectivity and stability.
73
2. Experimental
74
2.1. Material
75
Nanodiamond (ND) powders (99.9%) were purchased from Beijing Grish Hitech Co., and further purified by hydrochloric acid. Carbon nanotubes (CNTs) were supplied by Tsinghua University and treated with hydrochloric acid before use. Pd precursor (Pd(NO3 )2 solution), commercial Pd/C catalyst, phenylacetylene, styrene, ethylbenzene and ethanol were analytical regents and all purchased from Alfa Aesar without further purification. ND powders were annealed in Ar flow (100 mL/min) for 4 h at 90 0, 130 0 °C and then used as carbon supports.
49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71
76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99
2.2. Catalyst preparation Supported Pd catalysts were prepared by the depositionprecipitation method as our previous report [29]. First, 200 mg carbon support was dispersed into 30 mL deionized water in a 100 mL round-bottom flask, and the mixture was ultrasonically treated to obtain a homogenous suspension. Then, the pH value of carbon supports suspension was adjusted to about 10 by dropping 0.25 M Na2 CO3 solution. Second, the pH of Pd(NO3 )2 solution containing 16 mg/mL Pd was adjusted using 0.25 M Na2 CO3 to neutral. Subsequently, the Pd(NO3 )2 solution was added to above carbon support suspension dropwise under magnetic stirring at 100 °C, and then kept stirring at 100 °C in oil bath for one hour. At the end, the mixture was cooled to room temperature, collected by filter and washed several times with deionized water until it was free from Na+ , CO3 2− . Afterwards, the powders were dried at 60 °C for 12 h and reduced with H2 at 200 °C for 2 h.
100
2.3. Catalyst characterization
101
Transmission electron microscopy (TEM) images were obtained by using an FEI Tecnai G2 F20 microscope operated at 200 kV. High-angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) images were taken on an aberrationcorrected JEOL JEM-ARM200 operated at 200 kV. N2 physisorption were measured at −196 °C using a Micrometrics ASAP-2020 instrument. The X-ray photoelectron spectroscopy (XPS) was carried out
102 103 104 105 106 107
[m5G;August 30, 2018;4:50]
F. Huang et al. / Journal of Energy Chemistry xxx (2018) xxx–xxx
at ESCALAB 250 instrument with Al Kα X-rays (1489.6 eV, 150 W, 50.0 eV pass energy). UV-Raman spectroscopy was performed on powder samples by using and HORIBA LabRam HR Raman spectrometer. The excitation wavelength was 325 nm and a power of 0.2 mW. Temperature-programmed reduction (TPR) was performed on Catlab with a QIC-20 gas analysis system from Hiden Analytical in He and the heating rate was 10 K/min from room temperature to 360 °C. The loading amount of Pd was determined by inductively coupled plasma mass spectrometry (ICP-MS). The XRD patterns of the nanocarbon supported Pd catalysts were collected by using an X-ray diffractometer (D/MAX-2400) using a Cu Kα source at a scan rate of 2 °/min. 2.4. Catalytic reaction Liquid-phase selective hydrogenation of phenylacetylene was carried out in a 50 mL autoclave (silica vessel) under H2 pressure of 0.1 MPa at 303 K, and the stirring speed was set at 800 rpm. In a typical procedure, 5 mg catalyst, 1.85 mmol phenylacetylene and 10 mL ethanol were placed in the vessel. After the replacement of air with Ar for 3 times, the autoclave was filled with 1 bar H2 and then magnetically stirred. The reaction products were monitored and analyzed by gas chromatographic analysis (Agilent 7890A, internal standard n-octane). The conversion, selectivity, and reaction rate were calculated on basis of the following equations:
108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130
Conversion (% ) feed (mol )−Phenylacetylene residue(mol ) = PhenylacetylenePhenylacetylene × 100% feed (mol ) Selectivity (% ) Styrene product (mol ) = Phenylacetylene feed × 100% (mol )− Phenylacetylene residue (mol ) Reactionrate(r ) = − ddCt
3. Results and discussion
131
ND@G and OLC supporting materials with exceptional structures and properties were prepared by annealing ND powders at different temperatures according to our previous report [29]. As shown in Fig. 1(a), ND@G consists of a well-defined crystalline diamond with (111) planes covered by a distorted graphene layer. When annealed at 1300 °C, the curved graphene layers increase and enclose to form highly ordered spherical OLC with distances between neighboring shells close to 3.4 A˚ (Fig. 1b). Raman spectra of different carbon supports exhibited various main feature bands as shown in Fig. S1. The results from Raman spectra also confirmed that the content of sp2 -hybridized carbon atoms of ND@G is much lower than that of OLC. The ratio of ID1 /IG from Raman spectra still indicates that the much more defectiveness on ND@G (ID1 /IG = 0.70) with a thin, defective, curved graphene outer shell than that on OLC (ID1 /IG = 0.56) was obtained. BET surface area and pore volume increased slightly with increasing annealing temperature as shown in Table S1. The unchanged pore size distributions further confirmed that the integrity of the framework of ND-derived samples was not affected by the annealing treatment. Palladium NPs immobilized on carbon supports via a deposition-precipitation method. And the loading amount of Pd NPs on carbon supports was about 1 wt % respectively and was confirmed by inductively coupled plasma mass spectrometry (Table 2). Fig. 1(c and d) shows the STEM images of the fresh catalysts and the size distribution of Pd NPs. Pd NPs were uniformly dispersed on ND@G and OLC. The average size of Pd NPs on ND@G was about 1.87 nm, which was relatively smallest than that on OLC (about 2.3 nm) and CNTs (about 2.85 nm, Fig. S2). This result suggested that ND@G with more defectiveness benefits the dispersion and anchoring of Pd NPs. From XRD patterns of fresh catalysts (Fig. S3), no remarkable diffraction peaks attributed to Pd
132
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163
JID: JECHEM
ARTICLE IN PRESS
[m5G;August 30, 2018;4:50]
F. Huang et al. / Journal of Energy Chemistry xxx (2018) xxx–xxx
3
Fig. 1. HRTEM images of ND@G (a) and OLC (b); HAADF-STEM images of catalysts 1% Pd/ND@G (c) and 1% Pd/OLC (d), the insets are the corresponding particle size distribution histograms; High-resolution HAADF-STEM images of 1% Pd/ND@G (e) and 1% Pd/OLC (f).
164 165 166 167 168
crystalline were detected, which further indicated that little size Pd NPs were uniformly dispersed on all ND derived supports and CNTs. In addition, XRD patterns provide evidence of the change of graphitization degree in the surface of ND@G, OLC and CNTs. This result corresponded to Raman characterization (Fig. S1).
High-resolution HAADF-STEM images of 1% Pd/ND@G and 1% Pd/OLC were shown in Fig. 1(e and f). Quite different structures of Pd NPs were clearly observed. From Fig. 1(e), Pd NPs on ND@G show more irregular shape, lower crystallinity than that on OLC (Fig. 1f). This phenomenon was presumably attributed to SMSI in
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
169 170 171 172 173
JID: JECHEM 4
ARTICLE IN PRESS
Fig. 2. Yield and selectivity for hydrogenation phenylacetylene to styrene at full conversion over 1% Pd/ND@G, 1% Pd/OLC, 1% Pd/CNTs and Commercial Pd/C.
174 175 176 177 178 179 180 181 182 183 184 185 186 Q3 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201
[m5G;August 30, 2018;4:50]
F. Huang et al. / Journal of Energy Chemistry xxx (2018) xxx–xxx
Pd/ND@G. SMSI may also result in the formation of Pd-Cx carbide, which not only stabilizes Pd NPs but also helpfully modifies the selectivity of Pd-based catalysts in selective hydrogenation reactions. The catalytic activities of samples were evaluated by selective hydrogenation of phenylacetylene. As shown in Fig. S4, for the 1% Pd/ND@G catalyst sample, complete conversion of phenylacetylene occurred at 70 min. Accompanying with the increase of graphitization degree of different supports, the initial activity of catalysts rose. The conversion of phenylacetylene reached 100% at above 35 min over 1% Pd/OLC and above 30 min over 1% Pd/CNTs. The graphitization degree of supports may affect the specific activity of catalysts, due to the different adsorption capacity of supports to reactants through π -π stacking. However, the selectivity to styrene for 1% Pd/OLC and 1% Pd/CNTs droped straightly with prolongation of reaction time. Especially, the selectivity falled quickly after the full conversion and zero selectivity to styrene was obtained after 60 min (Fig. S4). In constract, the 1% Pd/ND@G can keep higher selectivity during the reaction. The selectivity to styrene in the same phenylacetylene conversion was in the order: 1%Pd/ND@G > 1% Pd/OLC > 1% Pd/CNTs (Table S2). Fig. 2 also clearly exhibited the catalytic performance of all the Pd nanocatalysts. Compared with other catalysts, 1% Pd/ND@G performed better catalytic selectivity and yield. Although the initial activity of 1% Pd/ND@G was lower than those of 1% Pd/OLC and 1% Pd/CNTs, the catalytic selectivity and yield over 1% Pd/ND@G was much more than those over other catalysts. In the selective hydrogenation of alkynes to alkenes, improving catalytic selectivity of catalysts could be much more important.
Detailed kinetic parameters were often applied to further quantitatively analyze and understand the influence of catalysts on the catalytic process [31,32]. The value of reaction rate r, reaction rate constant k and the ratio of k2 /k1 (phenylacetylene hydrogenation rate constant k1 , styrene hydrogenation rate constant k2 ) were shown in Fig. 3 and listed in Table 1. The ratio of k2 /k1 is used to quantitatively evaluate whether the catalyst facilitate styrene production and the lower ratio are more prone to maintain the higher styrene selectivity (see the ESI for detailed information). The 1% Pd/ND@G sample showed the lowest ratio k2 /k1 (0.74) indicating that the 1% Pd/ND@G may exhibit the highest selectivity of styrene in the semihydrogenation reaction of phenylacetylene due to the relative lower reaction rate in r2 . To further explore the origin of the catalysts sample for the partial reduction of alkynes to alkene, XPS and H2 temperature programmed reduction (H2 -TPR) were examined. As shown in Fig. 3(a), the binding energy (BE) of Pd 3d5/2 for 1% Pd/ND@G appears at 336.1 eV, which was higher compared to 1% Pd/ND@G, 1% Pd/OLC (BE value of Pd 3d5/ 2 = 335.7 eV) and 1% Pd/CNTs samples (BE value of Pd 3d5/ 2 = 335.9 eV). Higher BE shift of Pd 3d for 1% Pd/ND@G was contributed to SMSI between Pd NPs and supports. This result agreed with other previous reports [32,33]. Furthermore, the detailed XPS spectrum of the Pd lever was displayed in Fig. 3(a). Peaks observed at BE value of 335.1 and 340.5 eV correspond to the 3d5/2 and 3d3/2 levels of metallic state pallidium (Pd0 ) [34,35]. These peaks all shifted to lower BE value compared with 1% Pd/OLC and 1% Pd/CNTs in our work, indicating that the surfaces of Pd metal particles are more electric-rich. The peaks at 336.3 and 343.4 eV can be assigned to the oxidated state pallidium (Pd2+ ). It was obviously observed that the Pd was oxidized further in 1% Pd/ND@G. The ratio of Pd0 to Pd0 + Pd2+ , which was estimated from XPS data, decreased significantly in 1% Pd/ND@G than that in 1% Pd/OLC and 1% Pd/CNTs (Table 2). The much more existence of Pd2+ in 1% Pd/ND@G indicated that Pd NPs are bonded much stronger to the support [36]. As shown in Fig. 3(b), an obvious peak ranged from 30 °C to 360 °C occurred in the H2 -TPR profiles of both 1% Pd/ND@G and 1% Pd/OLC catalysts. The negative peak is attributed to the consumption of H2 and the location of consumption peak represents reduction temperature. Generally, a higher reduction temperature means a higher SMSI [21,37]. The reduction temperature of 1% Pd/ND@G was about 50 degree higher than that of 1% Pd/OLC. These results further confirmed that Pd NPs in sample 1% Pd/ND@G strongly bonded with carbon support. Diffusing carbonaceous from ND@G support into Pd lattices lead to the decrease of Pd crystallinity (forming Pd nanoclusters, as shown in Fig. 1e), and further hider the active hydrogen species diffusing
Fig. 3. Concentration of phenylacetylene (a) and styrene (b) vs. time during the phenylacetylene and styrene hydrogenation respectively. 1%Pd/ND@G, 1% Pd/OLC and 1% Pd/CNTs lines are derived from the linear fit of data at the beginning of reactions to obtain approximate initial reaction rate r.
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247
ARTICLE IN PRESS
JID: JECHEM
[m5G;August 30, 2018;4:50]
F. Huang et al. / Journal of Energy Chemistry xxx (2018) xxx–xxx
5
Table 1. The initial reaction rate and rate constant of phenylacetyrene and styrene hydrogenation over three different catalysts. Catalysts
r1 (mol·L−1 ·min−1 ·10−3 )
r2 (mol·L−1 ·min−1 ·10−3 )
k1 (mol·L−1 ·g−1 cat. ·MPa−1 ·min−1 )
k2 (mol·L−1 ·g−1 cat. ·MPa−1 ·min−1 )
k2 /k1
1% Pd/ND@G 1% Pd/OLC 1% Pd/CNTs
2.7 5.6 6.6
2.0 4.8 5.8
5.4 11.2 13.2
4 9.6 11.6
0.74 0.86 0.88
Table 2. Physical-chemical properties of supported Pd catalysts. Samples
Pd loading (%)a
Pd size (nm)b
Pd existence statesc
1% Pd/ND@G 1% Pd/OLC 1% Pd/CNTs
0.87 0.86 0.89
1.87 2.30 2.85
Pd0 ; Pd2+ Pd0 Pd0
a b c
Pd0 /(Pd2+ + Pd0 ) = 0.55 Pd0 /(Pd2+ + Pd0 ) = 0.90 Pd0 /(Pd2+ + Pd0 ) = 0.92
Determined by ICP; Average size observed by TEM; Determined by XPS.
Fig. 5. Reusability of 1% Pd/ND@G (a) and 1% Pd/CNTs (b) for six cycles.
crease of conversion and 10% decrease of selectivity) of both conversion and selectivity over 1% Pd/CNTs was observed after six cycles (Fig. 5b). This result further indicated that SMSI on surface of 1% Pd/ND@G still facilitates catalyst durability. Q4
257 258 259 260
Fig. 4. (a) Pd 3d XPS spectra of 1% Pd/ND@G, 1% Pd/OLC and 1% Pd/CNTs; (b) H2 TPR profiles of 1% Pd/ND@G and 1% Pd/OLC.
248 249 250 251 252 253 254 255 256
from Pd nanocrystal surface to subsurface. The formation of PdCx carbide through the strong interaction between Pd and carbon supports helpfully modifies the selectivity of 1% Pd/ND@G. On the contrary, the higher the reduction degree of Pd, the higher catalytic activity but the lower selectivity for hydrogenation [38] (Fig. 4). Besides, we further investigated the catalytic stability of 1% Pd/ND@G and 1% Pd/CNTs. As shown in Fig. 5(a), slightly decrease in the conversion and no clear selectivity over 1% Pd/ND@G was observed after six cycles. However, a large extent decrease (20% de-
4. Conclusions
261
In conclusion, as a heterogeneous catalyst support, ND@G material greatly influenced the properties of the resulting catalysts 1% Pd/ND@G through SMSI. These results further confirmed that formation of Pd-Cx carbide through the strong interaction between Pd and carbon supports could modify the catalytic selectivity of Pd-based catalysts in selective hydrogenation reactions. Besides, 1% Pd/ND@G demonstrated quite stable performance in recycling tests, which indicated that SMSI was conductive to the stability of catalysts.
262
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
263 264 265 266 267 268 269 270
JID: JECHEM 6 Q5 271
272
ARTICLE IN PRESS
Uncited references [28].
273
Acknowledgments
274
281 Q6 282
This work was supported by the Ministry of Science and Technology (2016YFA0204100), the National Natural Science Foundation of China (21573254, 21703261 and 91545110), the Youth Innovation Promotion Association (CAS), and the Sinopec China and Strategic Priority Research Program of the Chinese Academy of Sciences (XDA09030103), the Chongqing Research Program of Basic Research and Frontier Technology (cstc2016jcyjA0432), Scientific and Technological Research Program of Chongqing Municipal Education Commission (KJ1600328).
283
Supplementary materials
284 285
Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.jechem.2018.08.006.
286
References
287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307
[1] Y.X. Han, D. Peng, Z.Y. Xu, H.Q. Wan, S.R. Zheng, D.Q. Zhu, Chem. Commun. 49 (2013) 8350–8352. [2] L. Wang, F.X. Li, Y.J. Chen, J.X. Chen, J. Energy Chem. (2018) in press, doi:10. 1016/j.jechem.2018.02.001. [3] E. Esmaeili, A.M. Rashidi, Y. Mortazavi, A.A. Khodadadi, M. Rashidzadeh, J. Energy Chem. 22 (2013) 717–725. [4] D.S Deng, Y. Yang, Y.T. Gong, Y. Li, X. Xu, Y. Wang, Green Chem. 15 (2013) 2525–2531. [5] X.L. Li, W. Zhang, Y.S. Liu, R. Li, ChemCatChem 8 (2016) 1111–1118. [6] K. Tie, X.L. Pan, T. Yu, P. Li, L.M. He, X.H. Bao, J. Energy Chem. (2018) in press, doi:10.1016/j.jechem.2018.05.021. [7] W. Liu, C.O. Arean, S. Bordiga, E. Groppo, A. Zecchina, ChemCatChem 3 (2011) 222–226. [8] G. Vilé, D. Albani, M. Nachtegaal, Z. Chen, D. Dontsova, M. Antonietti, N. López, J. Pérez-Ramírez, Angew. Chem. 54 (2015) 11265–11269. [9] D. Teschner, J. Borsodi, A. Wootsch, Z. Révay, M. Hävecker, A. Knop-Gericke, S.D. Jackson, R. Schlögl, Science 320 (2008) 86–89. [10] L.B. Ding, H. Yi, W.H. Zhang, R. You, T. Cao, J.L. Yang, J.L. Lu, W.X. Huang, ACS Catal. 6 (2016) 3700–3707. [11] S.L. Yang, C.Y. Cao, L. Peng, J.L. Zhang, B.X. Han, W.G. Song, Chem. Commun. 52 (2016) 3627–3630.
275 276 277 278 279 280
[m5G;August 30, 2018;4:50]
F. Huang et al. / Journal of Energy Chemistry xxx (2018) xxx–xxx [12] L.D. Li, B.S. Zhang, E. Kunkes, K. Föttinger, M. Armbrüster, D.S. Su, W. Wei, R. Schlögl, M. Behrens, ChemCatChem 4 (2012) 1764–1775. [13] S.H. Wang, Z.L. Xin, X. Huang, W.Z. Yu, S. Niu, L.D. Shao, Phys. Chem. Chem. Phys. 19 (2017) 6164–6168. [14] M.B. Boucher, B. Zugic, G. Cladaras, J. Kammert, M.D. Marcinkowski, T.J. Lawton, E.C.H. Sykes, M. Flyzani-Stephanopoulos, Phy. Chem. Chem. Phy. 15 (2013) 12187–12196. [15] H.R. Zhou, X.F. Yang, L. Li, X.Y. Liu, Y.Q. Huang, X.L. Pan, A.Q. Wang, J. Li, T. Zhang, ACS Catal. 6 (2016) 1054–1061. ¨ J. Am. [16] M. Armbrus¨ ter, K. Kovnir, M. Behrens, D. Teschner, Y. Grin, R. Schlogl, Chem. Soc. 132 (2010) 14745–14747. [17] H. Tian, F. Huang, Y.H. Zhu, S.M. Liu, Y. Han, M. Jaroniec, Q.H. Yang, H.Y. Liu, G.Q.M. Lu, J. Liu, Adv. Funct. Mater. 28 (2018) 1801737–1801747. [18] H. Yan, H. Cheng, H. Yi, Y. Lin, T. Yao, C.L. Wang, J.J Li, S.Q. Wei, J.L. Lu, J. Am. Chem. Soc. 137 (2015) 10484–10487. [19] N.J. Divins, I. Angurell, C. Escudero, V. Pérez-Dieste, J. Llorca, Science 346 (2014) 620–623. [20] W. Shi, B.S. Zhang, Y.M. Lin, Q. Wang, Q. Zhang, D.S. Su, ACS Catal. 6 (2016) 7844–7854. [21] J. Liu, Y.Y. Yue, H.Y. Liu, Z.J. Da, C.J. Liu, A.Z. Ma, J.F. Rong, D.S. Su, X.J. Bao, H.D. Zheng, ACS Catal. 7 (2017) 3349–3355. [22] A.D. Johnson, S.P. Daley, A.L. Utz, S.T. Ceyer, Science 257 (1992) 223–225. [23] V. Ledentu, W. Dong, P. Sautet, J. Am. Chem. Soc. 122 (20 0 0) 1796–1801. [24] A. Michaelides, P. Hu, A. Alavi, J. Chem. Phys. 111 (1999) 1343–1345. [25] N.A. Khan, S. Shaikhutdinov, J.-H. Freund, Catal. Lett. 108 (2006) 159–164. [26] M. Wilde, K. Fukutani, W. Ludwig, B. Brandt, J.-H. Fischer, S. Schauermann, H.J. Freund, Angew. Chem. Int. Ed. 47 (2008) 9289–9293. [27] L.D. Shao, B.S. Zhang, W. Zhang, D. Teschner, F. Girgsdies, R. Schlögl, D.S. Su, Chem. Eur. J. 18 (2012) 14962–14966. [28] L.Y. Zhang, H.Y. Liu, X. Huang, X.P. Sun, Z. Jiang, R. Schlögl, D.S. Su, Angew. Chem. Int. Ed. 54 (2015) 15823–15826. [29] R. Wang, X.Y. Sun, B.S. Zhang, X.Y. Sun, D. Su, Chem. Eur. J. 20 (2014) 6324– 6331. [30] F. Huang, H.Y. Liu, D.S. Su, Sci. China Mater. 60 (2017) 1–19. [31] B.A. Wilhite, M.J. McCready, A. Varma, Ind. Eng. Chem. Res. 41 (2002) 3345– 3350. [32] Y.F. Wang, C. Zhu, Y.Y. Yang, Z.G. Zhao, J. Energy Chem. 27 (2018) 389–394. [33] J. Panpranot, K. Phandinthong, T. Sirikajorn, M. Arai, P. Praserthdam, J. Mol. Catal. A Chem. 261 (2007) 29–35. [34] L.J. Lei, Z.W. Wu, R.Y. Wang, Z.F. Qin, C.M. Chen, Y.Q. Liu, G.F. Wang, W.B. Fan, J.G. Wang, Catal. Sci. Tech. 7 (2017) 5650–5661. [35] X.M. Wang, G.J. Wu, N.J. Guan, L.D. Li, Appl. Catal. B 115 (2012) 7–15. [36] G. Spezzati, Y.Q. Su, J.P. Hofmann, A.D. Benavidez, A.T. DeLaRiva, J. McCabe, A.K. Datye, E.J.M. Hensen, ACS Catal. 7 (2017) 6887–6891. [37] H. Armendáriz, A. Guzmán, J.A. Toledo, M.E. Llanos, A. Vázquez, G. Aguilar-Ríos, Appl. Catal. A 211 (2001) 69–80. [38] J.W. Sun, Y.S. Fu, G.Y. He, X.Q. Sun, X. Wang, Catal. Sci. Tech. 4 (2014) 1742– 1748.
Please cite this article as: F. Huang et al., Palladium nanoclusters immobilized on defective nanodiamond-graphene core-shell supports for semihydrogenation of phenylacetylene, Journal of Energy Chemistry (2018), https://doi.org/10.1016/j.jechem.2018.08.006
308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355