Influence of surface strain on activity and selectivity of Pd based catalysts for the hydrogenation of acetylene: A DFT study

Similar documents
Effect of lengthening alkyl spacer on hydroformylation performance of tethered phosphine modified Rh/SiO2 catalyst

The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA 报告人 : 沈胤

Effect of promoters on the selective hydrogenolysis of glycerol over Pt/W containing catalysts

Synthesis of anisole by vapor phase methylation of phenol with methanol over catalysts supported on activated alumina

Fabrication of ultrafine Pd nanoparticles on 3D ordered macroporous TiO2 for enhanced catalytic activity during diesel soot combustion

Enhancement of the activity and durability in CO oxidation over silica supported Au nanoparticle catalyst via CeOx modification

Ni based catalysts derived from a metal organic framework for selective oxidation of alkanes

A new approach to inducing Ti 3+ in anatase TiO2 for efficient photocatalytic hydrogen production

available at journal homepage:

A highly efficient flower-like cobalt catalyst for electroreduction of carbon dioxide

Activity and selectivity of propane oxidative dehydrogenation over VO3/CeO2(111) catalysts: A density functional theory study

Growth of Cu/SSZ 13 on SiC for selective catalytic reduction of NO

Hydrothermal synthesis of nanosized ZSM 22 and their use in the catalytic conversion of methanol

Coating Pd/Al2O3 catalysts with FeOx enhances both activity and selectivity in 1,3 butadiene hydrogenation

Proton gradient transfer acid complexes and their catalytic performance for the synthesis of geranyl acetate

Photo induced self formation of dual cocatalysts on semiconductor surface

Species surface concentrations on a SAPO 34 catalyst exposed to a gas mixture

Single-atom catalysis: Bridging the homo- and heterogeneous catalysis

Source mechanism solution

On the Quark model based on virtual spacetime and the origin of fractional charge

Zinc doped g C3N4/BiVO4 as a Z scheme photocatalyst system for water splitting under visible light

The preload analysis of screw bolt joints on the first wall graphite tiles in East

Synthesis of PdS Au nanorods with asymmetric tips with improved H2 production efficiency in water splitting and increased photostability

d) There is a Web page that includes links to both Web page A and Web page B.

Magnetic Co/Al2O3 catalyst derived from hydrotalcite for hydrogenation of levulinic acid to γ-valerolactone

2012 AP Calculus BC 模拟试卷

Effects of Au nanoparticle size and metal support interaction on plasmon induced photocatalytic water oxidation

Atomic & Molecular Clusters / 原子分子团簇 /

系统生物学. (Systems Biology) 马彬广

Galileo Galilei ( ) Title page of Galileo's Dialogue concerning the two chief world systems, published in Florence in February 1632.

SiO2 supported Au Ni bimetallic catalyst for the selective hydrogenation of acetylene

Ho modified Mn Ce/TiO2 for low temperature SCR of NOx with NH3: Evaluation and characterization

NiFe layered double hydroxide nanoparticles for efficiently enhancing performance of BiVO4 photoanode in

Catalytic combustion of methane over Pd/SnO2 catalysts

三类调度问题的复合派遣算法及其在医疗运营管理中的应用

Cobalt nanoparticles encapsulated in nitrogen doped carbon for room temperature selective hydrogenation of nitroarenes

Visible light responsive carbon decorated p type semiconductor CaFe2O4 nanorod photocatalyst for efficient remediation of organic pollutants

Integrating non-precious-metal cocatalyst Ni3N with g-c3n4 for enhanced photocatalytic H2 production in water under visible-light irradiation

Influence of nickel(ii) oxide surface magnetism on molecule adsorption: A first principles study

USTC SNST 2014 Autumn Semester Lecture Series

Pore structure effects on the kinetics of methanol oxidation over nanocast mesoporous perovskites

Silver catalyzed three component reaction of phenyldiazoacetate with arylamine and imine

available at journal homepage:

Effect of the degree of dispersion of Pt over MgAl2O4 on the catalytic hydrogenation of benzaldehyde

There are only 92 stable elements in nature

Synthesis of Ag/AgCl/Fe S plasmonic catalyst for bisphenol A degradation in heterogeneous photo Fenton system under visible light irradiation

Key Topic. Body Composition Analysis (BCA) on lab animals with NMR 采用核磁共振分析实验鼠的体内组分. TD-NMR and Body Composition Analysis for Lab Animals

Preparation of LaMnO3 for catalytic combustion of vinyl chloride

Lecture 4-1 Nutrition, Laboratory Culture and Metabolism of Microorganisms

Wheat flour derived N doped mesoporous carbon extrudes as an efficient support for Au catalyst in acetylene hydrochlorination

天然气化学 Natural Gas Chemistry

Unsupported nanoporous palladium catalyzed chemoselective hydrogenation of quinolines: Heterolytic cleavage of H2 molecule

Surface treatment effect on the photocatalytic hydrogen generation of CdS/ZnS core shell microstructures

Promotional effects of Er incorporation in CeO2(ZrO2)/TiO2 for selective catalytic reduction of NO by NH3

In plasma catalytic degradation of toluene over different MnO2 polymorphs and study of reaction mechanism

Rigorous back analysis of shear strength parameters of landslide slip

Highly selective hydrogenation of furfural to tetrahydrofurfuryl alcohol over MIL 101(Cr) NH2 supported Pd catalyst at low temperature

能源化学工程专业培养方案. Undergraduate Program for Specialty in Energy Chemical Engineering 专业负责人 : 何平分管院长 : 廖其龙院学术委员会主任 : 李玉香

N doped ordered mesoporous carbon as a multifunctional support of ultrafine Pt nanoparticles for hydrogenation of nitroarenes

通量数据质量控制的理论与方法 理加联合科技有限公司

Effect of Gd0.2Ce0.8O1.9 nanoparticles on the oxygen evolution reaction of La0.6Sr0.4Co0.2Fe0.8O3 δ anode in solid oxide electrolysis cell

Synthesis and photocatalytic hydrogen production activity of the Ni CH3CH2NH2/H1.78Sr0.78Bi0.22Nb2O7 hybrid layered perovskite

Design, Development and Application of Northeast Asia Resources and Environment Scientific Expedition Data Platform

SnO2 based solid solutions for CH4 deep oxidation: Quantifying the lattice capacity of SnO2 using an X ray diffraction extrapolation method

Mesoporous polyoxometalate based ionic hybrid as a highly effective heterogeneous catalyst for direct hydroxylation of benzene to phenol

Concurrent Engineering Pdf Ebook Download >>> DOWNLOAD

Selective suppression of toluene formation in solvent free benzyl alcohol oxidation using supported Pd Ni bimetallic nanoparticles

GRE 精确 完整 数学预测机经 发布适用 2015 年 10 月考试

Resistance to SO2 poisoning of V2O5/TiO2 PILC catalyst for the selective catalytic reduction of NO by NH3

Steering plasmonic hot electrons to realize enhanced full spectrum photocatalytic hydrogen evolution

( 选出不同类别的单词 ) ( 照样子完成填空 ) e.g. one three

Preparation of N vacancy doped g C3N4 with outstanding photocatalytic H2O2 production ability by dielectric barrier discharge plasma treatment

Homogeneous boron doping in a TiO2 shell supported on a TiB2 core for enhanced photocatalytic water oxidation

Cu catalyzed deoxygenative gem hydroborylation of aromatic aldehydes and ketones to access benzylboronic esters

Synthesis of propylene glycol ethers from propylene oxide catalyzed by environmentally friendly ionic liquids

Geomechanical Issues of CO2 Storage in Deep Saline Aquifers 二氧化碳咸水层封存的力学问题

One step synthesis of graphitic carbon nitride nanosheets for efficient catalysis of phenol removal under visible light

澳作生态仪器有限公司 叶绿素荧光测量中的 PAR 测量 植物逆境生理生态研究方法专题系列 8 野外进行荧光测量时, 光照和温度条件变异性非常大 如果不进行光照和温度条件的控制或者精确测量, 那么荧光的测量结果将无法科学解释

Synthesis of FER zeolite with piperidine as structure directing agent and its catalytic application

available at journal homepage:

available at journal homepage:

Photocatalytic hydrogen evolution activity over MoS2/ZnIn2S4 microspheres

Tuning the growth of Cu MOFs for efficient catalytic hydrolysis of carbonyl sulfide

La doped Pt/TiO2 as an efficient catalyst for room temperature oxidation of low concentration HCHO

上海激光电子伽玛源 (SLEGS) 样机的实验介绍

Tailored one-pot production of furan-based fuels from fructose in an ionic liquid biphasic solvent system

Co(III)/Zn(II) catalyzed dearomatization of indoles and coupling with carbenes from ene yne ketones via intramolecular cyclopropanation

Principia and Design of Heat Exchanger Device 热交换器原理与设计

Biomolecule assisted, cost effective synthesis of a Zn0.9Cd0.1S solid solution for efficient photocatalytic hydrogen production under visible light

Highly enhanced visible-light photocatalytic hydrogen evolution on g-c3n4 decorated with vopc through - interaction

available at journal homepage:

Effect of the support on cobalt carbide catalysts for sustainable production of olefins from syngas

Low cost and efficient visible light driven microspheres fabricated via an ion exchange route

Solvent free thermal decomposition of methylenediphenyl di(phenylcarbamate) catalyzed by nano Cu2O

An efficient and stable Cu/SiO2 catalyst for the syntheses of ethylene glycol and methanol via chemoselective hydrogenation of ethylene carbonate

A review of the direct oxidation of methane to methanol

Numerical Analysis in Geotechnical Engineering

Enhanced visible light photocatalytic oxidation capability of carbon doped TiO2 via coupling with fly ash

Characterization and activity of V2O5-CeO2/TiO2-ZrO2 catalysts for NH3-selective catalytic reduction of NOx

Transcription:

Chinese Journal of Catalysis 39 (2018) 1493 1499 催化学报 2018 年第 39 卷第 9 期 www.cjcatal.org available at www.sciencedirect.com journal homepage: www.elsevier.com/locate/chnjc Article Influence of surface strain on activity and selectivity of Pd based catalysts for the hydrogenation of acetylene: A DFT study Ping Wang, Bo Yang * School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China A R T I C L E I N F O A B S T R A C T Article history: Received 1 March 2018 Accepted 7 April 2018 Published 5 September 2018 Keywords: Surface strain Pd Acetylene hydrogenation Selectivity Activity Subsurface Density functional theory Microkinetic modelling The effects of surface strain and subsurface promoters, which are both important factors in heterogeneous catalysis, on catalytic selectivity and activity of Pd are examined in this study by considering the selective hydrogenation of acetylene as an example. Combined density functional theory calculations and microkinetic modeling reveal that the selectivity and activity of the Pd catalyst for acetylene hydrogenation can both be substantially influenced by the effects of Pd lattice strain variation and subsurface carbon species formation on the adsorption properties of the reactants and products. It is found that the adsorption energies of the reactants and products are, in general, linearly scaled with the lattice strain for both pristine and subsurface carbon atom modified Pd(111) surfaces, except for the adsorption of C2H2 over Pd(111) C. The activity for ethylene formation typically corresponds to the region of strong reactants adsorption in the volcano curve; such an effect of lattice strain and the presence of subsurface promoters can improve the activity of the catalyst through the weakening of the adsorption of reactants. The activity and selectivity for Pd(111) C are always higher than those for the pristine Pd(111) surfaces with respect to ethylene formation. Based on the results obtained, Pd based catalysts with shrinking lattice constants are suggested as good candidates for the selective hydrogenation of acetylene. A similar approach can be used to facilitate the future design of novel heterogeneous catalysts. 2018, Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved. 1. Introduction Selective hydrogenation of organic compounds such as unsaturated hydrocarbons, α,β unsaturated aldehydes (ketones), and aromatic nitro compounds over catalysts is an interesting research area in heterogeneous catalysis and has been attracting increasing attention for a long time [1 9]. The current focus in this research area is to achieve high selectivity for relatively unstable intermediates; this requires that the hydrogenation only takes place for the proper bonds, and further hydrogenation should be prevented in some cases. The key aspect to solving this problem is to develop a proper catalyst exhibiting a high selectivity as well as high activity. Selective hydrogenation of acetylene in ethylene feed is very important in the olefin industry, as acetylene acts as an impurity in the olefin feed that poisons the catalysts used subsequently for the polymerization of ethylene. The selective hydrogenation of acetylene will both remove this impurity and increase the amount of ethylene produced. The most effective catalyst for this reaction was found to be Pd based owing to its relatively high selectivity and favorable activity. Previous investigations found that the presence of subsurface species such * Corresponding author. Tel: +86 21 20685147; E mail: yangbo1@shanghaitech.edu.cn This work was supported by the National Natural Science Foundation of China (21603142), the Shanghai Pujiang Program (16PJ1406800) and the Shanghai Young Eastern Scholar Program (QD2016049). DOI: 10.1016/S1872 2067(18)63081 5 http://www.sciencedirect.com/science/journal/18722067 Chin. J. Catal., Vol. 39, No. 9, September 2018

1494 Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 as carbon atoms formed from the dissociation of reactants and reaction intermediates could significantly promote the selective hydrogenation of acetylene over Pd [10,11]. Our previous density functional theory (DFT) calculations revealed that subsurface species could promote the reaction through the electronic modification effect, and the adsorption energy of alkenes over Pd surfaces would be decreased, thereby inhibiting the possibility of over hydrogenation reactions [12 16]. It was widely reported in the literature that the surface strain effect could influence the reactivity of metal surfaces and that the adsorption properties could be varied by changing the metal lattice size [9,17 37]. Nørskov and co workers found using DFT calculations that the surface reactivity of Ru(0001) towards the adsorption of carbon/oxygen and the dissociation of CO increased with lattice expansion. They further correlated the increased reactivity with the observed upshift in the metal d states during lattice expansion, and concluded that surface strain can be used to tailor the catalytic activities of metals [18]. Such a d band centric argument has been widely accepted by the catalysis community. A further study carried out by Mavrikakis group demonstrated that such lattice strains could induce the formation of subsurface species in some cases [20]. The authors employed hydrogen adsorption at the subsurface sites of Ni(111) as an example, and found that subsurface hydrogen would form more readily in stretched regions than in the unstretched regions of the Ni surface. Experimentally engineering the strain of catalyst surfaces has been attracting increased attention recently to tune the performance of these catalysts. However, the catalytic systems are far more complex, with many more variables to consider. One of the examples is that the surface of Pd during acetylene hydrogenation might be carbide like, and therefore, the adsorption process could be influenced by both the surface strain and presence of subsurface species. Meanwhile, the direct relationship between the surface strain and the selectivity and activity for acetylene hydrogenation over Pd under the reaction conditions is still missing, which is the focus of the current work. 2. Computational details All the DFT calculations in this work were performed with the VASP [38 41] code in the slab models. The exchange correlation functional PW91 [42] was used to calculate the electronic structure with generalized gradient approximation (GGA). The projector augmented wave (PAW) method was employed to describe the interactions between the atomic cores and electrons [43,44]. For all the Pd based surfaces, four layer 2 2 slabs with the upmost two layers relaxed during optimization were used to model the adsorption and reaction processes. The optimized equilibrium lattice constant was 3.95 Å [12], which is close to the experimentally measured 3.89 Å. Then, we changed the lattice constant by minus or plus one, two, or four percent, which are designated as m1p, m2p, m4p, p1p, p2p and p4p. A 5 5 1 k point sampling in the surface Brillouin zone was used in all the calculations. The vacuum height was set to be more than 12 Å to avoid any interactions between the slabs. An energy cutoff of 500 ev and a force threshold on each relaxed atom of 0.05 ev/å were used in the current work. The transition states were located using the constrained minimization method [45 47], and only one imaginary frequency was found for the transition states obtained. The adsorption and binding energies are defined as follows: Ead = Etotal (Eg + Eslab), where Etotal is the energy of the system after adsorption, Eg denotes the energy of the gas phase adsorbate, and Eslab is the energy of the slab. For the adsorption energy of the hydrogen atom, we used the energy of 1 /2 H2 as the reference. In this work, we used the simulation package CatMAP developed by Nørskov s group, which has been proven to be a powerful tool in predicting catalytic trends, to perform the microkinetic analyses of the hydrogenation activity [48,49]. In our calculations, the reaction conditions were set as: T = 350 K, P H 2 = 1 bar, PH 2 = 10 bar, and P H 4 = 0.01 bar. We used the fixed_entropy_gas and frozen_adsorbate models to deal with the thermochemistry of the gases and adsorbates by assuming that the entropy of all gases is 0.002 ev/k, except that of H2 (0.00135 ev/k). 3. Results and discussion The energies corresponding to the adsorption of C2H2, H, and C2H4 over pristine Pd(111) surfaces as a function of the variation in the lattice strain of Pd can be found in Fig. 1. It can be observed that the adsorption energies of all three species correlate linearly with the lattice strain, and a larger lattice results in stronger adsorption in all the cases; this is consistent with the d band center theory and the results presented by Mavrikakis et al. [18]. The optimized adsorption configurations of C2H2, C2H3, and C2H4, as well as the transition state structures for the hydrogenation of these C2 species over m1p Pd(111) are shown in Fig. S1 (Supporting Information, SI) as examples. Based on the linear relationships between the adsorption Adsorption energy (ev) 0.0-0.4-0.8-1.2-1.6-2.0-2.4 H 2 H 4 H -4-2 0 2 4 (d d eq )/d eq (%) Fig. 1. Adsorption energies of acetylene, ethylene, and hydrogen over pristine (solid lines) and subsurface carbon modified (dashed lines) Pd(111) as functions of the lattice strain. Here, d and deq represent the varied and equilibrium lattice constants, respectively, and this notation has been used throughout the paper.

Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 1495 Adsorption energy of H x (ev) -1.8-2.0-2.2-2.4-2.6-2.8-3.0-3.2-3.4 H 3 H 4 y = 0.58 x 1.18 y = 0.35 x 2.24-2.1-2.0-1.9-1.8-1.7-1.6-1.5 Adsorption energy of H 2 (ev) (a) Transition state energy (ev) -1.0-1.2-1.4-1.6-1.8-2.0-2.2 H 2 hydrogenation H 3 hydrogenation y = 0.89 x + 0.74 y = 1.04 x + 1.03-3.0-2.8-2.6-2.4-2.2-2.0 Initial state energy (ev) Fig. 2. Scaling relationships for the adsorption energies calculated for pristine Pd(111) surfaces with different strains. (a) Scaling relationship for the adsorption energies of C2H3 and C2H4 as a function of the adsorption energy of C2H2. (b) Transition state scaling relationship for the elementary reactions involved in the hydrogenation of C2H2 to C2H3 and further hydrogenation of C2H3 to C2H4. Here, the transition state energies are plotted against the corresponding initial state energies of the elementary reactions. (b) energies and lattice strain, the scaling relationship between the adsorption energies can be established using the adsorption energy of C2H2 and H as descriptors; this is shown in Fig. 2. It is found that the transition state (TS) energies are also linearly correlated with the corresponding initial state (IS) energies for the elementary steps involved in the hydrogenation of C2H2 to C2H4 over Pd(111) surfaces with different strains. All the TS IS scaling relations are also presented in Fig. 2. The corresponding activation barriers and reaction energies of the elementary reaction steps involved in the hydrogenation of acetylene to ethylene over the pristine Pd(111) surfaces with different lattice constants can be found in Table S1. Combining all the scaling relationships in Fig. 2 and using the microkinetic modeling approach developed by Nørskov s group [48,50,51], a three dimensional activity map can be obtained for the formation of ethylene, as shown in Fig. 3. It can be seen that although smaller lattice constants result in weaker adsorption of the reactants over the pristine Pd(111) surfaces, the activity for ethylene formation is increased. This may be Fig. 3. Activity for ethylene formation as a function of the adsorption energy of H and C2H2 over different surfaces. mip (pip) are the Pd(111) surfaces with i% smaller (larger) lattice, and eq represents the Pd(111) surface with the equilibrium lattice constant. This notation is used throughout the paper. due to the fact that the activity of Pd(111) for the hydrogenation of acetylene to ethylene corresponds to the strong adsorption side of the volcano curve, where, in general, the desorption of products from the surface is the rate determining step [52]. On the other hand, weakening of the adsorption strength of the surface species will increase the turnover rate by facilitating the desorption process, which is similar to the case of Ni catalyzed acetylene hydrogenation that we reported previously [53]. In order to further investigate the possible formation of subsurface species, we calculated the average adsorption energies of carbon and hydrogen atoms at the subsurface sites of Pd(111) surfaces with different lattice constants and subsurface coverages, and the corresponding trends are presented in Fig. 4. There are typically two possible sites for the adsorption of carbon and hydrogen at the subsurface site of Pd(111), namely, the subsurface octahedral and tetrahedral sites; the energies shown in Fig. 4 are those with lower energies at one site than the other. It is clearly demonstrated in this figure that the adsorption energies of both carbon and hydrogen become less negative with decreasing lattice constants, which is similar to the results reported by Greeley et al. [20]. However, the variation in the adsorption energies as a function of subsurface coverage is different between the carbon and hydrogen atoms. For carbon atoms, the adsorption becomes weaker with the increase in the subsurface coverages, while the adsorption strength of hydrogen does not vary much for different coverages. This observation is consistent with those that we reported recently for both Pd(111) and Pd(100) surfaces [12,15]. It should be mentioned that when the lattice of Pd is expanded by 4% compared to the equilibrated Pd lattice (p4p), the adsorption of subsurface hydrogen at the octahedral/tetrahedral sites becomes unstable, and these hydrogen atoms diffuse to the corresponding surface fcc/hcp sites after structural optimization (Fig. S2); therefore the corresponding adsorption energies are not included in Fig. 4. Here, we further consider Pd(111) surfaces with 0.25 monolayer (ML) subsurface carbon coverage, named Pd(111) C, as an example to investigate the effect of the presence of subsur

1496 Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 adsorption, resulting in a stronger deviation from the linear trend. This is believed to result from the slightly deformed Pd(111) C surface with smaller lattice constants due to the strong adsorption of C2H2, as shown in Fig. S3; such deformation will effectively release the stress within the lattice and lower the energy of the system. A similar deformation is not observed for the adsorption of C2H4 over Pd(111) C surfaces, which involve weaker binding than C2H2. Nevertheless, it will be shown later that such unusual adsorption behavior of acetylene observed still supports our argument on the trend for the catalytic activity for ethylene formation. Subsequently, the activity for ethylene formation is calculated based on all the activation energies and reaction energies obtained for all the surfaces studied, and the selectivity is also estimated by comparing the differences between the hydrogenation and desorption barriers of ethylene over the different catalyst surfaces, which is defined as ΔEa = Ea,hydr Ea,des. All of the energies are listed in the tables of the SI. According to our previous work [12,15,54 56], desorption barriers can be estimated based on the absolute value of the corresponding adsorption energies of ethylene, and therefore, higher ΔEa values result in higher ethylene selectivity. All the values of ΔEa and log(tof) are shown in Fig. 5, and the following trends can be readily observed. (1) ΔEa is always positive for both the pristine and subsurface carbon modified Pd(111) surfaces studied, indicating that Fig. 4. Average adsorption energy of carbon (a) and hydrogen (b) at the subsurface site as a function of lattice strain and coverage. These energies are referenced to the energy of (C2H2 H2)/2 and 1 /2H2 for the carbon and hydrogen atom, respectively. face atoms on the catalytic reactivity. The adsorption energies of C2H2, H, and C2H4 are calculated over Pd(111) C and plotted as a function of the lattice strain in Fig. 1. In general, the adsorption of all the species over the Pd(111) surfaces become weaker upon the formation of subsurface carbon atoms. In our previous work, we found that the weakened adsorption over Pd(111) C than over Pd(111) was due to the downshift of the metal d band upon the adsorption of carbon atoms at the subsurface sites [12]. We also find that the adsorption energies of H and C2H4 still linearly scale with the lattice strain, while the adsorption trend of C2H2 shows a peak at the equilibrium constant, indicating that changing the lattice constants of Pd(111) C from those of the equilibrated lattice will always result in stronger acetylene adsorption over the surface. After carefully examining the adsorption energies of C2H2 over all the Pd(111) C surfaces investigated, we found that the trend for larger lattice constants was the same as those observed for other species, while a smaller lattice would give rise to stronger Fig. 5. Selectivity (a) and activity (b) for ethylene formation over pristine and subsurface carbon modified Pd(111) surfaces with varying lattice strain.

Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 1497 the desorption of ethylene from the catalyst surface is preferred for all the surfaces studied; (2) The activity and selectivity of ethylene over Pd(111) C is always higher than those of the corresponding pristine Pd(111). It is found from the tables in the SI that the increased ethylene selectivity upon subsurface carbon modification is mainly due to the weakened adsorption of ethylene over these surfaces; (3) Decreasing the lattice constant will improve the activity and selectivity, in general, for ethylene formation over pristine Pd(111); (4) The m4p and m2p catalysts with subsurface carbon, which are the Pd(111) C surfaces with four and two percent smaller lattice constants than the equilibrated Pd lattice, reveal higher selectivity and activity for ethylene formation from acetylene hydrogenation than the normal Pd(111) and other Pd(111) C surfaces. As revealed above, the adsorption of acetylene over pristine Pd(111) is on the strong adsorption side of the volcano curve for ethylene formation; the enhancement in the activity for ethylene formation can be rationalized by the weaker adsorption of the reactants over the Pd(111) C surfaces. However, upon the formation of the subsurface carbon atoms, no clear trend can be observed in the activities of the Pd(111) C surfaces with smaller lattice constants than the equilibrated value (i.e., m4p, m2p, and m1p Pd(111) C surfaces) with the variation in the lattice constant. We ascribe this to the unusual scaling relation shown in Fig. 1, where a peak is observed in the trend of the acetylene adsorption energies over the subsurface carbon modified Pd(111) surfaces. Interestingly, we found that when the log(tof) over Pd(111) C surfaces were plotted as a function of the adsorption energy of C2H2+H, a linear relationship was obtained (Fig. S4); such a relationship further supports our argument that the adsorption energy of the reactants corresponds to the strong adsorption side of the volcano curve for ethylene formation. It is clearly shown in the current work that the effects of lattice strain and subsurface promoters can improve the performance of catalysts in the case of Pd catalyzed selective hydrogenation of acetylene. Although the activity of m2p Pd(111) C is the highest, its selectivity for ethylene formation is slightly lower than the m4p Pd(111) C catalyst, and we consider the latter as a good candidate for the selective acetylene hydrogenation reaction. Such shrinking of the lattice can be achieved experimentally through a number of catalyst preparation methods, as reported in the literature, including constructing core/shell type M/Pd structures, where M has a slightly smaller lattice constant than Pd, synthesizing Pd nanowires or octahedron nanoparticles with the first shell Pd Pd bond length possibly shorter than that of the Pd foil, or producing crystal twinning during the Pd catalyst synthesis process, wherein a lattice strain may result from the mismatch between the twinned structures [34 36,57,58]. 4. Conclusions We have thoroughly examined the effects of lattice strain and subsurface promoters, which are both important issues in heterogeneous catalysis, on the catalytic selectivity and activity of Pd by considering the selective hydrogenation of acetylene as an example. It is found that the adsorption energies of the reactants and products are, in general, linearly scaled with the lattice strain for both pristine and subsurface carbon atom modified Pd(111) surfaces, except for the adsorption of C2H2 over Pd(111) C. The activity for ethylene formation corresponds to the strong reactant adsorption side of the volcano curve, and such lattice strain effects and presence of subsurface promoters can improve the activity of the catalyst through the weakening of the adsorption of reactants. The activity and selectivity of Pd(111) C are always higher than those of the pristine Pd(111) surfaces for ethylene formation. These Pd based catalysts are suggested as good candidates based on the framework that we have developed for the selective hydrogenation of acetylene, and a similar approach can be used to facilitate the future design of novel heterogeneous catalysts. Acknowledgments We thank the HPC Platform of ShanghaiTech University for computing time. References [1] A. Borodziński, G. C. Bond, Catal. Rev. Sci. Eng., 2006, 48, 91 144. [2] A. Borodziński, G. C. Bond, Catal. Rev. Sci. Eng., 2008, 50, 379 469. [3] P. Gallezot, D. Richard, Catal. Rev. Sci. Eng., 1998, 40, 81 126. [4] B. Yang, D. Wang, X. Q. Gong, P. Hu, Phys. Chem. Chem. Phys., 2011, 13, 21146 21152. [5] H. G. Manyar, R. Morgan, K. Morgan, B. Yang, P. Hu, J. Szlachetko, J. Sa, C. Hardacre, Catal. Sci. Technol., 2013, 3, 1497 1500. [6] H. G. Manyar, B. Yang, H. Daly, H. Moor, S. McMonagle, Y. Tao, G. D. Yadav, A. Goguet, P. Hu, C. Hardacre, ChemCatChem, 2013, 5, 506 512. [7] Y. Gu, Y. H. Zhao, P. P. Wu, B. Yang, N. T. Yang, Y. Zhu, Nanoscale, 2016, 8, 10896 10901. [8] K. L. Wang, B. Yang, Catal. Sci. Technol., 2017, 7, 4024 4033. [9] J. J. Mao, W. X. Chen, W. M. Sun, Z. Chen, J. J. Pei, D. S. He, C. L. Lv, D. S. Wang, Y. D. Li, Angew. Chem. Int. Ed., 2017, 56, 11971 11975. [10] D. Teschner, J. Borsodi, A. Wootsch, Z. Révay, M. Hävecker, A. Knop Gericke, S. D. Jackson, R. Schlӧgl, Science, 2008, 320, 86 89. [11] F. Studt, F. Abild Pedersen, T. Bligaard, R. Z. Sørensen, C. H. Christensen, J. K. Nørskov, Angew. Chem. Int. Ed., 2008, 47, 9299 9302. [12] B. Yang, R. Burch, C. Hardacre, G. Headdock, P. Hu, J. Catal., 2013, 305, 264 276. [13] B. Yang, R. Burch, C. Hardacre, P. Hu, P. Hughes, J. Phys. Chem. C, 2014, 118, 3664 3671. [14] B. Yang, R. Burch, C. Hardacre, P. Hu, P. Hughes, J. Phys. Chem. C, 2014, 118, 1560 1567. [15] B. Yang, R. Burch, C. Hardacre, P. Hu, P. Hughes, Surf. Sci., 2016, 46, 45 49. [16] P. P. Wu, B. Yang, Phys. Chem. Chem. Phys., 2016, 18, 21720 21729. [17] M. Gsell, P. Jakob, D. Menzel, Science, 1998, 280, 717 720.

1498 Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 Graphical Abstract Chin. J. Catal., 2018, 39: 1493 1499 doi: 10.1016/S1872 2067(18)63081 5 Influence of surface strain on activity and selectivity of Pd based catalysts for the hydrogenation of acetylene: A DFT study Ping Wang, Bo Yang * ShanghaiTech University The effect of surface strain on the activity and selectivity of Pd and Pd carbide for the hydrogenation of acetylene is thoroughly examined. [18] M. Mavrikakis, B. Hammer, J. K. Nørskov, Phys. Rev. Lett., 1998, 81, 2819 2822. [19] A. Schlapka, M. Lischka, A. Gross, U. Kasberger, P. Jakob, Phys. Rev. Lett., 2003, 91, 016101/1 016101/4. [20] J. Greeley, W.R. Krekelberg, M. Mavrikakis, Angew. Chem. Int. Ed., 2004, 43, 4296 4300. [21] L. A. Kibler, A. M. El Aziz, R. Hoyer, D. M. Kolb, Angew. Chem. Int. Ed., 2005, 44, 2080 2084. [22] P. Strasser, S. Koh, T. Anniyev, J. Greeley, K. More, C. F. Yu, Z. C. Liu, S. Kaya, D. Nordlund, H. Ogasawara, M. F. Toney, A. Nilsson, Nat. Chem., 2010, 2, 454 460. [23] J. B. Wu, P. P. Li, Y. T. Pan, S. Warren, X. Yin, H. Yang, Chem. Soc. Rev., 2012, 41, 8066 8074. [24] X. M. Wang, Y. Orikasa, Y. Takesue, H. Inoue, M. Nakamura, T. Minato, N. Hoshi, Y. Uchimoto, J. Am. Chem. Soc., 2013, 135, 5938 5941. [25] I. Goikoetxea, J. I. Juaristi, R. D. Muino, M. Alducin, Phys. Rev. Lett., 2014, 113, 066103/1 066103/5. [26] S. Zhang, X. Zhang, G. M. Jiang, H. Y. Zhu, S. J. Guo, D. Su, G. Lu, S. H. Sun, J. Am. Chem. Soc., 2014, 136, 7734 7739. [27] Q. B. Deng, V. Gopal, J. Weissmuller, Angew. Chem. Int. Ed., 2015, 54, 12981 12985. [28] M. S. Du, L. S. Cui, Y. Cao, A. J. Bard, J. Am. Chem. Soc., 2015, 137, 7397 7403. [29] L. Li, F. Abild Pedersen, J. Greeley, J. K. Nørskov, J. Phys. Chem. Lett., 2015, 6, 3797 3801. [30] B. T. Sneed, A. P. Young, C. K. Tsung, Nanoscale, 2015, 7, 12248 12265. [31] M. F. Francis, W. A. Curtin, Nat. Commun., 2015, 6, 6261. [32] S. E. Temmel, E. Fabbri, D. Pergolesi, T. Lippert, T. J. Schmidt, ACS Catal., 2016, 6, 7566 7576. [33] K. Yan, T. A. Maark, A. Khorshidi, V. A. Sethuraman, A. A. Peterson, P. R. Guduru, Angew. Chem. Int. Ed., 2016, 55, 6175 6181. [34] C. H. Kuo, L. K. Lamontagne, C. N. Brodsky, L. Y. Chou, J. Zhuang, B. T. Sneed, M. K. Sheehan, C. K. Tsung, ChemSusChem, 2013, 6, 1993 2000. [35] M. F. Li, Z. P. Zhao, T. Cheng, A. Fortunelli, C. Y. Chen, R. Yu, Q. H. Zhang, L. Gu, B. V. Merinov, Z. Y. Lin, E. B. Zhu, T. Yu, Q. Y. Jia, J. H. Guo, L. Zhang, W. A. Goddard, Y. Huang, X. F. Duan, Science, 2016, 354, 1414 1419. [36] H. W. Huang, H. H. Jia, Z. Liu, P. F. Gao, J. T. Zhao, Z. L. Luo, J. L. Yang, J. Zeng, Angew. Chem. Int. Ed., 2017, 56, 3594 3598. [37] X. H. Bao, Acta Phys. Chim. Sin., 2017, 33, 651 652. [38] G. Kresse, J. Hafner, Phys. Rev. B, 1993, 47, 558 561. [39] G. Kresse, J. Hafner, Phys. Rev. B, 1994, 49, 14251 14269. [40] G. Kresse, J. Furthmuller, Comput. Mater. Sci., 1996, 6, 15 50. [41] G. Kresse, J. Furthmuller, Phys. Rev. B, 1996, 54, 11169 11186. [42] J. P. Perdew, Y. Wang, Phys. Rev. B, 1992, 45, 13244 13249. [43] P. E. Blöchl, Phys. Rev. B, 1994, 50, 17953 17979. [44] G. Kresse, D. Joubert, Phys. Rev. B, 1999, 59, 1758 1775. [45] A. Alavi, P. Hu, T. Deutsch, P. L. Silvestrelli, J. Hutter, Phys. Rev. Lett., 1998, 80, 3650 3653. [46] Z. P. Liu, P. Hu, J. Am. Chem. Soc., 2003, 125, 1958 1967. [47] A. Michaelides, P. Hu, J. Am. Chem. Soc., 2001, 123, 4235 4242. [48] A. J. Medford, C. Shi, M. J. Hoffmann, A. C. Lausche, S. R. Fitzgibbon, T. Bligaard, J. K. Nørskov, Catal. Lett., 2015, 145, 794 807. [49] P. P. Wu, B. Yang, ACS Catal., 2017, 7, 7187 7195. [50] N. Y. Yang, A. J. Medford, X. Y. Liu, F. Studt, T. Bligaard, S. F. Bent, J. K. Nørskov, J. Am. Chem. Soc., 2016, 138, 3705 3714. [51] A. A. Latimer, A. R. Kulkarni, H. Aljama, J. H. Montoya, J. S. Yoo, C. Tsai, F. Abild Pedersen, F. Studt, J. K. Nørskov, Nat. Mater., 2017, 16, 225 229. [52] B. Yang, R. Burch, C. Hardacre, G. Headdock, P. Hu, ACS Catal., 2014, 4, 182 186.

Ping Wang et al. / Chinese Journal of Catalysis 39 (2018) 1493 1499 1499 [53] B. Yang, R. Burch, C. Hardacre, G. Headdock, P. Hu, ACS Catal., 2012, 2, 1027 1032. [54] B. Yang, X. Q. Gong, H. F. Wang, X. M. Cao, J. J. Rooney, P. Hu, J. Am. Chem. Soc., 2013, 135, 15244 15250. [55] B. Yang, R. Burch, C. Hardacre, P. Hu, P. Hughes, Catal. Sci. Technol., 2017, 7, 1508 1514. [56] K. Yang, B. Yang, Phys. Chem. Chem. Phys., 2017, 19, 18010 18017. [57] A. X. Yin, X. Q. Min, W. Zhu, H. S. Wu, Y. W. Zhang, C. H. Yan, Chem. Commun., 2012, 48, 543 545. [58] J. Wu, L. Qi, H. You, A. Gross, J. Li, H. Yang, J. Am. Chem. Soc., 2012, 134, 11880 11883. 密度泛函理论计算研究表面应力对钯催化乙炔选择性加氢活性与选择性的影响 * 王萍, 杨波上海科技大学物质科学与技术学院, 上海 201210 摘要 : 在石油催化裂解过程中, 除了生成乙烯 丙烯及丁烯等烯烃, 还会产生部分炔烃. 目前工业上通常采用炔烃选择性加 氢转化为对应的单烯烃, 以除去其中炔烃. 由于产品烯烃中的炔烃等杂质含量需极低, 这就对用于加氢催化剂的活性和选 择性提出了很高的要求, 即催化剂需要选择性吸附炔烃并加氢, 而不损失其中的烯烃. 经过前期大量的基础研究工作, 目 前工业中炔烃选择性加氢应用最广泛的催化剂是负载型钯基催化剂. 然而, 单独的钯金属选择性并不理想, 因而对其选择 性以及活性进行调控成为了当前关注的研究课题. 本文采用密度泛函理论计算结合微观反应动力学模拟手段, 研究了钯金属表面应力存在条件下的活性与选择性, 以及 形成次表层物种的可能性和形成后的活性与选择性. 研究发现, 改变钯金属的晶格参数与表面应力, 反应物 表面反应中 间体和产物的吸附能都会产生相应的变化, 且吸附能与晶格参数的变化存在线性关系, 晶格参数越大, 吸附越强. 利用表 面反应过渡态能量与初始态能量之间的线性关系, 相应的乙炔加氢生成乙烯的反应速率可以通过微观反应动力学模拟得 到. 结果显示, 不同晶格参数的钯催化剂催化乙炔加氢生成乙烯的反应活性位于相应火山型曲线的强吸附侧, 即减弱乙炔 和氢的吸附强度可提高乙烯的生成速率. 在此基础上, 本文研究了不同表面应力的钯催化剂在次表面吸附不同覆盖度碳 原子和氢原子的情况, 发现晶格参数越大越有利于碳原子和氢原子在次表面的吸附. 同时, 研究发现在次表面碳掺杂的条 件下, 不同表面应力条件下的钯催化剂的活性均有所增强. 此外, 由于乙烯在所有研究的钯催化剂表面脱附比进一步加氢 容易, 因而乙烯都可以选择性生成. 关键词 : 表面应力 ; 钯 ; 乙炔加氢 ; 选择性 ; 活性 ; 次表面 ; 密度泛函理论 ; 微观反应动力学模拟 收稿日期 : 2018-03-01. 接受日期 : 2018-04-07. 出版日期 : 2018-09-05. * 通讯联系人. 电话 : (021)20685147; 电子信箱 : yangbo1@shanghaitech.edu.cn 基金来源 : 国家自然科学基金 (21603142); 上海浦江人才计划 (16PJ1406800); 上海高校青年东方学者 (QD2016049). 本文的电子版全文由 Elsevier 出版社在 ScienceDirect 上出版 (http://www.sciencedirect.com/science/journal/18722067).