Extending the σ-Hole Concept to Metals: An Electrostatic Interpretation

Aug 3, 2017 - Citation data is made available by participants in Crossref's Cited-by Linking service. For a more comprehensive list of citations to th...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV LAVAL

Communication

Extending the #-hole Concept to Metals: An Electrostatic Interpretation of the Nanostructural Effects in Gold and Platinum Catalysis Joakim Halldin Stenlid, and Tore Brinck J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b05987 • Publication Date (Web): 03 Aug 2017 Downloaded from http://pubs.acs.org on August 3, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Extending the s-hole Concept to Metals: An Electrostatic Interpretation of the Nanostructural Effects in Gold and Platinum Catalysis Joakim Halldin Stenlid and Tore Brinck* Applied Physical Chemistry, School of Chemical Science and Engineering, KTH Royal Institute of Technology, Sweden; [email protected] Supporting Information Placeholder ABSTRACT: Crystalline surfaces of gold are chemically inert,

whereas nanoparticles of gold are excellent catalysts for many reactions. The catalytic properties of nanostructured gold have been connected to increased binding affinities of reactant molecules to low-coordinated Au atoms. Here we show that the high reactivity at these sites is a consequence of the formation of s-holes, i.e. maxima in the surface electrostatic potential (Vs,max) due to the overlap of mainly the valence sorbitals when forming the bonding s-orbitals. The s-holes are binding sites for Lewis bases, and binding energies correlate with the magnitudes of the Vs,max. For symmetrical Au clusters, of varying size, the most positive Vs,max are found at corners, edges, and surfaces (facets) and decreasing in that order. This is in agreement with the experimentally and theoretically observed dependence of catalytic activity on local structure. The density of s-holes can explain the increasing catalytic activity with decreasing particle size also for other transition metal catalysts, such as platinum.

Gold has traditionally been considered the noblest of elements, an etiquette that is strongly connected to its chemical inertness. The discovery1 that nanoparticles of gold are efficient catalysts for carbon monoxide oxidation was groundbreaking and has paved the way for numerous applications of gold in catalysis.2,3 It is today established that the remarkable catalytic properties of nanostructured gold is strongly linked to the stronger binding of reactant molecules to lowcoordinated Au-atoms.4–7 The most catalytically active sites at Au nanoparticles are corners, edges and surfaces in decreasing order, and the binding strength generally decreases in that same order.5,6,8 Similarly, the binding energies and the catalytic activity of nanoporous gold catalysts are largely determined by the presence of kinks within surface steps on the inside of the pores.9 The strong correlation between catalytic activity and coordination number is not restricted to gold, but similar relationships are also found for other transition metals, such as platinum.7,10,11 However, the physical origin for the correlation between binding strength and coordination number has remained elusive. In this work, we will show

that the higher activity at low-coordinated Au sites is connected to the presence of s-holes, i.e. regions of depleted electron density indicated by positive surface electrostatic potential, which can be explained by the overlap of the singly occupied valence Au s-orbitals when forming the bonding sorbitals. The molecular electrostatic potential V(r) is a wellestablished tool for analyzing chemical bonding and intermolecular interactions,12,13 and it is rigorously defined by

V(r) =

(r')dr ∑ RZ − r − ∫ ρr'− r A

A

A

where ZA is the charge on nucleus A, located at RA, and ρ(r) is the electron density. In contrast to many other charge distribution descriptors, such as atomic partial charges, V(r) is a physical observable that can be computed from an experimental or theoretical electron density distribution. When analyzing intermolecular interactions it is common to compute and depict V(r) on the 0.001 au isodensity contour, which corresponds approximately to the van der Waals surface of a molecule.12–14 The V(r) of an atom is spherically symmetric, everywhere positive, and decreases monotonically toward zero when moving away from the nucleus. The formation of molecules leads to a redistribution of the electron density toward the more electronegative atoms and generation of areas of negative V(r). As an example, the surface V(r) [VS(r)] of ammonia has a strongly negative lone pair region on nitrogen and positive hydrogens. The minimum (VS,min) and the maximum (VS,max) in VS(r) of these regions reflect the hydrogen bond basicity and acidity, respectively.12 The charge distribution of an atom in a molecule is also polarized and VS(r) can simultaneously have negative and positive regions. Such polarization is particularly prominent for the heavier halogens (≥ Cl). As shown by Brinck et al. in 1992,15 the negative potential (negative VS,min) on the side of the halogen, and the positive end (positive VS,max ) of the atom at the extension of the bond, explain the tendencies of Lewis acids to interact with halogens in a side-on approach, and of Lewis bases to interact by an end-on approach, respectively (see Cl2 in Fig-

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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 49 50 51 52 53 54 55 56 57 58 59 60

ure 1). The latter type of interaction is today referred to as halogen bonding, and has become increasingly important in the design of supramolecular materials, catalyst, and drug molecules.16 Surface electrostatic potentials has become the standard tool to predict and analyze halogen bonding and are today routinely reported in the literature. The positive VS,max at the end of the chlorine atom ( Figure 1) is the result of a density depletion in this region, a !-hole, due to the polarization of the !-orbital towards the bond. In the case of fluorinated compounds, the VS,max at the end of the fluorine, which indicates a !-hole, often has a negative value.17 Clark et al. originally argued that the presence of a !-hole requires that the !-orbital is formed from a singly occupied p-orbital on the halogen.18 Murray et al. extended the concept of !-holes and !-hole bonding to the group IVVI elements.17 They found that a !-hole can have a significant s-contribution, and even suggested that hydrogen bonding should be considered a !-hole interaction. The valence configuration (d10s1) of the noble metals Cu, Ag, and Au, e.g. [Xe5d106s1] for Au, is similar to hydrogen in that there is a singly occupied s-orbital. It is consequently of interest to determine whether !-holes are formed at low-coordinated sites of noble metal clusters, and if the electrostatic potential at these sites can explain the experimentally observed correlation between catalytic activity and local structure.

Figure 1. Surface electrostatic potentials [VS(r)] on the 0.001 a.u. isodensity surfaces of Cl2, Au2 and Au13 (left). Red, followed by yellow, indicates the most positive sites, i.e. the sites most prone to interact with Lewis bases. The plot to the right shows the correlations between computed interaction energies with CO and H2O and the maxima (Vs,max) in VS(r) for Au13.

In order to investigate the existence of !-holes on gold, a number of Au clusters were studied by Kohn-Sham density functional theory (KS-DFT). The geometrical structures of Au2 and an Au13 cluster of CS symmetry were optimized at the PBE/Def2-SV(P) level of theory. In addition, icosahedral and cuboctahedral Aun (n=13-561) and Ptn (n=13,55) clus-

Page 2 of 5

ters were built with the interatomic distances matching the corresponding crystal structures. The VS(r) of the smaller clusters (n ≤ 137 ) at the 0.001 a.u. isodensity contour was evaluated from ρ(r) and V(r) obtained at the PBE/Def2SV(P) level using the Turbomole package,19 and was graphically depicted using the USF Chimera program.20 To analyze the variations in the VS(r) of the larger Au clusters, periodic KS-DFT computations were performed using the VASP program21,22 with the PBE functional and the projected augmented wave (PAW) method. Binding energies, without correction for vibrational effects, for the interaction between the Au13 cluster and CO and H2O were computed at the PBE0-D3/Def2-TVZPP//Def2-TVZP level. All computations using Gaussian basis sets described the inner electrons by an effective core potential (ECP) optimized to account for scalar relativistic effects.23 In addition, for comparison, the VS(r) of Cs Au13 were computed based on electron densities obtained with an all-electron basis sets and the DKH24 and ZORA25 approximations to describe relativistic effects using the ORCA26 program. The VS(r) thus obtained were found to be in good agreement with the VS(r) obtained with the ECP, but showed significant differences from all-electron non-relativistic computations (see Table S2 in S.I.). A more detailed description of the computational procedures is found in the S.I. In forming the Au2 molecule, mainly the valence s-orbitals combine to a !-orbital that is polarized toward the bond, and consequently VS(r) is negative over the center of the molecule and most positive (VS,max) over the end regions. The !hole at the end of each gold atom in Au2 is clearly seen from the VS(r) in Figure 1. The T-shaped geometry of the Au2 dimer is easily predicted by aligning the !-hole of one Au2 unit with the negative potential region (the bond VS,min ) of the other. Furthermore, in a similar manner as for the halogen molecules, Lewis bases preferentially interact with the !hole in the characteristic end-on approach. The charge distributions of larger Au clusters have a similar origin as for Au2, i.e. the valence s-overlap between atoms results in areas of positive potential (!-holes) at the lowcoordinated sites of the atoms and negative potential in the bonding regions. Figure 1 shows the VS(r) of a low energy Au13 cluster of CS symmetry, and there are VS,max at the sites on top of Au atoms. The magnitudes of the VS,max vary, and depend on the coordination number of the corresponding Au atom. More specifically, the VS,max positions indicate sites that are prone to bind Lewis bases, e.g CO or H2O, and the binding energies correlate with VS,max. Figure 1 shows the linear relationships between the interaction energies of CO and H2O and VS,max for the Au13 cluster.

2 ACS Paragon Plus Environment

Page 3 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Table 1. Surface electrostatic potential maxima (VS,max) in kcal/mol of the Aun (n=13,55,147) and Ptn (n=13,55) nanoclusters. Pt values in parentheses.

Cluster

Site

Icosahedral

Cuboctahedral

Au13

corner

9.2

8.1 (10.7)

Au55

corner

11.4

11.0 (12.8)

edge

4.5

4.5 (4.6)

facet (100)

(-)a

5.1 (-2.0)

corner

8.0

9.6

edge

5.8

6.1

facet (111)

2.4

3.1

Au147

facet (100)

(-)

a

1.8

a

There is no facet of this type.

Figure 2. VS(r) on the 0.001 a.u. isodensity surfaces of icosahedral and cuboctahedral clusters of Au13, Au55 and Au137. The same color scheme for VS(r) as in Figure 1 is used. The most positive Vs,max are found at the corners, followed by edges and facets, which corresponds to the ordering of catalytic activity.

It becomes further apparent that low-coordinated Auatoms are associated with areas of positive electrostatic potential when the icosahedral and cuboctahedral Au13, Au55 and Au137 clusters are analyzed in Figure 2. There are VS,max on top of the Au atoms, and their magnitudes generally match the well-known reactivity ordering at the different sites, i.e. decreasing binding affinity and catalytic activity in the order corners > edges > facets.5,6,8 Interestingly, the magnitudes of the VS,max are relatively independent of cluster size; the corner VS,max of the clusters in Figure 2 vary between 8.011.4 kcal/mol, with ico-Au55 having the most positive value and ico-Au137 the least positive value (Table 1). Furthermore, the periodic KS-DFT computations show that the large Au309 and Au561 clusters exhibit similar spatial variations in VS(r) as the smaller clusters (see S.I. for details), and confirm the observation that particularly corners, but also edges, has an increased affinity to interact with Lewis basis that is largely independent of particle size. Even the largest clusters, Au561, which are sufficiently large to have surface (facet) adsorption energies similar to crystalline surfaces of Au, 8,27 have corner VS,max of similar magnitudes as the smaller clusters. These observations are in agreement with earlier experimental and theoretical studies of Au nanoparticles that have shown that the increase in catalytic activity with decreasing size is mainly determined by the higher density of corners (and to some extent also edges) in the smaller nanoclusters.5,6

Our studies indicate that the strong dependence of the catalytic activity of gold on local structure is largely of electrostatic origin. The prototypical catalytic reaction of Au nanoclusters is the oxidation of CO to CO2, and it is not surprising that the electrostatic potential plays a key role in determining the adsorption affinity of the reactant and thereby the catalytic activity; CO is a Lewis base, and its negative lone pair regions are attracted to the Au sites with highest positive potential. Most catalytic reactions of gold follow the same mechanism with the catalytic activity being determined by the strength of the binding of a Lewis base reactant. The similar nanostructural effects on Ag and Cu catalysis as on Au catalysis can also be explained by the appearance of !holes at the low-coordinated atoms due mainly to s-orbital overlap.11,28 Platinum, which has the valence configuration (5d96s1), is another important metal catalyst that is known to exhibit increased catalytic activity with decreasing coordination number,7,11 and where we find that the positional selectivity is connected to areas of positive electrostatic potential. The surface electrostatic potentials of the cuboctahedral Pt13 and Pt55 clusters (Figure 3) show that the most positive VS,max are located at corners and edges and similarly to gold the corners, followed by edges, are the most active sites. However, the Pt corner VS,max are slightly larger in magnitude than the correponding values for the Au clusters (Table 1). Also for the Pt clusters the formation of !-holes at the lowcoordinated atoms is a consequence of the polarization of the electron density toward the bonding regions, but this effect is harder to deduce because of the mix of s, p and d orbital contributions. Clearly surface electrostatic potentials can become a very useful tool for characterizing the catalytically active sites of nanostructured catalysts, including not only nanoparticles but also nanoporous materials and defected surfaces. In particular, we see the potential for characterizing structurally complex particles and surfaces using periodic DFT with plane-wave basis sets, since determining the spatial

3 ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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 49 50 51 52 53 54 55 56 57 58 59 60

variations in VS(r) carries essentially no extra cost when performing such a DFT computation.

Page 4 of 5

calculations were performed at resources provided by the Swedish National Infrastructure for Computing (SNIC).

REFERENCES (1) (2) (3) (4) (5) (6)

Figure 3. VS(r) on the 0.001 a.u. isodensity surfaces of cuboctahedral clusters of Pt13 and Pt55. The same color scheme for VS(r) as in Figure 1 and Figure 2 is used. Similarly to the Au clusters, the most positive Vs,max are found at the corners, followed by edges, which corresponds to the ordering of catalytic activity.

In conclusion, we have shown that the intriguing nanostructural effects on gold catalysis can be explained by the appearance of areas of positive electrostatic potential, !holes, at low-coordinated gold atoms due to the overlap of singly occupied s-orbitals. The !-hole concept for catalytic activation is transferable to other transition metals, such as platinum, and the computation of surface electrostatic potentials has the potential to become an efficient tool in the characterization and design of nanostructured catalysts.

(7) (8) (9)

(10)

(11) (12)

(13) (14) (15) (16)

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website: VS(r) computed with VASP for Au clusters up to the size Au561; comparison of VS(r) for Au13 of CS symmetry computed with ECP and with all-electron basis set (with and withour relativistic corrections); computational details; Cartesian coordinates and electronic energies for all Au clusters.

(17) (18) (19)

(20)

(21) (22) (23)

AUTHOR INFORMATION Corresponding Author

[email protected]

(24) (25)

Funding Sources

The authors declare no competing financial interest.

ACKNOWLEDGMENT Funding from the Swedish Research Council (VR), the Swedish Nuclear Fuel and Waste Management Company (SKB) and the KTH CHE Excellence award is gratefully acknowledged. The

(26) (27)

(28)

Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Chem. Lett. 1987, 2, 405-8.! Bond, G. C.; Thompson, D. T. Catal. Rev. 1999, 41, 319-88.! Stratakis, M.; Garcia, H. Chem. Rev. 2012, 112, 4469-506.! Lemire, C.; Meyer, R.; Shaikhutdinov, S.; Freund, H. J. Angew. Chem. Int. Ed. Engl. 2004, 43, 118-21.! Hvolbæk, B.; Janssens, T. V. W.; Clausen, B. S.; Falsig, H.; Christensen, C. H.; Nørskov, J. K. Nano Today 2007, 2, 14-8.! Janssens, T. V. W.; Clausen, B. S.; Hvolbæk, B.; Falsig, H.; Christensen, C. H.; Bligaard, T.; Nørskov, J. K. Top. Catal. 2007, 44, 1526.! Cao, S.; Tao, F. F.; Tang, Y.; Li, Y.; Yu, J. Chem. Soc Rev. 2016, 45, 4747-65.! Li, H.; Li, L.; Pedersen, A.; Gao, Y.; Khetrapal, N.; Jónsson, H.; Zeng, X. C. Nano Lett. 2015, 15, 682-8.! Fujita, T.; Guan, P.; McKenna, K.; Lang, X.; Hirata, A.; Zhang, L.; Tokunaga, T.; Arai, S.; Yamamoto, Y.; Tanaka, N.; Ishikawa, Y.; Asao, N.; Yamamoto, Y.; Erlebacher, J.; Chen, M. Nat. Mater. 2012, 11, 775-80.! Mostafa, S.; Behafarid, F.; Croy, J. R.; Ono, L. K.; Li, L.; Yang, J. C.; Frenkel, A. I.; Cuenya, B. R. J. Am Chem. Soc 2010, 132, 157149.! Roldan Cuenya, B. Acc. Chem. Res. 2013, 46, 1682-91.! T. Brinck. In Theoretical Organic Chemistry; Párkányi, C. Ed.; Theoretical and Computational Chemistry; Elsevier: 1998; Vol. 5, pp 51 - 93.! Murray, J. S.; Politzer, P. WIREs Comput. Mol. Sci. 2011, 1, 153-63.! Bulat, F.; Toro-labbe, A.; Brinck, T.; Murray, J.; Politzer, P. J. Mol. Mod. 2010, 16, 1679-91.! Brinck, T.; Murray, J.; Politzer, P. Int. J. Quant. Chem. 1992, 57-64.! Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. Chem. Rev. 2016, 116, 2478-601.! Murray, J. S.; Lane, P.; Politzer, P. J. Mol Model 2009, 15, 723-9.! Clark, T.; Hennemann, M.; Murray, J. S.; Politzer, P. J. Mol. Mod. 2007, 13, 291-6.! TURBOMOLE V6.4 2012, a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989-2007, TURBOMOLE GmbH, since 2007; available from http://www.turbomole.com., ! Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. J. Comp. Chem. 2004, 25, 1605-12.! Kresse, G.; Furthmüller, J. J. Phys. Rev. B 1996, 54, 11169-86.! Kresse, G.; Furthmüller, J. J. Comp. Mater. Sci. 1996, 6, 15-50.! Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Theor. Chim. Acta 1990, 77, 123-41.! Reiher, M. Theor. Chem. Acc. 2006, 116, 241-52.! Lenthe, E. V.; Baerends, E.-J.; Snijders, J. G. J. Chem. Phys. 1993, 99, 4597-610.! Neese, F. WIREs Comput. Mol. Sci. 2012, 2, 73-8.! Kleis, J.; Greeley, J.; Romero, N. A.; Morozov, V. A.; Falsig, H.; Larsen, A. H.; Lu, J.; Mortensen, J. J.; Dułak, M.; Thygesen, K. S.; Nørskov, J. K.; Jacobsen, K. W. Catal. Lett. 2011, 141, 1067-71.! Gawande, M. B.; Goswami, A.; Felpin, F. X.; Asefa, T.; Huang, X.; Silva, R.; Zou, X.; Zboril, R.; Varma, R. S. Chem. Rev. 2016, 116, 3722-811.!

4 ACS Paragon Plus Environment

Page 5 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 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 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Insert Table of Contents artwork here

5

ACS Paragon Plus Environment