You are on page 1of 35

Subscriber access provided by UNIV OF DURHAM

Article
DFT-Computed Trends in the Properties of Bimetallic
Precious-Metal Nanoparticles With Core@shell Segregation
Andrew J. Logsdail, Lauro Oliver Paz-Borbon, and Christopher A. Downing
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b10614 • Publication Date (Web): 08 Feb 2018
Downloaded from http://pubs.acs.org on February 21, 2018

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 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 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.

The Journal of Physical Chemistry C 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 34 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 DFT-Computed Trends in the Properties of Bimetallic
9
10
11 Precious-Metal Nanoparticles with Core@Shell
12
13
14
15
Segregation
16
17
18
19
Andrew J. Logsdail,∗,† Lauro Oliver Paz-Bórbon,‡ and Christopher A. Downing¶
20
21
Cardiff Catalysis Institute, Cardiff University School of Chemistry, Main Building, Park Place,
22
23
Cardiff, CF10 3AT, UK, Instituto de Física, Universidad Nacional Autónoma de México,
24
25
Apartado Postal 20-364, México DF 01000, México, and University College London, Kathleen
26
27
Lonsdale Materials Chemistry, Department of Chemistry, 20 Gordon St., London, WC1H 0AJ, UK
28
29
30 E-mail: LogsdailA@cardiff.ac.uk
31
32
33
34
35 Abstract
36
37
Bimetallic precious-metal nanoparticles (NPs) are promising as components of heteroge-
38
39 neous catalytic systems, however the synergistic effects that occur when combining multiple
40
41 elements in these discrete systems are still being understood. Here, we present a system-
42
43 atic computational study that investigates the effect of geometry and compositional arrange-
44
45 ments on the properties of bimetallic core@shell precious-metal NPs. Combinations of the
46
47 elements Ag, Au, Pd and Pt are considered and, for surface coverages of just one monolayer,
48 ∗
To whom correspondence should be addressed
49 † Cardiff Catalysis Institute, Cardiff University School of Chemistry, Main Building, Park Place, Cardiff, CF10
50
3AT, UK
51 ‡ Instituto de Física, Universidad Nacional Autónoma de México, Apartado Postal 20-364, México DF 01000,
52
México
53 ¶ University College London, Kathleen Lonsdale Materials Chemistry, Department of Chemistry, 20 Gordon St.,
54
London, WC1H 0AJ, UK
55
56
57
58
59 1
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 34

1
2
3
we show that the electronic properties of the NP, such as the Fermi energy and d-band cen-
4
5 tre, are strongly correlated with the element occupying sites at the NP surface, in qualitative
6
7 but not quantitative agreement with previous two-dimensional slab calculations. In contrast,
8
9 strain effects from structural choices and elemental size mismatch play a minor role in altering
10
11 energetic levels. Charge transfer between the core and shell regions of the NPs is a balance
12
13 between the increased electronic degrees of freedom that implicitly exists at an NP surface
14
15 and the electronegativity of the constituent elements, with Au and Ag respectively good and
16
bad at retaining electron density. We use our results to rationalise recent experimental obser-
17
18 vations, and foresee that the outcomes will assist the rational atom-by-atom design of future
19
20 nanoparticle catalysts.
21
22
23
24
25 1 Introduction
26
27
28 Metallic nanoparticles (NPs) offer great promise for applications in chemical sciences due to their
29
30 novel properties when compared to bulk or atomic counterparts. Since the first reports of the re-
31
32 activity of gold NPs by Haruta 1 and Hutchings, 2 applications of precious-metal NPs have become
33
34 widespread through catalysis, 3–8 electronics 9–11 and optics. 12–14 However, the scarcity and high-
35
36 cost of Au and Pt demands a shift towards improvements in the design and controlled synthesis
37
38 of precious-metal NPs. 15 The use of bimetallic NPs, where two metals are combined, has thus
39
40 proven an excellent method for reducing costs whilst improving the catalytic chemistry feasible
41
42 with NPs. 15–18 Size, structure and chemical arrangements of the bimetallic NPs can be experimen-
43
44 tally controlled through careful synthesis protocols, 19,20 which allows fine tuning and maximisa-
45
46 tion of the the synergistic benefits available when combining multiple elements. 17 Bimetallic NPs
47
48 can be categorised broadly as having alloy, janus-segregated or core@shell chemical arrangements,
49
50 with the latter particularly appealing for catalysis as it is relatively facile to synthesise, with e.g.
51
52 sequential precipitation, and a reactive “first-choice” metal can be supported on a core of an alter-
53
54 native cheaper material, lowering the manufacturing cost of the catalyst and potentially improving
55
56 turnover frequencies due to compositional synergy. 21–25
57
58
59 2
60 ACS Paragon Plus Environment
Page 3 of 34 The Journal of Physical Chemistry

1
2
3
The discrete nature of NPs lends themselves towards computational modelling techniques
4
5
where forcefield or quantum mechanical methods can elucidate fundamental properties. 17 Many
6
7
comprehensive computational investigations have been considered for bimetallic NPs, or “nanoal-
8
9 loys”, 16,17 with emphasis on exploring either: (i) the structure, chemical arrangement and proper-
10
11 ties of energetic minima; 17,26–31 or (ii) the surface reactivity of idealised structures. 32–36 Whilst the
12
13 extrapolation of properties for single-element systems to large NP sizes are well established, 37,38
14
15 the same process is complicated for bimetallic NPs by the additional compositional parameters.
16
17 Similarly, the relationship of catalytic observables, such as reactivity, to the position of the d-
18
19 band centre 39–44 or structural features, such as edges and vertices, 45,46 have been investigated
20
21 thoroughly for monometallic NPs, but bimetallic NPs are structurally and compositionally more
22
23 complicated; 47 too simplify the analysis, d-band positioning, and chemical interactions, are often
24
25 considered using periodic slab models with dopants and/or monolayer coverages of a secondary
26
27 element probing the property changes. 23,25,48 For unrelaxed slab models, Ruban et al. documented
28
29 shifts in the d-band centre for a range of transition metals; 49 the effect of epitaxial strain on d-band
30
31 chemistry has been generalised since for bimetallic surfaces, 50,51 with the width of the surface
32
33 d-band noted as being responsive to structural factors, and the d-band centre downshifting in re-
34
35 sponse to band broadening. Such effects have been correlated to CO chemisorption on to Pt- and
36
37 Pd-coated Au surfaces by Pedersen et al. 52 and Roudgar et al., 53 respectively.
38
39 Despite the range of work covering bimetallic bulk surfaces, the properties of bimetallic NPs
40
41 remain less thoroughly studied. Tang and Henkelman showed that the binding energy of atomic
42
43 oxygen correlates to d-band levels of the surface layers in M@Pd NPs, where M ranged over Mo,
44
45 Co, Cu, Ag, Au and Pd. In particular, noble metal cores (Ag, Au) resulted in an upshift in the
46
47 surface d-band states, which led to stronger surface binding of the O species. 54 The authors note
48
49 that strain affects the position of the d-band centre in the NPs, but structural relaxation minimises
50
51 this effect when compared to single-crystal (i.e. slab) surfaces, at which point the redistribution
52
53 of electron density has a greater influence. However, Gorzkowski and Lewera recently reported a
54
55 disagreement between experiment and d-band theory for formic acid decomposition over Pd-core
56
57
58
59 3
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 34

1
2
3
Pt-shell (Pd@Pt) nanoparticles, 55 which was attributed to changes in the density of states close to
4
5
the Fermi level rather than the overall d-band width. Similarly, formic acid decomposition over
6
7
M@Pd NPs (M = Rh, Ru, Ag, Au, Pd, Pt) was correlated with influencing factors beyond the
8
9 standard d-band model, focusing instead on the work function of the core element, with Ag being
10
11 most effective, and surface charge accumulation used as rationale for the experimental results. 56
12
13 In contrast, Holmberg et al. produced recently an investigation of the Fermi energy (ε f ) position
14
15 for bimetallic AuAg nanoparticles, finding this was simply an average of ε f for monometallic
16
17 systems weighted by the percentage of the surface covered by each elemental species, e.g. ε f
18
19 for Au@Ag would be the same as for a purely Ag NP. 57 This behaviour was attributed to charge
20
21 equilibration across the bimetallic interface, and noted as being analogous to Vegard’s law for
22
23 the lattice parameter of binary bulk systems, 58 which opens up questions as to: (i) whether other
24
25 attributes of NPs are similarly dependent on the surface coverage by a particular element; and (ii)
26
27 whether these properties are affected by NP geometry.
28
29 To address some of the inconsistencies reported, and gaps in the literature with respect to the
30
31 d-band model for discrete bimetallic NPs, we have performed a density functional theory (DFT)
32
33 investigation of the energetic and electronic properties for core@shell segregated bimetallic NPs,
34
35 composed of the elements Ag, Au, Pd and Pt. We focus on NPs with 147-atoms, which allows us
36
37 to exclude spurious “quantum effects”. We study a range of structural motifs from close-packed
38
39 through to bulk fragments, which allows us to compare the structure-dependent properties for ele-
40
41 mentally equivalent systems. As for the periodic slab models highlighted from previous investiga-
42
43 tions, the chemical arrangement of the constituent elements affects electronic observables and we
44
45 discuss our observations with respect to the recent experimental and theoretical results outlined.
46
47
48
49
50
51
52
53
54
55
56
57
58
59 4
60 ACS Paragon Plus Environment
Page 5 of 34 The Journal of Physical Chemistry

1
2
3
4
2 Methodology
5
6
7 2.1 Structural Definitions
8
9
10 Three high-symmetry, 12-vertex geometries have been considered throughout this work, namely
11
12 the icosahedron (Ih), Ino-decahedron (I-Dh) and cuboctahedron (CO). Each of these motifs has the
13
14 same “magic number” nuclearities given by the third-order polynomial: 59
15
16
17
1
N(k) = [10k3 + 15k2 + 11k + 3] (1)
18 3
19
20
21 where k is the number of shells in the structure; the first few values of N ("magic numbers") in
22
23 this sequence are 13, 55, 147, 309 and 561. Increasing k leads to larger N but a lower proportion
24
25 of surface atoms, as explained via the liquid drop model. 60 The Ih, I-Dh and CO structures have
26
27 distinct atomic arrangements, which leads to differing geometrical attributes (Figure 1). The Ih
28
29 consists of 20 (111) faces, making it the most spherical of the three structures. This arrangement
30
31 minimises the surface area to volume ratio whilst maximising nearest neighbours contact, at the
32
33 expense of higher internal strain. 61 The Ih is energetically most stable for small clusters, where
34
35 the energy penalty of the large surfaces is too great unless exceptional electronic configurations
36
37 exist. 62 The minimisation of internal strain is more important for large clusters: The CO, with 8
38
39 (111) faces and 6 (100 faces), is a fragment of an FCC crystal and therefore internal strain is at
40
41 its lowest, meaning that CO is energetically preferable as N → ∞. 62 The I-Dh motif occupies the
42
43 transitional space between these two geometries, with its structure not containing the tight packing
44
45 of the Ih structure nor the FCC crystalline structure of the CO: it has 10 (111) faces and 5 (100)
46
47 faces.
48
49 In the following work, calculations have been performed on the Ih, I-Dh and CO structures
50
51 with N = 147. We have focused on model systems so as to characterise the effect of core@shell
52
53 segregation on NP properties. Thus, we have only considered local, rather than global optimisa-
54
55 tion for structures. This can be rationalised by considering how the formation and stabilisation
56
57
58
59 5
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 34

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
Figure 1: Views of the structural motifs investigated in this work: (a) Icosahedron (Ih), with the
39
40 (111) faces defined in red; (b) Ino-Decahedron (I-Dh), with the (111) faces defined in green; (c)
41 Cuboctahedorn, with the (100) and (111) faces defined in blue. The total density of states for Ag147
42 NPs with these different structures are presented in (d), relative to a vacuum level of 0 V, with the
43 colours of the plots matched to the structures defined in (a) - (c) i.e. Ih = red, I-Dh = green and CO
44
= blue. The Fermi energy, E f , relative to vacuum, is marked by a dashed vertical black line; this
45
46 was calculated as -3.99, -3.99 and -4.00 eV for the Ih, I-Dh and CO, respectively.
47
48
49
50
51
52
53
54
55
56
57
58
59 6
60 ACS Paragon Plus Environment
Page 7 of 34 The Journal of Physical Chemistry

1
2
3
of structures with varied geometry and chemical segregation can be performed in experimental
4
5
synthesis, due to e.g. kinetic trapping 63–66 and ligand-induced surface stabilisation. 17,67,68 The
6
7
range of structures considered allows us to identify variations with respect to topology as well as
8
9 composition.
10
11
12
13 2.2 Computational Methods
14
15
16 Energy calculations were performed using density functional theory (DFT), as implemented in
17
18 the real-space finite-difference DFT code GPAW. 69,70 GPAW uses the all-electron density, where
19
20 the core electrons are described using the projector augmented wave (PAW) in the frozen core
21
22 approximation, 71 with the setup based on a scalar-relativistic spin-paired all-electron calculation.
23
24 The PAW setup has recently been shown to be the most suitable ionic-core method for calculat-
25
26 ing the properties of large NPs as an alternative to more computationally-exhaustive all-electron
27
28 calculations. 72
29
30 Only the valence electrons are treated explicitly in our calculations, via soft pseudo valence
31
32 wavefunctions; for Ag and Au, this constitutes the outermost 11 electrons (4d 10 5s1 and 5d 10 6s1
33
34 configurations, respectively), whilst for Pd and Pt the outermost 10 electrons are included (4d 10
35
36 and 5d 10 , respectively). The Kohn-Sham wavefunctions are expanded directly on a real space
37
38 grid, with spacing of 0.18 Å, and the Hamiltonian is diagonalised after each geometry step using
39
40 the RMM-DIIS eigensolver. 73 The generalised gradient approximation (GGA) of Perdew, Burke
41
42 and Ernzerhof (PBE) 74 is used for the exchange-correlation (XC) energy functional. Fermi-Dirac
43
44 smearing of 0.1 eV was used to aid self-consistent field (SCF) convergence. No boundary condi-
45
46 tions were imposed and therefore all orbital energies are relative to the vacuum level (0 V). The
47
48 structure of bulk metals are calculated using a primitive FCC unit cell and a 14 × 14 × 14 k-point
49
50 mesh, with all other parameters as outlined above. Lattice parameters were minimised with respect
51
52 to energy using an equation of state.
53
54 Structural optimisation was performed using a quasi-Newtonian (BFGS) method 75 until the
55
56 maximum force on all atoms is below 0.01 eV/Å, and the total energy change between steps is less
57
58
59 7
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 34

1
2
3
than 10−4 eV. The Kohn-Sham orbitals are re-optimised after each ionic step, via self-consistent
4
5
evaluation, until the change in density is less than 10−5 electrons per valence electron and the in-
6
7
tegrated square of the residuals of the Kohn-Sham equations is less than 10−8 per state. Charge
8
9 redistribution is calculated using Bader charge analysis as implemented by Tang et al. 76 All den-
10
11 sity of states (DOS) are computed by directly projecting the wavefunctions on to atomic orbitals,
12
13 centred at the atomic coordinates of the optimised geometries.
14
15
16
17 2.3 Structural Analysis
18
19
20 The strain at each atomic site was calculated by considering the deformation of bonds in bimetallic
21
22 NPs compared to the equivalent monometallic systems, with results averaged as a function of like
23
24 atom species. The strain is expressed as a percentage of the total deformation:
25
26
27 l −L
28 e= (2)
L
29
30
31 where L and l are the initial and final lengths of the bond of interest. With this notation, contraction
32
33 of the bond leads to compressive strain whilst extension gives tensile strain.
34
35
36
37 2.4 Energetic Analysis
38
39
40 The binding energy per atom (Eb ) for a bimetallic cluster of composition XN−m Ym is defined as:
41
42
43 Etot − (N − m)EX − mEY
44 Eb = − (3)
45
N
46
47 where Etot is the total energy of the system, N is the total number of atoms, m is the number of
48
49 atoms of element Y, and EX and EY are the energies of the gas-phase atoms X and Y, respectively;
50
51 m = 0 for a monometallic system. Our definition is such that a more positive number indicates
52
53 stronger binding and, at the bulk limit (N → ∞), our Eb therefore relates to the bulk cohesive
54
55 energy (Ecoh ) as Ecoh = Eb .
56
57
58
59 8
60 ACS Paragon Plus Environment
Page 9 of 34 The Journal of Physical Chemistry

1
2
3
Etot , as reported in GPAW, is relative to the energy of the reference gas-phase atoms (EX and
4
5
EY ) calculated in a spin-paired and spherically symmetric state when generating the PAW setup.
6
7
However, the energy of the considered gas-phase atoms are dependent on their spin-state and the
8
9 calculation setup, and this is particularly important for Ag and Au, where S = 21 in the ground state.
10
11 Thus, the ground-state energies of the gas-phase atoms are calculated to be -0.19, -0.24, -0.04 and
12
13 -0.32 eV relative to the spin-paired and spherically symmetric reference state for Ag, Au, Pd and
14
15 Pt, respectively.
16
17
18
19 2.5 Electronic Analysis
20
21
22 The energy of the highest occupied molecular orbital (εHOMO ) and lowest unoccupied molecular
23
24 orbital (εLUMO ) are provided relative to the vacuum (0 V), having been taken directly from the
25
26 eigenvalues at the end of the self-consistent calculation. The Fermi energy (ε f ) is also reported
27
28 where appropriate: this includes smearing effects and indicates the level at which the orbitals are
29
30 equiprobable of containing an electron. The difference between εHOMO and εLUMO is known as the
31
32 energy gap, εg . The centre of mass for the d-electrons, ε̄d , is analogous to the d-band centre for
33
34 bulk 41 and given by: 33
35
36
37 Z εHOMO
ρd (ε)
38 ε̄d = dε − εHOMO (4)
39 −∞ ρd
40
41 where ε is the eigenvalue of an electronic state, ρd (ε) is the density of d-electrons for this
42
43 particular eigenvalue, and ρd is the total number of d-electrons. "Core" and "shell" atomic subsets
44
45 of the total systems were also identified by applying Eq. (4) explicitly to the projected orbitals
46
47 centred on these atoms, giving values for ε̄d,core and ε̄d,shell , respectively.
48
49
50
51
52
53
54
55
56
57
58
59 9
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 34

1
2
3
4
3 Results
5
6
7 3.1 Monometallic NPs
8
9
10 Structural, energetic and electronic results have been thoroughly investigated previously for 147-
11
12 atom Ag, Au, Pd and Pt NPs 61,63 and are therefore only discussed briefly here, with full analysis
13
14 provided in the Supporting Information (SI). In general, Eb does not vary significantly with struc-
15
16
ture, being greatest for the Ih and a difference of less than 0.03 eV/atom for the CO and I-Dh.
17
18
Eb is ordered Pt (5.03 eV) > Pd (3.08 eV) > Au (2.59 eV) > Ag (2.17 eV) for the Ih structures,
19
20
which is consistent with the bulk ordering (5.83, 3.75, 3.01 and 2.56 eV, respectively). Eb matches
21
22
experiment for Pd and Pt, reported as 5.84 and 3.89 eV, respectively, 77 but not for Au (3.81 eV)
23
24
and Ag (2.95 eV) due to bond-lengthening effects at the GGA level of theory. 78,79
25
26
The ordering of εHOMO inversely correlates with Eb , as Ag (-4.05) > Au (-4.83 eV) > Pd (-4.89
27
28
eV) > Pt (-5.36 eV) for the Ih motif. The variation in εHOMO for the other geometries was smaller
29
30
than 0.13 eV. εHOMO The bulk workfunction (Φ) can be related to our observables via Koopmans’
31
32
theorem 80,81 as Φ = −εHOMO . Φ is the energy required to remove an electron from a system and
33
34
can depend on surface geometry, ranging between 4.52 and 4.74 eV for bulk Ag (110) and (111)
35
36
surfaces; between 5.31 and 5.47 eV for Au (111) and (100) surfaces; between 5.22 and 5.6 eV
37
for Pd polycrystalline and (111) surfaces; and between 5.64 and 5.93 eV for the Pt polycrystalline
38
39
faces, respectively. 82 Our results for NPs are less than the bulk values by ≈ 0.5 eV, which we
40
41
believe is an effect of quantum confinement i.e. the discretisation of the electronic energy levels.
42
43
44
45 3.2 Bimetallic NPs
46
47
48 3.2.1 Structural and Energetic Properties
49
50
51 Eb for the bimetallic NPs (Table 1) can be related to the monometallic results through Vegard’s law,
52
53 which dictates that properties of a mixed system will be linearly dependent on the properties of the
54
55 constiuent elements. 58 For example, combining Pt with Ag reduces Eb (Ih) from 5.03 eV/atom
56
57
58
59 10
60 ACS Paragon Plus Environment
Page 11 of 34 The Journal of Physical Chemistry

1
2
3
(monometallic) to 3.85 and 3.32 eV/atom for Ag@Pt and Pt@Ag, respectively, with monometallic
4
5
Ag being 2.17 eV/atom. More interesting is the stability hierarchy of the structures as we vary the
6
7
chemical arrangements, with data for all structures given in the SI. In general, Ih is most stable
8
9 for all systems with an Ag core (Ag@M), and for an Ag shell with a Pd or Pt core (Pd@Ag and
10
11 Pt@Ag); and the CO is most stable for Au@Ag, Au@Pd and Pt@Pd. The I-Dh is not preferable
12
13 for any of the bimetallic nanopartices - it is joint lowest for only Pt@Pt, Ag@Pd, Au@Pd.
14
15
16 Table 1: Highest binding energy (Eb ) for monometallic and bimetallic core@shell NPs. The ele-
17 ment in the core (shell) of the NP is presented on the left (top). The values are taken for the highest
18 Eb , i.e. strongest bonded system, which was the Ih motif except for values given in bold, namely
19 those of Au@Ag, Au@Pd and Pt@Pd
20
21 Ag Au Pd Pt
22
Ag 2.17 2.47 2.48 3.85
23
24 Au 2.33 2.59 2.80 3.93
25 Pd 2.50 2.90 3.08 4.30
26 Pt 3.32 3.59 3.86 5.03
27
28
29
Instability for the Ih is sometimes related to excessive internal strain from shortened internal
30
31
bonds (Figure 2), where the strain can be reduced by positioning a smaller element in the core for
32
33
bimetallic systems; 83–85 considering our elements of interest, Pd and Pt, of Group 10, have ex-
34
35
perimental (DFT) bond lengths of 2.76 and 2.81 Å, respectively, with errors of ±0.01 Å, whereas
36
37 Ag and Au, of Group 11, have bond lengths (with similar errors) of 2.88 and 2.96 Å, respec-
38
39 tively. 82 Thus, the Ih core@shell could be rationalised as potentially particularly stable for (Group
40
41 10)@(Group 11) configurations, with the inverse being the lease stable. However, this hypothesis
42
43 of instability when the larger element is in the NP core is only valid for Au@Pd (Figure 2), where
44
45 the CO structure relieves internal strain; unaccounted charge transfer effects influence stability for
46
47 other compositions. We note that the stronger binding element is at the core of the structure for
48
49 other compositions favouring CO motifs (Au@Ag and Pt@Pd) and thus we suggest that the FCC
50
51 fragment facilitates ideal bond lengths for the stronger element (Figure 2), resulting in stronger
52
53 overall binding for the CO motif.
54
55 Calculation of the strain for the core (ecore ) and shell (eshell ) components of the system, as
56
57
58
59 11
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 34

1
2
3
presented in Figure 3, confirms there is an inverse correlation between core and shell strain for
4
5
most structures; however, whilst for Ih and I-Dh motifs there is typical not net gain in strain, net
6
7
positive strain is observed for CO, with the most prominent outliers being CO structures with Ag
8
9 shells, namely Au@Ag, Pd@Ag and Pt@Ag. We tentatively associate this observation with the
10
11 interatomic distances in the pure Ag nanoparticle being heavily contracted at specific sites on the
12
13 CO surface, such as vertices, as seen in Figure 2; however, this cannot be the complete explanation
14
15 as the effect is observed irrespective of the size of the element in the nanoparticle core, for both
16
17 smaller (Pd, Pt) and larger (Au) species, and so electronic effects may also play a role.
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 2: Cutplane views through the middle of monometallic and bimetallic core@shell struc-
42 tures with the internal bond lengths highlighted. Top, left to right: Pd147 (Ih); Pd147 (CO);
43 Au55 @Pd92 (Ih); Au55 @Pd92 (CO). Bottom, left to right: Ag147 (Ih); Ag147 (CO); Au55 @Ag92
44
(Ih); Ag55 @Pd92 (CO). Ag, Au and Pd atoms are illustrated in silver, gold and blue, respectively.
45
46 Bond lengths, given in Å, are highlighted in red for specific internal bonds to show the difference
47 in interaction distance in the core and shell regions.
48
49
50 More generally, three clustering of data points is observed in Figure 3, which can be attributed
51
52 to the relative Eb of the monometallic systems. For systems where the element on the nanoparticle
53
54 shell has a higher Eb than the core element, compression of the core and expansion of the shell is
55
56 observed; for the inverse, with the strongest binding element in the core, expansion is observed in
57
58
59 12
60 ACS Paragon Plus Environment
Page 13 of 34 The Journal of Physical Chemistry

1
2
3
the core in order for the system to maximise the energy available from these stronger bonds. Such
4
5
behaviour is most prominent for Pt, and to a lesser degree Pd; for Ag@Pt, the Pt shell compresses
6
7
the Ag core (ecore = -3.98 % in the case of the Ih motif); inversely, for Pt@Ag with the same motif,
8
9 the core expands (ecore = 2.59 %). For systems where the relative Eb of the composite elements are
10
11 similar, there is significantly less strain present in the core and shell regions of the nanoparticle,
12
13 hence the clustering of points where ecore ≈ eshell ≈ 0.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Figure 3: Relative strain on the core and shell atoms compared to the equivalent monometallic
35 structures. Purple squares, green circles and blue triangles represent CO, I-Dh and Ih structural
36 motifs, with the lines of best fit displayed as dashed lines as a guide to the eye. Significant outliers,
37 namely CO structures with net positive strain, are highlighted in red, whilst CO structures that are
38 the lowest energy motif for any composition are also highlighted in black.
39
40
41
42
43 3.2.2 Electronic Properties
44
45
46 The electronic properties of the bimetallic NPs have common trends: ε f , εHOMO and εLUMO are
47
48 defined by the element in the NP shell, agreeing with previous observations that ε f is dependent on
49
50 shell composition. 57 Presented in Table 2 are the values for εHOMO when considering the mono-
51
52 and bimetallic Ih structures, with clear correlations visible in each column, indicating that the
53
54 shell element dictates this particular electronic property. It is noted that for the weaker binding Ag,
55
56 εHOMO decreases when stronger binding core elements are considered; and for the stronger binding
57
58
59 13
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 34

1
2
3
Pt, an inverse correlation is observed i.e. εHOMO decreases as weaker binding core elements are
4
5
considered.
6
7
To illustrate further our point that nanoparticle properties depend heavily on the shell element,
8
9 Figure 4 compares the density of states (DOS) when considering an NP composed either as Ag@M
10
11 and M@Ag. For the Ag@M composition, ε f is clearly varied and the DOS outline for the Ag, Au,
12
13 Pd and Pt cores show little similarity, even if alligned to match by ε f . In contrast, the consistent
14
15 level in ε f for all M@Ag NPs is clearly visible in the lower graph of Figure 4, allowing us to
16
17 assert that the electronic properties are dominated by the shell element. Overall, εHOMO varies by
18
19 only 0.07 eV for M@Ag; variation of 0.26 and 0.27 eV is noted for M@Au and M@Pt, which
20
21 is attributed to the implicit strain of geometry and chemical variation. Furthermore, the lowest
22
23 and highest εHOMO for M@Pt are when M = Ag and Pt, respectively, which correlates not only
24
25 with binding energy, as noted above, but also with the Φ of the core metal (SI, Figure S4).The gap
26
27 between occupied and unoccupied states, εg , is ≈ 0 eV for all NPs, with a greatest value of 0.06
28
29 eV for Ih Pd@Ag and 0.03 eV for Ih Pd@Au.
30
31
32
Table 2: Energy of the highest occupied molecular orbital (εHOMO ) for monometallic and bimetal-
33 lic core@shell NP structures. The element in the core (shell) of the NP is presented on the left
34 (top). The values shown are for the Ih motif; full results are available in the SI for the I-Dh and
35 CO motifs, with the range of values greatest for Ag@M, and at most 0.22 eV for Ag@Pt.
36
37 Ag Au Pd Pt
38
Ag -4.05 -4.67 -4.99 -5.63
39
40 Au -4.03 -4.83 -4.95 -5.59
41 Pd -4.12 -4.96 -4.89 -5.46
42 Pt -4.12 -4.83 -4.78 -5.36
43
44
45 d ) and shell (ε̄ d ) atoms
The centre of mass for the d-band states was calculated for core (ε̄core
46 shell
47 and is presented in Table 3. The energetic position in each case (core or shell) depends heavily
48
49 d
on the elemental species, ranging for ε̄shell over -3.87 to -4.10 (0.33), -3.21 to -3.55 (0.34), -1.46
50
51 to -1.79 (0.33) and -1.84 to -2.28 (0.44) eV for Ag, Au, Pd and Pt, respectively, with the overall
52
53 variation given in brackets. These values are noted as being alignments with respect to εHOMO ,
54
55 d
in common with previous slab models, and when aligned absolutely the range of ε̄shell is actually
56
57
58
59 14
60 ACS Paragon Plus Environment
Page 15 of 34 The Journal of Physical Chemistry

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 Figure 4: Projected density of states (PDOS) for Ag containing core@shell NPs. Top: PDOS for
50 Ag atoms in the 55-atom core of a 147-atom Ih structure, with a 92-atom shell composed of Ag
51 (red), Au (green), Pd (blue) and Pt (yellow). Bottom: PDOS for Ag atoms in the 92-atom shell of
52 a 147-atom Ih structure, with a 55-atom core composed of Ag (red), Au (green), Pd (blue) and Pt
53
(yellow). For clarity, the PDOS of each composition are offset by a constant in the y-axis and, in
54
55 both graphs, the corresponding Fermi energy, ε f , is given with vertical dashed line of the matching
56 colour.
57
58
59 15
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 34

1
2
3
smaller: 0.26, 0.11, 0.21 and 0.28 eV for Ag, Au, Pd and Pt, respectively.
4
5
6 Table 3: Centre of mass for the occupied d-states (ε̄ d ) for monometallic and bimetallic core@shell
7 NP structures.The element in the core (shell) of the NP is presented on the left (top) of each Table.
8 The values are for the Ih geometry; full results for all structural motifs are available in the SI. For
9 d , the change relative to the monometallic system is given in brackets to aid comparison with
ε̄shell
10
11
literature.
12
13
ε̄ d Ag Au Pd Pt
14 Ag: -4.22 -3.92 -2.46 -2.80
15 Au: -4.06 -3.62 -2.28 -2.60
16 Pd: -3.19 -2.80 -1.97 -2.40
17
Pt: -3.46 -3.18 -2.28 -2.67
18
19
d Ag Au Pd Pt
20 ε̄core
21 Ag: -4.42 -4.54 -4.14 -4.30
22 Au: -4.01 -4.00 -3.70 -3.87
23
24
Pd: -2.06 -2.12 -2.43 -2.65
25 Pt: -2.67 -2.84 -3.11 -3.31
26
27 d
ε̄shell Ag Au Pd Pt
28 Ag: -4.10 -3.55 -1.46 -1.90
29
30 (-0.15) (0.24) (0.38)
31 Au: -4.09 -3.40 -1.43 -1.84
32 (0.01) (0.27) (0.44)
33 Pd: -3.87 -3.21 -1.70 -2.25
34 (0.23) ( 0.19) (0.03)
35
36 Pt: -3.93 -3.38 -1.79 -2.28
37 (0.17) (0.02) (-0.09)
38
39
40 Considering our observations alongside the slab calculations of Ruban et al., our results for
41
42 monometallic systems clearly match the previous work, with differences of 0.20, 0.16, 0.13 and -
43
44 d ; however, whilst qualitative agreement is found with experiment, quantitative dis-
0.03 eV for ε̄shell
45
46 agreements are noted in comparison to the XPS measurements of transition metal foils by Hofmann
47
48 et al., who observed similar differences between computation and experiment themselves, and as-
49
50 sociated errors with underlying physical faults in DFT; 44 given that our results for monometallic
51
52 systems are similar to those previously reported, we direct the reader to this previous work for a
53
54 more thorough discussion of the monometallic systems.
55
56 In the case of bimetallic NPs, we again compare our results to the fixed slabs of Ruban et
57
58
59 16
60 ACS Paragon Plus Environment
Page 17 of 34 The Journal of Physical Chemistry

1
2
3
al., where a monolayer "shell" of second species was included over a slab "core". 49 The relative
4
5 d
changes in ε̄shell (Table 3) are similar in magnitude for both ours and previous work, but there
6
7
is an absence of consistent direction of shifts for the d-band centre in the slab and NP models:
8
9 different signs are observed for Au@Ag (-0.21 vs. 0.01 eV in our work), Pd@Au (0.12 vs -0.15
10
11 eV), Au@Pd (-0.21 vs 0.27 eV), Pt@Pd (0.37 vs -0.09 eV) and Au@Pt (-0.08 vs 0.44 eV). We note
12
13 a common theme exists for Au cores, where slab models give a positive (negative) d-band shift for
14
15 Ag (Pd,Pt) shells, and the inverse occurs in our NP calculations. Furthermore, Gorzkowski and
16
17 Lewera comment that a small decrease in ε̄ d is expected for NPs relative to slabs and bulk, due
18
19 to size effects resulting in a decrease of εHOMO ; 55 however, we do not observe such an effect.
20
21 Considering also the results of Tang and Henkelman for 79-atom FCC M@Pd NPs, 54 positive
22
23 shifts in the d-band centre are observed for M = Ag and Au, relative to the Pd@Pd system, with
24
25 values of approximately 0.20 and 0.25 eV; this is very close to our results of 0.24 and 0.27 eV for
26
27 Ih, leading us to believe that morphology effects may be minimal.
28
29 To understand where the electrons have localised in our NPs, we analyse the atomic charge in
30
31 the NPs, which can be correlated with variation in e.g. reactivity for vertices, edges and faces. With
32
33 respect to charge equlibration alone, the most important information is whether charge accumulates
34
35 in the shell or the core of the NPs: the increased degrees of freedom at a surface will attract
36
37 electrons away from the core, though this could be countered by a strongly electronegative core
38
39 element. For example, charge transfer from Ag to Au is common in AuAg NPs as the Au (2.4) is
40
41 more electronegative than Ag (1.9), 86,87 whereas the electronegativity of Pd and Pt, which is 2.2
42
43 in both cases, differs less compared to the Group 11 elements, 77 and thus charge transfer may be
44
45 considered less of an energetic driving force.
46
47 Our results for charge transfer (Table 4) have clear trends: for Ag@M, the core has signifi-
48
49 cant charge depletion (q̄core > 0 e/atom), for all bimetallic NPs, with charge accumulation on the
50
51 surfaces (q̄shell < 0 e/atom) and particularly the vertex sites (SI). Furthermore, for an Ag shell
52
53 (M@Ag), charge transfer is also to the second element, indicating that, for the Ag containing NPs,
54
55 the low electronegativity is more influential than surface effects. For cores of Au, Pd and Pt, charge
56
57
58
59 17
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 34

1
2
3 Table 4: Mean change in the electron density for the core (q̄core ) and shell (q̄shell ) atoms in
4
monometallic and bimetallic core@shell NP structures. The element in the core (shell) of the
5
6 NP is presented on the left (top). The values shown are for the Ih motif, with full results for all
7 structures available in the SI.
8
9 q̄core Ag Au Pd Pt
10 Ag: -0.03 -0.12 -0.10 -0.16
11 Au: 0.06 -0.05 -0.01 -0.08
12
13
Pd: 0.04 -0.05 -0.04 -0.10
14 Pt: 0.09 -0.03 -0.00 -0.07
15
16
17 q̄shell Ag Au Pd Pt
18
19 Ag: 0.06 0.11 0.07 0.14
20 Au: 0.02 0.09 0.05 0.12
21 Pd: 0.02 0.09 0.07 0.13
22 Pt: 0.03 0.09 0.07 0.14
23
24
25 transfer is towards the shell in all cases (q̄shell > 0 e/atom) with the Pt shell for M@Pt proving most
26
27 prolific at accumulating charge. The outcome illustrates that the electronic degrees of freedom at
28
29 the surface play an important role in charge localisation for the bimetallic NPs when the elements
30
31 have similar electronegativities.
32
33 Referring back to the work of Tang and Henkelman, charge depletion was observed at the
34
35 surface for Ag@Pd and Au@Pd, which may be related to strain effects in the FCC motif, 54 though
36
37 no such effects were noted for our FCC fragment (CO) nor the I-Dh motif (SI). In contrast, we do
38
39 observe similar results to Holmberg et al.: 0.05 and 0.07 e/atom are transferred to the surface for
40
41 Au@Ag and Ag@Au, respectively. However, they also note the charge depletion from the core
42
43 is inhomogeneous in the latter case, being predominantly from the subsurface layer, which may
44
45 have some relation to the results we observe for an Ag core. In our simulations, the charge mostly
46
47 accumulates at the NPs vertices (Figure 5), which we attribute to the lower coordination of these
48
49 sites compared to edge and surface atoms; this trend is consistent for the Ih, I-Dh and CO structures
50
51 (SI). We also note the variation in charge of the surface atoms for Figure 5A and Figure 5B, which
52
53 is attributed to the breaking of symmetry in the structure geometry during the preceding structural
54
55 optimisation.
56
57
58
59 18
60 ACS Paragon Plus Environment
Page 19 of 34 The Journal of Physical Chemistry

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 Figure 5: Schematic representation of charge localisation for: (A) Au; (B) Ag; (C) Ag@Au; and
46 (D) Au@Ag 147-atom Ih NPs. Atoms highlighted in blue indicate those that have accummulated
47
48
electrons due to charge transfer, and atoms highlighted in red mark atoms that have lost electron
49 density; a scale is provided on the right hand side to aid the reader. Similar results were observed
50 for I-Dh and CO structures (See SI).
51
52
53
54
55
56
57
58
59 19
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 34

1
2
3
4
4 Discussion
5
6
Previous work has highlighted clearly the correlation between the d-band centre and adsorption
7
8
strength very clear for monometallic systems; 39–44 however, the inhomogeneous distribution of
9
10
the constituent elements in bimetallic NPs, such as the case of core@shell segregation, means
11
12
this standard model needs adaptation. The more complex parameter space, which now includes
13
14
interfacial strain and charge transfer effects, is most obvious when considering the inconsistencies
15
16
between our results and the previous slab calculations of Ruban et al.. However, our results do
17
18 show that the general behaviour of d-band centres on NPs is as for slab surfaces, which implies
19
20 that adsorption strength must be in-part correlated to the energetic position of the d-electrons on the
21
22 surface atoms, rather than the same measurements for the entire NP. This is most clearly seen by
23
24 d
noting the limited range of ε̄shell d
in comparison to ε̄core for the precious-metal elements discussed
25
26 in this manuscript.
27
28 With this in mind, we now consider the report of Gorzkowski and Lewera, 55 where ε̄ d was
29
30 measured as being particularly low for Pt@Pd NPs (-2.61 eV) and the previous computational
31
32 d . From our results
results were labelled erroneous because of an absence of a downshift in ε̄shell
33
34 in Table 3, we see that there is a downshift if we consider the d-band centre of the entire NP, ε̄ d ,
35
36 as might be measured via XPS; however, the d-band states on the surface are higher than for the
37
38 Pt core. Thus, our simulations differentiate between core and shell atoms, whereas the reported
39
40 experiment does not appear to do so. In addtion, we also note that the monometallic results from
41
42 Gorzkowski and Lewera, of -2.02 and -2.80 eV for Pd and Pt, are in very good agreement with our
43
44 ε̄ d results of -1.97 and -2.67 eV.
45
46 Similarly, let us consider the results of Tedsree et al. 56 for M@Pd, where higher reactivity is
47
48 attributed to a smaller work function for the core metal and the d-band model does not given a
49
50 similar linear correlation. We do not disagree with the observations for the work function, which
51
52 d , and this
are matched in our results, but we also observe that Ag@Pd has a relatively low ε̄core
53
54 increases for Au@Pd and Pt@Pd. Thus, the effects we observe can be related to interfacial charge
55
56 transfer effects, which are then dependent on the inherent material workfunctions. Additionally,
57
58
59 20
60 ACS Paragon Plus Environment
Page 21 of 34 The Journal of Physical Chemistry

1
2
3 d
though, we note that ε̄shell is relatively static in all our calculations, which implies other driving
4
5
forces may be worth considering for the surface chemistry observed in experiment.
6
7
As a final observation, we look in more detail at the role of strain on the change of NP proper-
8
9 d , or ε
ties. As seen in Figure 6, little to no absolute change is observed for ε̄shell HOMO (SI, Figure
10
11 S5), due to the strain on the surface elements when changing the core element. In contrast, strain
12
13 d
has more varied outcomes for the core elements: ε̄core shifts by a considerable amount relative to
14
15 the monometallic NP. Thus, further work is needed to systematically investigate these effects in
16
17 the regime that is beyond that of quantum confinement (N < 100) whilst not yet at the bulk, as it
18
19 seems the behaviour remains unpredictable at these sizes. Open questions remain with respect to
20
21 the role of strain and whether such effects could be harnessed for monolayer metallic deposition
22
23 on inorganic substrates, thus resulting in a shift of the d-band centre. We will investigate this in
24
25 our future work.
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 6: Strain for the core and shell elements relative to ∆ε d for different structures, i.e. the
43 change relative to monometallic NPs. ∆εshell d is given in absolute terms, thus implicitly incorporat-
44
45 ing changes in the εHOMO . Purple squares, green circles and blue triangles represent CO, I-Dh and
46 Ih structural motifs, with the lines of best fit displayed as dashed lines as a guide to the eye. The
47 poor quality of the fit is illutrated by the R2 of the fitted lines, which, for the core atoms, are 0.25,
48 0.56 and 0.74 for the CO, I-Dh and Ih, respectively; for the shell atoms, the R2 of the fitted lines
49 are 0.13, 0.02 and 0.26 for the same structures
50
51
52
53
54
55
56
57
58
59 21
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 34

1
2
3
4
5 Conclusions
5
6
We have investigated the structural, energetic and electronic properties of bimetallic NPs com-
7
8
posed of the precious-metal elements Ag, Au, Pd and Pt, focusing specifically on the properties
9
10
of core@shell nanostructures. The energetic properties are strongly correlated with the cohesive
11
12
energy of the composite elements, with most compositions favouring the Ih structure for the 147-
13
14
atom motifs investigated. Three specific chemical arrangements are more stable in a CO structure
15
16
at this nuclearity, namely Au@Ag, Au@Pd and Pt@Pd. For Au@Pd, the compressed nature of the
17
18 internal bonds in an Ih destabilises this motif relative to the CO, whereas perhaps such a structural
19
20 preference is not observed for Au@Pt due to the greater strength of the encapsulating Pt-Pt bonds.
21
22 For Au@Ag and Pt@Pd, the considered elements have similar radii but the stronger binding ele-
23
24 ment is also at the core of the structure, and thus we conclude that an ideal bond length for this
25
26 element, as facilitated by the FCC motif, ensures that the overall binding of the CO NP is stronger
27
28 than the more compressed Ih.
29
30 Analysis of the electronic structure shows that the energy of the highest occupied molecular
31
32 orbitals (εHOMO ) is strongly correlated to the element in the shell of the NP, with minor variance
33
34 depending on the core. εHOMO is lowest for systems with a Pt shell, and increases when Pt is
35
36 replaced by Pd, Au and Ag, respectively. For Ag@M, an energy gap in the electronic levels is
37
38 identified, which is also seen for Au@Au and Pd@Au; no energy gap is seen for systems with Pd
39
40 or Pt shells. Further review of the DOS offers insight into the positioning of the d-band centre of
41
42 mass, which plays an important part in the generic concepts of bonding during catalytic reactions.
43
44 The centre of mass for shell-centred d-band electrons is fairly consistent irrespective of the element
45
46 in the core of the NP, with ranges relative to εHOMO of 0.23, 0.34, 0.36 and 0.44 eV for shells
47
48 of Ag, Au, Pd and Pt, respectively; and only 0.33, 0.07, 0.16 and 0.11 eV in absolute terms.
49
50 This observation is extremely important as these are the physically accessible states for interacting
51
52 adsorbate molecules. Finally, we analysed the charge segregation across the NPs. In general
53
54 electrons accumulate at the NP surfaces irrespective of the core@shell composition, except in the
55
56 case of Ag: for M@Ag the core tends to gain electron density, which is explained due to the
57
58
59 22
60 ACS Paragon Plus Environment
Page 23 of 34 The Journal of Physical Chemistry

1
2
3
low electronegativity of Ag (1.9) with respect to other constituent elements (Au: 2.4; Pd: 2.2; Pt:
4
5
2.2). In all cases charge accumulates more prominently on the vertices of the NPs, especially for Pt
6
7
surfaces, and less so on the edges and surfaces, showing a distinct correlation between coordination
8
9 number and charge accumulation.
10
11
12
13 Supporting Information Available
14
15
16 The Supporting Information contains thorough analysis of the structure, energetic and electronic
17
18 properties for nanoparticles considered in this study, Tables of data containing all reported ob-
19
20 servables for the bimetallic nanoparticles discussed in this article, and Graphs of: (i) the pro-
21
22 jected density of states for NPs with a Pt shell and varying core elements; and (ii) the relation-
23
24 ship between strain and the change in the position of εHOMO . This material is available free of
25
26 charge via the Internet at http://pubs.acs.org. Information on the data underpinning the results pre-
27
28 sented here, including how to access them, can be found in the Cardiff University data catalogue
29
30 at http://doi.org/10.17035/d.2018.0045513133
31
32
33
34 Acknowledgement
35
36
37
The authors acknowledge N. Dimitratos, D. O. Scanlon, A. A. Sokol and C. R. A. Catlow for
38
39
several stimulating and useful conversations. Via our membership of the UK’s HEC Materials
40
41
Chemistry Consortium, which is funded by EPSRC (EP/L000202), this work used the ARCHER
42
43
UK National Supercomputing Service. This work also used the computational facilities of the
44
45
Advanced Research Computing @ Cardiff (ARCCA) Division, Cardiff University.
46
47
48
49
50
51
52
53
54
55
56
57
58
59 23
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 34

1
2
3
4
References
5
6 1. Haruta, M.; Tsubota, S.; Kobayashi, T.; Kageyama, H.; Genet, M. J.; Delmon, B. Low-
7
8 Temperature Oxidation of CO over Gold Supported on TiO2 , α-Fe2 O3 , and Co3 O4 . J. Catal.
9
10 1993, 144, 175–192.
11
12
13 2. Hutchings, G. Vapor phase hydrochlorination of acetylene: Correlation of catalytic activity of
14
15 supported metal chloride catalysts. J. Catal. 1985, 96, 292 – 295.
16
17
18 3. Haruta, M.; Daté, M. Advances in the Catalysis of Au Nanoparticles. Appl. Catal. A: Gen.
19
20 2001, 222, 427 – 437.
21
22
23
4. Turner, M.; Golovko, V. B.; Vaughan, O. P.; Abdulkin, P.; Berenguer-Murcia, A.;
24
25
Tikhov, M. S.; Johnson, B. F.; Lambert, R. M. Selective oxidation with dioxygen by gold
26
27
nanoparticle catalysts derived from 55-atom clusters. Nature 2008, 454, 981–983.
28
29 5. Corma, A.; Garcia, H. Supported gold nanoparticles as catalysts for organic reactions. Chem-
30
31 ical Society Reviews 2008, 37, 2096–2126.
32
33
34 6. Hvolbæk, B.; Janssens, T. V.; Clausen, B. S.; Falsig, H.; Christensen, C. H.; Nørskov, J. K.
35
36 Catalytic activity of Au nanoparticles. Nano Today 2007, 2, 14–18.
37
38
39 7. Narayanan, R.; El-Sayed, M. A. Shape-dependent catalytic activity of platinum nanoparticles
40
41 in colloidal solution. Nano letters 2004, 4, 1343–1348.
42
43
44 8. Wilson, O. M.; Knecht, M. R.; Garcia-Martinez, J. C.; Crooks, R. M. Effect of Pd nanoparticle
45
46 size on the catalytic hydrogenation of allyl alcohol. Journal of the American Chemical Society
47
48 2006, 128, 4510–4511.
49
50
51 9. Huang, D.; Liao, F.; Molesa, S.; Redinger, D.; Subramanian, V. Plastic-compatible low resis-
52
53 tance printable gold nanoparticle conductors for flexible electronics. Journal of the electro-
54
55
chemical society 2003, 150, G412–G417.
56
57
58
59 24
60 ACS Paragon Plus Environment
Page 25 of 34 The Journal of Physical Chemistry

1
2
3
10. Li, Y.; Wu, Y.; Ong, B. S. Facile synthesis of silver nanoparticles useful for fabrication of
4
5
high-conductivity elements for printed electronics. Journal of the American Chemical Society
6
7
2005, 127, 3266–3267.
8
9
10 11. Takei, K.; Yu, Z.; Zheng, M.; Ota, H.; Takahashi, T.; Javey, A. Highly sensitive electronic
11
12 whiskers based on patterned carbon nanotube and silver nanoparticle composite films. Pro-
13
14 ceedings of the National Academy of Sciences 2014, 111, 1703–1707.
15
16
17 12. Elghanian, R.; Storhoff, J. J.; Mucic, R. C.; Letsinger, R. L.; Mirkin, C. A. Selective colori-
18
19 metric detection of polynucleotides based on the distance-dependent optical properties of gold
20
21 nanoparticles. Science 1997, 277, 1078–1081.
22
23
24 13. Haynes, C. L.; Van Duyne, R. P. Nanosphere lithography: a versatile nanofabrication tool for
25
26 studies of size-dependent nanoparticle optics. The Journal of Physical Chemistry B 2001, 105,
27
28 5599–5611.
29
30
31 14. Kelly, K. L.; Coronado, E.; Zhao, L. L.; Schatz, G. C. The Optical Properties of Metal
32
33 Nanoparticles: The Influence of Size, Shape, and Dielectric Environment. J. Phys. Chem.
34
35 B 2003, 107, 668–677.
36
37
15. Toshima, N.; Yonezawa, T. Bimetallic nanoparticles—novel materials for chemical and phys-
38
39
ical applications. New Journal of Chemistry 1998, 22, 1179–1201.
40
41
42 16. Nanoalloys: From Theory to Applications. Faraday Discussions No.138, 2008.
43
44
45 17. Ferrando, R.; Jellinek, J.; Johnston, R. L. Nanoalloys: From Theory to Applications of Alloy
46
47 Clusters and Nanoparticles. Chem. Rev. 2008, 108, 845–910.
48
49
50 18. Johnston, R. L.; Wilcoxon, J. P. Metal Nanoparticles and Nanoalloys; Frontiers of
51
52 Nanoscience (Vol. 3); Elsevier: Oxford, 2012.
53
54
55 19. Sau, T. K.; Murphy, C. J. Room temperature, high-yield synthesis of multiple shapes of gold
56
57
58
59 25
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 26 of 34

1
2
3
nanoparticles in aqueous solution. Journal of the American Chemical Society 2004, 126, 8648–
4
5
8649.
6
7
8 20. Wegner, K.; Piseri, P.; Tafreshi, H. V.; Milani, P. Cluster beam deposition: a tool for nanoscale
9
10 science and technology. Journal of Physics D: Applied Physics 2006, 39, R439.
11
12
13 21. Su, R.; Tiruvalam, R.; Logsdail, A. J.; He, Q.; Downing, C. A.; Jensen, M. T.; Dimitratos, N.;
14
15 Kesavan, L.; Wells, P. P.; Bechstein, R. et al. Designer titania-supported Au–Pd nanoparticles
16
17 for efficient photocatalytic hydrogen production. ACS nano 2014, 8, 3490–3497.
18
19
20 22. Sankar, M.; Dimitratos, N.; Miedziak, P. J.; Wells, P. P.; Kiely, C. J.; Hutchings, G. J. De-
21
22 signing bimetallic catalysts for a green and sustainable future. Chem. Soc. Rev. 2012, 41,
23
24 8099–8139.
25
26
27
23. Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.; Liu, Z.; Kaya, S.; Nord-
28
29
lund, D.; Ogasawara, H. et al. Lattice-strain control of the activity in dealloyed core–shell fuel
30
31
cell catalysts. Nature Chem. 2010, 2, 454–460.
32
33 24. Villa, A.; Dimitratos, N.; Chan-Thaw, C. E.; Hammond, C.; Prati, L.; Hutchings, G. J. Glycerol
34
35 Oxidation Using Gold-Containing Catalysts. Accounts of chemical research 2015, 48, 1403–
36
37 1412.
38
39
40 25. Alayoglu, S.; Nilekar, A. U.; Mavrikakis, M.; Eichhorn, B. Ru-Pt core-shell nanoparticles for
41
42 preferential oxidation of carbon monoxide in hydrogen. Nature Mater. 2008, 7, 333–338.
43
44
45 26. Johnston, R. L. Evolving better nanoparticles: Genetic algorithms for optimising cluster ge-
46
47 ometries. Dalton Trans. 2003, 4193.
48
49
50 27. Ferrando, R.; Fortunelli, A.; Johnston, R. L. Searching for the optimum structures of alloy
51
52 nanoclusters. Chem. Phys. Phys. Chem. 2008, 10, 640–649.
53
54
28. Paz-Borbón, L. O.; Johnston, R. L.; Barcaro, G.; Fortunelli, A. Structural motifs, mixing, and
55
56
segregation effects in 38-atom binary clusters. J. Chem. Phys. 2008, 128, 134517.
57
58
59 26
60 ACS Paragon Plus Environment
Page 27 of 34 The Journal of Physical Chemistry

1
2
3
29. Rapallo, A.; Rossi, G.; Ferrando, R.; Fortunelli, A.; Curley, B. C.; Lloyd, L. D.; Tar-
4
5
buck, G. M.; Johnston, R. L. Global optimization of bimetallic cluster structures. I. Size-
6
7
mismatched Ag–Cu, Ag–Ni, and Au–Cu systems. The Journal of chemical physics 2005, 122,
8
9 194308.
10
11
12 30. Rossi, G.; Ferrando, R.; Rapallo, A.; Fortunelli, A.; Curley, B. C.; Lloyd, L. D.; Johnston, R. L.
13
14 Global optimization of bimetallic cluster structures. II. Size-matched Ag-Pd, Ag-Au, and Pd-
15
16 Pt systems. The Journal of chemical physics 2005, 122, 194309.
17
18
19 31. Kozlov, S. M.; Kovacs, G.; Ferrando, R.; Neyman, K. M. How to determine accurate chemical
20
21 ordering in several nanometer large bimetallic crystallites from electronic structure calcula-
22
23 tions. Chem. Sci. 2015, 6, 3868–3880.
24
25
26 32. Scamehorn, C. A.; Harrison, N. M.; McCarthy, M. I. Water chemistry on surface defect sites:
27
28 Chemidissociation versus physisorption on MgO(001). J. Chem. Phys. 1994, 101, 1547–1554.
29
30
31 33. Jennings, P. C.; Pollet, B. G.; Johnston, R. L. Electronic Properties of Pt–Ti Nanoalloys and
32
33 the Effect on Reactivity for Use in PEMFCs. J. Phys. Chem. C 2012, 116, 15241–15250.
34
35
34. Jennings, P.; Aleksandrov, H.; Neyman, K.; Johnston, R. O2 Dissociation on M@Pt Core–
36
37
Shell Particles for 3d, 4d, and 5d Transition Metals. J. Phys. Chem. C 2015, 119, 11031–
38
39
11041.
40
41
42 35. Song, C.; Ge, Q.; ; Wang*, L. DFT Studies of Pt/Au Bimetallic Clusters and Their Interactions
43
44 with the CO Molecule. The Journal of Physical Chemistry B 2005, 109, 22341–22350, PMID:
45
46 16853910.
47
48
49 36. Nilekar, A. U.; Alayoglu, S.; Eichhorn, B.; Mavrikakis, M. Preferential CO Oxidation in Hy-
50
51 drogen: Reactivity of CoreâĹŠShell Nanoparticles. Journal of the American Chemical Society
52
53 2010, 132, 7418–7428, PMID: 20459102.
54
55
56
57
58
59 27
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 34

1
2
3
37. Roldán, A.; Viñes, F.; Illas, F.; Ricart, J.; Neyman, K. M. Density functional studies of coinage
4
5
metal nanoparticles: scalability of their properties to bulk. Theor. Chem. Acc. 2008, 120, 565–
6
7
573.
8
9
10 38. Yudanov, I. V.; Genest, A.; Rösch, N. DFT Studies of Palladium Model Catalysts: Structure
11
12 and Size Effects. J. Clust. Sci 2011, 22, 433–448.
13
14
15 39. Hammer, B.; Nørskov, J. Electronic factors determining the reactivity of metal surfaces. Sur-
16
17 face Science 1995, 343, 211–220.
18
19
20 40. Nørskov, J. K.; Bligaard, T.; Rossmeisl, J.; Christensen, C. H. Towards the computational
21
22 design of solid catalysts. Nature Chem. 2009, 1, 37–46.
23
24
25 41. Nørskov, J. K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Density functional theory in surface
26
27 chemistry and catalysis. Proceedings of the National Academy of Sciences 2011, 108, 937–
28
29 943.
30
31
42. Crampton, A. S.; Rötzer, M. D.; Schweinberger, F. F.; Yoon, B.; Landman, U.; Heiz, U. Ethy-
32
33
lene hydrogenation on supported Ni, Pd and Pt nanoparticles: Catalyst activity, deactivation
34
35
and the d-band model. Journal of Catalysis 2016, 333, 51 – 58.
36
37
38 43. Lima, F. H. B.; Zhang, J.; Shao, M. H.; Sasaki, K.; Vukmirovic, M. B.; Ticianelli, E. A.;
39
40 Adzic, R. R. Catalytic ActivityâĹŠd-Band Center Correlation for the O2 Reduction Reaction
41
42 on Platinum in Alkaline Solutions. The Journal of Physical Chemistry C 2007, 111, 404–410.
43
44
45 44. Hofmann, T.; Yu, T. H.; Folse, M.; Weinhardt, L.; Bär, M.; Zhang, Y.; Merinov, B. V.; My-
46
47 ers, D. J.; Goddard, W. A.; Heske, C. Using Photoelectron Spectroscopy and Quantum Me-
48
49 chanics to Determine d-Band Energies of Metals for Catalytic Applications. J. Phys. Chem. C
50
51 2012, 116, 24016–24026.
52
53
54 45. Calle-Vallejo, F.; Loffreda, D.; Koper, M. T.; Sautet, P. Introducing structural sensitivity into
55
56
57
58
59 28
60 ACS Paragon Plus Environment
Page 29 of 34 The Journal of Physical Chemistry

1
2
3
adsorption–energy scaling relations by means of coordination numbers. Nature Chem. 2015,
4
5
7, 403–410.
6
7
8 46. Calle-Vallejo, F.; Tymoczko, J.; Colic, V.; Vu, Q. H.; Pohl, M. D.; Morgenstern, K.; Lof-
9
10 freda, D.; Sautet, P.; Schuhmann, W.; Bandarenka, A. S. Finding optimal surface sites on
11
12 heterogeneous catalysts by counting nearest neighbors. Science 2015, 350, 185–189.
13
14
15 47. Paz-Borbón, L.; Baletto, F. A DFT Study on the O2 Adsorption Properties of Supported PtNi
16
17 Clusters. Inorganics 2017, 5, 43.
18
19
20 48. Mott, D.; Luo, J.; Njoki, P. N.; Lin, Y.; Wang, L.; Zhong, C.-J. Synergistic activity of gold-
21
22 platinum alloy nanoparticle catalysts. Catal. Today 2007, 122, 378 – 385.
23
24
25 49. Ruban, A.; Hammer, B.; Stoltze, P.; Skriver, H. L.; Nørskov, J. K. Surface electronic structure
26
27 and reactivity of transition and noble metals. J. Mol. Catal. A Chem. 1997, 115, 421–429.
28
29
50. Greeley, J.; Nørskov, J. K. A general scheme for the estimation of oxygen binding energies on
30
31
binary transition metal surface alloys. Surface Science 2005, 592, 104 – 111.
32
33
34 51. Kibler, L. A.; El-Aziz, A. M.; Hoyer, R.; Kolb, D. M. Tuning reaction rates by lateral strain in
35
36 a palladium monolayer. Angewandte Chemie International Edition 2005, 44, 2080–2084.
37
38
39 52. Pedersen, M.; Helveg, S.; Ruban, A.; Stensgaard, I.; Lægsgaard, E.; Nørskov, J.; Besen-
40
41 bacher, F. How a gold substrate can increase the reactivity of a Pt overlayer. Surf. Sci. 1999,
42
43 426, 395–409.
44
45
46 53. Roudgar, A.; Groß, A. Local reactivity of metal overlayers: Density functional theory calcu-
47
48 lations of Pd on Au. Phys. Rev. B 2003, 67, 33409.
49
50
51 54. Tang, W.; Henkelman, G. Charge redistribution in core-shell nanoparticles to promote oxygen
52
53 reduction. J. Chem. Phys. 2009, 130, 194504.
54
55
56
57
58
59 29
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 30 of 34

1
2
3
55. Gorzkowski, M.; Lewera, A. Probing the limits of d-band Center Theory: Electronic and
4
5
Electrocatalytic Properties of Pd-Shell-Pt-Core Nanoparticles. J. Phys. Chem. C 2015, 119,
6
7
18389–18395.
8
9
10 56. Tedsree, K.; Li, T.; Jones, S.; Chan, C. W. A.; Yu, K. M. K.; Bagot, P. A. J.; Marquis, E. A.;
11
12 Smith, G. D. W.; Tsang, S. C. E. Hydrogen production from formic acid decomposition at
13
14 room temperature using a Ag-Pd core-shell nanocatalyst. Nat Nano 2011, 6, 302–307.
15
16
17 57. Holmberg, N.; Laasonen, K.; Peljo, P. Charge distribution and Fermi Level in bimetallic
18
19 nanoparticles. Phys. Chem. Chem. Phys. 2016, 18, 2924–2931.
20
21
22 58. Vegard, L. Die konstitution der mischkristalle und die raumfüllung der atome. Zeitschrift für
23
24 Physik A Hadrons and Nuclei 1921, 5, 17–26.
25
26
27 59. Martin, T. P. Shells of atoms. Phys. Rep. 1996, 273, 199–241.
28
29
60. Johnston, R. L. Atomic and Molecular Clusters; Taylor and Francis: London, 2002.
30
31
32 61. Baletto, F.; Ferrando, R.; Fortunelli, A.; Montalenti, F.; Mottet, C. Crossover among structural
33
34 motifs in transition and noble-metal clusters. J. Chem. Phys. 2002, 116, 3856–3863.
35
36
37 62. Xing, X.; Yoon, B.; Landman, U.; Parks, J. H. Structural Evolution of Au Nanoclusters: From
38
39 Planar to Cage to Tubular Motifs. Phys. Rev. B 2006, 74, 165423.
40
41
42 63. Baletto, F.; Ferrando, R. Structural properties of nanoclusters: Energetic, thermodynamic, and
43
44 kinetic effects. Rev. Mod. Phys. 2005, 77, 371–423.
45
46
47 64. Li, Z. Y.; Young, N. P.; Vece, M. D.; Palomba, S.; Palmer, R. E.; Bleloch, A. L.; Curley, B. C.;
48
49 Johnston, R. L.; Jiang, J.; Yuan, J. Three-dimensional Atomic-scale Structure of Size-selected
50
51 Gold Nanoclusters. Nature 2008, 451, 46–48.
52
53
54
65. Wells, D. M.; Rossi, G.; Ferrando, R.; Palmer, R. E. Metastability of the atomic structures of
55
56
size-selected gold nanoparticles. Nanoscale 2015, 7, 6498–6503.
57
58
59 30
60 ACS Paragon Plus Environment
Page 31 of 34 The Journal of Physical Chemistry

1
2
3
66. Tozer, D. J.; De Proft, F. Computation of the Hardness and the Problem of Negative Electron
4
5
Affinities in Density Functional Theory. J. Phys. Chem. A 2005, 109, 8923–8929, PMID:
6
7
16834296.
8
9
10 67. Yu, W. W.; Wang, Y. A.; ; Peng*, X. Formation and Stability of Size-, Shape-, and Structure-
11
12 Controlled CdTe Nanocrystals:âĂL’ Ligand Effects on Monomers and Nanocrystals. Chem-
13
14 istry of Materials 2003, 15, 4300–4308.
15
16
17 68. T. Rajh, .; Chen, L. X.; Lukas, K.; Liu, T.; Thurnauer, M. C.; ; Tiede, D. M. Surface Restruc-
18
19 turing of Nanoparticles:âĂL’ An Efficient Route for LigandâĹŠMetal Oxide Crosstalk. The
20
21 Journal of Physical Chemistry B 2002, 106, 10543–10552.
22
23
24 69. Mortensen, J. J.; Hansen, L. B.; Jacobsen, K. W. Real-space Grid Implementation of the Pro-
25
26 jector Augmented Wave Method. Phys. Rev. B 2005, 71, 035109.
27
28
29
70. Enkovaara, J.; Rostgaard, C.; Mortensen, J. J.; Chen, J.; Dulak, M.; Ferrighi, L.; Gavnholt, J.;
30
31
Glinsvad, C.; Haikola, V.; Hansen, H. A. et al. Electronic Structure Calculations with GPAW:
32
33
a Real-space Implementation of the Projector Augmented-wave Method. J. Phys.: Condens.
34
35
Matter 2010, 22, 253202.
36
37 71. Blöchl, P. E. Projector Augmented-wave Method. Phys. Rev. B 1994, 50, 17953–17979.
38
39
40 72. Marchal, R.; Yudanov, I. V.; Matveev, A. V.; Rösch, N. Scalable properties of metal clusters:
41
42 A comparative DFT study of ionic-core treatments. Chem. Phys. Lett. 2013, 578, 92–96.
43
44
45 73. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations
46
47 using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186.
48
49
50 74. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple.
51
52 Phys. Rev. Lett. 1996, 77, 3865–3868.
53
54
75. Shanno, D. F. Conditioning of Quasi-Newton Methods for Function Minimization. Math.
55
56
Comp. 1970, 24, 647–656.
57
58
59 31
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 32 of 34

1
2
3
76. Tang, W.; Sanville, E.; Henkelman, G. A Grid-based Bader Analysis Algorithm without Lat-
4
5
tice Bias. J. Phys.: Compute Mater. 2009, 21, 084204.
6
7
8 77. Kittel, C. Introduction To Solid State Physics 8th Edition; Wiley: New York, 2004.
9
10
11 78. Perdew, J. P.; Zunger, A. Self-interaction correction to density-functional approximations for
12
13 many-electron systems. Phys. Rev. B 1981, 23, 5048–5079.
14
15
16 79. Zarechnaya, E. Y.; Skorodumova, N. V.; Simak, S.; Johansson, B.; Isaev, E. I. Theoretical
17
18 study of linear monoatomic nanowires, dimer and bulk of Cu, Ag, Au, Ni, Pd and Pt. Compu-
19
20 tational materials science 2008, 43, 522–530.
21
22
23 80. Salzner, U.; Baer, R. Koopmans’ springs to life. J. Chem. Phys 2009, 131, 231101.
24
25
26
81. Bellafont, N. P.; Illas, F.; Bagus, P. S. Validation of Koopmans’ theorem for density functional
27
28
theory binding energies. Phys. Chem. Chem. Phys. 2015, 17, 4015.
29
30 82. Lide, D. R. CRC Handbook of Chemistry and Physics, 84th Edition; CRC Press: Boca Raton,
31
32 Florida, 2003.
33
34
35 83. Bochicchio, D.; Ferrando, R. Morphological instability of core-shell metallic nanoparticles.
36
37 Phys. Rev. B 2013, 87, 165435.
38
39
40 84. Laasonen, K.; Panizon, E.; Bochicchio, D.; Ferrando, R. Competition between Icosahedral
41
42 Motifs in AgCu, AgNi, and AgCo Nanoalloys: A Combined Atomistic–DFT Study. The Jour-
43
44 nal of Physical Chemistry C 2013, 117, 26405–26413.
45
46
47 85. Panizon, E.; Bochicchio, D.; Rossi, G.; Ferrando, R. Tuning the structure of nanoparticles by
48
49 small concentrations of impurities. Chemistry of Materials 2014, 26, 3354–3356.
50
51
52
86. Gould, a. L.; Heard, C. J.; Logsdail, a. J.; Catlow, C. R. a. Segregation effects on the properties
53
54
of (AuAg) 147. Phys. Chem. Chem. Phys. 2014, 16, 21049–21061.
55
56
57
58
59 32
60 ACS Paragon Plus Environment
Page 33 of 34 The Journal of Physical Chemistry

1
2
3
87. Cerbelaud, M.; Ferrando, R.; Barcaro, G.; Fortunelli, A. Optimization of chemical ordering in
4
5
AgAu nanoalloys. Phys. Chem. Chem. Phys. 2011, 13, 10232.
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 33
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 34 of 34

1
2
3
4
5
6
7
8
9
10
11
12
13 ...-...
14 >(l!J
-
-...
15 1
16 ...-I
17
18 1,/1,
(l!Ji
19
....,ro,
.i
20
21 V),
22 .......
23 o,
24 >,,1,..--.-------­
.,J
25 v,,
26 C:
27 (l!J1
28 Cl
29
30
31
32
33
34
35
36
...... 11 �10 �7 -.5
37 (e )
38
39
40
41
42
43
44
45
TOC GRAPHIC
ACS Paragon Plus Environment
46
47

You might also like