Electronic Properties of Pt–Ti Nanoalloys and the Effect on Reactivity

Jun 26, 2012 - School of Chemistry, The University of Birmingham, Edgbaston, Birmingham B15 2TT, U.K.. ABSTRACT: Theoretical investigations of Pt−Ti...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/JPCC

Electronic Properties of Pt−Ti Nanoalloys and the Effect on Reactivity for Use in PEMFCs Paul C. Jennings,† Bruno G. Pollet,‡ and Roy L. Johnston*,§ †

PEM Fuel Cell Research group, School of Chemical Engineering, Centre for Hydrogen and Fuel Cell Research, The University of Birmingham, Edgbaston, Birmingham B15 2TT, U.K. ‡ HySA Systems, SAIAMC, University of the Western Cape, Private Bag X17, Modderdam Road, Bellville 7535, Cape Town, South Africa § School of Chemistry, The University of Birmingham, Edgbaston, Birmingham B15 2TT, U.K. ABSTRACT: Theoretical investigations of Pt−Ti nanoparticles are presented using density functional theory (DFT) to study clusters up to sizes of 1.7 nm (201 atoms). Several compositions have been studied for varying sizes, and a projected density of states analysis has been performed. Changes in d-band properties have been correlated with differences in OH and CO binding energies with relation to changes in composition. It was found that a Pt-rich Pt−Ti alloy shows promise for potential use in a PEMFC with the alloy resulting in a downshift in d-center, compared to pure Pt clusters, which correlates with a weakening of the OH and CO adsorption energies. Furthermore, it was observed that varying the size of the cluster gives rise to changes in the d-band center with larger clusters typically having a more negative d-band center, although values obtained are not as negative as the d-band center for bulk Pt3Ti reported in the literature.



INTRODUCTION To develop and successfully market low-temperature (LT) and high-temperature (HT) proton exchange membrane fuel cells (PEMFCs), it is necessary to reduce the platinum (Pt) catalyst loading of both anode and cathode electrodes together with the associated cost without compromising the performance.1,2 The high intrinsic cost of Pt (£1,040.24 per troy ounce as of 20/02/ 123) forms a barrier to the widespread commercialization of PEMFCs. Current state-of-the-technology membrane electrode assemblies (MEAs) have a total Pt loading of 0.40 mg cm−2 comprising 0.05 mg Pt cm−2 used on the anode for the fast hydrogen oxidation reaction (HOR) and 0.35 mg Pt cm−2 for the cathode where the rate-limiting, four-electron oxygen reduction reaction (ORR) occurs. This gives a total of ∼30 g Pt per 100 kW PEMFC stack. A 10-fold decrease in Pt loading in PEMFC stacks is required to make PEMFCs a commercially viable option. There are several ways to do this: lower the Pt loading by using a Pt alloy (either binary or ternary4); replace Pt with a nonprecious electrocatalyst (inorganic or organic5,6); maximize the effective surface area of the Pt catalyst, i.e., the surface contact between the electrode catalyst layers (CL), the carbonaceous electronic conductor−gas diffusion layer (GDL), the polymer electrolyte membrane (Nafion, PEM) and the reactants (hydrogen and oxygen). The alloying of Pt to reduce loading7 and improve catalytic characteristics such as performance as well as durability and reactivity8−10 is a promising approach. We previously reported the results of theoretical calculations which suggested that © 2012 American Chemical Society

alloying Pt with titanium (Ti) to form Ti@Pt core−shell clusters results in weaker binding of hydroxyl (OH) and carbon monoxide (CO)11 with reduced Pt loading. Weaker binding of OH can lead to improvements in ORR kinetics, while weakening of CO binding lessens poisoning effects. CO is a common contaminant in H2 streams which are often produced through steam methane reforming (SMR). CO is responsible for blocking active sites and thus reducing the catalyst’s ability to oxidize the hydrogen molecule. In this work, in order to better understand the reasons for the changes in adsorption energies, a study of the electronic density of states (DOS) is performed. The average energy of the d-band (the d-band center) was linked to the adsorption properties of small molecules.12−15 Adsorption energy is dependent upon the filling of the d-band, where the occurrence of more unoccupied dstates allows for stronger binding of small molecules. Therefore, an upshift in the d-band center toward the Fermi energy (EF) gives rise to stronger binding; conversely, a downshift, away from the EF, results in weaker binding due to greater filling of highlying d-states. Simulations of nanoparticles present complex issues, especially when modeling transition metals, making it is necessary to perform ab initio simulations. Though it is computationally expensive to perform ab initio calculations on large clusters, density functional theory (DFT) calculations have been Received: April 13, 2012 Revised: June 18, 2012 Published: June 26, 2012 15241

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

analysis was performed, which we previously reported to give reliable results compared to Mulliken and Bader charge analysis for the Pt−Ti system.11 To simulate compositional and size effects on the ORR and susceptibility to CO poisoning, adsorption energies for single OH and CO molecules on a variety of clusters were calculated. The Atop-(111) site was chosen for chemisorption studies due to its presence in all cluster sizes and compositions. Furthermore, there is little probability of the small molecules migrating away from this site during relaxation calculations.

successfully run on a wide variety of clusters with several tens to hundreds of atoms.10,16,17 Furthermore, DFT calculations have been acknowledged as one of the best methods for performing accurate simulations on metallic clusters.18 DFT calculations have been performed on pure Pt, Ti, and bimetallic Pt−Ti clusters. A study of changes in the DOS associated with composition and size effects has been performed. Furthermore, selected cluster adsorption energies with OH and CO have been calculated in order to assess the effect of variations in the DOS on the utility of pure and bimetallic clusters as PEMFC electrocatalysts.





RESULTS AND DISCUSSION Structural Optimizations. Simulations have been carried out on truncated octahedral (TO) clusters of varying sizes and compositions. In this study, calculations have been restricted to high symmetry (Oh) core−shell and onion-like configurations, rather than exploring a wider range of chemical ordering patterns.10,20−23 It should be noted, however, that in a previous study of 38-atom TO clusters chemical ordering arrangements with some degree of surface mixing were also investigated.11 [For a wider discussion of chemical ordering in nanoalloys, see ref 10.] Complete TO structures are obtained for 38, 79, 116, 140, 201, 260, etc., atoms,24 where the odd-numbered TOs are atomcentered and the even-numbered TOs have a central M6 octahedron core. Complete core−shell structures were previously investigated for 38-atom TO clusters for composition A32B6.11 Here, larger TO clusters with 79, 116, 140, and 201 atoms are investigated. The 79-atom clusters have a complete core−shell structure at the A60B19 composition, although it is also possible to study the A68B1 composition for this atom-centered cluster. A78B38 and A110B6 compositions result in complete core−shell structures for the 116-atom clusters, with shells composed of one and two layers of species A, respectively. The 140-atom cluster has complete core−shell structures at A96B44 and A134B6 as well as an onion structure for A102B38 (A96B38A6). The A122B79 and A182B19 compositions result in complete core−shell structures for the 201-atom clusters. The A141B60 (A122B60A19) composition is an onion structure, although there is the possibility of creating an A140B61 (A122B60A18B1) onion structure as the cluster is atomcentered. Through the study of larger cluster sizes it is possible to investigate the effect of significantly reduced Pt loading in the electrocatalyst. Although the previously studied bimetallic Pt32Ti6 clusters still have around 84% Pt content, this is reduced to around 61% for the core−shell Pt122Ti79 cluster. When compared to the 75% Pt loading for the bulk Pt3Ti systems, this demonstrates the possibility to produce significant reductions through use of nanoparticles compared to bulk materials. The Pt60Ti19, Pt78Ti38, and Pt96Ti44 compositions correspond to 76%, 67%, and 68% Pt loading, respectively. Cluster binding energies (Eb) are used to give an indication of the relative cluster stability. More negative energies suggest that the structure is more stable than those with less negative energies. Figure 1a shows the binding energies associated with pure and alloyed core−shell TO structures for varying sizes and compositions. This shows that the most negative binding energies, and so the most stable structures, are obtained for the Pt-majority core−shell clusters, with the least negative, least stable structures, obtained for the pure Pt clusters. There is also a noticeable shift to more negative binding energies as size increases. This is due to the larger clusters studied in this work having increased average atomic coordination numbers when

METHODS In this work, all DFT geometry optimizations were performed using the PWscf plane wave DFT code in the quantum chemistry package Quantum-Espresso 4.2.1.19 The PWscf calculations were performed using the Perdew−Wang (PW91) exchangecorrelation functional with 10 (5d9, 6s1) and 12 (3s2, 3p6, 3d2, 4s2) valence electrons for Pt and Ti, respectively. A kinetic energy (Ek) wave function cutoff of 26.0 Ry was used, and the Ek cutoff for the charge density (ρ) was set at 208.0 Ry as ultrasoft pseudopotentials were used. Furthermore, Fermi−Dirac smearing was employed, with a Gaussian smearing of 0.007 Ry. The supercell dimensions were taken to be sufficiently large to remove the possibility of interactions with neighboring clusters, however not so large as to be detrimental to calculation speed. When setting up the calculations, all parameters were optimized to best utilize the computational resources while obtaining as accurate results as possible. The values of all cutoffs, smearing, and cell sizes were optimized to ensure maximum efficiency, which becomes particularly important for the larger cluster sizes. The effects of electron spin were also studied but seen to have little effect on the results, so spin-polarized calculations were not investigated further. The binding energy per atom (Eb) of a cluster is given by Eb =

1 [Etot(A mBn) − mEtot(A) − nEtot(B)] N

(1)

where N is the total number of atoms, Etot(AmBn) is the total electronic energy of the cluster, and Etot(A) is the total electronic energy of the single atom. The energy of the d-band center (εd) is given by εd =

∫ ρ E dE ∫ ρ dE

(2)

where ρ is the d-band density, E is the d-band energy, and ρ dE is the number of d-states. The root-mean-squared d-bandwidth (Wd) is given by Wd =

∫ ρ E 2 dE ∫ ρ dE

(3)

When calculating εd and Wd for a specific fraction of a cluster (e.g., the core or shell atoms), eqs 2 and 3 were applied for the relevant subset of atoms, i.e., by summing over atoms in the appropriate fraction of the cluster. DFT explicitly treats electron interactions to effectively model catalyst substrate interactions. Furthermore, explicit treatment of electrons is required for DOS and charge analysis. The projwfc module of the Quantum-Espresso package was used to perform DOS and charge analysis. In this investigation, Löwdin charge 15242

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

DOS Analysis. DOS calculations have been performed to determine the d-band characteristics for the varying TO cluster sizes and compositions. Following geometry optimization, a projected DOS (PDOS) analysis is performed on the bare clusters. In conjunction with the PDOS calculations, Löwdin charge analysis is also performed to give an indication of electron transfer within the cluster. This allows us to link charge effects to changes in the PDOS, which is discussed later. In this study, when quoting PDOS values, only the outer shell of atoms is considered, unless otherwise stated. The values stated are not divided up to give data per atom; the values are for the entire outer shell. This is because it is at the outer layer of atoms that the electrochemical reactions will occur, and so changes in shell properties can have significant effects on the catalytic activity of a cluster. Furthermore, by averaging over the entire outer shell, it removes variations due to site specific changes, e.g., centroid or edge sites gaining greater charge. Figure 2 shows a plot of d-center (εd) and d-width (Wd) for the 38-, 79-, 116-, 140-, and 201-atom clusters, for which the raw data

Figure 1. Binding energies (Eb) calculated for (a) pure Pt and Ti and bimetallic Pt−Ti monolayered core shell clusters, studying compositional as well as size effects, as well as for (b) the onion and bilayered core shell structures for the 140- and 201-atom clusters. Figure 2. Relationship between d-center (εd) and d-width (Wd) for the 38 (blue), 79 (red), 116 (green), 140 (orange), and 201 (purple) atom pure Pt (diamond), Pt-rich monolayer (square), Ti-rich monolayer (triangle), and Ti pure (circle) clusters.

compared to the small 38-atom clusters. The stability of the Ptmajority core−shell isomer is advantageous for PEMFCs which benefit from a Pt shell on which the ORR and HOR can take place. Figure 1b shows the calculated binding energies for all 140and 201-atom cluster compositions investigated. This reveals that the Pt-rich core−shell and onion structures had very similar stabilities. For the Ti-rich clusters, the onion structure is slightly more stable than both core−shell structures. However, the Ptrich clusters remain most favorable for the onion or monolayered core−shell composition. It can also be seen that the bilayer Pt shell (Pt134Ti6 and Pt182Ti19) clusters are less stable than for the Pt-rich monolayer and onion structures. However, for the Ti-rich clusters, the bilayer conformation is slightly more favorable than the monolayer. This suggests that the larger Pt core induces more strain within the clusters, making them slightly less stable. We previously reported that for the 38-atom TO clusters the presence of a Ti shell produced significantly stronger interactions with OH and CO molecules.11 This would inhibit the ORR; in other words, the electrocatalyst would be less efficient than pure Pt. Also, the electrocatalyst would be more susceptible to CO poisoning, thus leading to more blocking of active sites on the catalyst surface. Therefore, in this work emphasis is placed on the energetically favorable bimetallic Pt-rich clusters, although Tirich clusters are also investigated for completeness. Comparisons are made with the currently used pure Pt clusters to establish any advantages which may occur due to alloying.

can be found in Table 1. This reveals the expected trend that a more negative d-center coincides with an increase in d-band width. As d-band occupancy is constant, the d-center shifts to more negative values (relative to the Fermi energy, EF) as a consequence of the wider d-band. If the d-band were to become more narrow, this would result in the d-center shifting to a more positive value. The d-center of the pure and majority Pt clusters are found to be more negative than for the pure and majority Ti clusters. This is because Pt has nine d-band valence electrons while Ti has only two. The d-band center of the bimetallic Pt-rich clusters is also found to be slightly more negative than for the pure Pt clusters. This trend is observed for all cluster sizes, suggesting that the formation of a Ti@Pt cluster results in broadening of the d-band. Although the Ti@Pt bimetallic system has fewer electrons than the pure Pt system, the Ti levels are located higher in energy than those of Pt (as Ti is more electropositive than Pt), and so broadening occurs upon alloying. This broadening then results in the observed downshift in d-center. It is also possible to see size effects with, in general, a slight downshift in d-center and a broadening of the d-width as the cluster size increases. However, the relationship between d-center and d-width for the Ti-rich and pure Ti clusters is less regular. Pt@Ti bimetallic 15243

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

Table 1. Calculated Values of Fermi Energy (EF), d-Center (εd), d-Width (Wd), and Average Charge of Various Clusters (Data for Outer Shell Atoms) as Well as OH and CO Adsorption Energies Eads (eV) total no. atoms

chemical ordering

no. Pt atoms

EF (eV)

εd (eV)

Wd (eV)

av shell charge

OH

CO

38

pure monolayer monolayer pure pure monolayer monolayer pure pure monolayer monolayer pure pure monolayer monolayer pure pure monolayer monolayer pure

0 6 32 38 0 19 60 79 0 38 78 116 0 44 96 140 0 79 122 201

−2.750 −2.707 −4.589 −4.950 −2.479 −2.401 −3.979 −4.346 −2.141 −1.950 −3.435 −3.801 −1.828 −1.732 −3.044 −3.488 −1.189 −0.998 −2.264 −2.642

−0.381 −0.514 −2.320 −2.020 −0.390 −0.398 −2.379 −2.110 −0.162 0.041 −2.460 −2.082 −0.392 −0.407 −2.596 −2.160 −0.329 −0.317 −2.548 −2.115

1.048 1.341 2.907 2.715 1.051 1.403 2.976 2.817 1.005 1.393 3.013 2.781 1.089 1.441 3.148 2.893 1.044 1.473 3.085 2.826

0.033 0.058 −0.017 0.022 0.033 0.071 −0.032 0.022 0.042 0.094 −0.036 0.023 0.029 0.083 −0.042 0.015 0.041 0.099 −0.045 0.020

−5.117 −5.139 −2.258 −2.775 −5.853 −5.955 −2.440 −2.787 −6.045 −6.154 −2.379 −2.792 −6.107 −5.259 −2.540 −2.686 −5.116 −6.066 −2.344 −2.637

−2.871 −3.095 −0.788 −1.848 −3.478 −3.246 −1.075 −1.773 −3.286 −3.527 −1.037 −1.800 −3.249 −3.778 −0.923 −1.606 −3.458 −1.740 −0.710 −1.521

79

116

140

201

clusters have slightly broader d-widths than the pure Ti clusters. This can again be explained by considering the effect of alloying with Pt. The Pt d-band character is found below that for the Ti dband. Therefore, upon alloying with Pt, broadening occurs due to interactions of the lower Pt d-band character with the higher Ti dband. For the pure Ti and Ti-rich clusters size effects seem to have little effect on d-band characteristics. The lack of correlation between εd and Wd for the pure Ti and Ti-rich clusters is most likely due to structural effects. Contraction of Ti−Ti bonds would result in a downshift of the d-center, while elongation would lead to an upshift.25 The relationship between the d-band character for Pt and Ti at the 116-cluster size can be seen in Figure 3. This clearly demonstrates that, upon alloying Pt with Ti, the d-band levels below EF are generally dominated by the Pt orbitals. Conversely, the levels above the EF tend to be dominated by Ti. This is likely due to two effects: charge transfer and orbital overlap. Charge transfer resulting in donation of electron density from Ti to Pt will lead to filling of the higher d-band states, causing an increase in EF and a consequent downward shift of the d-band center related to EF. Furthermore, it is likely that orbital overlap will increase due to interactions between the Pt and Ti, resulting in broadening of the d-band and thus a downshift in d-center. Through the study of larger clusters, it is also possible to investigate compositional effects in more detail. The 38-atom clusters allows for four high symmetry configurationstwo pure and two core−shellas does the 79-atom cluster if changes to the single central atom are ignored. The larger 116-atom cluster allows for the study of six high-symmetry structures: two pure and four core−shell. The 140- and 201-atom clusters allow for the generation of further structures: two pure, four core−shell, and two onion. A breakdown of the d-band characteristics for these larger clusters is given in Table 2. The bimetallic Pt78Ti38 cluster has a more negative d-center than the pure Pt116 cluster. However, the Pt110Ti6 cluster has a slightly less negative d-center than the pure Pt116 cluster. Splitting the cluster into layers (inner core, outer core, and shell) reveals

the reason for this. For the pure Pt116 cluster the ordering of the d-center, with ascending negativity, is inner core (Pt) > outer core (Pt) > shell (Pt), with a difference of 0.69 eV between the outer core and shell. The Pt110Ti6 cluster has altered ordering due to the presence of the Ti core to outer core (Pt) > shell (Pt) > inner core (Ti). There is a slightly larger difference between the outer core and shell of 0.92 eV in this case. This shows that the Pt d-band broadening occurs most significantly at the Pt−Ti interface; in the Pt110Ti6 case the outer core is able to shield the shell from the same broadening effects as with the Pt78Ti38 cluster. Analyses recently reported by Chen et al.26 demonstrated experimental results showing favorable ORR properties when Pt forms a near monolayer on the underlying nonprecious metal. This supports the results presented here where the most significant changes in d-band character happen at the Pt−Ti interface. The same trend can be seen for the larger clusters in all cases. The relationship between d-center, d-width, and composition for the Ti-rich clusters is slightly more complex. The pure Ti116 and the bimetallic Pt6Ti110 clusters have similar d-centers. This suggests that the six inner core Pt atoms in the bimetallic Pt6Ti110 cluster have little effect on the shell due to the shielding effects of the outer core. For the pure Ti116 cluster the d-center ordering, with ascending negativity is outer core (Ti) > inner core (Ti) > shell (Ti). For the bimetallic Pt6Ti110 the ordering is inner core (Pt) > outer core (Ti) > shell (Ti). However, the shell of the bimetallic Pt38Ti78 cluster has a slightly positive d-center coupled with a larger d-width; the ordering in this case is outer core (Pt) > inner core (Pt) > shell (Ti). A narrowing of the d-width and a slight upshift in d-center were expected due to interactions between the Pt core and Ti shell. However, there is a broadening of the d-band and a dramatic upshift in d-center to a slightly positive value. This suggests that charge transfer from the Ti shell to the Pt core lead to a positive d-center. Once again, the same trends are true for the larger clusters. The 140 (A96B38A6) and 201 (A122B60A19) atom onion clusters result in the cluster shells having similar d-centers to those for the 15244

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

Figure 3. Relationship between d-band character for Pt116 (a), Pt110Ti6 (b), Pt78Ti38 (c), Pt38Ti78 (d), Pt6Ti110 (e), and Ti116 (f). d-band PDOS is evaluated for fractions of the clusters, comparing the inner core, outer core, and shell. This demonstrates that the higher orbitals, closer to the Fermi energy (EF), have predominantly Ti character, whereas the lower-lying orbital are mostly Pt in character.

trends observed in this study; as the cluster size increases, tending toward the bulk, the d-center becomes more negative. For the pure Pt clusters, the d-center ranges from −2.02 eV (38 atoms) to −2.16 eV (140 atoms). For the monolayered bimetallic Pt-rich clusters, the d-center ranges from −2.32 eV (38 atoms) to −2.60 eV (140 atoms). Charge Effects. The results of the Löwdin charge analysis are shown in Figure 4, giving composition, average outer shell charge, and a color-coded image of the charge distribution. This shows the charges on clusters of constant size plotted on the same scale to investigate how the charge changes with composition. This clearly demonstrates that most clusters have an outer shell with an average positive charge, apart from the Ti@

A96B44 and A122B79 clusters. For the Pt96Ti38Pt6 and Pt122Ti60Pt19 clusters, the d-center ordering, with ascending negativity, is inner core (Pt) > shell (Pt) > outer core (Ti). For the Ti96Pt38Ti6 and Ti122Pt60Ti19 clusters, the d-center ordering, with ascending negativity, is outer core (Pt) > inner core (Ti) > shell (Ti). This is in line with the previously investigated core−shell structures, showing large changes in d-center at the Pt−Ti interface. A search of the literature shows that the bulk Pt surface has a calculated d-center of −2.44 eV while bulk Pt3Ti has a calculated d-center of −3.16 eV.13 This demonstrates the trend that there is a downshift in the d-center when Pt is alloyed with Ti, also suggesting weaker adsorption of e.g. OH to the bulk Pt3Ti system compared to the pure Pt bulk. These results fit well with the 15245

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

Table 2. Calculated Values of d-Center (εd) and d-Width (Wd) for Pure, Monolayer, Bilayer, and Onion Compositions at Varying Sizes Showing the Relationship between Cluster Inner Core, Outer Core, and Shell Atoms εd (eV)

Wd (eV)

total no. atoms

chemical ordering

no. Pt atoms

inner core

outer core

shell

inner core

outer core

shell

116

pure bilayer monolayer monolayer bilayer pure pure bilayer onion monolayer monolayer onion bilayer pure pure bilayer onion monolayer monolayer onion bilayer pure

0 6 38 78 110 116 0 6 38 44 96 102 134 140 0 19 60 79 122 141 182 201

−0.461 −3.075 −2.193 0.144 0.536 −2.899 −0.443 −3.404 −0.820 −2.539 0.039 −3.233 0.503 −2.734 −0.543 −3.193 −0.551 −2.495 0.043 −3.165 0.644 −2.727

−0.521 −0.544 −2.970 −0.220 −2.930 −2.775 −0.553 −0.695 −3.469 −3.424 −0.258 −0.242 −2.860 −2.582 −0.637 −0.712 −3.438 −3.312 −0.290 −0.325 −2.927 −2.715

−0.162 −0.111 0.041 −2.460 −2.010 −2.082 −0.392 −0.313 −0.354 −0.407 −2.596 −2.604 −2.154 −2.160 −0.329 −0.276 −0.285 −0.317 −2.548 −2.640 −2.070 −2.115

1.177 3.379 2.866 1.250 2.476 3.648 1.155 3.703 2.179 3.118 1.221 3.652 2.412 3.478 1.197 3.537 1.909 3.110 1.250 3.592 2.353 3.483

1.245 1.425 3.452 2.118 3.692 3.559 1.255 1.464 3.859 3.844 2.096 2.169 3.596 3.357 1.305 1.574 3.828 3.763 2.041 2.188 3.656 3.509

1.005 0.994 1.393 3.013 2.681 2.781 1.089 1.045 1.394 1.441 3.148 3.165 2.857 2.893 1.044 1.040 1.423 1.473 3.085 3.178 2.735 2.826

140

201

the system becomes positive, it is the d-band that loses the most electron density. Figure 6 shows the relationship between charge and d-center for various cluster sizes with varying compositions. For the Ptrich clusters, a loss of charge density results in the d-center becoming more negative. Also, as the cluster size increases, there is greater transfer of negative charge to the outer shell. The same is not true for the Ti-rich clusters, where gain of positive charge results in a downshift in the d-center. This is likely due to structural effects, (e.g., bond strain) canceling out the various charge effects through opposing changes in d-band characteristics. Chemisorption Studies. In these chemisorption studies, calculated adsorption energies are linked to previously discussed DOS values. This enables an indication of how changes in the composition and size of clusters affects adsorption energies brought about by changes in the electronic structure of the cluster. The adsorption energies reported in Table 1 are for the atop-(111) surface site only. It should be noted that this site is not actually the most favorable for these types of systems.11,25 In the previous study11 it was found that the site of strongest binding (the edge-bridge site between the (111) and (100) planes) demonstrated little difference between absorption on the pure or bimetallic clusters. However, when comparing adsorption energies for the other sites, more significant differences were observed. As the cluster size increases, the number of strongly binding edge-bridge sites decreases relative to the other sites where differences in the adsorption energies were observed. It was therefore thought that this site would be most suitable to allow relevant comparisons between the pure and bimetallic clusters. This site also has the advantage of being a locally stable binding site; no barrierless migration from the site was observed in the previous study.11 Figure 7a shows the OH adsorption energy decreasing, as the d-center became more negative, shifting down relative to the EF.

Pt core−shell structures. The shell of the Ti majority clusters becomes significantly more positive than other compositions. This shows that alloying Pt and Ti results in a loss of electron density from the electropositive Ti atoms and a gain of charge density by the electronegative Pt atoms. It also shows that for the pure clusters there is slight donation from the shell to the core of the cluster. It is expected that loss of electron density results in a downshift of the EF and a narrowing of the DOS. This is coupled with the previously discussed observation that a narrowing of the d-band would result in an upshift in the d-center. But, the overall charge of the clusters remains neutral (with only the core or shell atoms gaining charge), and the DOS is plotted with respect to the EF. This normalization results in only narrowing or broadening of the d-band a due to changes in atomic charge having an observable effect on the d-center. This is clearly demonstrated through studying charged Pt38 TO clusters. The DOS characteristics were studied for the entire cluster as electron density was altered for the whole system. The neutral cluster was compared with Pt38x clusters where x = −2, −1, +1, and +2. Figure 5a shows the relationship between the EF and d-center, demonstrating that changes in the EF due to charge have no effect on the d-center. Instead, the d-center changes due to broadening or narrowing of the d-band, seen in Figure 5b. It is apparent that as the system becomes more positive, the EF decreases, the d-width decreases, and there is an upshift in the dcenter energy. The reverse is found for negative clusters. We also plotted changes in charge distribution, as shown in Figure 5c. The figure shows the average number of electrons in the s-, p-, and d-bands. The partial charges are normalized to the values calculated for the neutral cluster. A negative value indicates gain of electrons and positive values represent loss of electrons. When the system becomes negative, the p-band gains more electron density than the s- or d-bands. Conversely, when 15246

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

Figure 4. Charge density plots showing net atomic charges, with negative charge in blue and positive charge in red. The scale remains constant across varying compositions and sizes. This plot demonstrates the donation of electron density from Ti to Pt.

This trend is generally observed across varying cluster sizes and compositions suggesting a close relationship between the calculated d-center and the adsorption energy which has been extensively discussed in the literature.13,15,27 This explains the

slight weakening of OH adsorption on the bimetallic Pt-rich clusters compared to other cluster compositions. The d-center characteristics help to explain the differences in binding energy for the pure and bimetallic clusters investigated in 15247

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

Figure 6. Relationship between d-center (εd) and charge for the 38 (blue), 79 (red), 116 (green), 140 (orange), and 201 (purple) atom pure Pt (diamond), Pt-rich (square), Ti-rich (triangle), and Ti pure (circle) clusters.

Figure 5. Effect of altering the cluster charge on the Fermi energy (EF) (a), d-width (Wd) (b), and band filling (c) with relation to the d-center (εd).

this study. Stronger binding is observed for the pure Pt clusters compared to the bimetallic Pt-rich clusters. This can be attributed to the downshift observed in d-center when comparing the pure to the bimetallic. This downshift is due to filling of orbitals in the Pt shell through donation of d-band density from the Ti core. The adsorption energy associated with CO bonding has also been studied in order to assess the susceptibility of the electrocatalyst to CO poisoning. From previous studies, it was found that reducing the Pt−C bond strength results in a reduced susceptibility to CO poisoning.28 This study is therefore aimed at establishing the strength with which a CO molecule binds to the cluster surface.

Figure 7. Relationship between d-center (εd) and (a) OH adsorption energy as well as (b) CO adsorption for the 38 (blue), 79 (red), 116 (green), 140 (orange), and 201 (purple) atom pure Pt (diamond), Ptrich (square), Ti-rich (triangle), and Ti pure (circle) clusters to the atop(111) site.

Previous investigations demonstrated unusual CO bonding to the Ti-rich clusters, where both the C and O are bound to the Ti surface.11 This is once again observed for the larger cluster sizes. This unusual binding motif results in stronger CO bonding than would be observed for the more common axial binding motif, 15248

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C



ACKNOWLEDGMENTS P.C.J. thanks the Centre for Hydrogen and Fuel Cell Research at The University of Birmingham and the RCUK CDT in Hydrogen, Fuel Cells and their applications for PhD funding, as well as financial support from COST Action MP0903: “Nanoalloys as Advanced Materials: From Structure to Properties and Applications”. The calculations described in this paper were performed using The University of Birmingham’s BlueBEAR HPC service,29 which was purchased through HEFCE SRIF-3 funds. Via our membership of the UK’s HPC Materials Chemistry Consortium, which is funded by EPSRC (EP/ F067496), this work made use of the facilities of HECToR, the UK’s national high-performance computing service, which is provided by UoE HPCx Ltd at the University of Edinburgh, Cray Inc. and NAG Ltd., and funded by the Office of Science and Technology through EPSRC’s High End Computing Programme.

with M−C binding only. This binding motif was attributed to the particularly strong Ti−O bonds that formed. Results for CO adsorption energies are presented in Figure 7b. A similar trend is observed for CO adsorption as with OH adsorption. As the d-center upshifts towards EF, the adsorption energy becomes more negative, resulting in stronger binding. Once again this suggests electronic effects contribute significantly to the weaker binding energy of CO to the bimetallic Ptrich cluster compared to other compositions. The CO adsorption energy for the Pt79Ti122 cluster is significantly higher in energy than for the other Ti-rich clusters. This is due to the calculation reaching convergence with the CO still in the axial position. When setting up all chemisorption calculations, a slightly longer M−CO bond length was generated for the initial structure. The idea behind this was to avoid favoring the planar binding motif. In all other calculations, the M−CO bond length reduced throughout the bfgs cycles until the oxygen was close enough to the cluster surface that a switchover between the axial and planar geometries occurred. While this did not occur for the Pt79Ti122 cluster, with the oxygen remaining too far away from the Ti surface, it does demonstrate the significantly stronger binding exhibited by the CO in the planar conformation compared with the axial conformation.



REFERENCES

(1) Litster, S.; McLean, G. J. Power Sources 2004, 130, 61−76. (2) Wee, J. H.; Lee, K. Y.; Kim, S. H. J. Power Sources 2007, 165, 667− 677. (3) www.wolframalpha.com, accessed 20/02/2012. (4) Mani, P.; Srivastava, R.; Strasser, P. J. Power Sources 2011, 196, 666−673. (5) Lee, K.; Zhang, L.; Zang, J. Non-noble Electrocatalysts for the PEM Fuel Cell Oxygen Reduction Reaction. In PEM Fuel Cell Electrocatalysts and Catalyst Layers; Springer: London, 2008. (6) Othman, R.; Dicks, A. L.; Zhu, Z. Int. J. Hydrogen Energy 2012, 37, 357−372. (7) Gasteiger, H. A.; Panels, J. E.; Yan, S. G. J. Power Sources 2004, 127, 162−171. (8) Yoo, J. S.; Kim, H. T.; Joh, H. I.; Kim, H.; Moon, S. H. Int. J. Hydrogen Energy 2011, 36, 1930−1938. (9) Liu, G.; Zhang, H.; Zhai, Y.; Zhang, Y.; Xu, D.; Shao, Z. G. Electrochem. Commun. 2007, 9, 135−141. (10) Ferrando, R.; Jellinek, J.; Johnston, R. L. Chem. Rev. 2008, 108, 845−910. (11) Jennings, P. C.; Pollet, B. G.; Johnston, R. L. Phys. Chem. Chem. Phys. 2012, 14, 3134−3139. (12) Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G.; Ross, P. N.; Lucas, C. A.; Markovic, N. M. Science 2007, 315, 493−497. (13) Kitchin, J. R.; Nørskov, J. K.; Barteau, M. A.; Chen, J. G. J. Chem. Phys. 2004, 120, 10240−10246. (14) Nørskov, J. K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 937−943. (15) Stamenkovic, V.; Mun, B. S.; Mayrhofer, K. J.; Ross, P. N.; Markovic, N. M.; Rossmeisl, J.; Greeley, J.; Nørskov, J. K. Angew. Chem. 2006, 118, 2963−2967. (16) Paz-Borbon, L. O.; Johnston, R. L.; Barcaro, G.; Fortunelli, A. Eur. Phys. J. D 2009, 52, 131−134. (17) West, P. S.; Johnston, R. L.; Barcaro, G.; Fortunelli, A. J. Phys. Chem. C 2010, 114, 19678−19686. (18) Derosa, P. A.; Seminario, J. M.; Balbuena, P. B. J. Phys. Chem. A 2001, 105, 7917−7925. (19) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo, I.; Dal Corso, A.; Fabris, S.; Fratesi, G.; de Gironcoli, S.; Gebauer, R.; Gerstmann, U.; Gougoussis, C.; Kokalj, A.; Lazzeri, M.; Martin-Samos, L.; Marzari, N.; Mauri, F.; Mazzarello, R.; Paolini, S.; Pasquarello, A.; Paulatto, L.; Sbraccia, C.; Scandolo, S.; Sclauzero, G.; Seitsonen, A. P.; Smogunov, A.; Umari, P.; Wentzcovitch, R. M. J. Phys.: Condens. Matter 2009, 21, 395502−395521. (20) Ismail, R.; Johnston, R. L. Phys. Chem. Chem. Phys. 2010, 12, 8607−8619. (21) Núñez, S.; Johnston, R. L. J. Phys. Chem. C 2010, 114, 13255− 13266.



CONCLUSIONS From the results presented in this study, it was found that compositional and size effects can have a significant effect on the d-band properties of the Pt−Ti nanoparticles. Shifts are observed in the d-center when comparing pure and bimetallic clusters, with downshifts observed for the Pt-rich bimetallic clusters compared to the pure clusters. For Pt-rich clusters, this downshift has been previously noted in other systems such as the case of bulk Pt3Ti. However in our studies, when comparing the nanoparticles to the bulk system, this shift seems to vary with size and composition. We ascribe changes in the d-center to changes in electron density, with electron donation from Ti to Pt. This accounts for the downshift observed for the shell of the Pt-rich cluster with donation from the Ti core to the Pt shell, resulting in greater filling of the d-band. Furthermore, these changes in d-band characteristics were found to have an effect on the binding energies of OH and CO. It was also shown that where a downshift in d-center occurs, leading to more occupied orbitals, this results in weaker binding. Conversely, an upshift in d-center, suggesting more unoccupied orbitals, results in stronger binding. This suggests that while the bulk Pt3Ti system may results in a slightly weaker Pt−O bond than would be desired, nanoparticles may result in slightly stronger binding. From the calculations performed here, it is possible to see that while slightly stronger adsorption is observed compared to the bulk system, weaker binding is observed compared to the pure Pt clusters. This means that the nanoparticle Pt−Ti alloy system should present advantages over both the bulk Pt3Ti and pure Pt nanoparticle systems.



Article

AUTHOR INFORMATION

Corresponding Author

*Tel +44 (0) 121 414 7477; Fax +44 (0)121 414 4403; e-mail r.l. [email protected]. Notes

The authors declare no competing financial interest. 15249

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250

The Journal of Physical Chemistry C

Article

(22) Rossi, G.; Rapallo, A.; Mottet, C.; Fortunelli, A.; Baletto, F.; Ferrando, R. Phys. Rev. Lett. 2004, 93, 105503−105507. (23) Ferrando, R.; Fortunelli, A.; Rossi, G. Phys. Rev. B 2005, 72, 085449−085458. (24) Shao, X.; Wu, X.; Cai, W. J. Phys. Chem. A 2009, 114, 29−36. (25) di Paola, C.; Baletto, F. Phys. Chem. Chem. Phys. 2011, 13, 7701− 7707. (26) Chen, Y.; Liang, Z.; Yang, F.; Liu, Y.; Chen, S. J. Phys. Chem. C 2011, 115, 24073−24079. (27) Hammer, B.; Nørskov, J. K. Surf. Sci. 1995, 343, 211−220. (28) Pereira, L. G.; Paganin, V. A.; Ticianelli, E. A. Electrochim. Acta 2009, 54, 1992−1998. (29) www.bear.bham.ac.uk, accessed 29/02/2012.

15250

dx.doi.org/10.1021/jp303577t | J. Phys. Chem. C 2012, 116, 15241−15250