Category Archives: Student Post

Possible transition processes of Pt adatom on Pt(100) surface

by Hepeng Ye

Introduction:

Surface diffusion is a common process for solid materials, and in our case, the Pt adatom could diffuse along the surface of Pt(100). Diffusion is governed by the thermal energy from the environment and the property of such behavior is important for understanding surface phase formation, heterogeneous catalysis and other aspects.

In our case, Pt adatom is hypothesized to undergo a hopping between two adjacent four-fold sites or undergoes concerted substitution. To study the process, a model needs to be built and in our case the model is fairly simple since neither diffusion process has linear interpolation issue, and as consequence, one just need an initial and final two-state model for the transition state study. By connecting the initial and final state of the system, one could locate the TS (transition state) of that process and use the energy barrier from the transition search to further study the kinetic of process.

Experimental:

First cleave the Pt(100) lattice, and create the slab and a five-layer slab of 2X2 super-cell with 5Å vacuum space was set as the starting point. As mentioned the goal of the experiment is to identify the more likely transition process from the two, and in order to get reasonable comparison between the two processes, one need to first optimized the system(slab) to make sure that the super-cell is large enough for boundary conditions and enough vacuum space to prevent interaction from the next periodic slab along the Z-axis. Layers of the slab will not be optimized due to time limit, and based upon experience, five layers should be enough.

For vacuum space optimization, build three slabs with 5Å, 10Å and 15Å vacuum space, and calculate energy with CASTEP, functional used is GGA PBE, pseudo-potential is OTFG-ultrasoft, and all other parameters are set to default (272.1eV energy cutoff, 6x6x1 K-points, core radii is 1.27A with electron configuration to be 5s2 5p6 5d9 and 6s1).

Figure 1. Example of the slab used for vacuum space optimization. (5layer, with 10A vacuum)

Use the lowest energy vacuum space to continue slab size optimization.

To do this, one will first identify the most stable height of the Pt adatom on the surface. Sitting on top of the fourfold site, it is obvious that the adatom cannot be either too close or too far from the surface where too close to each other there will be repulsion and too far to each other will make Pt adatom isolated from the surface and could interact with the next periodic slab. Consider the short range dispersion, there should be a local energy minimum for the Pt to sit on top of the surface and bring down the system’s energy. So the height is tuned manually by putting the adatom between 1Å to 3Å above the center of the four-fold site.

When Pt adatom transit from one four-fold site to another, the energy for it being reactant should be the same as it being the product. Using this idea, I will separately calculate the energy for Pt adatom sitting at different fourfold sites, for different size of slabs.

The lowest energy height optimized previously will be used, and start with 3×3 super-cell.

Figure 2. Position of initial and final state of fourfold site.

 

 

As shown in the Figure 2, in 3×3 super-cell, the starting point of Pt, say, at the center, and could diffuse into the nearby fourfold sites. Due to symmetry, the energy barriers for all four direction should be the same, so just consider one of the pathways. In this case, energy of the system is calculated for Pt on the center site and on the edged site.

To get energies of each initial and final state, all calculations are performed with CASTEP, functional used is GGA PBE, pseudo-potential is OTFG-ultrasoft, and all other parameters are set to be the same as above when doing slab-size optimization.

After calculations are done, and comparing the two energies and check the relative difference, and based upon that difference, one will decide whether it necessary to keep using 3×3 super-cell or increase the slab size to 4×4 super-cell or even larger.

Then, in order to find out the transition state, one need to build the initial and final state or the reactant and product state respectively. There are two possible diffusion processes, one is the hopping between two fourfold sites, another is the concerted substitution of Pt adatom with Pt on the top layer. And in the calculations, two assumptions will be made to simplify the problem and reduce computational cost:

  1. In either process, Pt atom will only diffuse with the closest Pt atom or fourfold site.
  2. Due to the fact that system used has already optimized its size, and due to the symmetry, only consider Pt adatom diffuse in one direction, instead of four.

Based upon these assumptions, two sets of reactant and product slabs will be built for corresponding diffusion pathways.

 

Figure 3. Transition process scheme for hopping between fourfold states.

Figure 4. Transition process of concerted substitution of Pt adatom with Pt atom on the surface.

After determining the initial and final state of the process, DMol3 TS-search will be used to find the transition state/energy for each process while comparing transitional energy calculated with different search protocols.

Data analysis and discussion:

For the slab size optimization, calculations were performed for 3×3 super-cell where the Pt adatom sits in the fourfold site in the middle, and the fourfold site on the edge, and the energies are close enough to say they are in the same environment.

For the manually adjusted Pt adatom positions, three-point-adjust method is used to iteratively adjust and narrow down the height range and get to the lowest energy and its corresponding height/distance. One could see from the plot that at 1.8Å above the surface, the energy is the lowest.

Figure 5. Energy diagram of Pt adatom with relative heights to the slab surface.

After that, a 5 layer, 3×3 super slab is applied for the Transition-State study.

In Castep transition state search, there are some different synchronous transit methods[3], and here, some of these methods were used and were compared their differences. For the fourfold to fourfold hopping, the transition energies for different TS-search protocols are shown in the Table 1 below:

Table 1. Transition energy of hopping between fourfold sites calculated from different TS-search protocol.

For concerted substitution, the transition energy for each TS-search protocol is also listed in the table2.

Table 2. Transition energy of concerted substitution calculated from different TS-search protocol.

Different transition searching protocols have different calculation process, but for each process, they are all in agreement with the transition energies are on the same magnitude. While for Pt substitution process, the transition energy barrier is ten times larger than that of hopping process, which makes sense that such process requires breaking Pt metallic bonds with the surrounding Pt atoms and then another Pt atom comes in and forms new bonds, there should be a lot more energy required for such reaction.

Conclusion:

Based upon the energy barrier calculated from the Transition-Search, it is obvious that the concerted substitution has way larger energy barrier than the fourfold sites hopping, which means at same conditions(say same temperature) it will be much less likely for Pt adatom diffuse into the Pt(100) surface and undergoes the concerted substitution process.

There are some experimental setups that could be improved, say the slab I used has 5 layers, which I did not optimize. Also, I stopped at 3×3 supercell size and did not look at 4×4 supercell simply because the energy between two fourfold sites are very close to each other(~0.005 hartree for the total system) for 3×3 supercell, but maybe 4×4 will give even more identical energy, I do not know and maybe someone could try 4×4 disregarding the computational cost.

 

Reference:

  1. J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 77 (1996)

2. J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C. Fiolhais,      Atoms, Molecules, Solids, And Surfaces – Applications of the Generalized Gradient  Approximation for Exchange and Correlation, Phys. Rev. B, 46 (1992)

3. CASTEP GUIDE, Transition state search task. Date of access: April. 27. 2019

https://www.tcm.phy.cam.ac.uk/castep/documentation/WebHelp/content/modules/castep/tskcasteptss.htm

A brief introduction to the theory, parameterization and application of Projector Augmented Wave (PAW) datasets to plane wave DFT calculations

by Brandon Bocklund

Introduction

Many of the observable properties that makes materials interesting are linked to the electronic structure, so it is desireable with to perform DFT calculations with the most accurate computational methods that are affordable. Calculations within DFT scale as \( \mathcal{O}(n^3) \) for n electrons [1], so many technologically important materials, including transition metals such as Fe, are computationally demanding to calculate when considering all the electrons. This is exacerbated for core electrons, since the wave functions near the nuclei tend to oscillate, requring a large number of plane waves required to describe the wave functions near the nuclear cores in extended systems. Since many key properties are influenced by the outermost valance electrons that control bonding, it is sensible to fix the core electrons and treat them approximately to make calculations of the same level of theory more affordable or to make more complex calulations affordable. Treating core electrons as fixed is known as the frozen core approximation.

In this post, different cutoff radii will be used to generate PAW datasets for Ti. Differences in the generated PAW datasets will be demonstrated by comparing the cutoff energy required to converge the energies of hcp Ti and by comparing the formation energy of TiO2.

Treating core electrons

Core and valence electrons

Typically there are two main parameters to consider when treating electrons as core or valence that effect how the wave functions will be constructed: the core radius and the number of core vs. valence electrons. The core radius is the spherical distance from the center of the atom that forms the outer bound for the approximation to the wave functions. The number of electrons in the core vs. valance is simply the partition of how many electrons of charge density an atom can contribute to bonding. For example, Ti has 22 electrons that must be partitioned between core and valance. Both of these parameters are typically chosen by the person developing the new description.

A basic description of some methods for treating core electrons in extended systems follows.

Pseudopotentials

Pseudopotentials replace the core wave functions with a pseudo wave function with the goal of defining a smooth function to represent the effective potential of the core. The pseudo wave function are constructed so that the value and the derivative of the pseudo wave function matches the value and derivative of the true wave functions at the core radius. The advantage of pseudopotentials is their formal simplicity, however the pseudo wave function prevents the recovery of the full core electron wave functions. James Goff published a post on pseudopotentials [2] that describes their construction in a detailed, introductory way.

LAPW

Another method for treating core electrons is the Linear Augmented Plane Wave (LAPW) method. The main idea of this method is that two partial waves, which can be used to describe the all-electron wave functions, are joined at the core radius and matched to the all-electron solution. Thus, the main advantage is the core wave functions can be recovered, but this method comes at a higher computational cost than the pseudopotential method. A more complete description of the pseuodpotential and LAPW methods can be found in [3].

PAW

The Projector Augmented Wave (PAW) method presented by Blöchl [4] as a formalism that generalizes and combines favorable aspects of the pseudopotential and LAPW methods through projector functions. The basis of the PAW method is the combination of partial wave functions from the isolated atom with pseudo partial waves. The PAW method defines a transformation operator that maps the pseudo wave functions onto the all-electron wave functions,
\begin{equation} \left| \Psi \right> = \mathcal{T} \left| \tilde\Psi \right> \end{equation}
for the all-electron wave function, \( \left| \Psi \right> \), the transformation operator, \( \mathcal{T} \), and the pseudo wave functions \( \left| \tilde\Psi \right> \). The transformation operator is defined as
\begin{equation} \mathcal{T} = 1+ \sum_i \left( \left| \phi \right> – \left| \tilde\phi \right> \right) \left< \tilde p_i \right|\end{equation}
where \( \left| \phi \right> \) and \( \left| \tilde\phi \right> \) are the all-electron partial waves and pseudo partial waves, respectively, and \( \left< \tilde p_i \right| \) are the projectors that are constructed to match the all-electron solution. In the PAW method, PAW datasets are generated that can recover the full core electron wave functions, like the LAPW method, but at a lower cost than the LAPW method, closer to the computational cost of pseudopotentials. More information can be found for the PAW method in [4-6] and more detailed description of the relationship between pseudopotentials, APW, and the PAW methods can be found in [4].

Calculation details

Calculations are performed using plane waves and the PBE functional implemented by Quantum Espresso [8] in the pwscf code. PAW datasets [4] are used to describe the core states. For Ti, PAW datastets are generated as part of this work with the ld1 code distributed with Quantum Espresso, based on the PS Library 1.0 parameterizations [9]. The PS Library 1.0 parameterization for O is used directly. The reference states for Ti and O are chosen as hcp Ti and the O2 molecule in a 10x10x10 fcc box. The size of the O2 box was converged to within 0.1 mRy. Detailed discussion of the convergence of the box size is out scope for this post.

The cutoff energy convergence of the different potentials are compared in the following sections below for Ti and TiO2. Oxygen cutoff energy was converged to 0.1 mRy at a wavefunction cutoff energy of 90 Ry and a kinetic energy cutoff of 720 Ry. For the calculation of formation energy, the maximum converged cutoff energy for each set of potentials is used across all structures. In all cases, Ti required a larger cutoff energy for convergence than O, so it was used for the calculations of TiO2.

The kpoint grids were converged to within 0.1 mRy. For hcp Ti, an 11x11x7 Monkhorst-Pack grid was used. For the O2 box, a 3x3x3 Monkhorst-Pack grid, and for rutile TiO2 a 4x4x7 grid. Showing the detailed kpoints convergence tests are out of scope for this study. Even though the different PAW datasets can lead to different electronic structures and kpoints convergence, a single set of kpoints for each set of structures were used across all of the different PAW datasets for simplicity.

For comparing the cutoff energy of hcp Ti, calculations with each PAW potential used the same structure, a hexagonal primitive cell with a lattice parameter of 5.8 a.u. and a c/a ratio of 1.58 (corresponding to the experimental c/a ratio). All static calculations were performed using tetrahedron integration and an energy convergence of 1e-8 Ry. For calculations of the formation energy, the cells and ionic positions for all cells (O, Ti for each potential, TiO2 for each potential) were relaxed with a force convergence criteria of 1e-9 Ry/bohr. The Brillouin zone integration for relaxations was treated with Gaussian smearing with a smearing width of 0.001 Ry.

 

Generating PAW datasets for Ti

The ld1 program was used to generate the PAW datasets. The documentation can be found online [10]. The starting point for the tests here is the input file for generating a PAW dataset for the spin polarized, scalar relativistic [12] Ti from the PS Library version 1.0.0 [9]. The input file for the PS Library starting point follows below. The settings here create a PAW dataset for Ti with Troullier-Martins pseudization [11] via Bessel functions. The the key inputs that will be changed is `rcore`, which is the matching radius in atomic units for the core charge smoothing. The bottom section describes the number of wave functions (6) that will be pseudized and which are described by the lines below. Taking the first line (3S) as the example, the first number is the principle quantum number (n=1, starting with 1 for the lowest s, 2 for the lowest p), then the angular momentum quantum number (l=0), followed by the number of occupying electrons, the energy used to pseudize the state (=0.00), the matching radius for the norm-conserving pseudopotential (=0.85), the matching radius for the ultrasoft pseudopotential (=1.30) and finally the total angular momentum (=0.0). Changing the norm-conserving or ultrasoft pseudopotential matching radii do not change the PAW dataset, since the radii are controlled by the input parameters above. Note that the energies for the unbound states (4p=4 Ry and unoccupied 3d=0.05 Ry) need to be non-zero.

&input
title='Ti',
zed=22.,
rel=1,
config='[Ar] 4s2 4p0 3d2',
iswitch=3,
dft='PBE'
/
&inputp
lpaw=.true.,
pseudotype=3,
file_pseudopw='Ti-PSL.UPF',
author='ADC',
lloc=-1,
rcloc=1.6,
which_augfun='PSQ',
rmatch_augfun_nc=.true.,
nlcc=.true.,
new_core_ps=.true.,
rcore=0.8,
tm=.true.
/
6
3S 1 0 2.00 0.00 0.85 1.30 0.0
4S 2 0 2.00 0.00 0.85 1.30 0.0
3P 2 1 6.00 0.00 0.80 1.60 0.0
4P 3 1 0.00 4.00 0.80 1.60 0.0
3D 3 2 2.00 0.00 0.80 1.50 0.0
3D 3 2 0.00 0.05 0.80 1.50 0.0

To demonstrate how the cutoff radius affects the calculation accuracy and expense (through the cutoff energy), two new PAW datasets will be generated with core radii of 0.4 a.u. (much smaller, further referred to as “small core”) and 2.0 a.u. (much larger, further referred to as “large core”). Note that these are note expected to be accurate, but instead to demonstrate how the parameters affect might affect calculations and impress the care and detail that should be given to generating proper potentials.

Energy cutoff convergence of Ti and formation energy TiO2

Figure 1a and 1b compre the relative and absolute energy convergence with respect to the wave function energy cutoff for the differently generated Ti PAW datasets. In Figure 1a, the PS Library converges at a wave function cutoff energy of 120 Ry, the large core at 110 Ry, and the small core does not converge to less than 0.1 mRy/atom even at 140 Ry. Computational demands and the instability of the PAW dataset limited the full convergence.

The cutoff energies follow the trend that would be expected. Compared to the default PS Libary PAW dataset, increasing the core radius allows for more of the wave function to be described by the core, and therefore the dataset is softer (requires smaller energy cutoffs). In addition, the small core is not converged even at 140 Ry because the smaller radius requires a harder potential with more plane waves being required to describe the wave function. Figure 1a also shows unusual behavior for the energy to not decrease variationally with the cutoff energy and instead shows that there was some osciallation in the energy with increasing energy cutoff. Note that the large core and the the PS Library default are quite close in energy, so the differences may be hard to see in both figures. Figure 1b aims to demonstrate that the energies for different pseudopotentials are different even for the same calculation settings, so they cannot be compared.

Figure 1a: Energy cutoff convergence relative to the energy calculated at a cutoff energy of 140 Ry. Energy differences are taken from 10 Ry to 140 Ry. Calculations not shown are outside the bounds of the plot.

Figure 1b: Energy convergence for cutoff energies ranging from 10 Ry to 140 Ry. Missing results at lower energy cutoff are off the scale of the plot.

As discussed previously, energies between different pseudopotentials cannot be directly compared, therefore a metric that can cancel the differences is desired. The formation energy of TiO2 can cancel the differences caused by the different treatment of Ti and give some insight into the practical implications of different treatments of the electronic structure. Using the cutoff energies found for Ti (120 Ry for PS Library, 140 Ry for small core, 110 Ry for the large core) the energies per atom of hcp Ti, O2 in a vaccuum box, and TiO2 are calculated. The formation energy of TiO2 is calculated in the usual way:
\begin{equation} E^\mathrm{form} = E_{\mathrm{TiO_2}} – (E_{\mathrm{Ti}} + E_{\mathrm{O_2}}) \end{equation}
Since in each case, the cutoff energies are convered for all calculations within each formation energy calculation, the formation energies can be safely compared even with different cutoff energies used between them.

StructurePS LibrarySmall coreLarge core
hcp Ti (Ry/atom)-183.6910-183.6974-183.6975
O (Ry/atom)-41.4776-41.4776-41.4776
TiO2 (Ry/atom)-89.1361-89.1363-89.1357
TiO2 formation (Ry/atom)-0.2541-0.2521-0.2515

The total energies and formation energies are shown in Table 1. The differences in the formation energy are larger than the convergence criteria (0.1 mRy/atom, 0.0001 Ry/atom) for the formation energies of the PS Library to the small core PAW dataset and the PS Library to the large core PAW dataset. The largest difference is 0.0026 Ry/atom for the small core and PS Library formation energies, which corresponds to above 3 kJ/mole-atom. Differences that large could impact the relative stability of TiO2 in a more comprehensive study that used inaccurate PAW datasets.

Conclusion

PAW datasets were reviewed in context of other types of frozen core approximations, then PAW datasets were generated for Ti with larger and smaller core radii than the ones parameterized in the PS Library v1.0.0. It was shown that the different core radii can affect the energy cutoff convergence, the total energy, and the formation energy of a compund, TiO2.

References

[1] Section 10.5, D.S. Sholl, J.A. Steckel, Density Functional Theory, John Wiley & Sons, Inc., Hoboken, NJ, USA, 2009. doi:10.1002/9780470447710.
[2] J. Goff, Transferability of Cu Pseudopotentials in Cu/CuO Systems, Density Funct. Theory Pract. Course. (2019). https://sites.psu.edu/dftap/2019/03/31/transferability-of-cu-pseudopotentials-in-cucuo-systems/.
[3] D.J. Singh, L. Nordstrom, Planewaves, Pseudopotentials and the LAPW Method, Second Edi, Springer US, 2006. doi:10.1007/978-0-387-29684-5.
[4] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B. 50 (1994) 17953–17979. doi:10.1103/PhysRevB.50.17953.
[5] X. Gonze, F. Finocchi, Pseudopotentials Plane Waves-Projector Augmented Waves: A Primer, Phys. Scr. T109 (2004) 40. doi:10.1238/physica.topical.109a00040.
[6] R.M. Martin, Electronic Structure: Basic Theory and Practical Methods, Cambridge University Press, 2004. doi:10.1017/CBO9780511805769.
[7] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett. 77 (1996) 3865–3868. doi:10.1103/PhysRevLett.77.3865.
[8] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G.L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A.P. Seitsonen, A. Smogunov, P. Umari, R.M. Wentzcovitch, QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials., J. Phys. Condens. Matter. 21 (2009) 395502. doi:10.1088/0953-8984/21/39/395502.
[9] A. Dal Corso, Pseudopotentials periodic table: From H to Pu, Comput. Mater. Sci. 95 (2014) 337–350. doi:10.1016/j.commatsci.2014.07.043.
[10] http://web.mit.edu/espresso_v6.1/amd64_ubuntu1404/qe-6.1/Doc/INPUT_LD1.html
[11] N. Troullier, J.L. Martins, Efficient pseudopotentials for plane-wave calculations, Phys. Rev. B. 43 (1991) 1993–2006. doi:10.1103/PhysRevB.43.1993.
[12] Koelling, D. D. & Harmon, B. N. A technique for relativistic spin-polarised calculations. J. Phys. C Solid State
Phys. 10, 3107–3114 (1977).

Calculating transition state and activation energy of Ag diffusion via adjacent threefold sites on Cu(111) surface

Author: Stephen Holoviak

Introduction:

Diffusion of adatoms on a surface is an important phenomena in many processes; from thin film deposition (1) to catalysis. This study is an attempt to understand the activation energy of an Ag atom diffusing between the adjacent threefold sites on the Cu(111) surface.

Figure 1: Threefold sites on Cu(111) surface.

Method:

All calculations were carried out in the Dmol3(2) package in materials studio.

Exchange-correlation functional used was LDA-PWC (3).

Kpoint density: 1x1x1, all meshes were distributed via the Monkhorst-pack method (4). If more computational resources were available, more k-points would be added.

SCF Tolerance: 10E-4 [eV]

Geometry Optimization Tolerance: 10E-4[Ha] = 0.0272[eV]

Orbital Cutoff: 3.4[Å] with an all electron core treatment and a double numeric (DN) basis set (5).

Convergence Testing:

The k-point mesh and SCF tolerance used in these calculations lead to unstable results that do not provide very accurate energies. Unfortunately, the computational resources available for this project were extremely limited and in order for the calculations to reach any result at all these extremely coarse parameters had to be used. The instability was confirmed by comparing the calculations against partial results with a SCF tolerance of 10E-6[eV] and a 4x4x1 k-point mesh. The system energies between the coarse calculations and the finer calculations varied by nearly 0.19[eV].

Calculations:

The reactant and product states of the reaction were defined as the hcp and fcc sites respectively.

A 3-layered 1×1 slab of Cu(111) with a vacuum spacing of 10[Å] was used for the calculations. An Ag atom was placed at the hcp site and the geometry of the Ag and all Cu layers were optimized in order to define the reactant state. An Ag atom was placed at the fcc site and the geometry of the system was optimized in order to define the product state. With more computational resources, this cell should probably be increased in size, with more unit cells and more layers, fixing the geometry of the bottom layers to represent the bulk material.

Figure 2: Reactant and product vacuum slabs used for transition state searching.

The transition state of the reaction was found using the complete LST/QST method as implemented in Dmol3 (6).

The energy of each image of the reaction was defined as follows:

E = E_{Ag|Cu(111)} - E_{Cu(111)} - E_{Ag}

Figure 3: Complete LST/QST transition state search.

The transition state was found to be the bridge site between the two threefold sites. While the exact numbers of the frequencies and energy at the transition state will not be very accurate, there was one imaginary frequency along the reaction coordinate that supports its identity as a transition state.

Figure 4: Reactant state, transition state, and product state structures.

The energy barrier found for the hcp to fcc transition was 0.051[eV], but the level of convergence for these calculations is on the same order of magnitude as this barrier, so the exact barrier cannot be determined from this study. The height of this barrier reported in literature is 0.023[eV] (7), giving the same order of magnitude as this study.

The energy of the two threefold sites was found to be slightly different, with the fcc site being 0.003[eV] more stable, however the calculations from this study cannot tell which one is actually more stable.

Conclusion:

Due to computational resource limitations, an exact value for the activation energy of Ag diffusing between adjacent threefold sites on Cu(111) could not be found, however the order of magnitude was determined to be 10E-3[eV]. The location of the transition state was found to be the bridge site in between the threefold sites, this location is supported by the presence of an imaginary frequency along the reaction coordinate. It was also found that the energy of the threefold sites is different, due to the presence(hcp) or absence(fcc) of a Cu atom directly below the site in the next layer, but it could not be determined which of these sites is more energetically favorable.

References:

(1)     M. Ohring (2001) “Materials Science of thin Films”, Section 7.4: “Kinetic Processes in Nucleation and Growth”, 386-400

(2)     B. Delley (2000) “From molecules to solids with the DMol3 approach” J. Chem. Phys. 113

(3)     J. Perdew, Y. Wang (1992). “Accurate and simple analytic representation of the electron-gas correlation energy”. Phys. Rev. B. 45 (23): 13244–13249.

(4)     H. Monkhorst, J. Pack (1976) “Special points for Brillouin-zone integrations.” Physical Review B, 13(12): 5188-5192

(5)     B. Delley (1990). “An All-Electron Numerical Method for Solving the Local Density Functional for Polyatomic Molecules”. J. Chem. Phys. 92: 508–517.

(6)     T. Halgren, W. Lipscomb (1977) “The synchronous-transit method for determining reaction pathways and locating molecular transition states.” Chem. Phys. Letters 49: 225-232

(7)     A. Kotri, E. El Koraychy, M. Mazroui, Y. Boughaleb (2017) “Static investigation of adsorption and hetero‐diffusion of copper, silver, and gold adatoms on the (111) surface” Surface and Interface Analysis, 49 (8)

Determination of the Activation Energy and Frequencies for the Diffusion of Ag on Cu(111) via Three-fold Hollow Sites

by Angela Nguyen

1. Introduction

The purpose of this project is to calculate the activation barrier for the diffusion of a Ag atom on a Cu(111) surface via adjacent three-fold hollow sites through the use of DFT with a plane wave basis set as implemented by the CAmbridge Serial Total Energy Package (CASTEP) module [1] in Materials Studio. To accomplish this task, the transition state for the diffusion must be determined and then verified via a vibrational frequency calculation and visual observation of the transition state. The vibrational frequency calculation was implemented using the Dmol module [2] in Materials Studio. The Cu(111) surface was chosen as it is the most stable surface of Cu. For the purpose of this project, the Ag atom will be diffusing from the fcc hollow site to the hcp hollow site on the Cu(111) surface.  From the DFT calculations, the transition state was determined to be where the Ag atom is crossing the bridge site on the Cu(111) surface. The activation barrier was then calculated to be 0.31 eV with an imaginary vibrational frequency at -31.37 \(cm^{-1}\).

2. DFT Parameters

Listed below are the parameters used for the optimization of the Cu(111) surface as implemented in CASTEP.

Exchange-Correlation Functional TypeGeneralized Gradient Approximations [3]
Exchange-Correlation FunctionalPerdew Burke Ernzerhof [4]
Relativistic TreatmentKoelling-Harmon [5]
Psuedopotential"On the fly" generated (OTFG) ultrasoft [6]
K point GridMonkhorst-Pack [7]
SpinUnrestricted
Energy Convergence2.0 x 10^-5 eV/atom
Max Force0.05 eV/A

For Cu, the cutoff radius for the above psuedopotential is 2.0 Bohr (1.06 Å) with 11 valence electrons in the following configuration \(3d^{10} 4s^{1}\). For Au, the cutoff radius for the above pseudopotential is 1.59 Bohr (0.84 Å) with 19 valence electrons in the following configuration \(4s^{2} 4p^{6} 4d^{10} 5s^{1}\).

Below is a list of parameters for the vibrational calculations of the transition state as implemented in Dmol.

Exchange-Correlation Functional TypeGeneralized Gradient Approximations [3]
Exchange-Correlation FunctionalPerdew Burke Ernzerhof [4]
Relativistic TreatmentKoelling-Harmon [5]
Core TreatmentAll Electron
K point GridMonkhorst-Pack [7]
Basis SetDouble Numerical plus d-functions (DND) v3.5

3. Methodology

3.1 Construction of Cu(111) Vacuum Slab and Convergence Check

A vacuum slab of Cu(111) consisting of 5 layers was constructed from bulk Cu with a vacuum spacing of 12 Å between each slab in which the bottom 3 layers of the slab was constrained to model the bulk. The construction of 5 layers for the vacuum slab was chosen to ensure that the bulk could be modeled by the bottom 3 layers. A more rigorous study in the future could be performed to determine whether varying the number of layers in the slab affects the activation energy. A vacuum spacing of 12 Å was used to ensure that the slabs were isolated from each other due to the periodic nature of DFT. Single point energy calculations were performed on the vacuum slab to ensure convergence with respect to k points and the cutoff energy. A kpoint mesh of 2x2x1 (2 irreducible kpoints) and an energy cutoff of 400 eV was chosen to be sufficient for convergence of future calculations. The Cu(111) surface was then geometrically optimized using the chosen kpoint grid and cutoff energy.

Figures 1 and 2 Figure 2 display plots of the cohesive energy per atom as a function of the number of irreducible k points and cutoff energy respectively as determined by Materials Studio. The black dotted lines represent the +/- 0.01 tolerance used to determine which kpoint grid and energy cutoff to use.

Figure 1. Cohesive energy (eV/atom) vs. Number of irreducible k points. The blue circles represent the data points obtained from DFT calculations in Materials Studio. The blue line is to help guide the eye across the plot. The black dashed lines represent the +/- 0.01 eV tolerance used to determine the optimal number of irreducible k point mesh. The red square represents optimal k point mesh that will be used in future calculations.

Figure 2. Cohesive energy (eV/atom) vs. Energy cutoff (eV). The blue circles represent the data points obtained from DFT calculations in Materials Studio. The blue line is to help guide the eye across the plot. The black dashed lines represent the +/- 0.01 eV tolerance used to determine the optimal energy cutoff. The red square represents optimal energy cutoff that will be used in future calculations.

Two different versions of the optimized Cu(111) surface was constructed with the Ag atom in either the fcc hollow site or the hcp hollow site. These will represent the initial and final states for the following transition state search respectively.

3.2 Transition State Search and Vibrational Frequency Calculation 

With the initial and final states for the diffusion of Ag on Cu(111) built, a linear interpolation of 10 images between the two states was performed. A transition search was then conducted using the above outlined CASTEP parameters using the “Complete LST/QST” algorithm [8] in Materials Studio. An option was chosen in Materials Studio that the initial and final states should be optimized before any transition state search calculation was performed. This was done because linear interpolation between optimized initial and final states failed. Once the transition state was found, a vibrational frequency calculation was performed on the transition state to determine if the transition state was reasonable or not. If it is reasonable, the vibrational calculation should result in one imaginary (negative) frequency in which Ag atom is moving along the reaction coordinate towards the hcp hollow site. In addition, visual inspection of the transition state was done to make sure it was reasonable.

4. Results

Figures 3 and 4 below depict the optimized initial and final states of Ag adsorbed onto Cu(111) on the fcc and hcp hollow site respectively.

Figure 3. Ag adsorbed onto the fcc hollow site of Cu(111). The light blue sphere represents Ag and the orange spheres represents Cu.

Figure 4. Ag adsorbed onto the hcp hollow site of Cu(111). The light blue sphere represents Ag and the orange spheres represent Cu.

Figure 5 shows the transition state for the diffusion of Ag on Cu(111) via adjacent threefold hollow sites. From the figure, it can be seen that the transition state for the diffusion of Ag across Cu(111) is located along the reaction coordinate where the Ag atom is crossing the bridge site. It is reasonable for the transition state to be located here, since the Ag atoms will have to maneuver over two Cu atoms before settling in the adjacent hcp hollow site. The activation energy was determined to be 0.31 eV.

Figure 5. Transition state for the diffusion of Ag on Cu(111). The Ag atom is located along the bridge site of the surface. The light blue sphere is Ag and the orange spheres are Cu.

Lastly, in order to confirm the above transition state, a vibrational calculation was performed, which resulted in an imaginary frequency at -31.37 \(cm^{-1}\). To verify that this was a plausible vibrational frequency for the transition state, a visualizer in Materials Studio was used to view this vibration, which can be seen in Figure 6. From the figure, it can be seen that the Ag atoms is moving along the reaction coordinate towards the hcp hollow site while crossing the bridge site. This confirms that a proper transition state was found for the diffusion of Ag on Cu(111).

The table below shows the activation energy and imaginary frequency as calculated in this post.

Activation Energy [eV]Frequency [\(cm^{-1}\)]
"Theoretical"0.31-31.37
Literature [9]0.023

5. Conclusions

For the diffusion of Ag on Cu(111) via adjacent hollow sites, it was determined that the transition state is located at the bridge site with an activation energy of 0.31 eV. This was further confirmed through visual observation and by checking for one vibrational frequency for the transition state. One imaginary vibrational frequency was located at -31.31 \(cm^{-1}\) corresponding to the Ag atoms crossing the bridge site towards the adjacent hollow site.

From literature, Kotri et al. stuided the hetero-diffusion of metal atoms on other metal surfaces via adjacent hollow sites [9]. One such surface that they tested was the hopping of Ag on Cu(111) using Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) [10]. To determine the activation energy, Kotri et al. used the drag method in which the adatom is dragged from its inital site to the nearest empty site [9]. At each step, the adatom is then allowed to relax with the respect to the plane perpendicular to its motion. In their work, it was determined that the activation energy for Ag on Cu(111) was 0.023 eV, an order of magnitude less than the activation energy determined by this project. Interesting to note is that the activation barrier for the Ag atom hopping from the fcc hollow site to the hcp hollow site and vice versa was the same. For this project, the activation energy varied by 0.08 eV depending on which path the Ag atom took.

Reasons for the deviation between activation energies can be due to the difference in methodologies on how the transition state was obtained, how much “optimized” the initial and final states were, and the number of kpoints sampled. The “Complete LST/QST” algorithm could not be the most accurate method to use to find the transition state as it only optimizes the image with the highest energy, while leaving the other images untouched. Even though the transition state was “confirmed” to be correct through vibrational frequency analysis, it would be useful to run another transition state search method to verify if the current activation energy is correct. One such algorithm to use the nudged elastic band (NEB) method in which all images are optimized rather than the image with the maximum amount of energy in the “Complete LST/QST” method. When viewing the optimized inital and final images (Figures 3 and 4), it can be seen that the Ag atom is not located in the middle of each hollow site, where the highest symmetry of the site is. This could lead to higher energy values calculated as this would be the Ag would have to diffuse a longer distance between the fcc and hcp hollow sites. Lastly, a number of kpoints sampled in the Brillouin zone could be increased to sample any missed phenomenon in the system that was not detected before.

6. References

[1] Clark Stewart J et al., “First principles methods using CASTEP ,” Zeitschrift für Kristallographie – Crystalline Materials , vol. 220. p. 567, 2005.
[2] B. Delley, J. Chem. Phys. 92, 508 (1990).
[3] J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made Simple,” Phys. Rev. Lett., vol. 77, no. 18, pp. 3865–3868, Oct. 1996.
[4] J. P. Perdew et al., “Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation,” Phys. Rev. B, vol. 46, no. 11, pp. 6671–6687, Sep. 1992.
[5] D. D. Koelling and B. N. Harmon, “A technique for relativistic spin-polarised calculations,” J. Phys. C Solid State Phys., vol. 10, no. 16, pp. 3107–3114, Aug. 1977.
[6] D. Vanderbilt, “Soft self-consistent pseudopotentials in a generalized eigenvalue formalism,” Phys. Rev. B, vol. 41, no. 11, pp. 7892–7895, Apr. 1990.
[7] H. J. Monkhorst and J. D. Pack, “Special points for Brillouin-zone integrations,” Phys. Rev. B, vol. 13, no. 12, pp. 5188–5192, Jun. 1976.
[8] Halgren, T. A.; Lipscomb, W. N. Chem. Phys. Lett., 49, 225 (1977).
[9] A. Kotri, E. El koraychy, M. Mazroui, and Y. Boughaleb, “Static investigation of adsorption and hetero-diffusion of copper, silver, and gold adatoms on the (111) surface,” Surf. Interface Anal., vol. 49, no. 8, pp. 705–711, Aug. 2017.
[10] LAMMPS User Manual 30, Sandia National Laboratories (Oct 2014) 2.

Calculation of Activation Energy and Frequencies for Ag Diffusion between Adjacent Threefold Sites on Cu (111) Surface.

Author : Shyam Deo

Introduction

In order to understand formation of nanostructures, it is necessary to understand the underlying microscopic processes, such as adsorption, desorption, and diffusion. In the current study, we focus on the dynamics of a single adatom of silver (Ag) for diffusion between adjacent three-fold sites on Cu (111) surface. The activation energy for diffusion has also been calculated along with the frequencies of diffusion.

Methods

Plane-wave Density Functional Theory (DFT) calculations were carried out using the Vienna Ab Initio Simulation Package (VASP), version 5.4.4. The electron-electron exchange and correlation energies were computed using the Perdew, Burke, and Ernzerhof functional (PBE) [1]. The projector augmented-wave (PAW) [2] method was used to represent the ion-core electron interactions. For Cu, the cutoff radius for the above psuedopotential is 2.3 Bohr (1.22 Å) with 11 valence electrons in the following configuration 3d10 4s1. For Ag, the cutoff radius for the above pseudopotential is 2.4 Bohr (1.27 Å) with 11 valence electrons in the following configuration 4d10 5s1. The structural convergence criteria were 0.05 eV Å-1 for all unconstrained atoms, while the convergence criteria defining self-consistency of the electron density was 10-5 eV.

Figure 1: (a) Cu (111) 3×3 unit cell with 5 layers and vacuum. (b) Convergence test for k-points. (c) Convergence test for ENCUT

A Cu FCC experimental bulk lattice constant of 3.615 Å was used for building Cu (111) surface [3]. A five layered 3 x 3 unit slab of Cu (111) was used for surface calculations where the bottom three layers were frozen during the optimization and the top two layers were unconstrained (Figure 1a). A plane wave energy cutoff of 450 eV and Monkhorst-Pack [4] k-point mesh of 5 x 5 x 1 was used for all surface slab calculations after proper convergence tests described in the next section. To minimize spurious interslab dipole interactions between the periodic slabs, a vacuum space of 15 Å was used and dipole corrections were added in the direction perpendicular to the surface. Transition state was located using the climbing image nudged elastic band (CI-NEB) method [5] (5 images) by relaxing the reaction tangent force below 0.05 eV Å-1.Transition state was verified to contain a single imaginary frequency along the reaction coordinate. These vibration frequencies were calculated by the tag IBRION = 7 implemented in VASP. It determines the Hessian matrix (matrix of the second derivatives of the energy with respect to the atomic positions) using density functional perturbation theory [6-7].

Convergence Tests 

Convergence test for K-points mesh sampling the brilliouin zone was done at a plane wave energy cut off of 550 eV by varying the k-points against the single point energy of Cu (111) surface. The relative energies are plotted against the number of irreducible k-points as shown in Figure 1b. The energy varied by an amount 0.006 eV between 13 and 24 irreducible K-points. However, the energy fluctuated when the K-points were raised further. To reduce the computation cost, lower K-points were chosen due to limited availability of computational resources which should be a reasonable starting point because we are mostly concerned with a difference in energies between different systems and not the absolute numbers themselves. Then for ENCUT convergence, the cut-off energy for plane wave was varied from 350 eV to 550 eV as plotted in Figure 1c. The energy changed only by 0.017 eV between 450 eV and 550 eV. Finally, the k-points mesh of 5x5x1 with 13 irreducible K-points and a plane wave energy cut off of 450 eV was considered suitable for all subsequent surface calculations of adsorption energies.

Results and Discussions

Adsorption of Ag Atoms over Cu (111)

A Cu (111) surface has two types of three-fold binding sites available for adsorption by a silver atom – hollow-fcc and hollow-hcp sites shown in Figure 2a. The figure shows adjacent hollow-fcc and hollow-hcp sites. A hollow fcc site is a three fold site with no atom directly below  but with an atom in the layer next to it. A hollow hcp site is a three fold site with an atom directly below  and  with no atom in the layer next to it. We probed the adsorption of silver atom over 3×3 Cu (111) unit cell over these two adjacent sites.

The adsorption energy for Ag was calculated as:

Eads,Ag = EAg* – EAg – E*

where Eads,Ag is the adsorption energy of Ag, EAg* is the energy of Ag bound to the surface, EAg is the energy of single Ag atom and E* is the energy of the bare Cu(111) surface. The adsorption energies for the two adjacent three-fold sites are labelled in Figure 2b. Both sites show the same adsorption energy of 2.38 eV (Figure 2b and 2c). These two states obtained were then studied for diffusion between the two adjacent sites and the activation energy was probed for the same.

Figure 2: a) Cu (111) surface showing two adjacent hollow-fcc and hollow-hcp sites of adsorption, b) Ag atom adsorbed on a hollow-fcc site on Cu (111) and c) Ag atom adsorbed on a hollow-hcp site on Cu (111). Adsorption energies for the two states of adsorption are also shown for comparison.

Activation Energy for Diffusion

The two states with adsorption of Ag over hollow-fcc site and hollow-hcp site was used as two end points for NEB calculation. The former was used as the initial state while the later was taken as the final state after diffusion from the hollow-fcc site to the hollow-hcp site. The intermediate images for NEB run were obtained using linear interpolation to give five images along the straight line trajectory for diffusion between the two adjacent sites. The energy profile for images along the reaction coordinate obtained from BEB run is shown in figure 3. The transition state is also shown. The activation energy obtained is 0.0328 eV for diffusion from hollow fcc site to hollow hcp site while the activation energy for reverse diffusion obtained is 0.0305 eV. A bridge site adsorption of Ag was obtained as the transition state (one with the highest energy). This was confirmed for single imaginary frequency (with value -1.89 THz) along the reaction coordinate (from hollow-fcc to hollow-hcp and back vibration). The other two vibration frequencies obtained were 4.12 THz (vibration towards the plane of surface and away) and 1.01 THz (vibration parallel to the plane of adsorption). A single imaginary frequency along the reaction coordinate confirms the transition state for diffusion from hollow-fcc to the adjacent hollow-hcp site. All vibrational frequencies are reported in Figure 4 (a).

Figure 3: Energy profiles computed using NEB (5 images) for diffusion of Ag atom from hollow-fcc site to hollow-hcp site of adsorption. The TS state is labelled and the transition state structure is shown.

The frequencies of hopping was then calculated by the following expression as a function of temperature (Figure 4) where ΔE is the activation energy for hopping, ν1 to νN are  the vibrational frequencies associated with the minimum initial state (hollow-fcc site adsorption) and ν†1 to ν†N-1 are  the real vibrational frequencies associated with the transition state. The rate of hopping has been plotted in Figure 4b as a function of temperature.

Figure 4: (a) Vibrational frequencies in THz for Ag atom adsorption over hollow fcc site and transition state (bridge site)  on Cu (111) and (b) hopping rate for Ag atom on Cu (111) as a function of temperature (K)

Conclusion

NEB calculations for diffusion from the hollow-fcc site adsorption of Ag to the adjacent hollow-hcp site was studied and an activation barrier of 0.0328 eV was obtained. These two sites showed same adsorption energies of 2.38 eV. As a further study, diffusion coefficient can be calculated by ab initio methods from the calculated activation energies and adsorption energies as done earlier by Minkowski et al. for Cu adatoms over Cu (111) surfaces [8]. The calculated adsorption energies and activation barrier for Ag adatom matches reasonably with the reported values of 2.43 eV and 0.023 eV, respectively [9]. The coefficients for surface diffusion and rate for hopping can help understand aggregation of atoms on surfaces to form clustered particles of metals on the catalyst support.

References

  1. Perdew, J.P. and Y. Wang, Accurate and simple analytic representation of the electron-gas correlation energy. Physical Review B, 1992. 45(23): p. 13244-13249.
  2. Blöchl, P.E., Projector augmented-wave method. Physical Review B, 1994. 50(24): p. 17953-17979.
  3. Straumanis, M.E. and L.S. Yu, Lattice parameters, densities, expansion coefficients and perfection of structure of Cu and of Cu-In [alpha] phase. Acta Crystallographica Section A, 1969. 25(6): p. 676-682.
  4. Monkhorst, H.J. and J.D. Pack, Special points for Brillouin-zone integrations. Physical Review B, 1976. 13(12): p. 5188-5192.
  5. Nudged-elastic band used to find reaction coordinates based on the free energy. The Journal of Chemical Physics, 2014. 140(7): p. 074109.
  6. S. Baroni, P. Giannozzi and A. Testa, Phys. Rev. Lett. 58, 1861 (1987).
  7. X. Gonze, Phys. Rev. A 52, 1086 (1995); X. Gonze, Phys. Rev. A 52, 1096 (1995).
  8. Mińkowski, M. and M.A. Załuska-Kotur, Diffusion of Cu adatoms and dimers on Cu(111) and Ag(111) surfaces. Surface Science, 2015. 642: p. 22-32.
  9. Kotri, A.El koraychy, E.Mazroui, M., and Boughaleb, Y. ( 2017Static investigation of adsorption and hetero‐diffusion of copper, silver, and gold adatoms on the (111) surfaceSurf. Interface Anal.49705– 711. doi: 10.1002/sia.6211.

 

Evaluating activation barrier of Ag hopping on Cu(111)

Author: Sharad Maheshwari

Introduction

Surface diffusion can play an important role in creating a topological modification on the surface as in the case of nanoparticle growth. In the following work we aim to identify the kinetic barrier for the diffusion of a Silver (Ag) atom as it hops from a three-fold fcc site to the adjacent three-fold hcp site on (111) surface of Copper (Cu). We will use density functional theory-based methods to evaluate the activation barrier for the hopping process. We will also calculate the frequency of this hopping process.

Calculation Details

Electronic Calculations

Electronic structure calculations were performed using the Vienna Ab initio Simulation Package (VASP)[1, 2] a plane wave basis set pseudo-potential code. We used the projector augmented wave (PAW) [3] method for core-valence treatment. The exchange and correlation energies were calculated using the Perdew, Burke, and Ernzerhof (PBE) [4] functional described within the generalized gradient approximation (GGA) [5]. A plane-wave basis set cutoff energy of 450 eV was used. The ionic convergence limit was set to 0.01 eV Å-1 while the electronic convergence limit was set to 10-5 eV. The Fermi level was smeared with the Methfessel-Paxton [6] scheme using a smearing width (σ) of 0.2 eV.

Surface Model

A 3 x 3 surface slab model was used to construct periodic surfaces of Cu (111) using an experimental bulk lattice constant of 2.556 Å. The slab models were comprised of 4 layers of metal atoms.  The top two layers of the slab along with the adsorbate were allowed to relax until the force on atoms in these layers was within 0.01 eV Å-1 while the bottom two layers were kept fixed to imitate their bulk arrangement.  A vacuum region of 10 Å was inserted in the models to exclude periodic interaction between the slabs. Dipole corrections were also added in the direction normal to the surface.

The sampling of the Brillouin zone for all 3 x 3 surface cells was conducted with a k-point mesh of 5 x5 x 1 generated automatically using the Monkhorst Pack method[7]

Transition States

To locate the transition state, a series of 5 linearly interpolated images were formed between the reactant and the product state. The nudged elastic band method (NEB) [8] was used to locate the transition state. Once an approximate transition state was obtained, it was further refined using the climbing image nudged elastic band method (CI-NEB) [9]. Vibrational frequency calculations were then performed using IBRION = 7 in VASP to confirm the transition state found the first order saddle point. It was further ensured that the single imaginary frequency obtained was along the reaction pathway.

Results

Figure 1 illustrates the reactant, transition and product state for the hopping of Ag atom from the fcc to hcp site of Cu(111) surface. The activation barrier for the hopping process is 0.035 eV.

Figure 1. Top and Side view of the a) reactant b) transition c) product states for the Ag hopping from fcc site to the hcp site on Cu (111)

For the barrier evaluated using the CI-NEB method, the reactant and the product states were used to create a series of images. The energy profile obtained from the calculation for the images is shown below in Figure 2. (All energies are referenced to the reactant state and does not include zero- point vibrational energy corrections).

Figure 2. Reaction energy diagram for the hopping of Ag atom from fcc to hcp site on Cu(111)

One important thing that was observed in a trial calculation is that the NEB method missed the transition state when 4 linearly interpolated intermediate images were used as shown in Figure 3. Using small even number of images can result in missing the saddle point and thus if using the NEB method, a higher density of images are needed near the saddle point. This highlights the need for caution when evaluating the activation barrier using the NEB method and using linear interpolation to obtain the images.

Figure 3. Reaction energy diagram obtained from the calculations using NEB method with 4 images obtained through linear interpolation

On an absolute scale, the zero-point vibrational energy (ZPVE) corrections were found to be very small. However, as the activation barrier itself is very small, we corrected the activation barrier by including the ZPVE corrections to the initial and the transition state. The corrected activation barrier after including the ZPVE corrections is 0.027 eV.

To evaluate the frequency of this hopping, we use the following equation:

  

where υi and υi†  are the frequencies evaluated for reactant and transition state respectively, kB is the Boltzmann’s constant, ΔEact is the evaluated activation barrier and T is the temperature in Kelvin.  As the transition state is first-order saddle point, it is energy minimum in all direction except for the reaction coordinate, along which it is maximum. Figure 4 shows the hopping frequency of Ag atom from fcc to hcp site on Cu(111) surface at different temperatures.

Figure 4. Frequency of hopping of Ag atom from fcc to hcp site on Cu(111) surface as a function of temperature

Conclusion

We used DFT methods to estimate the kinetic barrier for the hopping of Ag atom from the fcc to hcp site on Cu(111) surface. The activation barrier for this process after including the ZPVE corrections is found to be 0.027 eV. The value found is in reasonable agreement with the reported values [10]. This work illustrates the DFT methods can be used to evaluate kinetic parameters of the process. It also highlights that while using methods to evaluate the activation barrier, one should be careful and must ensure that the transition state found is the correct transition state. Vibrational calculations must be performed and a single imaginary frequency along the reaction coordinate must be confirmed to ensure that the transition state found is indeed the transition state we were looking for.

 

References

[1] G. Kresse, J. Furthmuller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci., 6 (1996) 15-50.

[2] G. Kresse, J. Furthmuller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B, 54 (1996) 11169-11186.

[3] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B, 59 (1999) 1758-1775.

[4] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 77 (1996) 3865-3868.

[5] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C. Fiolhais, Atoms, Molecules, Solids, And Surfaces – Applications of the Generalized Gradient Approximation for Exchange and Correlation, Phys. Rev. B, 46 (1992) 6671-6687.

[6] M. Methfessel, A.T. Paxton, High-precision sampling for Brillouin-zone integration in metals, Physical Review B, 40 (1989) 3616-3621.

[7] H.J. Monkhorst, J.D. Pack, Special Points for Brillouin-Zone Integrations, Phys. Rev. B, 13 (1976) 5188-5192.

[8] G. Henkelman, H. Jónsson, Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points, The Journal of Chemical Physics, 113 (2000) 9978-9985.

[9]  G. Henkelman, B.P. Uberuaga, H. Jónsson, A climbing image nudged elastic band method for finding saddle points and minimum energy paths, The Journal of Chemical Physics, 113 (2000) 9901-9904.

[10] Kotri, E. El koraychy, M. Mazroui, Y. Boughaleb, Static investigation of adsorption and hetero-diffusion of copper, silver, and gold adatoms on the (111) surface, Surface and Interface Analysis, 49 (2017) 705-711.

Study of O-H bond stretching frequency at different adsorption environments of OH adsorbed on Pt(111)

Author: Hoang (Bolton) Tran

Problem Statement

Previously (Post 2), the energetic and tilt angle of hydroxyl (OH) adsorption on Pt (111) was studied. In this study (Post 3), the OH bond stretching vibration \((\nu_{OH})\) is of primary interest. Firstly, the zero-point energy (ZPE) will be considered for adsorption energy. Secondly, the adsorption energy \((E_{ad})\) and \(\nu_{OH}\) will be computed and analyzed at different OH adsorption environments:

  • Adsorption sites: Atop vs. bridge (see Fig. 1 in Post 2)
  • Surface coverage: 1.0 ML vs. 0.25 ML (Monolayer – number of adatom per surface atom)
  • Water solvation: introduce water molecules above the adsorbed molecule.

Methodology

Software:

DFT was implemented in both CASTEP and DMol3, a plane wave basis set software embedded inside BIOVIA Material Studio (MS). [1]

Calculation:

Geometric optimization (CASTEP): Please see Post 2

Vibrational analysis (DMol3):

  • Exchange-Correlation Functional: Generalized Gradient approximations – Perdew-Burke-Ernzerhof (GGA-PBE) [2]
  • Self-Consistent Field tolerance: 1E-05 eV
  • Monkhost-Pack [3] grid size: 10x10x1 (50 irreducible kpoints) for 1×1 supercell & 5x5x1 (13 irreducible kpoints) for 2×2 supercell.
  • Self-consistent dipole correction in z-direction (normal to surface)

The procedure of optimizing, cleaving and adsorbing OH onto Pt(111) surface are the same as outlined in Post 2. This procedure was carried out again for each adsorption environments outlined below. The optimized structure of each adsorption environment was then used as input for the vibrational frequency calculation.

ZPE correction:

The adsorption energy is defined as:

\begin{equation} E_{ad}=E_{OH/slab}-E_{OH}-E_{slab} \end{equation}

To calculate the ZPE correction for \(E_{ad}\), ZPE was calculated for both the free OH radical and the adsorbed OH. Since the free OH only has O-H stretching vibration, the ZPE of adsorbed OH also considered only the O-H stretching \(\nu_{OH}\).

Adsorption environments:

1.  Adsorption sites:

The two adsorption sites of interest are the high symmetry atop and bridge sites. Values of \(E_{ad}\) at 1.0 ML coverage were already calculated from previous study. See Post 2 for more details.

2. Surface coverage:

Adsorption energy at the atop site for two difference OH coverage values are considered: 1.0 ML and 0.25 ML.  In Post 2, \(E_{ad}\) was already calculated for 1.0 ML. To achieve 0.25 ML, 2×2 supercell with one absorbed OH was built (Fig. 1).

1.0 ML
0.25 ML

Figure 1. Visualization of OH surface coverage of 1.0 ML and 0.25 ML (click on tab to view)

3. Water solvation:

Solvation effect, or the interaction of water molecules near surface with the adsorbed OH was explored. Atop sites at 0.25 ML was used with 1 or 2 water molecules floating just above the adsorbed OH (Fig. 2).

1 water
2 water

Figure 2. Optimized structure of water floating above adsorbed OH (click on tab to view)

The adsorption energy for the system with extra water molecules were calculated differently:

\begin{equation} E_{ad}=E_{OH/slab/water}-E_{OH}-E_{slab/water} \end{equation}

In this case, \(E_{slab/water}\) was calculated by optimizing a bare slab with floating water(s).

Results and discussion

1. ZPE corrections:

The adsorption energies, calculated across different adsorption environments, with and without ZPE are shown in table 1 below.

Table 1. Comparing \(E_{ad}\) with and without ZPE:

Adsorption environments\(E_{ad}\) without ZPE (eV)\(E_{ad}\) with ZPE (eV)
1.0 MLBridge-2.14-2.15
Atop-3.02-3.03
0.25 MLAtop 0 water-2.90-2.90
Atop 1 water-2.99-2.99
Atop 2 waters-3.21-3.22

The difference in ZPE are all 0.01 eV or less, which is a lot smaller than the difference between \(E_{ad}\) at different environment. Therefore, it is concluded that ZPE does not affect the adsorption energy of OH on Pt (111) significantly.

This result also has significant implementation later.

2. Adsorption environments:

Calculated values of \(\nu_{OH}\) and \(E_{ad}\) for each environments are shown in Figure 3 and 4 below:

Figure 3. Frequency of OH stretching at different adsorption environments and of free OH.

Figure 4. Adsorption energy of OH at different environments.

The hypothesis of this study is: The adsorption energy is stronger (more negative) when the adsorption environment (site, coverage, solvation) allows the adsorbed OH to form stronger hydrogen bonds.

The strength of hydrogen bond is characterized by \(\nu_{OH}\), weaker OH stretching (lower \(\nu_{OH}\)) means stronger hydrogen bonds. The presented results put this hypothesis to the test.

Adsorption sites:

Looking first at the 1.0 ML (blue section) in Fig. 3, it is observed that the OH vibrations at atop and bridge sites are both smaller than that of free OH radical, indicating some degree of hydrogen bonds formed at the sites. The hydrogen bond strength is greater at the atop site. Looking now at 1.0 ML section in Fig. 4, atop adsorption is stronger than that of the bridge site (results of Post 2).

Surface coverage:

Results at “1.0 ML Atop” and “0.25 ML Atop 0w” represent the direct comparison for surface coverage. \(\nu_{OH}\) at 0.25 ML is about the same as free OH, indicating the absence hydrogen bonds. Fig. 4 then shows that OH adsorbs weaker at 0.25 ML.

Water solvation:

Focusing now at the 0.25 ML sections (orange sections). Surprisingly, adding one water (Atop 1w) does not lower \(\nu_{OH}\), indicating that there are still no hydrogen bonding. This, however, makes sense by observing the position of this water molecule in Fig. 2 above. The placement of this water molecule did not seem to produce hydrogen bonding, which would be water’s O atom pointing toward OH’s H atom.

Nevertheless, OH adsorption is still slightly stronger with one water. This can be attributed to the solvation effect on the slab itself, because this effect is the only difference left between the two equations for \(E_{ad}\) (\(E_{slab}\) and \(E_{slab/water}\)).

Considering now the case with two water, \(\nu_{OH}\) is now lower, indicating formation of hydrogen bonding, which is also confirmed by looking at position of the second added water in Fig. 2. Finally, the adsorption with 2 waters is much stronger than with both 0 water and 1 water by looking at Fig. 4.

Conclusion

Despite the above analysis seems to indicate that stronger hydrogen bonds lead to stronger adsorption, the aforementioned hypothesis is false. The most obvious evidence is from the ZPE correction. It was already shown that ZPE does not contribute significantly to the adsorption energy. The ZPE correction only comes from the difference in frequency of OH stretching \(\nu_{OH}\). Therefore, \(\nu_{OH}\), despite varying between different environments, does not contribute significantly to \(E_{ad}\).

In other words, the most that one can conclude from these results is that there is a correlation between hydrogen bond strength \(\nu_{OH}\) and adsorption energy \(E_{ad}\). The reason behind this correlation can be thought of as hydrogen bonds formation lowering the energy of adsorbed OH \(E_{OH/slab}\) or \(E_{OH/slab/water}\), which subsequently lowers adsorption energy \(E_{ad}\). However, it would be wrong to conclude that \(\nu_{OH}\) is solely responsible for the difference in \(E_{ad}\), because the magnitude does not check out.

Finally, another interesting takeaway is that water solvents increase the adsorption tendency, not merely by interacting directly with the adsorbed species (through hydrogen bonding), but potentially through other electronic interaction with the surface as well. This is why understanding of the solvation effects on solid-liquid interfacial processes (e.g. electrocatalysis) is of great interest.

References

[1] “First principles methods using CASTEP”, Zeitschrift fuer Kristallographie 220(5-6) pp. 567-570 (2005) S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson, M. C. Payne

[2] J. P. Perdew, K. Burke, M. Enzerhof (1996). Generalized Gradient Approximation Made Simple, Phys, Rev. Lett., 77, 3865.

[3] H. J. Monkhorst and J. D. Pack (1976), Special points for Brillouin-zone integrations, Phys. Rev. B 13, 5188.

Comparing Two Transition Processes for Platinum Adatom Diffusion on Pt (100)

By Nathan McKee

Introduction

This post examines the diffusion of a platinum adatom on the (100) surface of platinum. Platinum has an fcc structure, so the (100) surface has a fourfold hollow site to which a platinum adatom can bind. An adatom in this position has multiple ways to diffuse to an adjacent fourfold hollow site, two of which are considered in this post. One is the simple hopping of the adatom to an adjacent site. The other is the atomic exchange of the adatom with an atom on the top layer of the surface, such that the surface atom travels to the adjacent site while it is replaced by the adatom. These processes are illustrated in Figure 1. DFT calculations were performed to find the transition states for both of these processes, and to estimate the frequency of each. The DFT calculations were carried out with the plane-wave based code CASTEP. The GGA PBE functional was used1, as well as on-the-fly generated (OTFG) ultrasoft pseudopotentials2. These pseudopotentials include 32 valence electrons per platinum atom, in the 4f14 5s2 5p6 5d9 6s1 configuration, with a cutoff radius of 1.27 Å. The convergence tolerance was set at 2.0*10-4 eV per atom.

Figure 1: Illustrations of surface diffusion in the cases of a) hopping, and b) exchange. All atoms shown are Pt, with some being highlighted to indicate that they play the biggest part in the transition process.

Cell Construction

To represent the (100) surface of platinum, a vacuum slab was used, with a vacuum distance of 12 Å. The thickness of the slab was chosen to be 2 layers, so as to represent the surface sufficiently without surpassing the computational limits of the system. In choosing the x and y dimensions of the cell, the interaction between adatoms must be considered. The use of plane-wave DFT necessitates the construction of a periodic crystal, but the goal of this post is to examine the diffusion of only a single adatom. This requires the cell to be large enough that the adatoms in adjacent cells do not affect each other significantly. A 3 x 3 cell was chosen to this end. The resulting cell is shown in Figure 1.

Cutoff Energy and k Points

To ensure that the calculations converge properly, an analysis of the selection of k points and the cutoff energy was performed on the slab, including the adatom. The k points were chosen to be an NxNx1 Monkhorst-Pack grid3 of evenly spaced points in reciprocal space, as is conventional for slab models in which the “a” and “b” lattice constants are equal. This analysis revealed that using a 6x6x1 k point grid in conjunction with a cutoff energy of 350 eV yields the energy within 0.1 eV. While this uncertainty is substantial, it is less than the difference between the transition state energies of the diffusion processes being considered, as will be shown later.

Transition State Search

Before considering transition states, the product and the reactant had to be identified through geometry optimization. The bottom layer of the slab was held in place to represent the bulk platinum crystal, while the top layer and the adatom were allowed to relax and find the optimal configuration. The force tolerance was set to 0.05 eV/Å, and the energy tolerance was set to 2.0 * 10-4 eV per atom. In general, one would expect to have to perform this geometry optimization for both the reactant and the product, but in the case the reactant and product are identical except for a constant translation. Therefore, only the configuration of the reactant was optimized, and the product was created by manually shifting the reactant atoms along the x axis.

With the reactant and product complete, the search for transition states begins by matching corresponding atoms in the reactant and product configurations. For the hopping transition, the adatoms in the reactant and product were matched with each other, while the surface atoms were matched according to position so that they barely moved. For the atomic exchange transition, the adatom in the reactant was matched with one of the surface atoms nearest to the hopping path in the product. Likewise, that surface atom in the reactant was matched with the adatom in the product.

The algorithm used in the transition state searches was the complete LST/QST approach. This algorithm begins with a linear synchronous transit (LST)4, followed by a conjugate gradient mnimization. The state found in this way is then used as the intermediate state in a quadratic synchronous transit (QST)4. If necessary, the last two steps are repeated until a transition state is found.

Results

These calculations yield a transition state for each diffusion process. When the reactant energies are subtracted from the transition state energies, the result is called the activation energy, ΔE. The transition rate can then be calculated using the Arrhenius equation5:

where k is the transition rate, and the prefactor nu is the atomic vibration frequency.

The activation energies obtained from the calculations are 1.04 ± 0.10 eV and 0.82 ± 0.10 eV for the hopping transition and the exchange transition, respectively. Estimating the vibrational frequency to be 5E-12 Hz, predictions can be made for the transition rates at any temperature. At 300 K, for instance, the rates are calculated to be:

khop = 1.7E-29 Hz

kexch = 9.3E-26 Hz

From these results, it is apparent that a small difference in activation energy can have a large impact on transition rates. To predict transition rates more precisely, calculations with smaller uncertainties would need to be performed. Nonetheless, these calculations indicate with some confidence that the atomic exchange transition is more frequent than the hopping transition. This preference is supported experimentally6, albeit for a slightly different exchange process in which both atoms move along the same diagonal direction.

References

  1. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
  2. Gonze, X. & Finocchi, F. Pseudopotentials Plane Waves–Projector Augmented Waves: A Primer. Phys. Scr. 2004, 40 (2004).
  3. Monkhorst, H. J. & Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B 13, 5188–5192 (1976).
  4. Halgren, T. A.; Lipscomb, W. N. Chem. Phys. Lett.49, 225 (1977)
  5. Arrhenius, S. A. (1889). “Über die Dissociationswärme und den Einfluß der Temperatur auf den Dissociationsgrad der Elektrolyte”. Z. Phys. Chem. 4: 96–116.
  6. G. L. Kellogg and P. J. Feibelman, Phys. Rev. Lett. 64 (1990) 3147.

Determination of the transition state and the activation energy for Ag atom diffusion on Cu111 surface via the adjacent threefold sites

Author : Junseok Kim

 

Introduction

Understanding a diffusion process of an adatom on a metal surface is very important in many catalytic fields, especially for nanocrystal growth. One way for studying the diffusion process is to find a transition state with corresponding activation energy. Regarding to this, we studied Ag atom diffusion on Cu111 surface in this project. Especially, we focused on the hopping of Ag between adjacent threefold sites (fcc and hcp hollow) on Cu111 surface. And, we used Density functional theory (DFT) for all energy calculations since it is a powerful tool for calculating the energy of an atomic surface or slab [1]. Lastly, we determined the transition state for the hopping of Ag atom and the activation energy.

 

Method

All DFT energy calculations were performed using the Vienna Ab initio Simulation Package (VASP) [2, 3]. For an exchange-correlation functional, the generalized gradient approximation (GGA) – Perdew Burke Ernzerhof (PBE) was used [4]. Projector augmented wave (PAW) was also employed to describe the interaction of ionic core and valance electron [5]. To generate the pseudopotential, 11 valance electrons (3d10 4s1 of the electron configuration) with 2.3 Bohr cutoff radius was used for Cu, while 11 valence electrons (4d10 5s1 of the electron configuration) with 2.4 Bohr cutoff radius was used for Ag.

We employed 3.615 Å from experiment [6] as the lattice constant of bulk Cu for building Cu111 surface. We used a four layered 2 x 2 unit slab for Cu111 surface with fixing two bottom layers and 15 Å of vacuum spacing during the surface optimization and energy calculation, and a single Ag atom was placed on the surface. Due to asymmetric slab, self-consistent dipole correction was included for all energy calculations.

The energy convergence test was carried for the bare Cu111 surface without Ag atom with respect to k-point grid and cut off energy. K-point grid was varied from 5x5x1 to 12x12x1 and cut off energy performed varying the cut off energy from 300 to 550 eV. The convergence test was continued until the difference with respect to the result at the highest k-point grid and cut off energy is less than 0.02 eV.

Then, the geometry optimization for Ag atom on adjacent threefold sites (fcc and hcp sites) of Cu111 was performed with tolerances of 0.01 eV/Å for the force and 10^-6 eV for the energy before determining the transition state. To find the transition state, we used the climbing image nudged elastic band (CI-NEB) method with the reaction tangent force less than 0.05 eV [7]. Also, we used linear interpolation to get 5 intermediate images. To confirm that the obtained transition state is valid, we calculated the vibration frequency for the transition state.

 

Result

Convergence test

      

Table 1. The relative total energy of the bare Cu111 surface varying k-point grid (Left) and cut off energy (Right)

Table 1 shows the convergence test for the bare Cu111 surface with respect to k-point grid (Left) and cut off energy (Right). From the convergence test, we obtained the optimal values of k-point grid (8 x 8 x 1) and cut off energy (450 eV), which are used for all calculations.

 

Geometry optimization of Ag atom for adjacent threefold sites of Cu111

 

Figure 1. The optimized configuration of Ag on fcc (Left) or hcp (Right) sites on Cu111 surface ; Cu (Orange) and Ag (Light blue).

After the geometry optimization, it is found that there is no significant difference for the Ag-Cu bond length of fcc (2.593 Å) and hcp (2.595 Å) sites as well as their total slab energy, -58.436 eV for fcc site and -58.434 eV for hcp site. Even though the energy difference for Ag on two adjacent sites exists, it is hard to conclude which one is more favorable than the other, since the difference is too small. We may need to calculate the energy for other sites that Ag atom could be also placed on, such as atop or bridge, to determine the most favorable site, which is not important for this project. We used these two optimized states as the end point of transition state for Ag atom diffusion on Cu111 surface.

 

Transition state and vibration frequencies

Figure 2. Reaction energy profiles using Cl-NEB for the hopping of Ag atom between the adjacent threefold site of surface, fcc and hcp sites.

Figure 3. The configuration (Bridge) of Ag atom on Cu111 surface at the transition state ; Cu (Orange) and Ag (Light blue).

Figure 2 shows the reaction energy profiles for the hopping of Ag atom between the adjacent fcc and hcp sites of Cu111 using Cl-NEB. From Cl-NEB calculation with 5 images along the reaction coordinate, we could find the prospective transition state that has the highest energy. To confirm it is the transition state, we calculated the vibration frequencies for the state, only allowing Ag atom to move during the calculation. A single imaginary frequency (-1.049 THz) was obtained with other two real vibration frequencies, 3.676 and 1.776 THz. As a result, we finally determined the transition state where Ag atom is placed on bridged site of Cu111 surface as shown in Figure 3.

It is found that the activation energy is 0.0568 eV for the hopping from fcc to hcp site, while the reverse activation energy for the hopping from hcp to fcc is 0.0545 eV. The activation energies are quite larger than the reported values (0.023 eV) [8], Shyam’s work (0.0328 eV) and Sharad’s work (0.035 eV). The main reason for these deviations seems because of using a different size of unit cell. We used a relatively smaller size of unit cell (2×2), which leads to a higher surface coverage of Ag than that of others. Even though the coverage effect on Ag atom diffusion on Cu111 surface has not been reported, a high coverage may negatively affect the hopping of Ag on Cu111 and thus lead to the increase in the activation energy [9]. With respect to this, it is worth to note that our value would be a good case for the comparison about the coverage effect on Ag atom diffusion on Cu111. Furthermore, it has been expected that Cu atom can diffuse fast on Cu111 even at room temperature since the activation energy for the self-diffusion (0.03 ~ 0.04 eV) has not much difference with the Boltzmann’s factor (kBT ~ 0.025 eV) at room temperature [10]. Regarding to this, we can also expect that Ag atom would easily diffuse as well as the self-diffusion on Cu111 at room temperature.

 

Conclusion

In this project, we optimized the adsorption of Ag atom on fcc and hcp sites of Cu111 surface with a very small total energy difference between two sites. And, we successfully determined the transition state for the diffusion of Ag atom on Cu111 surface with Cl-NEB method using DFT. Finally, the activation energy for Ag atom hopping from fcc to hcp site on Cu111 surface is found to be 0.0568 eV, which is not well matched to the reported value but is meaningful. Thus, we believe that this DFT calculation would contribute to understanding the diffusion process of Ag atom on Cu111 surface.

 

Reference

[1] “Density-Functional Theory of Atoms and Molecules”, R.G. Parr, W. Yang, Y. Weitao (1994), Oxford University Press.

[2] G. Kresse and J. Furthmuller, Comput. Mater. Sci., 1996, 6, 15–50.

[3] G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186.

[4] Perdew, J. P; Burke, K; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865-3868

[5] P. E. Blo¨chl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979.

[6] “Crystallography and Surface Structure: An Introduction for Surface Scientists and Nanoscientists. 2nd edition”, Hermann, K. Wiley-VCH: Berlin, 2016.

[7] G. Henkelman, B.P. Uberuaga, H. Jónsson, the Journal of Chemical Physics, 113 (2000) 9901-9904.

[8] Kotri, E. El koraychy, M. Mazroui, Y. Boughaleb, Surface and Interface Analysis. 2017, 49, 705–711

[9] Inderwildi et al, “Coverage dependence of oxygen decomposition and surface diffusion on
rhodium (111): A DFT study”, J. Chem. Phys. 122, 034710 (2005)

[10] Jian Wang et al, “Diffusion barriers on Cu surfaces and near steps”, 2004 Modelling Simul. Mater. Sci. Eng. 12 1209

Exploring the Effects of Surface Coverage on the Binding Site and Adsorption Energy of Atomic O on Pt(111)

by Angela Nguyen

1. Introduction

The purpose of this project is to explore the effect of surface coverage on the preferred binding site and adsorption energy of atomic O on Pt(111) through the use of DFT with a plane wave basis set as implemented by the CAmbridge Serial Total Energy Package (CASTEP) module [1] in Materials Studio. The Pt(111) surface chosen as it is the most stable surface for Pt. Four different binding sites  (atop, bridge, fcc hollow, and hcp hollow) and two different surface coverages (1 monolayer (ML) and 0.25 ML) will be investigated. For the purpose of this project, 1 ML refers to there being one adsorbate atom for every one metal atom on the surface. The reason why surface coverage is also investigated is due to the periodic nature of the DFT code. If the O atoms are too close (1 ML), they will feel a repulsive effect from each other and the total energy of the system will be affected. In addition, the 0.25 ML surface was chosen, because in literature, this was the preferred surface coverage of O on Pt(111). In order to determine which binding site most optimal for O, the adsorption energy for O on each binding site, where the lowest adsorption energy will be indicative of the preferred site. From the DFT calculations, the O atom preferred the fcc hollow site for both surface coverages. For the 1 ML surface coverage, the adsorption energy was determined to be -0.49 eV at a height of 1.29 Å above the surface and for the 0.25 ML surface, the adsorption energy was -1.23 eV at a height of 1.36 Å.

2. DFT Parameters

Listed below are the parameters used for the optimization of each vacuum slab of Pt(111) and oxygen.

Exchange-Correlation Functional TypeGeneralized Gradient Approximations [2]
Exchange-Correlation FunctionalPerdew Burke Ernzerhof [3]
Relativistic TreatmentKoelling-Harmon [4]
Psuedopotential"On the fly" generated (OTFG) ultrasoft [5]
K point GridMonkhorst-Pack [6]
SpinUnrestricted
Dipole CorrectionsSelf-consistent
Energy Convergence2.0 x 10^-5 eV/atom
Max Force0.05 eV/A

For O, the cutoff radius for the above psuedopotential is 1.10 Bohr (0.58 Å) with 6 valence electrons in the following configuration 2s2 2p4. For Pt, the cutoff radius for the above pseudopotential is 2.40 Bohr (1.27 Å) with 32 valence electrons in the following configuration 4f14 5s2 5p6 5d9 6s1.

3. Methodology

3.1 Optimization of Oxygen

In order to calculate the adsorption energy of oxygen on Pt(111), the energy of diatomic oxygen must first be calculated in the gas phase. To do this, an oxygen molecule was constructed in Materials Studio and was geometry optimized using CASTEP with the above parameters. The energy cutoff used for this calculation was 400 eV with 1 (1x1x1) irreducible k points. For the calculation,molecular oxygen was placed in a 10x10x10 vacuum cube.

3.2 Construction and Geometry Optimization of Pt(111) Vacuum Slabs

Two different vacuum slabs were constructed in order to provide a surface coverage of 1 and 0.25 ML of O. For the 1 ML surface coverage, a 1x1x5 unit cell of Pt(111) was used and for the 0.25 ML surface coverage, a 2x2x5 unit cell of Pt(111) was used. Each vacuum slab consisted of 5 layers, which the bottom 3 layers were fixed to mimic the bulk crystal and the top 2 layers were allowed to relax, with 12.0 Å vacuum between each slab. Single point energy calculations were performed on both slabs to ensure convergence with respect to k points and the cutoff energy. The table below shows the k point mesh and energy cutoff used for each vacuum slab. Using the respective k points mesh and energy cutoff from the table, each vacuum slab was optimized to allow the top 2 layers to relax to their minimum energy configuration. From this point on, the 1x1x5 unit cell will be referred as the 1 ML slab and the 2x2x5 unit cell will be referred as the 0.25 ML slab.

Figure 1. 1×1 unit cell of Pt(111) for 1 ML surface coverage. a) Top view of the unit cell where the O atom will be adsorbed onto the surface. b) Side view of the unit cell showing which layers were fixed. The blue sphere represents a Pt atom. The red sphere represents a Pt atom that was fixed during optimization, and the grey sphere represents a Pt atom that was able to relax to the lowest energy configuration.

Surface CoverageK Point MeshEnergy Cutoff [eV]
1 ML7 x 7 x 1400
0.25 ML7 x 7 x 1400

3.2 Adsorption of O Adsorbed onto the Surface

On the Pt(111) surface, there are four different possible binding sites that O could adsorb onto the surface: atop, bridge, fcc hollow, and hcp hollow. Figure 2 depicts the location of the various binding sites for O on Pt(111). Each configuration was then geometrically optimized to determine the lowest energy configuration for each binding site with different surface coverages. The adsorption energy for oxygen was then calculated for each binding site configuration using the following formula:

\begin{equation}E_{ads}=E_{surf+ads}-\frac{E_{ads(g)}}{2}-E_{surf}\end{equation}

where \(E_{ads}\) is the adsorption energy, \(E_{surf+ads}\) is the energy of the surface with an adsorbate, \(E_{ads(g)}\) is the energy of the adsorbate in the gas phase, and \(E_{surf}\) is the energy of the bare surface. The adsorption energies can then be compared for each surface coverage to determine which is the optimal binding site of O.

Figure 2. Potential binding sites for O atom on Pt(111) surface. The blue sphere represent the Pt atom and the red sphere represents the O atom.

4. Results

When optimizing the geometry of each binding site configuration with CASTEP, it was found the when O was placed on a bridge or hcp hollow site, the O atom moved to an fcc hollow site. This shows that the fcc hollow binding site is the more stable binding site as compared to the bridge and hcp hollow site. In order to verify this hypothesis, the adsorption energies were calculated for the bridge and hcp hollow site for 1 ML surface coverage before the O atom was moved to the fcc hollow site. The results for these calculations are shown in the table below and will be discussed later.

In addition, the atop binding site did not geometrically optimize to the fcc hollow site because of the high symmetry around the site. Even though the O atom was placed off symmetry to the atop site, the forces on the atom pulled it towards the atop site rather than to an fcc hollow site. Therefore, only the results for the binding site configurations where the O atom started in the atop or fcc hollow site will be considered. The results of the optimization of the bridge site configuration are shown below in Figure 3 and Figure 4 for 1 ML and 0.25 ML surface coverage respectively.

Figure 3. Optimized bridge site configuration for O on Pt(111) with 1 ML surface coverage. a) Initial configuration for O binding on the bridge site of Pt(111). b) Finial optimized configuration for the “bridge site” binding of O on Pt(111). Note that the O atom has moved to the fcc hollow binding site indicating that the fcc hollow site is more stable than the bridge site. The blue sphere represent the Pt atom and the red sphere represents the O atom.

Figure 4. Optimized bridge site configuration for O on Pt(111) with 0.25 ML surface coverage. a) Initial configuration for O binding on the bridge site of Pt(111). b) Finial optimized configuration for the “bridge site” binding of O on Pt(111). Note that the O atom has moved to the fcc hollow binding site indicating that the fcc hollow site is more stable than the bridge site. The blue sphere represent the Pt atom and the red sphere represents the O atom.

4.1 Adsorption of O on 1 ML Slab

Figure 5 shows the final optimized atop and fcc hollow site configuration for O on Pt(111) with 1 ML surface coverage.

Figure 5. Optimized binding site configuration for O on Pt(111) with 1 ML surface coverage. a) Binding site configuration for the atop site where the O atom is 1.82 Å above the Pt surface. b) Binding site configuration for the fcc hollow site where the O atom is 1.29 Å above the Pt surface. The larger blue sphere represent the Pt atom and the smaller red sphere represent the O atom. The red sticks represent the bonding between the O atom and the Pt atom.

The table below lists the results for adsorption energy and the distance the O atom is above the Pt(111) surface for the atop, fcc hollow, bridge, and hcp hollow sites on the 1 ML slab. As it can be seen, the estimated adsorption energies from the bridge and hcp hollow sites are greater than that of the fcc hollow site. Because of this, the geometry optimization will try to move the O atom to most stable site possible, which is the fcc hollow site. From the adsorption energy of the atop and fcc hollow sites, it can be seen that at 1 ML surface coverage, the O atom favors the fcc hollow site of the Pt(111) surface as there is a 0.23 eV difference between the two binding sites. Because of the close proximity of the O atoms on the surface, they are able to feel the repulsive forces of the other O atoms causing for a higher adsorption energy as compared to the 0.25 ML surface coverage.

Binding SiteAdsorption Energy [eV]O Distance to Pt Surface [Å]
Atop-0.251.82
FCC Hollow-0.491.29
Bridge*-0.32N/A
HCP Hollow*-0.35N/A

4.2 Adsorption of O on 0.25 ML Slab

Figure 6 shows the final optimized atop and fcc hollow site configuration for O on Pt(111) with 0.25 ML surface coverage.

Figure 6. Optimized binding site configuration for O on Pt(111) with 0.25 ML surface coverage. a) Binding site configuration for the atop site where the O atom is 1.82 Å above the Pt surface. b) Binding site configuration for the fcc hollow site where the O atom is 1.36 Å above the Pt surface. The larger blue sphere represent the Pt atom and the smaller red sphere represent the O atom. The red sticks represent the bonding between the O atom and the Pt atom.

The table below lists the results for adsorption energy and the distance the O atom is above the Pt(111) surface for the atop and fcc hollow site on the 0.25 ML slab. From the adsorption energy, it can be seen that at 0.25 ML surface coverage, the O atom also favors the fcc hollow site of the Pt(111) surface as there is a 0.93 eV difference between the two binding sites. Here, the O atom is far away enough from each other to not feel any repulsion from other O atoms and both unpaired electrons are able to form bonds to be stabilized.

Adsorption Energy of O on Pt(111) with 0.25 ML

Binding SiteAdsorption Energy [eV]O Distance to Pt Surface [Å]
Atop-0.301.82
FCC Hollow-1.231.36

5. Conclusions

In previous experimental studies found in literature, many researchers agree that on the Pt(111) surface the O atom tends to have a 0.25 ML surface coverage and prefers to bind on the fcc hollow site which the O atom 1.29 Å above the surface [7] [8]. Unfortunately, not much has been studied on a 1ML surface coverage of O on Pt(111). From this project, it can be seen that for the 1 ML slab the qualitative results match experimental data where the adsorption energy of O on the fcc hollow site is lower than that of the atop site and the O atom is also 1.29 Å above the surface. The DFT calculations also verify the reason that O prefers to bind to the Pt(111) surface at a 0.25 ML surface coverage due to the repulsive interactions felt at a 1.0 ML surface coverage. In addition, the adsorption energies for the 0.25 ML surface coverage corroborates with the experimental data stating that the O atom prefers to bind to the fcc hollow site with 0.25 ML surface coverage. The distance of the O atom above the Pt(111) surface for the 0.25 ML is slightly higher as compared to the experimental results, but this could be due to the different conditions for the computational calculations. For example, for these computations, the Pt(111) slab was placed in a vacuum environment with no other impurities. In an experiment, this type of environment is not easily achievable.

6. References

[1] Clark Stewart J et al., “First principles methods using CASTEP ,” Zeitschrift für Kristallographie – Crystalline Materials , vol. 220. p. 567, 2005.
[2] J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made Simple,” Phys. Rev. Lett., vol. 77, no. 18, pp. 3865–3868, Oct. 1996.
[3] J. P. Perdew et al., “Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation,” Phys. Rev. B, vol. 46, no. 11, pp. 6671–6687, Sep. 1992.
[4] D. D. Koelling and B. N. Harmon, “A technique for relativistic spin-polarised calculations,” J. Phys. C Solid State Phys., vol. 10, no. 16, pp. 3107–3114, Aug. 1977.
[5] D. Vanderbilt, “Soft self-consistent pseudopotentials in a generalized eigenvalue formalism,” Phys. Rev. B, vol. 41, no. 11, pp. 7892–7895, Apr. 1990.
[6] H. J. Monkhorst and J. D. Pack, “Special points for Brillouin-zone integrations,” Phys. Rev. B, vol. 13, no. 12, pp. 5188–5192, Jun. 1976.
[7] A. Kokalj, A. Lesar, M. Hodošček, and M. Causà, “Periodic DFT Study of the Pt(111): A p(1×1) Atomic Oxygen Interaction with the Surface,” J. Phys. Chem. B, vol. 103, no. 34, pp. 7222–7232, 2002.
[8] S. G. A. Materer N., Barbieri A., Doll R., Heinz K., Van Hove M.A., “Reliability of detailed LEED structural analyses : Pt ( 111 ) and Pt ( 111 ) -p ( 2×2 ) -O,” vol. 325, pp. 207–222, 1995.