DOI:
10.1039/C6RA09274H
(Paper)
RSC Adv., 2016,
6, 69670-69676
Interactions of Pt nanoparticles with molecular components in polymer electrolyte membrane fuel cells: multi-scale modeling approach†
Received
11th April 2016
, Accepted 16th July 2016
First published on 18th July 2016
Abstract
In this study, a three-phase interfacial system of a fuel cell is simulated using a multi-scale simulation approach consisting of quantum mechanical density functional theory and molecular dynamics simulations. Through these simulations, the structural and transport properties of the three-phase system are investigated. The molecular interactions among the components of the three-phase interfacial system are examined by density functional theory and parameterized for potential energy functions of force field. First, we investigate the interactions of the Pt clusters with various molecules as a function of distance using the density functional theory method with dispersion correction. Based on the results of these calculations, a non-bonded interaction curve is built for each Pt–molecule pair. Such non-bonded interaction curves are reproduced by potential energy functions with optimized parameters. Based on these investigations, we develop a force field to describe the structures and transport properties of the Nafion–Pt–carbon (graphite) three-phase interfacial system using molecular dynamics simulations.
Introduction
There has been strong interest in nanosized metal particles, owing to their engineering importance,1 which stems from their distinctly different properties compared to the bulk phase. This size-dependent difference in properties is also valid for various transition metals.2 In particular, despite their high price, Pt nanoparticles are considered as the catalyst of choice for fuel cells owing to their excellent catalytic activities. Therefore, theoretical and computational investigations3 as well as experimental studies4 have been performed extensively to obtain a fundamental understanding on the size-dependent properties of Pt nanoparticles, with the aim of maximizing the benefits from the use of such precious metals.
Compared to experimental methods, theoretical and computational approaches may have a unique advantage in tackling small systems with dimensions of nanometer-scale. In the latter methods, the atomic structures and corresponding properties are examined directly. Lee et al.3g investigated Pt2 molecules using Kramers restricted ab initio calculations and determined that dynamic and non-dynamic electron correlation effects are required to obtain reliable spectroscopic data. Varga et al.3a found that relativistic effects should be considered in order to calculate accurate binding energy values for diatomic molecules of heavy metals using relativistic density functional theory (DFT). Grönbeck and Andreoni3b studied neutral and anionic Pt clusters of various shapes up to Pt5 using SLDA and BLYP. Based on their DFT calculations, they found that 3-D geometries are unfavorable up to Pt5 clusters. Similarly, through a DFT study on 70 Pt clusters of various sizes and shapes using PW91 and plane wave basis set, Xiao and Wang3e found that the planar shape is preferred for small clusters of up to 9 atoms. In addition, they found that the layered Pt clusters are very stable compared to closed-packed spherical shapes. This observation has also been validated by several research groups.3f,5 However, by examining the effect of spin–orbit interactions on Ptn (n = 4–6) using B3LYP, Sebetci6 reported that the 3-D geometries are energetically preferred, which conflicts with previous reports.3b,d,e,5 Using BPW91 and effective core potential, Aprà and Fortunelli3d reported that Ptn (n = 13, 38, and 55) starts developing metallic characteristics and Pt55 favors the icosahedral geometry (Ih), whereas Pt13 favors the D4h geometry. Chepulskii and Curtarolo7 calculated the energies of various common Pt nanoparticle shapes as a function of size up to 400 atoms using various GGA functionals and plane wave basis set. From such large cluster calculations, they found that the surface energy depends on the lattice parameters. They also suggested that the bulk surface energies are valid only for nanoparticles with diameters greater than 1.46–1.57 nm.
Although there are numerous studies on Pt nanoparticles, the interaction of Pt nanoparticles with other molecular species has received relatively little attention. This may be mostly due to the difficulties associated with experimentally probing molecular interactions at the nanometer scale. Langenbach et al.8 found that water molecules are adsorbed on Pt (111) surfaces via oxygen using IR-reflection and UV-photoemission, and Michaelides et al.9 confirmed that water adsorption occurs on the on-top sites of the Pt (111) surface using PW91 and plane wave basis set. Huang and Balbuena10 used molecular dynamic (MD) simulations to study Pt nanoparticles with spherical and cubic shapes containing 256–260 atoms on graphite. They found that in the case of nanoparticles, a melting-like structural transformation is observed at a temperature that is much lower (∼1000 K) than the bulk melting temperature (2045.15 K).
Thus, our primary objective in this simulation study is to model the Pt nanoparticle with various molecular species in a polymer electrolyte membrane fuel cell (PEMFC) and to elucidate the detailed interactions between the main components of PEMFC. First, we used density functional theory (DFT) to study Pt nano-clusters as used in PEMFCs. We tested the various Pt nanoparticles to obtain their most probable energy state with normalized cohesive energy (NCE) calculations. Based on these results, we developed a coordination number (CN) model to evaluate the energy of a cluster. Our model correctly predicted the energy of various facets of Pt surface. The CN model captures the surface energy reasonable well, especially for the (111) surface. Consequently, we have probed the various moieties of the electrolyte with Pt (111) surface to obtain the interaction, position and conformation of the adsorbed species. From this, we developed a force field for use in molecular dynamics (MD) simulations describing the three-phase interface (Pt–graphite–electrolyte) of the PEMFC system. Using the newly developed a force field, we successfully described diffusion of oxygen in three-phase model which contains a Pt cluster on top of graphite.
Computational details
In order to investigate the electronic structures and interactions of Pt nanoparticle with various sizes and shapes, we used spin-unrestricted density functional theory (DFT) through Jaguar11 by employing three widely used functionals, namely M06,12 PBE,13 and B3LYP14 with LACVP** basis set containing 6-31G** and LANL2DZ effective core basis sets15 (Fig. 1). We also used spin-unrestricted periodic DFT through CASTEP16 and DMol3 (ref. 17) implemented in Materials Studio18 with PBE functional and plane wave basis set. All the geometries shown in Fig. 1 are optimized by DFT calculations. The electronic structure of the Pt particles depends on both the size as well as the shape of the particles. Consequently, different clusters have different multiplicities and energy values. For such DFT calculations, we started with an initial guess value for the spin state, using the method described by the spin guesser from Goddard.19 Then, we simulated various multiplicities around this initial guess until we arrived at the lowest energy value, which turned out that various functionals agree with the multiplicity showing the lowest energy.
 |
| Fig. 1 Various Pt nanoparticle (PtX–Y–Z) models used for DFT calculations in this study. X, Y, and Z denote the number of Pt atoms in the first, second, and third layers of the nanoparticles, respectively. | |
From these results, we found that that the stability of the Pt nanoparticle increases with increasing the particle size, and that the two-layer and three-layer Pt nanoparticles have nearly the same stability to each other. Using this information, we selected a highly stable model of Pt nanoparticles to develop a force field describing the interaction between the Pt nanoparticles and molecules. To obtain force fields describing the interaction of the molecular species with Pt, we optimized the force field parameters in the Morse potential to reproduce the DFT energies and structures.
After developing a force field for the Nafion–Pt interactions, we built a model for the three-phase system of the PEMFC for the MD simulation. For the MD simulations, we utilized the DREIDING force field20 and F3C water model,21 which have been used widely to describe hydrated polymeric systems such as hydrogels,22 fuel cells,23 epoxy molding compounds,24 and so on. The total energy (Etotal) is calculated using eqn (1):
|
Etotal = EvdW + EQ + Ebond + Eangle + Etorsion + Einversion
| (1) |
where
EvdW,
EQ,
Ebond,
Eangle,
Etorsion, and
Einversion are the van der Waals, electrostatic, bond stretching, angle bending, torsion, and inversion components, respectively. The atomic charges of the molecules in the MD simulations were calculated using Mulliken population analysis, while the atomic charges of the water molecule were from the F3C water model.
21 For this study, the annealing and equilibrium MD simulations were performed using LAMMPS (Large-scale Atomic/Molecular Massively Parallel Simulator) from Plimpton at Sandia National Laboratory.
25 The equations of motion were integrated using the velocity Verlet algorithm
26 with a time step of 1.0 fs. The Nose–Hoover temperature thermostat
27 was used with a damping relaxation time of 0.1 ps. After the annealing simulations, NPT simulations were performed during 5 ns for equilibration. The data were then collected from another 5 ns NPT MD simulation for analysis.
Results and discussion
Analysis of Pt nanoclusters
In order to obtain the most probable energy state, we performed spin-unrestricted calculations with various spin multiplicities. The results of these calculations are presented in Fig. 2. Here, the relative energies are calculated with respect to the minimum energy value (i.e., energy deviation from the minimum energy) and the optimal spin state is defined as the spin state resulting in the minimum energy value. After specifying the minimum energy value of Pt nanoparticles with the optimal spin state, we calculated the normalized cohesive energy (NCE) using eqn (2): |
NCE = [E(PtX–Y–Z[m]) − n × E(Pt)]/n
| (2) |
where m and n denote the spin multiplicity and total number of Pt atoms in the nanoparticles, respectively, and X, Y, and Z denote the number of Pt atoms in the first, second, and third layers of the nanoparticle, respectively. Please note that the sum of X, Y and Z is equal to n. E(PtX–Y–Z[m]) and E(Pt) are the energy values of the Pt nanoparticles and single Pt atoms, respectively. According to eqn (2), the stability of Pt nanoparticles increases as the NCE decreases towards more negative values.
 |
| Fig. 2 Change in the relative energy of Pt nanoparticles as a function of spin multiplicity. Relative energy denotes the energy deviation from the minimum energy with optimal spin multiplicity. Other cases are not presented due to similarity. | |
In Fig. 3, we summarize the NCEs calculated for the Pt nanoparticles shown in Fig. 1. First of all, it is commonly observed from the calculations involving three different functionals that the NCE decreases as a function of the number of Pt atoms in a nanoparticle, approaching the experimental NCE value for bulk Pt (−5.85 eV).28 This implies that the stability of the Pt nanoparticles increases with increase in the particle size. Although the general trend of NCE is the same regardless of the functional used in this study, it should be noted that the NCE values themselves are different depending on the functionals used and may be ranked in the following order: NCE (PBE) < NCE (M06) < NCE (B3LYP). In other words, PBE results in the most stable NCE value, whereas B3LYP yields the least stable NCE for the same Pt nanoparticle.
 |
| Fig. 3 Change in the normalized cohesive energy (NCE) at various layers as a function of the number of Pt atoms in a cluster. The NCE values of the Pt clusters are higher than that of bulk Pt. | |
Another observation from Fig. 3 is that the two- and three-layer Pt nanoparticles have nearly the same stability and the data points from these nanoparticles appear to be identical for each functional. On the other hand, the single-layer nanoparticles are relatively less stable with higher NCE values, since the Pt atoms in the single-layer model lack the coordination with the neighboring atoms. From these results, we presume that the two-layer model with a high number of Pt atoms as well as three-layer model would be sufficient for describing the Pt nanoparticles with any given number of Pt atoms. Thus, our result appears in agreement with Sebetci's findings.6 Using this information, we selected a highly stable model of Pt nanoparticles to develop a force field describing the interaction between the Pt nanoparticles and molecules.
Coordination number (CN) model for Pt nanoparticles
Based on the DFT calculations for various Pt nanoparticles, we developed a model to evaluate the energy of the Pt nanoparticles based on their size and shape. For this purpose, we counted the number of Pt atoms corresponding to each CN for all the Pt nanoparticle models in Fig. 1. The example in Fig. 4(a) shows that four Pt atoms in Pt6–3–1 nanoparticle have CN of 3 while six Pt atoms have CN of 6. After determining the number of Pt atoms for each CN, we calculated the atomic cohesive energy for each CN ((ACE)CN) in eqn (3) using the least-squares method: |
 | (3) |
where CE and NCN denote the cohesive energy and number of Pt atoms for each CN, respectively. In Fig. 4(b), it is observed that (ACE)CN decreases with increasing CN and this decrease roughly follows the line connecting the two reference points (zero for CN = 0, which is obviously defined from the concept and 5.85 eV (ref. 28) for CN = 12, which is experimentally reported). This implies that the stability of Pt atoms in the nanoparticles increases with increasing CN, indicating increasing the number of bonds (or the number of bonded neighboring atoms).
 |
| Fig. 4 Pt nanoparticles have atoms with various coordination numbers (CNs): (a) model of the Pt6–3–1 nanoparticle, which has 4 Pt atoms with CN = 3 and 6 Pt atoms with CN = 6; (b) change in atomic cohesive energy of Pt as a function of CN reported in Table 1. | |
However, in Fig. 4(b), it is also important to note the deviations in the values of (ACE)CN from the dashed line, particularly for lower CN values. The dashed line between CN = 0 and CN = 12, which predicts the stability of the Pt atoms, increases linearly as a function of CN on the assumption that each bond has a constant strength regardless of the number of bonds. However, such negative deviation from the dashed line indicates that the bond strength is not constant. Rather, it decreases as a function of CN. In other words, the characteristics of bond between two Pt atoms depend on the number of the neighboring Pt atoms. This interpretation could be clearly confirmed by plotting ACE/CN as a function of CN, which is shown in Fig. 4(b). The contribution of individual Pt–Pt bonds to ACE is most significant for CN = 1, and becomes weaker with increasing CN.
In order to validate the CN model in eqn (3) with the ACE contributions of each CN in Table 1, we calculated the surface energy with various facets using the CN model. As shown in Table 2, we compared the surface energies calculated using the CN model with the values reported in the literature.7,29 For this purpose, we calculated the surface energies for several different facets such as (111), (100), and (730) as shown in Fig. 5. Although the surface energies for the (100) and (730) surfaces calculated from the CN model slightly deviate from the reported surface energy values, in general, the CN model captures the surface energy reasonably well, especially for the (111) surface (Table 2).
Table 1 Atomic cohesive energy (ACE) as a function of coordination number (CN). As expected, more coordinated atoms form more stable Pt nanoparticles by lowering the cohesive energy
CN |
1 |
2 |
3 |
4 |
5 |
6 |
Energy (eV) |
−0.93 |
−1.77 |
−2.29 |
−2.91 |
−2.64 |
−2.72 |
CN |
7 |
8 |
9 |
10 |
11 |
12 |
Energy (eV) |
−3.13 |
−4.41 |
−4.76 |
−4.82 |
−4.90 |
−5.80 |
Table 2 Comparison between the surface energies calculated from the CN model and reference values from the literature
 |
| Fig. 5 Various facets of Pt surfaces: (a) (111) direction; (b) (100) direction; (c) (730) direction. The colors are used to clarify the structural repetition of the atomic layers. | |
Force field development
To obtain force fields describing the interaction of the molecular species with Pt, we used the Morse potential whose parameters obtained are summarized in Table 3. These parameters were developed with a pragmatic approach in mind, namely that the force field would include all the DFT interaction energies.
Table 3 Force field parameters for the ionomer moieties interacting with platinum using the Morse potential function

|
Pairs |
D (kcal mol−1) |
R0 (Å) |
γ |
O(H2O)–Pt |
2.500 |
2.416 |
11.648 |
H(H2O)–Pt |
1.364 |
3.054 |
9.830 |
O(H3O+)–Pt |
6.433 |
3.473 |
7.746 |
H(H3O+)–Pt |
0.231 |
1.609 |
12.000 |
O(O2)–Pt |
0.429 |
3.470 |
7.024 |
C(Nafion)–Pt |
0.338 |
4.476 |
13.789 |
F(Nafion)–Pt |
0.205 |
3.674 |
5.995 |
Oether(Nafion)–Pt |
0.409 |
3.470 |
7.024 |
S(Nafion)–Pt |
3.041 |
3.413 |
14.672 |
OSO3(Nafion)–Pt |
2.351 |
2.511 |
5.144 |
C(graphite)–Pt |
0.118 |
4.421 |
7.887 |
Interaction of Pt (111) with water and hydronium ions. To describe the interaction of water (H2O) and hydronium (H3O+) molecules with a Pt (111) surface, we calculated the adsorptive binding energy using M06 functional with LACVP** basis set. For this purpose, the structures of H2O–Pt14–13–8 and H3O+–Pt12–7 nanoparticles were optimized (Fig. S1(a) and (b)†). The binding energy was then calculated as a function of distance. As shown in Fig. S1(c),† the binding energies are ∼−15.01 kcal mol−1 at a distance of 2.411 Å and ∼−40.43 kcal mol−1 at a distance of 2.899 Å for water and hydronium ion, respectively. To check whether the size of the nanoparticles affects the binding energy, we also calculated the binding energy of water on a periodic Pt (111) surface with PBE-D3 and DNP basis sets in DMol3.17 First, the magnitude of the binding energy of water clearly indicates that the water molecules are physisorbed on the Pt (111) surface. On the other hand, we observe that the binding of the hydronium ion with the Pt surface is much stronger than that of water, corresponding to chemisorption. We think that this type of strong adsorption of charged molecular species on a metal surface is due to the polarization of the metal surface induced by the charges in the molecule. Thus, the DFT calculations of the molecular interactions indicate that the adsorption of hydronium ion is preferred to water molecule on the hydrated Pt nanoparticles even regardless of the electrochemical reactions.
Interaction with oxygen molecule. By investigating various configurations of oxygen (O2) molecules adsorbed on the Pt (111) surface (Fig. S2†) using DFT calculations, it is found that an O2 molecule lies along the Pt–Pt bond on the Pt (111) surface (Fig. S3(a)†). As shown in Fig. S3(b),† the O2 molecule is adsorbed at a distance of 2.12 Å from the Pt surface with a binding energy of −7.71 kcal mol−1 (S = 11, where S is the spin multiplicity). On the other hand, a slightly weaker adsorption occurs at a distance of 2.72 Å with a smaller binding energy of −7.07 kcal mol−1 (S = 13), which has longer range of interaction. As shown in Fig. S3(c),† this implies that O2 molecule may approach the Pt surface with S = 13 and then change its spin state to S = 11 in order to gain more stabilization. We think that this is a preparatory step for the oxygen reduction reaction, which is not in the scope of this study.
Interaction with polymer electrolyte (Nafion). Although the surface of electrodes is generally coated with a polymer electrolyte, typically Nafion, the interactions of such coatings with the Pt catalyst have not been thoroughly understood, owing to the lack of proper instrumental analysis techniques for probing the structure and dynamic motion of polymer chains on the metal surface within a few nanometer thickness. In order to investigate the interactions between Nafion and Pt, we first prepared three small model molecules by fragmenting the Nafion unit structure (Fig. S4(a)†). The three structures prepared are perfluoroethane (CF3CF3), perfluorodimethyl ether (CF3OCF3), and perfluoromethyl sulfonate (CF3SO3−), which represent the backbone chain, spacer, and sulfonate group in the side chain of Nafion, respectively. We performed DFT geometry optimization for these model molecules on Pt (111) surface and the results are shown in Fig. S4(b)–(d).† From these optimization calculations, the binding energies were calculated to be −11.88 kcal mol−1 at 3.70 Å, −13.25 kcal mol−1 at 3.25 Å, and −51.33 kcal mol−1 at 2.71 Å for CF3CF3–Pt12–7 (Fig. S4(b)†), CF3OCF3–Pt12–7 (Fig. S4(c)†), and CF3SO3−–Pt12–7 (Fig. S4(d)†), respectively. This indicates that CF3CF3 and CF3OCF3 are physisorbed, whereas CF3SO3− is chemisorbed. From the binding energy values for the model molecules, it is inferred that the surface of Pt nanoparticles may be occupied by the sulfonate groups of Nafion, hydronium ions, and water molecules rather than CF3CF3 and CF3OCF3.
Interaction with graphite. To describe the interactions between graphite and Pt cluster, we calculated the binding energy of Pt6–3–1 cluster model on a periodic graphite surface (Fig. S5(a)†) with PBE-D3 functional and DNP basis set. The size of the periodic graphite model was set as 9.84 × 9.84 × 33.00 Å and z layers of graphite were used to calculate the interactions with the Pt6–3–1 cluster. Fig. S5(b)† shows that the binding energy is ∼−26 kcal mol−1 at a distance of 3.650 Å indicating that chemisorption occurs between the Pt6–3–1 cluster and graphite.
O2 diffusion in three phase model. After developing a force field for the Nafion–Pt interactions, we built a model for the three phase system of the fuel cell as shown in Fig. 6(a). This model consists of an electrolyte, electrode, and Pt particle in the cathode region. The simulation box size was 51.38 × 51.91 × 509.00 Å, which allows 4 layers of graphite to fit unstrained in the x and y directions. In the z direction, the Nafion ionomer was around 100 Å thick, whereas the graphite layer was around 13 Å thick (0.335 nm × 4 layer). The bottom layer of the graphite was fixed. Next, we placed a Pt cluster consisting of 170 atoms on top of the graphite. We also introduced 16 Nafion chains with a molecular mass of 11
468 Da and ten sulfonate groups per chain, whose equivalent weight is ∼1100. The number of water molecules was 2240 (λ = 14, 14 water molecules per sulfonate group).
 |
| Fig. 6 (a) Three phase model consisting of Pt nanoparticle, Nafion, and graphite with oxygen, water, and hydronium ions; (b) mean square displacement of O2 molecules as a function of time in the three phase model at 353 K. Gray, purple, white, red, cyan, orange, green, and yellow colors indicate carbon, Pt, hydrogen, oxygen, oxygen of hydronium, oxygen of O2 molecules, fluorine, and sulfur, respectively. | |
In our previous study,30 we analyzed the surface of Pt nanoparticles and the distribution of O2 molecules in a three-phase interfacial system using Connolly surface analysis and the pair correlation function, respectively. It was found that ∼60% of the surface of Pt nanoparticles is available for electrochemical reaction in the three-phase interfacial system. We also found that O2 molecules tend to be located around the Nafion phase rather than the water phase.
The diffusion of O2 molecules in the three-phase system is another interesting property of PEMFCs, because the transport of O2 molecules would affect the performance of PEMFC. The mean square displacement (MSD) of O2 molecules in the three-phase system was obtained from the last 5 ns of the NPT MD simulations at 353 K (Fig. 6(b)) to calculate the diffusion coefficients (D) defined in eqn (4)
|
 | (4) |
where
r(
t) and
r(0) are the positions of the target molecules at a time
t greater than 0 and
t = 0, respectively. The calculated diffusion coefficient of O
2 molecules in the three-phase system at 353 K is 6.5 × 10
−3 cm
2 s
−1, whereas that in the bulk hydrated Nafion phase at 353 K is 1.8 × 10
−5 cm
2 s
−1. This result indicates that the diffusion of O
2 molecules in the three-phase system is significantly greater than in the bulk Nafion system.
Conclusions
In this study, we characterized the relationship between the CN and energy of the Pt particles. This understanding would be useful in the preparation of model structures with various numbers of Pt atoms in a three-phase system. Our CN-based model of stable Pt particles is suitable for predicting the ACE of a Pt particle consisting of n atoms in any arbitrary shape. We also developed a force field describing the interactions between Pt particle and other components such as graphite, Nafion, water, hydronium, and oxygen in the three-phase system by optimizing the parameters of potential energy functions to reproduce the structures and energies from the DFT calculations. The force field developed in this study can be used to perform extensive MD simulations to investigate the structures and transport in the three-phase interfacial system.
Acknowledgements
This research was supported by Ballard via the Department of Energy (DOE) grant No. DE-EE0000466. This research was also supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2015M1A2A2057129).
Notes and references
-
(a) K. Koga, T. Ikeshoji and K. Sugawara, Phys. Rev. Lett., 2004, 92, 115507 CrossRef PubMed;
(b) Y. Sun and Y. Xia, Science, 2002, 298, 2176 CrossRef CAS PubMed;
(c) R. Sardar, A. M. Funston, P. Mulvaney and R. W. Murray, Langmuir, 2009, 25, 13840 CrossRef CAS PubMed;
(d) S. Link, C. Burda, B. Nikoobakht and M. A. El-Sayed, J. Phys. Chem. B, 2000, 104, 6152 CrossRef CAS.
-
(a) G. Hodes, Adv. Mater., 2007, 19, 639 CrossRef CAS;
(b) M. J. Yacaman, J. A. Ascencio, H. B. Liu and J. Gardea-Torresdey, J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.--Process., Meas., Phenom., 2001, 19, 1091 CrossRef CAS.
-
(a) S. Varga, B. Fricke, H. Nakamatsu, T. Mukoyama, J. Anton, D. Geschke, A. Heitmann, E. Engel and T. Bastug, J. Chem. Phys., 2000, 112, 3499 CrossRef CAS;
(b) H. Grönbeck and W. Andreoni, Chem. Phys., 2000, 262, 1 CrossRef;
(c) J. Wang, G. Wang and J. Zhao, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 66, 035418 CrossRef;
(d) E. Aprà and A. Fortunelli, J. Phys. Chem. A, 2003, 107, 2934 CrossRef;
(e) L. Xiao and L. Wang, J. Phys. Chem. A, 2004, 108, 8605 CrossRef CAS;
(f) W. Q. Tian, M. Ge, B. R. Sahu, D. Wang, T. Yamada and S. Mashiko, J. Phys. Chem. A, 2004, 108, 3806 CrossRef CAS;
(g) D.-K. Lee, G.-H. Jeung and Y. S. Lee, Int. J. Quantum Chem., 2009, 109, 1975 CrossRef CAS.
-
(a) W. Eberhardt, P. Fayet, D. M. Cox, Z. Fu, A. Kaldor, R. Sherwood and D. Sondericker, Phys. Rev. Lett., 1990, 64, 780 CrossRef CAS PubMed;
(b) Z. L. Wang, J. M. Petroski, T. C. Green and M. A. El-Sayed, J. Phys. Chem. B, 1998, 102, 6145 CrossRef CAS;
(c) M. B. Airola and M. D. Morse, J. Chem. Phys., 2002, 116, 1313 CrossRef CAS.
-
(a) M. N. Huda, M. K. Niranjan, B. R. Sahu and L. Kleinman, Phys. Rev. A, 2006, 73, 053201 CrossRef;
(b) K. Bhattacharyya and C. Majumder, Chem. Phys. Lett., 2007, 446, 374 CrossRef CAS.
- A. Sebetci, Phys. Chem. Chem. Phys., 2009, 11, 921 RSC.
- R. V. Chepulskii and S. Curtarolo, ACS Nano, 2011, 5, 247 CrossRef CAS PubMed.
- E. Langenbach, A. Spitzer and H. Lüth, Surf. Sci., 1984, 147, 179 CrossRef CAS.
- A. Michaelides, V. A. Ranea, P. L. de Andres and D. A. King, Phys. Rev. Lett., 2003, 90, 216102 CrossRef CAS PubMed.
- S.-P. Huang and P. B. Balbuena, Mol. Phys., 2002, 100, 2165 CrossRef CAS.
- Jaguar, v7.5, Schrödinger, LLC, New York, 2008 Search PubMed.
-
(a) Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, 215 CrossRef CAS;
(b) Y. Zhao and D. G. Truhlar, J. Chem. Theory Comput., 2008, 4, 1849 CrossRef CAS PubMed.
-
(a) J. P. Perdew, K. Burke and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 16533 CrossRef CAS;
(b) J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
-
(a) A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS;
(b) P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623 CrossRef CAS.
- P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 299 CrossRef CAS.
- M. C. Payne, M. P. Teter, D. C. Allan, T. A. Arias and J. D. Joannopoulos, Rev. Mod. Phys., 1992, 64, 1045 CrossRef CAS.
- B. Delley, J. Chem. Phys., 1990, 92, 508 CrossRef CAS.
- Materials Studio, v5.0, Accelrys Software Inc., San Diego, 2009 Search PubMed.
-
(a) J. Kua and W. A. Goddard, J. Phys. Chem. B, 1998, 102, 9481 CrossRef CAS;
(b) J. Kua and W. A. Goddard, J. Phys. Chem. B, 1999, 103, 2318 CrossRef CAS.
- S. L. Mayo, B. D. Olafson and W. A. Goddard, J. Phys. Chem., 1990, 94, 8897 CrossRef CAS.
- M. Levitt, M. Hirshberg, R. Sharon, K. E. Laidig and V. Daggett, J. Phys. Chem. B, 1997, 101, 5051 CrossRef CAS.
-
(a) S. G. Lee, G. F. Brunello, S. S. Jang and D. G. Bucknall, Biomaterials, 2009, 30, 6130 CrossRef CAS PubMed;
(b) S. G. Lee, G. F. Brunello, S. S. Jang, J. H. Lee and D. G. Bucknall, J. Phys. Chem. B, 2009, 113, 6604 CrossRef CAS PubMed;
(c) S. G. Lee, W. Koh, G. F. Brunello, J. I. Choi, D. G. Bucknall and S. S. Jang, Theor. Chem. Acc., 2012, 131, 1206 CrossRef;
(d) S. G. Lee, T. A. Pascal, W. Koh, G. F. Brunello, W. A. Goddard III and S. S. Jang, J. Phys. Chem. C, 2012, 116, 15974 CrossRef CAS.
-
(a) G. Brunello, S. G. Lee, S. S. Jang and Y. Qi, J. Renewable Sustainable Energy, 2009, 1, 033101 CrossRef;
(b) G. F. Brunello, W. R. Mateker, S. G. Lee, J. I. Choi and S. S. Jang, J. Renewable Sustainable Energy, 2011, 3, 043111 CrossRef.
-
(a) S. G. Lee, J. I. Choi, W. Koh, S. S. Jang, J. Kim and G. Kim, IEEE Trans. Compon., Packag., Manuf. Technol., 2011, 1, 1533 CrossRef CAS;
(b) S. G. Lee, S. S. Jang, J. Kim and G. Kim, IEEE Trans. Adv. Packag., 2010, 33, 333 CrossRef CAS.
- S. Plimpton, J. Comput. Phys., 1995, 117, 1 CrossRef CAS.
- W. C. Swope, H. C. Andersen, P. H. Berens and K. R. Wilson, J. Chem. Phys., 1982, 76, 637 CrossRef CAS.
-
(a) S. Nose, Mol. Phys., 1984, 52, 255 CrossRef CAS;
(b) S. Nose, J. Chem. Phys., 1984, 81, 511 CrossRef CAS;
(c) S. Nose, Mol. Phys., 1986, 57, 187 CrossRef CAS.
- D. R. Lide, CRC handbook of chemistry and physics, CRC Press, Inc., Boca Raton, FL, 71st edn, 2005 Search PubMed.
-
(a) N. E. Singh-Miller and N. Marzari, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 235407 CrossRef;
(b) H. Iddir, V. Komanicky, S. Ogut, H. You and P. Zapol, J. Phys. Chem. C, 2007, 111, 14782 CrossRef CAS;
(c) S. Baud, C. Ramseyer, G. Bihlmayer, S. Blugel, C. Barreteau, M. C. Desjonqueres, D. Spanjaard and N. Bernstein, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 70, 235423 CrossRef;
(d) J. L. F. Da Silva, C. Stampfl and M. Scheffler, Surf. Sci., 2006, 600, 703 CrossRef CAS;
(e) G. Boisvert, L. J. Lewis and M. Scheffler, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, 1881 CrossRef CAS.
- G. F. Brunello, J. I. Choi, D. A. Harvey and S. S. Jang, ECS Trans., 2012, 50, 155 CrossRef.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09274h |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.