Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.

1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

Computational study of bulk and surface


properties on ruthenium oxide (RuO2)
Mmeshi J. Hiine1, Brian Ramogayana1, Phuti E. Ngoepe1, and Khomotso P. Maenetja1
1 MaterialsModelling Centre, School of Physical and Mineral Sciences, University of Limpopo,
Private Bag x1106, Sovenga 0727, South Africa.

Abstract. Metal oxides are widely used in lithium-air batteries to improve


the formation of stable discharge products and improve lifespan and
electrochemical performance. Despite the intense studies on metal oxides
catalysts, ruthenium oxide attracted the most attention since it doesn’t only
catalyse the redox processes but reduces the over-potential and stabilizes the
Li cyclability. Hence, in this work we discuss the bulk and low Miler index
surfaces of RuO2 using the first principle density functional theory
calculations. It was found that the lattice parameters are in good agreement
with the reported results, with less than 1.4% difference. Furthermore, RuO2
was also found to be mechanically stable with all positive independent
elastic constants (Cij) obeying the mechanical stability criteria and a positive
tetragonal shear modulus (C’> 0). The bulk to shear ratio indicates that the
structure is ductile. The density of states shows a slight pseudo gap for RuO 2
at the Fermi energy, which suggests that the structure is stable. Finally, low
Miller index surfaces (i.e. (110), (010), (001), (111), and (101)) were
modelled using METADISE code, and the most stable facet was in
agreement with the reported literature.

1 Introduction
The whole world is moving away from the reliance on fossil fuels and the quest of finding
alternative energy sources which are environmentally friendly, clean and affordable
continues. Various alternative energy sources are considered, which includes solar energy
[1], biomass [2], hydropower [3], geothermal energy [4] and etc. Among the listed energy
sources, lithium-ion battery continues to grow and advance for use in electronic devices,
electric cars, and large-scale stationary storage systems. However, their output capacity is
limited to 150 mAh/g which is due to the cathode-electrolyte interactions, resulting in
transitional-metal dissolution and irreversible structural transformation upon cycling [5, 6].
Lithium-air batteries (LABs) appeared as an alternative next-generation energy
source with high energy density value of 3505 Wh/kg, which is close to that of a gasoline
engine and 10 times greater than that of lithium-ion batteries [7, 8]. Unlike Li-ion batteries,
the LABs have its positive electrode as metal (i.e. lithium) and a negative porous electrode
that allows oxygen in from the surrounding (hence it’s called the breathing battery). During
discharge processes, the oxygen drawn from the atmosphere through the porous cathode is
reduced by the lithium ions to form discharge products such as Li2O2 and Li2O, then the

© The Authors, published by EDP Sciences. This is an open access article distributed under the terms of the Creative
Commons Attribution License 4.0 (http://creativecommons.org/licenses/by/4.0/).
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

oxides produced further decompose to form back the lithium ions and oxygen during the
charging process [9, 10]. Several studies demonstrated that the performance of the battery
depends on which oxide is produced as a discharge product [11, 12, 13].
However, the fundamental challenge that limits the use of metal-air battery
technology is the production of unstable discharge products which further interact with other
components of the battery resulting in capacity fading. Several catalysts have been proposed
to promote the formation and decomposition of stable discharge products during
charge/discharge cycle, which are the transitional metal-oxides [14, 15], noble metal/metal
oxides [16] and carbon-based materials [17, 18, 19]. Among all, metal oxide catalyst gave an
improved capacity retention and improved electrochemical performance [20]. These metal
oxide catalysts include manganese dioxide (MnO2) [21], cobalt oxide (Co3O4) [22],
ruthenium dioxide (RuO2) [23], and titanium dioxide (TiO2) [24], which can significantly
reduce charge over-potential and increase the cycling performance [25].
Metal oxides are widely considered for catalysis in various applications, such as
carbon conversion/energy conversion [26], water splitting [27], etc. In lithium-air batteries,
the use of metal oxides nanostructure showed an improved formation and decomposition of
stable Li2O2 discharge product [28]. Subsequently, Maenetja et al [29] studied the formation
of both Li2O2 [30] and NaO2 [31]stable discharge products on the MO2 (M = Mn, Ti, V),
where it was indicated that the formation of Li2O2 gas-phase is more favored in the presence
of 𝛽-MnO2 catalyst.
One other catalyst is the ruthenium dioxide, which resulted in low charge over-
potentials due to its high catalytic activity during cycling, hence long-term cycle stability
[32]. O'Dwyer et al. [33] indicated that the RuO2 catalyst can still decompose Li2O2, implying
that RuO2 has strong OER activity. Liao et al [34] RuO2 was designed and studied as a 2D
nanosheet structure, which had a high surface-to-mass ratio and low weight, allowing the
material to have a higher electrolyte-electrode interaction and enhance O2 diffusion.
Additionally, because most solid-state catalyst materials are metal-based and have a high
density, the porous structure and large surface area of the carbon material may help reduce
the catalyst material's loading mass and raise the battery's specific energy.
Here, we discuss the density functional theory calculations of ruthenium dioxide
bulk and surfaces. We analyze the bulk structural properties, elasticity, density of states, low
Miller index surface and particle morphology with reference to the reported literature.

2 Method
Spin-polarised density functional theory (DFT) calculations were performed with the Vienna
Ab-initio Simulation Package (VASP) program [35], using the generalized gradient
approximation (GGA) in the form of the Perdew–Burke–Ernzerhof (PBE) exchange-
correlation functional [36]. The cut-off energy was fixed at 500 eV and the k-mesh of 6×6×8
and 6×6×1 was set for the bulk and surfaces, respectively. The semi-empirical method of
Grimme with the Becke-Johnson damping [D3-(BJ)] [37, 38] was included to model the long-
range dispersion interactions and describe the surfaces properly [39, 40, 41, 42]. To model
the Ru 3d electrons, the effective parameter, Ueff = 3.9 eV was used, which was consistent
with our previous studies [43]. Gaussian smearing with a width of 0.05 eV was included [38]
to improve the convergence of the Brillouin zone integrations during geometry optimizations.

3 Results and discussion

2
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

3.1 Bulk properties

3.1.1 Equilibrium lattice parameters

The crystal structure of ruthenium oxide (RuO2) – rutile phase has a tetragonal symmetry
with a space group P42/mnm (no. 136). It has a body-centred structure with geometry
coordination of octahedral tetravalent ruthenium (Ru4+) ions and trigonal planar oxygen
coordinated (O2-) ions. Upon full optimization, the predicted lattice parameters were a = b =
4.49 Å and c = 3.11 Å, which are in agreement with the listed literature data within <1.4%
[44, 45, 46, 47]. The heat of formation of this structure is -380.66 kJ/mol which indicate that
RuO2 is thermodynamically stable and the calculated ∆Hf was in agreement with the reported
literature.

Table 1. Equilibrium lattice parameters and heats of formation for RuO2 with reference to the
reported literature.
This work Ref.
GGA+U
Mehtougui Tse et al. Wang et al. Sun et al.
et al [44] [45] [46] [47]
Lattice a=b 4.49 4.449 4.554 4.54 4.499
parameter
c 3.11 3.099 3.137 3.13 3.107
(Å)
∆Hf -380.66 Cordfunke Nell et al
(kJ/mol) et al [48] [49]
-314.15 -313.92

3.1.2 Mechanical properties: Elasticity


Table 2 summarises the calculated elastic constants, moduli, and the Pugh's ratio at the strain
of 0.005 for RuO2. The elastic constants can be used to gain insight on the mechanical
stability of solids. A crystal structure is said to be mechanically stable if strain energy is
positive (which implies that all the elastic constants should be positive) and it should satisfy
the stability criteria for its respective crystal. For tetragonal phases, there are six independent
elastic constants, which are C11, C12, C13, C33, C44, C66. The stability following are the
tetragonal system's mechanical stability requirements [50]:
2
2𝐶13
C44 > 0, C66 >0, C66 > |𝐶12 | and C66 + C66 - (1)
𝐶33
𝐶11 −𝐶12 𝐶11 −𝐶12 2𝐶44
B= 3
, C'= 2
,A=𝐶 (2)
11 −𝐶12
Table 2 summarizes the calculated and their respective reference elastic constants
obtained from different reported data. The elastic constants in table 2 comply with the above
mentioned mechanical stability requirements, proving that they are mechanically stable since
they are all positive. We observed a major difference with the listed available reference data
since we have included the U-parameter of 3.9 eV to model the localized 3d Ru electrons.
RuO2 is mechanically stable as shown by our structure's expected tetragonal shear modulus
(C'), which is positive and equal to 33.5. The B/G ratio of 1.94, which is higher than 1.75,
indicates that the structure is ductile.

Table 2. The calculated elastic constants (Cij), moduli, tetragonal shear modulus (C'), Pugh’s ratio
(B/G) and the anisotropy (A).

3
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

This work Ref.


DFT+U-D3
LDA method CASTEP Exp. And
[44] GGA-PBE DFT(GGA and
[51] LDA)
[52]
Elastic C11 328.39 364.99 303 308.38
constants
C12 221.39 294.24 240 214.49
(GPa)
C13 180.58 224.86 190 171.05
C33 520.70 626.69 536 508.95
C44 123.84 150.61 117 108.63
C66 231.13 258.67 221 208.22
Moduli B 214.75 - - -
G 110.68 124.96 104 -
Y 283.37 330.73 275 -
C' 33.5 - - -
Pugh’s ratio k = B/G 1.94 - - -

3.1.3 Density of states

In order to correlate the structural and mechanical stability of the RuO2 system, we compare
their total density of states (tDOS) plots in Figure 1. It is noted from literature that the DOS
of structures of the same composition can be used to mimic the stability trend with respect to
their behaviour at the Ef (Fermi level), in which the structure with the highest and lowest
density of states at Ef is considered the least and most stable, respectively [53, 54]. The total
DOS and partial DOS obtained from bulk RuO2 oxide are plotted in figure 1. The valence
band width in the previous work was -8.1 eV, while the overall valance band width in this
study is -8 eV. Additionally, there is a partially filled electronic band located right at the
Fermi level [55]. The partial DOS shape demonstrates that the O-2p state dominates the
valence band at lower energy (O atoms adopt the sp2 hybridization), but the top section
predominantly arises from the Ru–4d state, with the Ru4+ peak occurring at -1 eV. The
conduction band began with greater Ru-4d than O-2p states [56]. These PDOS features are
in qualitative agreement with results from previous studies [57, 58]. We observe a sight
pseudo-gap at the Fermi energy (Ef), indicating that our structure is electrically stable.

4
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

Fig. 1. The total and partial density of states for RuO 2.

3.2 RuO2 Surfaces

3.2.1 Surface models

All the surface terminations were generated by cutting the geometry-optimized (2×2)
supercell bulk structure [59], using the Tasker [60] dipole method, as implemented in the
Minimum Energy Techniques Applied to Dislocations, Interfaces and Surface Energies
(METADISE) code [61]. When cleaving the surfaces, symmetry and stoichiometry were
preserved, thus no electric dipole moments. When constructing the surface terminations, we
considered the stacking sequence for low Miller index facets. We have modelled the two
possible terminations for each surface orientation using stoichiometric, non-polar, and
symmetric slabs along the z-direction. We then considered the stabilities of five low-index
surface orientations ((110), (010), (001), (111) and (101)) by performing periodic
calculations of slabs with stoichiometric composition, and vacuum gaps (see Figure 2). The
total number of layers and surface areas of each surface are showed in Table 3. All the

5
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

surfaces have 6 atoms with 2 formula units. In each surface, a vacuum slab of 15 Å was added
perpendicular to the slab to avoid periodical interactions.

Fig. 2. Side view of the low-index (110), (100), (101), (001) and (111) surfaces of rutile RuO2.

3.2.2 Surface energies and particle morphology


Table 3 summarizes the calculated surface energies of RuO2 using the expression:
𝐸𝑠𝑙𝑎𝑏 −𝐸𝑏𝑢𝑙𝑘
𝛾= (3)
2𝐴
where 𝐸𝑠𝑙𝑎𝑏 and 𝐸𝑏𝑢𝑙𝑘 are the total energy of the slab and relaxed bulk, and A is the area
of the slab. The (110) surface with the lowest surface energy of 𝛾 = 0.045 eV/Å2 was found to
be the most stable facet, which is in agreement with the reported literature data of 0.0645
eV/Å2 , [62]. and 0.078 eV/Å2 [56]. Similar to other rutile phases, such as IrO2 [63] and TiO2
[64], the (110) surface was previously observed to be the most stable facet. As compared to
other modelled facets, we observed an increasing surface energy, and decreasing stability,
i.e. (110) < (101) < (111) < (001) < (100).
Table 3. Terminations, Surface area (Asurface), the number of layers (Nlayers) and surface
energy for low miller index RuO2 surfaces
Surfaces Terminations (Nlayers) Surface Surface energy
(Å2) (eV/Å2)
(110) -Ru-O- 9 19.502 0.045
(101) -O- 10 25.012 0.075
(111) -O- 7 24.001 0.132
(001) -Ru-O- 9 21.607 0.170
(100) -Ru-O- 11 15.738 0.482

6
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

These calculated surface energies were utilized to determine the equilibrium RuO2 particle
morphology using Wulff's method [65]. Similarly to isostructural transition metal oxides like
TiO2 [66], MnO2 [67]and VO2 [68], the most enhance plane is for (110) surface. The (001),
(101) and (111) surface does not appear in the Wulff morphology because of its higher
surface free energy with respect to the (110) and (011) planes.

Fig. 3. Wulff's construction of the equilibrium morphology for a RuO 2.

Conclusion
The ruthenium dioxide bulk and surfaces stability were successfully evaluated using density
functional theory (DFT). It was found that the lattice parameters are in good agreement with
the reported results, with less than 5% difference. Furthermore, RuO 2 was also found to be
mechanically stable with all positive independent elastic constants (Cij) obeying the
mechanical stability criteria and a positive tetragonal shear modulus (C’> 0). The bulk to
shear ratio indicates that the structure is ductile. The density of states shows a slight pseudo
gap for RuO2 at the Fermi energy, which suggests that the structure is stable. The electrical
conductivity of RuO2 bulk can therefore be correlated with the number of free d-electrons of
the metal ion, confirming that RuO2 to be a metallic-like conductor. The most stable surface
was found to be the (110) surface and we observed the surface stability trend of (110) < (101)
< (111) < (011) < (100). The morphology of RuO2 is strongly consistent with previous studies
and other rutile phase structures, like MnO2.

References

[1] R. Prăvălie, I. Sîrodoev, J. Ruiz-Aria and M. Dumitraşcu, Renew.


enrgy, 193, (2022)
[2] E.H. Kalyesubula, Ceramics and Chemical Engineering Divisio,,
(2006)
[3] W. Yuan, S. Zang, C. Su, D, Yan and Z. Wu, 252 (2022).
[4] T.H. Çetin, J. Zhu, E. Ekici and M. Kanoglu, Energy Convers and
Manage, 260 (2022)
[5] J. Wandt, A. Freiberg, R. Thomas, Y. Gorlin A. Seibel, R. Jung, H.A.
Gasteiger and M. Tromp, J. Mat Chem. A,47 (2016).
[6] S. Choi and G. Wang, Ad. Mat. tech, 3, (2018)

7
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

[7] D. Capsoni, M. Bini, S. Ferrari and P. Mustarelli, Carbon Nanomater


Ad. Energy Sys, 1, (2015)
[8] S.P.P. Badwal, S.S. Giddey, C. Munnings, S.I. Bhatt and A.F.
Hollenkamp, Front. Chem, 2 (2014).
[9] T. Ogasawara, A. Débart, M. Holzapfel, P. Novák and P.G. Bruce, J.
Am. Chem. Soc.,128 (2006)
[10] J. Read, J. Electrochem. Soc., 149 (2002)
[11] S. Kang, Y. Mo, S.P. Ong and G. Ceder, Chem. Mater., 25(2013).
[12] J.S. Hummelshøj, A.C. Luntz and J.K. Nørskov, J. Chem. Phys. 138
(2013)
[13] Y. Mo, S.P. Ong and G. Ceder, Phys. Rev. B, 84 (2011)
[14] Y.C. Lu, H.A. Gasteiger and Y. Shao-Horn, J. Am. Chem. Soc., 133
(2011)
[15] W. Choi, S. Jee, H.K. Cho, C.H. Ahn, G. Yang, J. Lee and C. Yu, J.
Mater. Chem. A, 3 (2015)
[16] E. Yilmaz, C. Yogi, K. Yamanaka, T. Ohta and H.R. Byon, Nano
Lett., 13 (2013)
[17] X. Lin, Y. Zhang, L. Li and A Yu, ACS nano, 5, (2011)
[18] D. Zhai, H.H. Wang, J. Yang, K.C. Lau, K. Li, K. Amine and L.A.
Curtiss, Curtiss, J. Am. Chem. Soc., 135 (2013)
[19] S. Wang, S. Dong, J. Wang, L. Zhang, P. Han, C. Zhang, X. Wang, K.
Zhang, Z. Lan and G. Cui, J. Mater. Chem., 22 (2012)
[20] M. Lee, Y. Hwang, K-H. Yun and Y-C, Chung, J. power Sources,
288, (2015)
[21] N.V. Boas, J.B.S. junior, L.C. Varanda, S.A. Machado and M.L.
Calegaro, Appl Catal B: Environ, 258, (2019)
[22] S-H. peng, T-H. Cheng, C-H. Lee, H-C. Lu and S.J. Lue, j. power
Sources, 471 (2020)
[23] Y. Liu, B. Li, Z Cheng, C. Li, X. Zhang, S. Guo, P. He and H. Zhou,
J. Power Sources, 396 (2018)
[24] G. Zhao, Y. Niu, Li. Zhang and K. Sun, J. Power Sources, 270 (2014)
[25] A. Eftekhari and B. Ramanujam., J. Mater. Chem., 5 (2017)
[26] W. Zhua, X. Chen, J. Jin, X. Di, C. Liang and Z. Liu, Chinese J. Catal,
41(2020)
[27] C. Wang, L. Jin, H. Shang, H. Xu, Y. Shiraishi and Y. Du, Chin Chem
Lett, 32 (2021)
[28] A. Débart,A.J. Paterson,Ji. Bao,P.G. Bruce , Angew. Chem. Int. Ed.,
47 (2008)

8
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

[29] T.A. Mellan, K.P. Maenetja, P.E. Ngoepe, S.M. Woodley, C. Richard,
A. Catlow and R. Grau-Crespo , R. Soc. Chem, 47 (2013)
[30] K.P. Maenetja and P.E. Ngoepe, ACS Omega, 7 (2022)
[31] K.P. Maenetja and P.E. Ngoepe, “First Principles Study of Oxygen
Adsorption on Li-MO2 (M = Mn, Ti and V) (110) Surface,” Journal
of The Electrochemical Society, vol. 168, p. 1, 2021.
[32] C. Li, X. Zhang, S. Guo,Z. Cheng, P. He and H. Zhou, “J. Power
Sources, 395 (2018
[33] S. Vankova, C. Francia, J. Amici, J. Zeng, S. Bodoardo, N. Penazzi,
G. Collins, H. Geaney and C. O'Dwyer, ChemSusChem., 10 (2017).
[34] K. Liao, X. Wang, Y. Sun, D. Tang, M. Han, P. He, X. Jiang, T.
Zhang and H. Zhou, Energy & Environ Science, 8 (2015)
[35] G. Kresse and J. Futhmuller., Phys. Rev., 54 (1996)
[36] J.P Perdew, Phys. Rev. B, 72, (1992)
[37] S. Grimme, S. Ehrlich and L. Goerigk, J. com chem, 32 (2011)
[38] S. Grimme, J. Antony, S. Ehrlich and H. Krieg, The J. chem phys, 132
(2010)
[39] D. Santos-Carballal, A. Roldan, R. Grau-Crespo and N.H. de Leeuw,
PhysChem,16 (2014)
[40] D. Santos-Carballal, A. Roldan and N.H. de Leeuw, The J. Phys Chem
C, 120 (2016)
[41] D. Santos-Carballal, A. Roldan, N.Y. Dzade and N.H. de Leeuw, Ser
A, Mathl, phys, and eng sci, 376 (2018)
[42] V. Postica, A. Vahl, J. Strobel, D. Santos-Carballal, O. Lupan, A.
Cadi-Essadek, N.H. de Leeuw, F. Schütt, O. Polonskyi, T. Strunskus,
M. Baum, L. Kienle, R. Adelung and F. Faupel, J.f Mat Chem A, 6
(2018)
[43] S.C. Abrahams and J.L. Bernstein, j. physi chem, 55 (1971)
[44] N. Mehtougui, D. Rached, R. Khenata, H. Rached, M. Rabah and S.
Bin-Omran, Mats sci in semicon proces,15 (2012)
[45] J.S. Tse, D.D. Klug, K. Uehara, Z.Q. Li, J. Haines and J.M. Léger,
Geochim Cosmochim Act, 61(1997)
[46] J. Wang, C. Cheng, Q. Yuan, H. Yang, F. Meng, Q. Zhang, L. Gu, J.
Cao, L. Li, S.C. Haw and Q. Shao, Chem, vol. 8 (2022)
[47] S.C. Sun, H. Jiang, Z.Y. Chen, Q. Chen, M.Y. Ma, L. Zhen, B. Song
and C.Y. Xu, Angew Chem Int Ed 61 (2022).
[48] E.H.P. Cordfunke and R.J.M. Konings, Thermochimica Acta, 129
(1988)

9
MATEC Web of Conferences 370, 02003 (2022) https://doi.org/10.1051/matecconf/202237002003
2022 RAPDASA-RobMech-PRASA-CoSAAMI Conference

[49] St. C. Hugh and J. Nell, Geochim Cosmochim Act, 61(1997)


[50] M.J. Mehl, B.M. Klein and D.A. Papaconstantopoulos, Intermetal
Compd: Principles and Practice, 1 (1995).
[51] Y. Li and Z. Zeng, Inter J.f Modern Phys C, 19 (2008)
[52] K.P. Bohnen, R. Heid, O. de la Pena Seaman, B. Renker, P. Adelmann
and H. Schober, Phys Review B, 75 (2007)
[53] Yu.N. Gornostyrev, O.Yu. Kontsevoi, A.F. Maksyutov, A.J. Freeman,
M.I. Katsnelson, A.V. Trefilov, et al., Phys. Rev. B.,(2004).
[54] D.A. Pankhurst, D. Nguyen-Manh, D.G. Pettifor, Phys. Rev. B., 69
(2004)
[55] J. Riga, C. Tenret-noel, J.-J, pireaux, R. Caudano, J. Verbist and Y.
Gobillon, Phys. Scr, 16 (1977)
[56] F. Creazzo and S. Luber, App Surf SCi, 570 (2021)
[57] J. De Almeida and R. Ahuja, Phys. Rev. B, 73 (2006)
[58] N. Mehtougui, D. Rached, R Khenata, H. Rached, M. Rabah and S.
Bin-Omran, Mater. Sci. Semicon. Proces, 15 (2012)
[59] C. Y. Ouyang , S. Q. Shi and M. S. Lei, J. Alloys Compd., 474 (2009)
[60] P.W. Tasker, J. Phys. C: Solid State Phys., 12(1972)
[61] G. W. Watson , E. T. Kelsey , N. H. de Leeuw , D. J. Harris and S. C.
Parker, J. Chem. Soc., Faraday Trans., 92 (1996)
[62] H. Wang and W.F. Schneider, J. Chem Phys, 127 (2007)
[63] S. Albertin, L.R. Merte, . Lundgren. R. Martin, J.F. Weaver, A-C.
Dippel, O. Gutowski and U. Hejral Phys Chem, vol. 126 (2022).
[64] M.J.S. Abb, T. Weber, L. Glatthaar,and H. Over, Langmuir, 35,
(2019)
[65] G. Wulff Z. and K. Miner., 34 (1901)
[66] R. R. Maphanga, S. C. Parker and P. E. Ngoepe, Surf Sci, vol. 603
(2009)
[67] T.A. Mellan, K.P. Maenetja, P.E. Ngoepe, S.M. Woodley. R.A.
Catlow and R. Grau-Crespo, Surf Sci, vol. 603 (2009)
[68] Mellan, T.A. and Grau-Crespo R, J. Chem. Phys, 137 (2012)

10

You might also like