Vous êtes sur la page 1sur 50

Article

DiffusionTheKineticsJournalofPhysical
Chemistry C is published
by the American

of Gold and Copper


Chemical Society. 1155
Sixteenth Street N.W.,
Atoms on Pristine
Washington, DC 20036
Published by American
Subscriber access provided by University of
Chemical Society.

Groningen
Copyright © American
Chemical Society.
and Reduced Rutile
TiO (110) TheSurfacesJournlofPhyical
Chemistry C is published
Mathilde Iachella,bythe TanguiAmerican Le
Bahers, and DavidChemicalLoffredaSociety.1155
Sixteenth Street N.W.,
Washington, DC20036
J. Phys. Chem. C ,JustAccepted
Subscriber access provided PublishedbybyUniversityAmerican of Manuscript • DOI: 10.1021/
Chemical Society.

Groningen
Copyright © American
Chemical Society.
acs.jpcc.7b08183 • Publication
Date (Web): 12 Dec 2017
The Journal of Physical
Downloaded from http://
Chemistry C is published pubs.acs.org on December 17, 2017
by the American
Chemical Society. 1155
Sixteenth Street N.W.,
Washington, DC 20036
Published by American
Subscriber access provided by University of
Chemical Society.

Groningen
Copyright © American
Chemical Society.
Just Accepted
The Journal of Physichavel
“Just Accepted” manuscripts been p
Chemistry C is published
online prior to technicalbytheAmediting,rican formatting f
Chemical Society .1155
Society provides “Just Accepted” as a
Sixteenth Street N.W.,
dissemination of scientific material as soon
Washington, DC 20036
appearSubscriberin accessfullin providedPDFPublishedformatbybyUniversityAmeriaccompaniedof b
Chemical Society.
Groningen
Copyright © American
Chemical Society.
fully peer reviewed, but should not be consi readers
and citable by the Digital Object Ide to authors.
TheJournal of Physical
Therefore, the “Just Accepted
Chemistry C is published
in the journal. After a manuscript is technic
by the American
Accepted” Web siteChemicaland Societypublished.1155 as an A
Sixteenth Street N.W.,
changes to the manuscript text and/or gra
Washington, DC 20036
Published by American
Subscriber access provided by University of
Chemical Society.

Groningen
Copyright © American
Chemical Society.
and ethical guidelines that apply to the jo or
consequences arising from the use of in
The Journal of Physical
Chemistry C is published
by the American
Chemical Society. 1155
Sixteenth Street N.W.,
Washington, DC 20036
Published by American
Subscriber access provided by University of
Chemical Society.

Groningen
Copyright © American
Chemical Society.
Page 1 of The41 Journal of Physical Chemistry

1
2
3
4 ACS Paragon Plus Environment
5

6
7
The Journal of Physical Chemistry Page 2 of 41

1
2
3
4
5
6
7
8 Diffusion Kinetics of Gold and Copper Atoms on
9
10
11
12
Pristine and Reduced Rutile TiO2 (110) Surfaces
13
14
15 Mathilde Iachella, Tangui Le Bahers, and David Loffreda
16
17
18 Univ Lyon, Ens de Lyon, CNRS UMR 5182, Université Claude Bernard Lyon 1, Laboratoire de
19
20 Chimie, F-69342, Lyon, France.
21
22
23 E-mail: david.loffreda@ens-lyon.fr
24
25 Phone: +33 (0)472728843. Fax: +33 (0)472728860
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
To whom correspondence should be addressed
57
58
59 1
60 ACS Paragon Plus Environment
Page 3 of 41 The Journal of Physical Chemistry

1
2
3
Abstract
4
5
6 Statistical mechanics and transition-state theory have been used to investigate diffusion
7
8 kinetics of gold and copper atoms on pristine and various reduced surfaces of rutile TiO 2 (110). A
9
10 DFT+U approach has been employed to calculate potential energy maps and to evaluate the
11
12 required diffusion activation barriers. The role of the support reducibility has been examined on
13
adsorption properties (optimal structures, energetics and spin polarization) and diffusion kinetics,
14
15 especially for the reduced support presenting a single sub-surface oxygen vacancy. This
16
17 approach has allowed us to demonstrate key discrepancies between Au and Cu atoms and to
18
sketch out a comparative scenario for the early stage nucleation of Au and Cu nanoparticles on
19
20
21 the various surface states of TiO2 (110).
22
23
24
25
26 1. Introduction
27
28
29 In the domains of materials science and heterogeneous catalysis, the nucleation and
30
31
the growth of supported metallic nanoparticles are difficult processes to comprehend
32 1–4
33 from experimental and theoretical standpoints. In fact, little is known regarding
34
35 adsorption thermodynamics (metal-lic atoms, nanoclusters, nanoparticles) or diffusion,
36
nucleation and growth kinetics, especially at the atomic scale. For instance, those
37
38 elementary processes remain mostly unsolved for the nu-cleation of gold nanoparticles
39
40 5–7 8–11
41 supported on rutile TiO2, although the experimental literature is abundant.
42
43 Since gold and copper nanoparticles supported on rutile TiO 2 (110) are of particular interest for
44
45 CO oxidation, they have been examined with various experimental techniques such as scanning
46
12–15 16
47 tunneling microscopy (STM), high-resolution scanning electron microscopy (HRSEM), high-
48
49 17,18
angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), and
50
51 3,4,19
52 grazing incidence small-angle X-ray scattering (GISAXS). From those measurements,
53 information about the distribution of the nanoparticle size, the general shape of nanoparticles, their
54
55 number of atomic planes and their aspect ratio is obtained from low temperature ultra-high vacuum
56
57 (UHV) to operando conditions. In contrast, very little is known about diffusion kinetics. In addi-
58
59 2
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 41

1
2
3 20 21 22
4 tion, experimental results (LEED, SXRD, STM ) concerning copper nanoparticles
5
6 deposited on rutile TiO2 (110) are really scarce in the literature.
7 In spite of the accuracy of the experimental methods, there are still open questions
8
9 related to the nucleation and the growth mechanisms and kinetics of Au and Cu
10
11
12
nanoparticles on TiO2 (110), especially for the interpretation of STM measurements. Among
13
the related questions, the influence of the heterogeneity and the defects of the support
14
15
16 (surface and subsurface oxygen vacancies, steps, TiO x clusters, hydroxylation, interstitial Ti
17
18
atoms, etc) is often evoked in the literature but rarely solved at the atomic scale.
19 23–42
20
In order to advance on these questions, density functional theory (DFT) studies have been
21 proposed during the last twenty years and compared to available measurements. The proposed
22
23 theoretical models concern the interaction of Au or Cu on rutile TiO 2 (110) from the single atom to
24
25 30–32,35,36,39,41,42
clusters (less than 25 atoms). Regarding the adsorption of Au on stoichiometric
26
27 TiO2 (110), there is no clear consensus so far, neither for the adsorption site nor for the adsorption
28
29 26,40 24,25,27,29,31,38,39
energy. For the structure, various top and bridge sites have been reported.
30
31 1 24–27,29,31,38,40
The adsorption energy spans from -10 to -100 kJ.mol . In contrast, for Au adsorption
32
33
on a singly-reduced TiO2 (110) surface, the best structure is proposed on the oxygen vacancy and
34
35 1 24–27,32,37,38,40
the adsorption energy ranges from -175 to -269 kJ.mol .
36
37
38 For the adsorption of Cu on stoichiometric TiO 2 (110), there is an agreement regarding
39 23,29,40
40 the most stable structure (a bridging position on oxygen terminal rows). The
41 1 23,29,40
42
adsorption energy varies in the range -230 to -265 kJ.mol . On a singly-reduced
43
TiO2 (110) surface, the adsorption properties have been examined on the oxygen vacancy
44
45 1 40
(for a density of 1/8 ML, ML=monolayer); the adsorption energy being -185 kJ.mol .
46
47
A few results are available in the literature about diffusion kinetics for Au and Cu on rutile TiO 2
48
49 (110). On the basis of low energy ion scattering (LEIS) and an Arrhenius plot of Au critical coverage
50
51 against temperature, the activation energy for Au monomer diffusion on TiO 2 (110) has been
52
53 1 43
evaluated to 8.8 kJ.mol . In addition, the difference of adsorption and diffusion energies has been
54
55 1 44
measured (41 kJ.mol ), hence leading to an estimation of the adsorption energy of
56
57
58
59
3
ACS Paragon Plus Environment
60
Page 5 of 41 The Journal of Physical Chemistry

1
2
3 1
49.8 kJ.mol for Au on TiO2 (110). From a combined STM and DFT study, a strong Au-vacancy
4
5 8
6 interaction has been proposed, defining Au atom-vacancy complexes that can diffuse. When ana-
7 lyzing sequences of STM images, no evidence for diffusion of single Au atom-vacancy complexes
8
9 has been found at any investigated temperatures. The potential energy maps of Au adsorption and
10
11 diffusion on stoichiometric and surface reduced TiO2 (110) surfaces have been addressed in the
12
27–29
13 literature. The authors have demonstrated that there is no preferential direction for Au diffu-sion
14
15 1
on the stoichiometric support, whereas the calculated activation barrier is 40 kJ.mol on the reduced
16
17 28,29
surface (along Ti5C row for a Au atom close to the free oxygen vacancy located in the surface).
18
19 1
This barrier is decreased to 25 kJ.mol , if the vacancy is occupied by an additional Au atom. Those
20
21 1
22 results are compatible with another DFT study concluding to barriers of 16 and 19 kJ.mol for Au on
23 27
24 stoichiometric TiO2 (110). The authors have proposed a larger difference for the diffusion barriers
25
1
26 of Au on surface reduced TiO 2 (110) depending on the direction (from 58 to 178 kJ.mol ). These
27
28 calculations have been confirmed by other authors on both stoichiometric and surface reduced
29 37,39
30 supports. The isotropy of diffusion commonly seen at zero temperature by DFT studies for Au
31
32 on stoichiometric TiO2 (110) has been questioned by a more recent ab initio molecular dynamics
33
38,45 1 40
34 approach and a static DFT study (15 and 55 kJ.mol ).
35
36
Concerning the diffusion of Cu on stoichiometric TiO 2, Pillay et al. have estimated
37
38 1 29
the diffu-sion barrier to 75 kJ.mol . This result is in fair agreement with a more recent
39
1 40
40 DFT investigation (62 and 140 kJ.mol on stoichiometric surface).
41
42 Despite the numerous references available in the literature concerning the adsorption and diffu-
43
44 sion of Au and Cu atoms on stoichiometric and surface reduced TiO 2 (110), the role of subsurface
45
46 oxygen vacancies has not been investigated yet. In addition, there are still controversies regarding
47
48 Au and Cu adsorption properties on this support, from a theoretical standpoint.
49
50 In this work, we present a comprehensive study of the adsorption and diffusion properties of Au
51
52 and Cu atoms, on the stoichiometric, surface reduced and sub-surface reduced TiO 2 (110). Potential
53
54 energy maps are presented systematically for all those chemical systems on the basis of an
55
56 unrestricted DFT+U approach. From these maps, the activation energy barriers for Au and Cu
57
58
59 4
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 41

1
2
3 along two orthogonal directions over the three different support states have been extracted and
4
5 used to compute diffusion coefficients from transition state theory and Arrhenius plots. A
6
7 comparative scenario of early-stage nucleation is discussed for Au and Cu nanoparticles on this
8
9
support on the basis of diffusion kinetics and adsorption thermodynamics. The role of the
10
11
12
support reducibility and associated defects is addressed in this analysis. Finally those original
13
14
results are discussed and compared to state-of-the-art measurements and theoretical works.
15
16
17
18 2. Methodology
19
20
21
22
2.1 Computational Details
23
24 Calculations have been performed by using density functional theory (DFT) in periodic
25
26 46–48
boundary conditions, with the Vienna ab-initio simulation package (VASP ). The
27
28 generalized gradient approximation (GGA) has been used with the Perdew-Burke-Ernzerhof
29
30 49
31
exchange-correlation func-tional (PBE ), and the projector augmented-wave method
32
50
33 (PAW ). An energy cut-off of 400 eV has been considered for the expansion of the plane-
34
35 wave basis set. The kinetic energy cut-off for the augmentation charges has been set at 604
36 eV. Spin-polarized calculations have been performed throughout the study. Accurate
37
38 geometry optimizations have been ensured by tight criteria for the convergence of both total
39
40 7 1
electronic energy ( 10 eV) and residual forces acting on the nuclei ( 0.01 eV.Å ).
41
42
43
The theoretical description of rutile TiO2 (110) non-polar pristine and surface reduced surfaces
44 51
45 leads to two main issues. First, the band gap width is underestimated at the GGA-PBE level (1.7
46
47 eV vs 3.05 eV experimentally). Second, for surface reduced TiO 2 (110), i.e. surfaces exhibiting
48
49
oxygen vacancies, two electrons are released back to the oxide surface per removed oxygen atom.
50 The formation of oxygen vacancies creates an electronic state in the band gap, usually called gap
51
52 52
state (lying 2.3 eV above the valence band and 0.7 eV below the conduction band, hence at 3=4
53
54 3+
of the band gap forming a Ti ). The usual GGA functionals fail to describe the corresponding
55
56 electron localization at a metallic center. A probative semi-empirical approach, namely DFT+U,
57
58
59
5
ACS Paragon Plus Environment
60
Page 7 of 41 The Journal of Physical Chemistry

1
2
3 has been proposed to improve this spurious delocalization. An additional term U, corresponding
4
5 to electron on-site repulsion and based on a Hubbard model is considered to cure the self-
6
53
7 interaction error. In this study, the Dudarev’s approach (PBE+U) has been selected with an
8
9 optimal value Ueff of 4.2 eV, offering the best compromise between the opening of the band gap
10
11 51
(2.16 eV) and the localization of the gap state (at 3=5 of the band gap).
12
13 o
14 TiO2 rutile has a tetragonal unit cell (a = b 6=c and a = b = g = 90 ) belonging to the P42=mnm
15
54
16 space group. Our PBE+U results (with a Monkhorst-Pack k-points mesh of (9 9 15), corresponding
17
18 55
to 120 irreducible points) are in fair agreement with experimental measure-ments : atheo = 4:615 Å
19
20
21 (aexp = 4:593 Å); ctheo = 3:001 Å (cexp = 2:958 Å) and utheo = 0.304 (uexp = 0.305). The systematic
22
23 3 1
error on the band gap has been evaluated by changing the extension of the Ti atom valence (3d 4s
24
25 2 6 3 1
26 vs 3s 3p 3d 4s ), the pseudopotentials (PAW-PBE vs GW) and the choice of the non-local
27
56 57
28 dispersion corrected functional (vdW-DF2 and BEEF-vdW ). The value varies from 2.16 eV
29
30 (PBE+U, 4 valence electrons per Ti) to 2.33 eV (PBE+U, 12 valence electrons per Ti), to 1.75 eV
31
(BEEF-vdW, 12 valence electrons per Ti), to 2.2 eV (vdW-DF2, 12 valence electrons per Ti) and to
32
33 2.35 eV (PBE+U with GW pseudopotentials, 12 valence electrons per Ti). The optimal lattice
34
35 parameters are weakly perturbed by the change of the functional, the choice of the valence or the
36
pseudopotentials. According to our tests, the inclusion of dispersion makes worse the prediction of
37
38 the band gap, whereas GW pseudopotentials agree quantitatively with PAW-PBE pseudopotentials.
39
40 In this context of DFT+U, the extension of the valence up to 12 electrons offers the best compromise
41
for the band gap prediction. More sophisticated approaches with hybrid functionals or DFT+GW
42
43 58–63
44 have been considered recently to enhance the description of the band gap. Hence for an
45
46 optimal prediction of charge transfer, excess electron distribution and electronic properties, those
47
58 59 61
48 functionals (B3LYP, HSE06, screened-exchange hybrid func-tional sX, full-range hybrid
49
50 63 3+
functional ) work better. In this study, we have chosen to describe the localized states of Ti in
51
52 interaction with adsorbed metallic atoms with the DFT+U approach: PBE+U, PBE-PAW and 12
53
54
valence electrons per Ti atom. For Au adsorption on stoichiometric TiO 2 (110), the fair agreement
55
56 between our calculated DFT+U adsorption energy and the mea-
57
58
59 6
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 41

1
2
3
surements validates our approach (see section 3.3). For bulks the numeric error has been
4
5 estimated by changing the k-point grids and corresponding number of irreducible k-points. A
6
7 4
8 convergence on the total electronic energy of 2 10 eV is ensured by a (9 9 15) mesh.
9
10
11
12 2.2 Surface Models
13
14
Various surfaces of TiO2 (110) have been modeled in this work (see Figure 1). The pristine non-polar
15
16 termination (presented in Figures 1a and 1b) is described by a (2 1) supercell and terminated by
17
18 bridging oxygen rows along c. Two different reduced surfaces have been considered with a single
19
20 oxygen vacancy: either located in the surface (cf. Figures 1c and 1d) or in the first oxygen plane in
21
22
the sub-surface (cf. Figures 1e and 1f). Both reduced systems have been described by a (3 1)
23 supercell (corresponding to a vacancy surface density of 1=3 ML). The vacuum space between two
24
25 equivalent slabs along the z direction has been set at 20 Å. In all those surface models, one oxide
26
27
layer contains 3 atomic planes (following the sequence O-Ti 2O2-O). The associated k-points grids
28
29
are defined from those of the bulk: (7 7 1) and (5 7 1) for pristine and surface reduced terminations,
30
31 respectively. One important characteristic of the surface models is the choice of the thickness of the
32
33 symmetric slab based on the convergence of the surface energy. A detailed analysis of surface
34
35 energy calculations for pristine surface model as a function of the slab thickness has shown that
36
2
37 absolute convergence is obtained for a 13-layer thick slab (362 mJ.m with a systematic error of 10
38
39 2
40
mJ.m ). To make the theoretical study possible, a good compromise has been done by keeping 7-
41
42 2
layer thick slabs (leading to a larger systematic error of 70 mJ.m ). For the two models of surface
43
44 reduced systems, the convergence of the surface energy is more difficult, where more than 13-layer
45
46 thick symmetric slabs are required to reach an equivalent accuracy. This delicate question is rarely
47
mentioned in the literature. For a sake of consistency with pristine surfaces, we have modeled the
48
49
2
50 reduced surfaces with 7-layer thick slabs (with respective systematic errors of 80 and 23 mJ.m by
51
52 comparison with 11-layer thick slabs). The surface energies for the surface reduced surfaces are
53
54 2
1109 and 1331 mJ.m with a single oxygen vacancy in the surface and sub-surface layer,
55
56 respectively. As a conclusion, the presence of a vacancy in the sub-surface
57
58
59
7
60
ACS Paragon Plus Environment
Page 9 of 41 The Journal of Physical Chemistry

1
2
3 is possible for clean supports but much less favorable.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure 1:
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
Top and lateral views of rutile TiO2 (110) surfaces: (a), (b) stoichiometric surface described
56
57 by a (2 1) supercell; (c), (d) surface reduced surface described by a (3 1) supercell; (e), (f) sub-
58 surface reduced slab described by a (3 1) supercell. Black and green dashed circles in the top
59 and lateral views of the reduced TiO 2 slabs localize the oxygen vacancies in the surface and in
60 the subsurface, respectively. c parameter is defined along y axis.
Regar of which the coverage qM is 1=4 ML and 1=6 ML, respectively. In the geometry optimizations,
ding the
the bottommost oxide layer is frozen to the bulk-like optimal structure, while the 3 uppermost
adsorption
layers are allowed to relax completely.
properties
In all our models of clean TiO 2 surfaces and of adsorbed systems, the spin polarization has been
of Au and
relaxed simultaneously with the geometry optimizations. For converged results, the spin polarization
Cu on
is essentially held by the Ti atoms of the surface layers (never on the Au or Cu atoms). The spin
those TiO2
distribution depends on each system and there is no simple general trend or rule among all
surfaces,
non-
symmetric 8
ACS Paragon Plus Environment
slabs
composed
of 4 layers
or 12
atomic
planes
(equivalen
t to 7-layer
thick
symmetric
slabs)
have been
built up.
These
adsorption
s have
been
considere
d both on
pristine
and
reduced

TiO2
surfaces
with
associated
supercells
of (2 1)
and (3 1),
The Journal of Physical Chemistry Page 10 of 41

1
2
3
the considered models. Tests imposing the spin polarization to small integer values (ranging
4
5
from 0 to 3) have been performed on the optimized structures to check whether the direct
6
7
minimization had led to the most stable magnetic state. By following such a procedure, we
8
9
10
have never registered a more stable spin-polarized state than the one obtained by VASP.
11
12
13
14
2.3 Potential Energy Maps and Diffusion Kinetics
15
16 The supercells used to built the Potential Energy Maps (PEMs) of Au and Cu adsorption on various
17
18 TiO2 surfaces are those defined previously in Figure 1. The adsorption has been modeled on
19
20 2
specific points of regular grids covering the TiO 2 surfaces. The point densities are 2 points.Å and
21
22 2
0.8 points.Å for pristine and reduced surfaces, respectively. For each specific point, all the degrees
23
24
25 of freedom of the uppermost TiO 2 layers and the z coordinate of the adsorbed metal have been
26
allowed to relax during the geometry optimizations, whereas x and y adsorbate coordinates have
27
28 been frozen. The total spin polarization is also relaxed throughout the PEMs.
29
30 The adsorption energy (DEads) of the metal M (Au or Cu atom) on the various TiO 2
31
32 surfaces is defined as follows:
33
34
35
DEads = E(M/TiO2(x )) E(M) E(TiO2(x )) (1)
36
37
38
Where E(M/TiO2(x )) is the total electronic energy of the adsorbed system, E(M) being the total
39
40
electronic energy of the gas phase metal M reference and E(TiO 2(x )) being the one of the bare
41
42
support surface. x refers to the three different TiO 2 surfaces : 0 for pristine (Sto), 1 for surface
43
44 2
reduced (Red) and sub-surface reduced (sRed). The areas of these supercells are A =39.2 Å for
45
46 2
stoichiometric surfaces and A =58.8 Å for surface reduced and sub-surface reduced slabs.
47
48 For each metal and each particular surface, a PEM is built by an interpolation over the series of
49
50 adsorption energies calculated for each point of the exploration grid. Afterwards, the geometry of
51
52 each point of the grid is fully relaxed, in order to determine all the stationary points in the map (local
53
54 minima, local maxima and saddle-points). The nature of all the corresponding states (lo-
55
56
57
58 9
59 ACS Paragon Plus Environment
60
Page 11 of 41 The Journal of Physical Chemistry

1
2
3 cal minimum, transition state or second-order saddle-point) is validated by a vibrational analysis,
4
5 exposed in details in the Supporting Information (SI). In this analysis, the phonons of the relaxed
6
7
8 surface layer of the TiO 2 (110) surfaces are calculated at the G point (periodically in-phase
9
10 vibra-tions), as well as all the relaxed degrees of freedom of the adsorbed metal. In particular,
11
12 the local minima show only positive force constants and real frequencies while the transition
13
14 states exhibit one negative force constant or a single imaginary frequency. The difference of
15
16 energy between the fully relaxed geometries and constrained structures coming from the PEMs
17
allows the estimation of the systematic error related to the use of those maps ( 0:01 eV). Once
18
19 the most stable adsorp-tion structure is found, the closest point is defined on the grid and the
20
21 diffusion activation barriers of Au and Cu atoms toward the next stable position are calculated in
22
23 each case by considering their diffusion either through the bridging oxygen rows (named
24
25 "perpendicular diffusion", along x axis) or along those oxygen rows (named "parallel diffusion",
26
27 along y axis). These effective barri-ers have been evaluated by calculating the energy difference
28
29 between the highest saddle-point and the global minimum along those two directions. The
30
31 64
32
corresponding diffusion coefficients of each metal M have been calculated as follows:
33
34
35
36 DEdiff
37 D= act
t (2)
38 k
D = ln kBT A t B (3)
39 4h:d
; =

1
T

40
41
42
43 2 1
44 Where d is the diffusion coefficient constant (cm .s ), D being the Neperian
45
46 logarithm re-duced variable of d, t being the reduced variable of temperature T (K) , A
47 diff
48 being the area of the unit cell and DE act being the diffusion activation barrier along x
49
50 or y direction, evaluated from the PEMs.
51
52
53
54
55
56
57
58
59 10
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 41

1
2
3
4
2.4 Electronic Analysis
5
6 For the calculation of the density of states (DOS) and the Mulliken population analysis, unrestricted
7
8 single point calculations, performed with the localized basis set, have been carried out with the
9
10 65,66 67
range-separated hybrid functional HSE06 along with the ab initio CRYSTAL14 code. Self-
11
12 consistent solutions to the Hartree-Fock and Kohn-Sham equations are obtained by developing the
13
14 one-electron wave function on a localized (Gaussian) basis sets, thus allowing the efficient use of
15 hybrid functionals. For the Ti atoms, the Hay and Wadt small core pseudopotential is used along with
16
17 68
18 the (6s6p5d)/[3s3p3d] basis set for the 12 valence electrons. For the O atom, the all electron basis
19
69
20 set 6-31G(d) of Gatti et al. is used. For the Au atom, an energy-consistent relativistic 19-electron-
21
22 70
pseudopotential along with the basis set (9s4p3d)/[3s2p2d] optimized by Doll et al. is used. The
23
24 reciprocal space has been sampled with a (7 7) k-point mesh. The convergence criterion for the SCF
25
26 7
27
cycle has been set at 10 Ha per unit cell.
28
29
30
31 3. Results and Discussion
32
33
34 As explained previously, the nature of TiO 2 surface during the nucleation and the growth of Au
35
36 and Cu NPs is rather complex. Various different surface states coexist: stoichiometric and
37
38 reduced areas, where Au and Cu can adsorb, diffuse and nucleate NPs. In addition more
39
40
complex defects such as steps and titanium oxide clusters may play also an important role in
41
this nucleation. At a first glance, one would like to include all those components in a unique DFT
42
43 model. Unfortu-nately, such an approach cannot be developed in a reasonable time, due to the
44
45 complexity of the corresponding system. In the following, we thus propose to investigate the
46
47 nucleation of the NPs by examining the adsorption and diffusion of Au and Cu on various states
48
49
50 of TiO2 surfaces. In order to decouple the role of each surface state, those systems have been
51
52
considered separately. Diffusion kinetics calculated at the working temperature will determine
53
the most likely nucleation sites. At this stage, the influence of the support reducibility will be
54
55 studied by the presence of surface and sub-surface oxygen vacancies.
56
57
58
11
59
ACS Paragon Plus Environment
60
Page 13 of 41 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 2: Top views of the various adsorption sites on rutile TiO 2 (110) surfaces: (a) stoichiometric surface
19 described by a (2 1) supercell; (b) surface reduced surface described by a (3 1) supercell; (c) sub-surface
20 reduced slab described by a (3 1) supercell. Onefold top sites are noted "Top", while twofold bridge sites are
21 marked "Br". The coordination of surface oxygen and titanium atoms is defined in Figure 1
22 .
23
24
25 3.1 Adsorption Properties
26
27 An overview of the adsorption properties is obtained by calculating complete PEMs for Au
28
29
30 and Cu atoms on pristine, surface and sub-surface reduced TiO 2 supports. The method
31
32 used to generate those maps is addressed in the Methodology section. The definition of all
33
34 the considered top and bridge positions between the surface oxygen and titanium atoms is
35
36 presented in Figure 2 and re-used in Figures 3-5. The PEMs contain various stationary
37
38 points corresponding to specific x and y constrained positions of Au and Cu atoms. All those
39
geometries have then been completely optimized and the related results have been
40
41 reported in Table 1. These optimal structures are either local minimum (m) or N-order
42
43 saddle-point (s). After the full geometry optimization, some of these constrained structures
44
45 labeled (u) in Table 1 are unstable. The optimized geometries of the most stable adsorption
46
47 positions for Au and Cu atoms on these two surfaces have been depicted in Figure 6.
48
49
50
51 3.1.1 Au Adsorption on Pristine and Reduced TiO 2 (110) Surfaces
52
53 In the case of Au adsorption on the pristine TiO 2 (110) support, there exist one minimum and 4
54
55 saddle-points (see Figures 3a and 3b). After the full geometry optimization, the stable structure
56
57
58 12
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 14 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14 1
15 Table 1: Adsorption energies (kJ.mol ) for Au and Cu atoms on the stoichiometric, surface
16
reduced (Red) and sub-surface reduced (sRed) TiO 2 (110) supports. Two differents
17
18
methods have been considered: unrestricted GGA (UGGA) formalism for Au and Cu on the
19 stoichio-metric support and unrestricted GGA+U method in all the cases. All the geometries
20 of the stationary points obtained in the PEMs (see Figures 3-5) have been completely
21 relaxed af-terwards and the corresponding adsorption energies reported in this table. There
22 exist two types of structures: top and bridge (Br). All the adsorption sites are defined in
23 Figures 3-5. The stationary points can be local minima (m) or N-order saddle-points (s).
24 Saddle-points used to calculate diffusion activation barriers are noted s . Those connecting
25
26
local minima along x (y) axis are noted s x (sy ). In a few cases, saddle-points are very close
27 to local minima noted m on the potential energy maps, so that the activation barriers can be
28 estimated from those states (s m ). When the adsorption structure extracted from the PEMs
29 was unstable during the geometry optimization, the used notation is (u).
30
31 Adsorbate Au Cu
32 Support Stoichiometric Reduced Stoichiometric Reduced
33
Red sRed Red sRed
34
Method UGGA UGGA+U UGGA+U UGGA UGGA+U UGGA+U
35
36 Structure
37 Top Ti5C -33 (s) -29 (s) -120 (sx ) -125 (m) -47 (s) -41 (s) -64 (mx ) -76 (s)
38 Top O2C -33 (s) -20 (sx ) -118 (s) -24 (s) -112 (s) -96 (sy ) -95 (my ) -104 (sy )
39 Br Ti5C-O2C -40 (m) -46 (m) (u) (u) -112 (s) -102 (s) (u) (u)
40 Br O2C-O2C -13 (s) -18 (s) 1 (sy ) -11 (s) -163 (m) -149 (m) -135 (m) -179 (m)
41
Br O3C-O3C -25 (s) -26 (sy ) -87 (s) -90 (sy ) -63 (s) -36 (sx ) -39 (s) -56 (sx )
42
Top Vacancy - - -213 (m) -17 (sx ) - - -161 (m) -120 (s)
43
44 Exp.43,44 49.8 - - - -
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 13
60 ACS Paragon Plus Environment
Page 15 of 41 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
Figure 3: Potential energy maps (PEM) for the adsorption of Au (a-c) and Cu (d-f) atoms on
41
42 stoichiometric TiO2 (110) surface at the (a), (d) PBE and (b), (e) PBE+U levels. Adsorption energies
43 1
(DEads) are expressed in kJ.mol . The definitions of x and y direct coordinates and corresponding
44 ranges are given in (c). All the particular adsorption sites used in (c) and (f) with colored crosses are
45 defined in Figure 2. This set of colors has been transferred in (a), (b), (d) and (e), where colored
46 circles indicate local minima and colored squares, N-order saddle-points.
47
48
49
50
51
52
53
54
55
56
57
58
59 14
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure 4: Potential energy map (PEM) for the adsorption of (a) Au and (b) Cu atoms on the surface reduced
23 1
TiO2 (110) support at the PBE+U level. Adsorption energies (DE ads) are expressed in kJ.mol . The definitions
24
25 of x and y direct coordinates and corresponding ranges are given in (c). All the particular adsorption sites used
26 in (c) with colored crosses are defined in Figure 2. There is a single oxygen vacancy in the surface in the
27 supercell. This set of colors has been transferred in (a) and (b), where colored circles indicate local minima,
28 colored squares, N-order saddle-points and triangles the unstable positions.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Figure 5: Potential energy map (PEM) for the adsorption of (a) Au and (b) Cu atoms on the sub-surface
50
1
51 reduced TiO2 (110) support at the PBE+U level. Adsorption energies (DE ads) are expressed in kJ.mol . The
52 definitions of x and y direct coordinates and corresponding ranges are given in (c). All the particular adsorption
53 sites used in (c) with colored crosses are defined in Figure 2. There is a single oxygen vacancy in the sub-
54 surface in the supercell. This set of colors has been transferred in (a) and (b), where colored circles indicate
55 local minima, colored squares, N-order saddle-points and triangles the unstable positions.
56
57
58
59
15
60
ACS Paragon Plus Environment
Page 17 of 41 The Journal of Physical Chemistry

1
2
3 is a non-symmetric bridge position between a surface Ti atom (Ti5C) and a surface bridging
4
5 1 1
6 atom (O2C) (see Figures 6a and 6b) with an adsorption energy of -40 kJ.mol or -46 kJ.mol ,
7
8 at the PBE or PBE+U level, respectively (cf. Table 1). Since, from a general point of view, the
9
GGA+U results are consistent with the GGA ones, qualitatively (optimal positions) and
10
11 quantitatively (ad-sorption strength), only the DFT+U analysis will be considered in the following
12
13
14 discussions. For Au adsorption on the surface reduced TiO 2 (110), there exist one minimum (m)
15
16 1
and 4 saddle-points (s) (cf. Figure 4a): the minimum at -213 kJ.mol corresponding to a top
17
18 position above the oxygen vacancy (cf. Figures 7a and 7b). The adsorption strength is five times
19
20 larger on the sur-face reduced support than on the pristine termination. For the adsorption on
21
22 the last surface state (sub-surface reduced support, see Figure 5a), one minimum (m) is also
23
24 1
obtained with an adsorp-tion energy of -125 kJ.mol . This structure corresponds to a top Ti5C
25
26 site, at the closest position to the sub-surface oxygen vacancy (cf. Figures 7e and 7f)). Likewise
27
28 the two previous support states, four saddle-points are determined on the PEM. The structure
29
30 determined on the sub-surface reduced support is different from the two other surface states
31
32
(stoichiometric and surface reduced). Moreover, its adsorption energy is more moderate than
33 the one on surface reduced support with an intermediate value.
34
35 Table 2: Spin polarization properties for Au and Cu best adsorption structures on
36 stoichio-metric, surface reduced (Red) and sub-surface reduced (sRed) TiO 2 (110)
37 1
38 supports. The adsorption energy (DE ads, kJ.mol ) is recalled. The total spin polarization
39 (M, mB) is ad-dressed with the spin properties projected on the adsorbates (Au or Cu)
40 and the Ti atoms layer by layer.
41
42 Adsorbate Clean support Au Cu
43
Support Stoichiometric Reduced Stoichiometric Reduced Stoichiometric Reduced
44
Red sRed Red sRed Red sRed
45 DE
46 ads - - - -46 -213 -125 -149 -161 -179
47 M 0 2 2 0 1 1 1 1 3
48 Au or Cu - - - 0 0 0 0 0 0
49 st
Ti (1 ) 0 0 1 0 0 0 0 0 1+1
50 nd
Ti (2 ) 0 1+1 1 0 1 1 1 1 1
51 rd
Ti (3 ) 0 0 0 0 0 0 0 0 0
52 th
Ti (4 ) 0 0 0 0 0 0 0 0 0
53
54
55
56
57
58
59 16
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 41

1
2
3 Regarding the magnetic properties, the gold atom (doublet state) loses its spin polarization once
4
5 chemisorbed on stoichiometric, surface reduced and sub-surface reduced supports (cf. Table 2). The
6
7 total spin polarization of the adsorption state on the stoichiometric surface is zero (diamagnetic
8 state). Hence, the adsorption process has not modified the spin properties of the stoichiometric
9
10 clean support (diamagnetic state, as well). After adsorption on the surface reduced support, the best
11
12 3+
adsorption state exhibits a total magnetic moment of 1 m B (ferromagnetic state) localized on a Ti
13
14 1
sub-surface atom (second layer). The strong chemisorption of Au (-213 kJ.mol ) on this surface
15
16 reduced support partially quenches the spin polarization, since the clean system has two reduced
17
18 3+
sub-surface Ti atoms (Ti ) and a total spin polarization of 2 mB. The moderate adsorption on the
19
20 1
21 sub-surface reduced support (-125 kJ.mol ) does not change this picture (a ferromagnetic state with
22
3+
23 a total spin polarization of 1 mB localized on a Ti sub-surface atom; the spin polarization being
24
25 partially quenched with respect to the bare sub-surface reduced support). The presence of excess
26
71
27 electrons on subsurface Ti sites is compatible with previous theoretical works and resonant photo-
28
29 72
electron diffraction measurements for clean surfaces.
30
31
Concerning the electronic analysis, single-point energy calculations have been performed with
32
33 CRYSTAL04 (as detailed in the methodology section), for the most stable adsorption structure of Au
34
35
on the pristine TiO2 (110) surface. HSE06 functional has frequently been assessed for repro-ducing
36
37
73,74
38 electronic structure of semiconductors and more specifically titanium oxides. The DOS
39
40 computed for the system containing one Au atom on a bridge position on the TiO 2 (110) surface with
41
42 HSE06 is presented in Figure 8 along with the spin density of the system. It clearly appears that the
43
44 unpaired electron remains localized on the gold atom with only a minor delocalization on the closest
45
Ti atom. As matter of fact, the Mulliken partitioning of the spin density indicates 0.69 unpaired
46
47 electron on Au, 0.15 unpaired electron on the closest Ti, 0.05 unpaired electron on the closest O
48
49 atom and less than 0.05 electron on the remaining atoms. In other words, the elec-tron transfer from
50
51
Au to the TiO2 (110) surface is rather low, in agreement with previous studies suggesting the same
52
53 23,25,38
54 adsorption form. The system remains a semiconductor, with a valence band near the Fermi
55
56
level and a gap state localized on the Au atom and a conduction band being a
57
58
59 17
60 ACS Paragon Plus Environment
Page 19 of 41 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
Figure 6: Fully relaxed adsorption structures of (a, b) Au and (c, d) Cu atoms on stoichiometric TiO 2
44 (110) surface at the PBE+U level. Optimal distances are given in Å. The color labels are blue for Ti,
45 red for O, dark yellow for Au and light brown for Cu.
46
47
48
49
50
51
52
53
54
55
56
57
58
59 18
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 41

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
Figure 7: Fully relaxed adsorption structures of (a, b, e, f) Au and (c, d, g, h) Cu atoms on reduced TiO 2
24 (110) surfaces at the PBE+U level. Single oxygen vacancies (depicted with dotted circles) are
25 localized in the support surface (a-d) and in the sub-surface (e-g). Optimal distances are given
26 in Å. The color labels are blue for Ti, red for O, dark yellow for Au and light brown for Cu.
27
28
29 state delocalized state on Ti atoms.
30
31
32 3.1.2 Cu Adsorption on Pristine and Reduced TiO 2 (110) Surfaces
33
34
35 The adsorption properties of Cu on both pristine and reduced TiO 2 (110) supports have been ex-
36
37 amined in a similar way as Au. The corresponding PEMs have been presented in Figures 3d, 3e, 4b
38
39
and 5b, for stoichiometric, surface reduced and sub-surface reduced supports, respectively. From a
40
qualitative point of view, the PEM of adsorbed Cu on the pristine support is very simi-lar to the
41
42
equivalent one for Au (identical number of stationary points, one minimum (m) and 4 saddle-points
43
44 (s), as recalled in Table 1). This being said, after the full geometry optimization, the most stable
45
46 adsorption structure for Cu is different from Au (bridge O2C-O2C, cf. Figures 6c and 6d). From a
47
48 quantitative standpoint, the adsorption strength is approximately three times larger for Cu (-149
49
50 1 1
kJ.mol ) than Au (-46 kJ.mol ). As expected, the Cu-O bond length (1.9 Å) is much shorter than the
51
52
Au-O distance (2.3 Å), hence reflecting the difference of stability and electronegativity variation
53
54
between Cu, Au and O atoms. Concerning now the adsorption of Cu
55
56
57
58 19
59 ACS Paragon Plus Environment
60
Page 21 of 41 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 8: Density of states of Au adsorbed on pristine TiO 2 (110) surface computed with the HSE06
43 functional. The total DOS is plotted in black whereas the red, blue and yellow curves correspond to the
44
projections on O, Ti and Au, respectively. To facilitate the interpretation, the area below the gold projection
45
was colored in yellow. In inset, the spin density if presented in semi-transparent green color (isovalue
46
0.003 a.u.). The dashed vertical line indicates the energy of the top of valence band.
47
48
49
50
51
52
53
54
55
56
57
58
59
20
ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 22 of 41

1
2
3 on the surface reduced support, the number of stationary points in the PEM is again equivalent
4
5 to the chemisorption of Au, with a subtle difference regarding the number local minima (2
6
7 minima for Cu versus 1 for Au). Likewise the case of Au, when comparing stoichiometric and
8
9
surface reduced supports, the best adsorption position for Cu is different (top vacancy with an
10
1
11 adsorption energy of -161 kJ.mol , cf. Figures 7c and 7d). The bridge O2C-O2C structure is a
12
13 1
14
metastable local minimum on the PEM with an adsorption energy of -135 kJ.mol . Contrary to
15
1
16 the analysis developed for Au (-167 kJ.mol ), the increase of adsorption strength between
17
18 1
19
stoichiometric and surface reduced supports is much weaker for Cu (-12 kJ.mol ). Regarding
20
now the adsorption on the sub-surface reduced support, the only similarity with Au is the
21
22 identical number of stationary points. However, the minimum is different in terms of structure
23
24 1
25
(bridge O2C-O2C, cf. Figures 7g and 7h) and adsorption energy (-179 kJ.mol ), since this time,
26
the adsorption strength has increased compared to surface reduced support. This is in line with
27
28 the larger geometric relaxation registered in the surface layers of the support (see Figure 7h
29
30 compared to Figure 7f). Although the Cu-O optimal bonds are rather long (2.7 Å), the adsorption
31
of Cu promotes a concomitant lift of the symmetry of the sub-surface reduced support and an
32
33 increase of the net spin polarization (see here after) which appear both stabilizing.
34
35 Concerning the magnetic properties, the copper atom (doublet state) loses its spin
36
37 polarization upon adsorption whatever the surface oxide state, likewise Au adsorption (cf.
38
39 Table 2). The total spin polarization of the adsorption state on the stoichiometric surface is 1
40
41 3+
mB (ferromagnetic state), localized on a Ti sub-surface atom (second layer). This results
42
43 differs from Au on stoichiometric surface. The adsorption of Cu on surface reduced support
44
45 leads to similar spin polarization prop-erties as Au on the same surface. In contrast, the
46
47 adsorption on the sub-surface reduced support increases the total spin polarization (a
48
49 3+ 3+
ferromagnetic state of 3 mB localized both on two Ti sur-face atoms and on one Ti sub-
50
51 surface atom). As shown in Table 2, the net spin has increased in the surface layer of the
52
53 support where the geometric distortion has been promoted by Cu adsorp-tion.
54
55
56
57
58
59
21
ACS Paragon Plus Environment
60
Page 23 of 41 The Journal of Physical Chemistry

1
2
3
4
3.2 Diffusion Kinetics
5
6 Among the key elementary steps related to the nucleation and the growth of Au and Cu nanoparti-
7
8 cles on rutile TiO2 (110), we examine in the following section the diffusion properties of Au and Cu
9
10 atoms on the corresponding pristine and surface reduced support. The PEMs provide the required
11
12 information in terms of saddle-points adsorption structures and energies, and diffusion activation
13
14 barriers (cf. Table 3). To quantify diffusion kinetics, the calculation of the activation barriers has
15
16 been carried out on the basis of the PEMs. Those estimated activation barriers (differences of
17
18 energies between constrained saddle-point and global minima, along the x and y directions) are
19
20 1
converged in the range 0-4 kJ.mol (the numeric error being determined by comparison with val-
21
22 ues resulting from completely relaxed saddle-point and global minimum structures). This justifies
23
24 the quality of the PEMs regarding diffusion kinetics.
25
diff 1
26 Table 3: Activation energy barriers (DE act ) of diffusion (kJ.mol ) for Au and Cu atoms on the
27
stoichiometric, surface reduced (Red) and sub-surface reduced (sRed) TiO 2(110) sup-ports at
28
29
reference zero temperature. The diffusion along x (y) axis is defined in Figures 3-5. These
30
barriers have been extracted from the corresponding PEMs, by connecting the most stable
31 adsorption structures (see Table 1 for the definitions of those saddle-points sx and sy ).
32
33 Adsorbate Au Cu
34 Support Stoichiometric Reduced Stoichiometric Reduced
35 Red sRed Red sRed
36 diff
DEact (x) 25 95 108 117 97 128
37 diff
DEact (y) 22 214 34 55 67 73
38
39 PBE28,29 - 40 - 75 - -
27
40 PBE 16-19 58-178 - - - -
40 15-55 - - 62-140 - -
41 PBE+U
42 Exp.43,44 8.8 -
43
44
45
46
In the case of Au adsorption on the stoichiometric surface, the diffusion is easy whatever the
47
48 1
49 direction (similar activation barriers in the range 22-25 kJ.mol , cf. Table 3). Therefore large
50
diffusion coefficients have been calculated in the range 300-500 K (a typical range of working
51
52 15,43,44
temperature ), as shown in Figure 9a. The nature of the saddle-points (s x ) and (sy ) is different
53
54
along x and y direction (top O2C and bridge O3C-O3C, respectively). Regarding Au on surface
55
56 1
and sub-surface reduced TiO2, the diffusion becomes globally either moderate (34-108 kJ.mol
57
58
59 22
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 41

1
2
3 1
4 on the sub-surface reduced support) or difficult (95-214 kJ.mol on the surface reduced
5
support), hence decreasing significantly the diffusion coefficients by several orders of magnitude
6
7 in the range 300-500 K (cf. Figures 9b and 9c). In the case of the presence of a single-oxygen
8
9 vacancy in the surface layer, the smallest activation barrier occurs for the diffusion along x (95
10
11 1
12
kJ.mol , (sx )). The comparison between the three terminations (stoichiometric, surface
13
reduced and sub-surface reduced supports) shows a monotonous increase of the diffusion
14
15 barrier along x and a maximum along y for the support exhibiting surface oxygen vacancies.
16
17 For the diffusion of Cu on the stoichiometric surface, the diffusion is more difficult than for Au
18
19 on the stoichiometric support, whatever the direction since the activation barriers are in the
20
1
21 range 55-117 kJ.mol . A direct consequence is the significant decrease (several orders of
22
23 magnitude) of the corresponding diffusion coefficient in the range 300-500 K, as depicted in
24
25 Figure 9a. The diffusion of Cu along the y axis is much easier (the saddle-point (s y ) being top
26
27
28 O2C) than along x (bridge O3C-O3C for (s x )). Regarding now the diffusion on the surface
29
30
reduced and sub-surface reduced supports, there are qualitative and quantitative discrepancies
31 with respect to Au. The presence of single-oxygen vacancy induces globally a more moderate
32
33 variation of the diffusion barriers and opposite trends regarding the evolution of the activation
34
35 energies with respect to the nature of the support terminations. Along x, the diffusion barriers
36
1
37 are in the same order (117, 97 and 128 kJ.mol ), while there is a progressive increase along y
38
39 1
(from 55 to 73 kJ.mol ). In the case of the surface reduced support, the change of diffusion
40
41 between the two directions is less important. This analysis is reflected on the variation of the
42
43 diffusion coefficients (cf. Figures 9b and 9c). Note in passing that for Cu diffusion on the surface
44
45 reduced oxide, the vibrational analysis has shown that the N-order saddle-points appearing in
46
47 the PEM correspond to highly metastable local minima named mx and my which are close in
48
49 energy to the effective saddle-points (as explained in Table 1).
50
51 As a preliminary conclusion, the diffusion of a Au atom is easy in the range 300-500 K what-
52
53 ever the direction on stoichiometric TiO 2 (110) and it becomes more difficult on surface reduced
54
55 1 1
titania (from 25 to 95 kJ.mol ) and sometimes impossible (214 kJ.mol ). In contrast, the dif-
56
57
58
23
59
ACS Paragon Plus Environment
60
Page 25 of 41 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 9: Diffusion diagrams for Au and Cu atoms on (a) stoichiometric, (b) surface reduced and (c) sub-
2 1
17 surface reduced TiO2(110) supports. The reduced diffusion coefficients D (ln(cm .s )) are plotted against the
18 1
inverse of temperature t (K ). For a sake of simplicity, the associated reverse x axis (temperature,
19 K) is also presented and the interesting range of working temperature 300-500 K is marked with
20
vertical black dotted lines. The diffusion of each metal M (red plots for Au and blue plots for Cu)
21
along x (y) axis is depicted with full (dotted) lines, and noted M x (M y). The areas of fast
22
diffusion in the range 300-500 K are defined with gray rectangles.
23
24
25
fusion of a Cu atom shows more moderate variations between pristine and reduced
26
27 1
28 surfaces (from 55 to 73 kJ.mol ). Hence, the presence of single-oxygen vacancy in the
29
30 surface of the support strongly slows down the diffusion of Au whereas the diffusion of
31
Cu is much less perturbed (no-tably along y axis, in both cases).
32
33 In the temperature range 500-800 K, many competitive events can occur and compete with the
34
35
36 diffusion of atoms (cluster growth and diffusion, diffusion from the defects, formation of TiO x clusters
37 and stripes, restructuration of the support and migration of metallic clusters towards these defects,
38
39 75
40 etc). Note in passing that several of these processes already happen at 300 K. According to our
41 analysis of diffusion kinetics, in the range 500-800 K, Au and Cu atoms start to diffuse competitively
42
43
44 on the various surfaces of rutile TiO 2 (110) (stoichiometric or reduced support). In particular, the
45
anchor effect of the oxygen vacancy disappears in agreement with previous obser-
46
47 76
48 vations.
49
50
51 3.3 Discussion
52
53
54 Before examining the possibilities for the nucleation sites for Au and Cu on TiO 2 (110), let us
55
56 compare our PBE+U calculations with previous experimental and theoretical works, both for ad-
57
58 24
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 26 of 41

1
2
3
sorption energetics and diffusion barriers. In Table 4, our most stable adsorption structures for Au
4
5
6 and Cu on stoichiometric and surface reduced TiO 2 (110) are reported with previous theoretical
7 9,23–27,29,31,35,37–40,43,44
8 and experimental investigations. For Au adsorption on stoichiometric TiO2
9 1
10 (110), our stable structure bridge O2C-Ti5C presents an adsorption energy of -46 kJ.mol at the
11 1 43,44
12 unrestricted PBE+U level, which is in fair agreement with the measurements (50 kJ.mol ).
13
14
Throughout the years, various GGA functionals have been employed in the theoretical studies, first
15 23–25,31 26,27,29,35,37 9,35
16 PW91, then PBE, RPBE and more recently, in the context of the DFT+U
17 38–40
18 methodology implemented in the VASP code, the PBE+U functional with several values of U.
19 1
20
Table 4: Adsorption energies (DEads) (kJ.mol ) for Au and Cu atoms on stoichiometric, surface
21 reduced TiO2 (110) supports. In this table, our PBE+U values are reported and com-pared with
22 several previously published results, using various GGA functionals and DFT+U methods. The
23 experimental value is also recalled. In all the cases, the energies of the most sta-ble adsorption
24
sites (top, bridge noted Br and oxygen vacancy noted Ov) have been addressed.
25
26
a In this work, the oxygen vacancy was located in the sub-surface, in the second layer of
b
27 the oxide. In this study, the Cu atom was adsorbed on a reduced support presenting a
28 surface oxygen vacancy.
29
30 Adsorbate Au Cu
31 Support Stoichiometric Surface reduced Stoichiometric
GGA-Method Site DE Site DE Site DE
32 ads ads ads
33 O
This work Br O2C-Ti5C -46 v -213 Br O2C-O2C -148
34 PW91
23
- - - - Br O2C-O2C -230
35 24 O
PW91 Br Ti5C-Ti5C -58 v -268 - -
36 25
PW91 Br O2C-Ti5C -43 Ov -238 - -
37 26
38 PBE Top Ti5C -10 Ov -184 - -
27
39 PBE Br O2C-Ti5C -53 Ov -210 - -
29
40 PBE Br O2C-O2C -58 - - Br O2C-O2C -249
31 O
41 PW91 Br O2C-Ti5C -43 v -175 - -
9 O
42 RPBE Top O2C -59 v -110 - -
43 35 a
PBE Br O2C-Ti5C -34 Top Ti5C -126 - -
44 37
PBE - - Ov -245 - -
45 38
PBE+U, U=4.2 eV Br O2C-Ti5C -56 Ov -148 - -
46 39 O
47 PBE+U, U=4.2 eV Br O2C-Ti5C - v - - -
40 O
48 PBE+U, U=5.8 eV Top O2C -98 v -246 Br O2C-O2C -265
77 - - -234 b -170
49 PBE+U, U=4.0 eV Ov Ov
50 50 - -
Exp.43,44
51
52
53
54
55 Our result for the adsorption structure of Au on stoichiometric TiO 2 (110) agrees with the
56
57
58 25
59 ACS Paragon Plus Environment
60
Page 27 of 41 The Journal of Physical Chemistry

1
2
3 25,27,31,35,38,39
4 majority of the previous studies, for both position (bridge O2C-Ti5C) and energy
5 1
6 (from -34 to -56 kJ.mol for a similar coverage of 0.5 ML), independently from the choice of the
7
24,29
8 functional or methodology. Other highly symmetric structures (bridge Ti5C-Ti5C, bridge
9
29 26 9,40
10 O2C-O2C, top Ti5C and top-O2C ) belonging to a wider range of adsorption energies
11
12 1
(from -10 to -98 kJ.mol ) have been proposed in the literature as stable forms, although they
13
14 correspond to saddle-points in our PEMs. For the adsorption of Au on the reduced support, the
15 comparison can be done only for single oxygen vacancies present in the surface (no available
16
17
data in the literature for the adsorption of Au on sub-surface reduced TiO 2 (110)). There is a
18
19 global consensus on the best adsorption form (top position on the vacancy), whereas our
20
1
21 adsorption energy of - 213 kJ.mol belongs to a very broad range of previously published
22
23 1 9,24–27,31,35,37–40,77
values (from -110 to -268 kJ.mol ). Our result is in a narrow agreement
24
25 27
with Iddir et al.’s value. In the case of Cu, elements of comparison are available mostly for the
26
27
28 adsorption on stoichiometric TiO 2 (110). Here again, our stable structure bridge O2C-O2C
29 23,29,40 1
30 agrees with the previous works, although our adsorption energy (-148 kJ.mol ) is more
31 1
32 moderate than the reported values (from -230 to -265 kJ.mol for a coverage of 0.5 ML).
33
77
34 Recently, the adsorption of a Cu atom has been considered on surface reduced TiO 2 (110).
35
36 1
The proposed adsorption site (top on the vacancy) and adsorption strength (-170 kJ.mol ) are
37
38 1
entirely compatible with our results (top position and -161 kJ.mol ). Unfortunately, there is no
39
40 available measurement for Cu heat of adsorption on this support.
41 77,78
42 In Refs., the importance of the chemical hardness of the Lewis acid/base has been com-
43
44 pared to the electron accepting/donating capacities in order to understand the binding of metal
45
46 atoms to oxide supports. In the case of rutile TiO 2, the authors have explained that metal atoms can
47
48 act as either Lewis bases (cations) on unsaturated oxygen atom sites or Lewis acids (anions) on
49 unsaturated Ti atom sites. Since absolute hardness of Au (3.5) is larger than that of Cu (3.25), Au is
50
51
52 expected to bind more strongly than Cu on reduced rutile TiO 2 (Lewis acid in interaction with an
53 oxygen vacancy surface site). Conversely, Au adsorption strength should be weaker than that of Cu
54
55
56 on stoichiometric TiO2 (Lewis base in interaction with bridging oxygen surface site).
57
58
59 26
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 41

1
2
3
Those predicted trends are thus consistent with our DFT+U calculations.
4
5 Regarding diffusion barriers, our calculated values at the unrestricted PBE+U level are com-
6
7 27–29,40,43,44
pared to previous theoretical and experimental results in Table 3. For Au on the
8
9 1
10
sto-ichiometric support (22-25 kJ.mol ), the agreement is once again fair with Iddir et al.’s
11 27 1
12
study (16-19 kJ.mol ). Those activation barriers are slightly larger than the one evaluated
13 1 43,44
14 from the measurements (8.8 kJ.mol ). Other authors have proposed a range of larger and
15 40 1
16 anisotropic barriers (15-55 kJ.mol ). The comparison for Au on the reduced support can be
17
18 done only for surface reduced TiO2 with previously published theoretical works at the PBE
19
27–29 1 27
20 level. Although different in extent (58-178 kJ.mol in Ref. ), the anisotropy registered for
21
1
22 our diffusion barriers (95-214 kJ.mol ) is consistent with the authors’ findings. For Cu on TiO 2
23
28,29,40
24 (110), we can com-pare our results with previous theoretical investigations only for the
25
1
26 stoichiometric support. Our range of diffusion barriers (55-117 kJ.mol ) is compatible with those
27
1
28 studies, in terms of anisotropy and extent (62-140 kJ.mol ). The energetic order of the diffusion
29
30 barriers found for Au and Cu on the stoichiometric and surface reduced support is consistent
31 23 77,78
32
with the previously published analyses of electronic structures and chemical hardness for
33 adsorption states. As discussed here above, on the stoichiometric support, Au adsorption is
34
1 1
35 much weaker (-46 kJ.mol ) than that of Cu (-149 kJ.mol ). Au diffusion barriers (22-25 kJ.mol
36
1 1
37 ) are thus smaller than Cu ones (55-117 kJ.mol ) and associated kinetics are consequently
38
39 1
faster. At the opposite, on the surface reduced support, the adsorption of Au (-213 kJ.mol ) is
40
41 1
much stronger than the one of Cu (-161 kJ.mol ). Then, depending on the direction, Au
42
43 1 1
diffusion barriers are found either equiva-lent (95 kJ.mol ) or larger (214 kJ.mol ) than those of
44
45 1
Cu (67-97 kJ.mol ), thus meaning that diffusion rates are globally slower. In the case of the
46
47 sub-surface reduced support, the adsorptions are more competitive between the two metallic
48
1 1
49 atoms, Cu being more stable (-179 kJ.mol ) than Au (-125 kJ.mol ). The calculated diffusion
50
1
51 barriers for Au are thus globally slightly lower (34-108 kJ.mol ) than those of Cu (73-128 kJ.mol
52
1
53 ). Hence the corresponding diffusion rates are of the same order of magnitude.
54
55 In summary, the comparative analysis of our adsorption and diffusion properties of Au and
56
57
58
59 27
60 ACS Paragon Plus Environment
Page 29 of 41 The Journal of Physical Chemistry

1
2
3
Cu on the various terminations of rutile TiO2 (110) shows interesting discrepancies between the
4
5
6 two metals (cf. Tables 1,3 and Figure 9). The adsorption strength and structure of Au on TiO2
7 (110) strongly depend on the nature of the surface state, with an adsorption energy rather weak on
8
9 1
the stoichiometric surface (-46 kJ.mol ) and really strong on the surface reduced support (-213
10
11
1
12 kJ.mol ). The metastability of the sub-surface oxygen vacancy does not promote the adsorption of
13
14 1
Au on this surface state, since the chemisorption remains moderate (-125 kJ.mol ) and lower than
15
16
17 on surface reduced TiO2 (110). Hence, the presence of defects on the support such as surface and
18
19 sub-surface oxygen vacancies tends to stabilize significantly the Au atom. Diffusion kinet-ics support
20 this analysis (cf. Figure 9). On the clean stoichiometric surface, Au atom can diffuse very easily
21
22 along the surface bridging oxygen rows (y axis) or perpendicularly, whereas diffusion is rather
23
24 difficult on the reduced support, even at 500 K (one exception along the surface oxygen rows of the
25 sub-surface reduced support). According to this analysis, defects such as oxygen vacancies should
26
27 promote the nucleation of Au nanoparticles, at least in the early stage. This conclusion is compatible
28
29 8,9,79
with several experimental investigations. For Cu, the situation is different. Both the adsorption
30
31 strength and structure is weakly perturbed by the change of the surface state and the presence of
32
33 oxygen vacancies. These defects increase the adsorption energy rather moderately
34
1
35 (from -149 to -179 kJ.mol ), while the bridge O2C-O2C position is the competitive form. In-
36
37 1
terestingly, the adsorption strength is larger on the sub-surface reduced support (-179 kJ.mol ) than
38
39 1
on the surface reduced state (-161 kJ.mol ). This shows thus that the presence of mestable sub-
40
41 surface oxygen vacancies may be promoted by Cu adsorption. So the nucleation sites of Cu
42
43 nanoparticles can be multiple on TiO2 (110) with a preference for defects, although the role of
44
45 the reducibility of the support is less marked than for Au nucleation. This analysis is
46
47 15,80
entirely supported by diffusion kinetics and is consistent with measurements. In
48
49 fact, the Cu atom can diffuse with more difficulty than the Au one, especially on the
50
51 stoichiometric support, hence slowing down the nucleation kinetics.
52
53
54
55
56
57
58 28
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 30 of 41

1
2
3
4
4. Conclusion
5
6 In this work, we have studied the adsorption properties and diffusion kinetics of Au and Cu atoms on
7
8
9 three different surface states of rutile TiO 2 (110) (stoichiometric, surface reduced and sub-surface
10
11 reduced support) from a DFT+U unrestricted approach. A systematic and detailed analysis of these
12
13 properties has been reported by exploring and plotting potential energy maps with geo-metric
14
15 constraints. The various possible adsorption forms and diffusion saddle-points have been extracted
16
17 from these plots and refined with complete geometry optimizations. The magnetic prop-erties upon
18
19 adsorption have been carefully examined and addressed. The diffusion activation barri-ers have
20
21
been evaluated from these potential energy maps along the main directions of the support surface.
22
The corresponding diffusion rates have been calculated on the basis of transition-state theory and
23
24
the diffusion diagrams have been plotted as a function of temperature.
25
26 Interesting discrepancies have been registered between the two metals. Au adsorption
27
28 struc-ture and energetics strongly depend on the reducibility of the support whereas Cu
29
30 chemisorption is much less sensitive to this property. For instance, the most stable structure of
31
32 Au on the sto-ichiometric surface is first different from those on the reduced support, and
33
34 second it exhibits a much weaker adsorption energy. In contrast, the best adsorption form of Cu
35
36 is almost the same (bridge position above terminal oxygen rows of the support) whatever the
37
38 surface state. In addi-tion, Cu adsorption energy is weakly perturbed by the reducibility of the
39
40 support. So far, Au is best chemisorbed on the surface reduced support on top of the oxygen
41
42 vacancy, while Cu adsorption is the strongest one on the sub-surface reduced system.
43
44 Concerning magnetic properties, Au chemisorption tends to quench partially the total net
45
46
47
spin polarization of the TiO 2 (110) reduced support wherever the location of the single
48
oxygen vacancy. In contrast, Cu adsorption induces an increase of the total spin
49
50 polarization, except for the surface reduced system. In the majority of the optimal structures,
51
52 3+
53
the excess electrons are located in the sub-surface Ti atoms, leading to Ti species, in
54
agreement with the state of the art. In the case of Cu adsorption on the sub-surface reduced
55
56 support, the excess electron is however located on the support surface.
57
58
59
29
ACS Paragon Plus Environment
60
Page 31 of 41 The Journal of Physical Chemistry

1
2
3
From a general standpoint, diffusion properties are also different between Au and Cu
4
5
6 on the various surface states of rutile TiO 2 (110). Au atom diffuses easily in all the
7
8 directions on the stoichiometric support, whereas Cu, first diffuses with more difficulty
9
10 and preferentially along the terminal bridging oxygens rows of the support. On the
11
12 reduced surfaces, Au diffusion is really slown down, except along the direction of the
13
14 oxygen rows for the case of the sub-surface reduced support. For Cu on the reduced
15
systems, diffusion kinetics are globally similar to the one on the stoichiometric surface
16
17 (moderate diffusion). This occurs essentially along the terminal bridging oxygen rows.
18
19 From those analyses of adsorption thermodynamics and diffusion kinetics, a scenario
20
21 has been sketched out regarding the possible nucleation sites for Au and Cu nanoparticles
22
23
24 on rutile TiO2 (110). In the early stage of this process, Au nucleation appears to be sensitive
25
to the reducibility of the support (presence of defects such as single oxygen vacancies),
26
27 whereas Cu nucleation seems to be less promoted by those defects.
28
29 This work opens interesting perspectives for the understanding of the early-stage nucleation
30
31
32 sites for Au and Cu nanoparticles on rutile TiO 2 (110). Our DFT+U results are useful as input
33
34 parameters for Monte Carlo simulations aiming to determine average surface coverage and
35
36 average diffusion kinetics that could be also compared to measurements. Our systematic
37
38 investigation of Au and Cu adsorption and diffusion properties on the rutile support should thus
39
encourage the experimentalists to explore these systems and provide new measurements. Our
40
41
approach is general and can be transferred to the study of any atomic adsorbate on various
42
43 inorganic supports including oxides, sulfides, metals, carbides, nitrides. This paves also the way
44
45 to the theoretical study of more complex support defects, such as steps, titanium oxide clusters,
46
47 hydroxylated domains or other subsurface defects.
48
49
50
51 Acknowledgement
52
53
54 The authors thank Dr Axel Wilson, Dr Geoffroy Prévot (INSP, Paris, France) and Dr Françoise
55
56 Delbecq (ENS, Lyon, France) for helpful discussions. They also thank IDRIS in Paris, CINES in
57
58
59
30
ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 32 of 41

1
2
3 Montpellier (project 609, GENCI/CT8) and PSMN in Lyon for CPU time and assistance. Mathilde
4
5 Iachella thanks Ecole Doctorale de Chimie de Lyon, France (ED 206) for the PhD funding.
6
7
8
9 Supporting Information Available
10
11
12 In the Supporting Information, we expose the harmonic vibrational analysis
13
14 characterizing the nature of the stationary points related to the diffusion pathways of Au
15
16 and Cu atoms on pristine, surface and subsurface reduced TiO 2 (110). This material is
17
18 available free of charge via the Internet at http://pubs.acs.org/.
19
20
21
22 References
23
24 (1) Campbell, C. T. The Energetics of Supported Metal NanoParticles: Relationships to SIntering
25
26 Rates and Catalytic Activity. Accounts of Chemical Research 2013, 46, 1712–1719.
27
28
29 (2) Grzelczak, M.; Perez-Juste, J.; Mulvaney, P.; Liz-Marzan, L. M. Shape Control in
30
31 Gold Nanoparticle Synthesis. Chem. Soc. Rev. 2008, 37, 1783–1791.
32
33
(3) Renaud, G.; Lazzari, R.; Leroy, F. Probing Surface and Interface Morphology with Grazing
34
35
Incidence Small Angle X-Ray Scattering. Surface Science Reports 2009, 64, 255 – 380.
36
37
38 (4) Lazzari, R.; Renaud, G.; Jupille, J.; Leroy, F. Self-similarity During Growth of the
39
40
41
Au=TiO2(110) Model Catalyst as Seen by the Scattering of X-Rays at Grazing-
42
Angle In-cidence. Phys. Rev. B 2007, 76, 125412.
43
44
45 (5) Diebold, U. The Surface Science of Titanium DiOxide. Surface Science Reports
46
47 2003, 48, 53–229.
48
49
50 (6) Ganduglia-Pirovano, M. V.; Hofmann, A.; Sauer, J. Oxygen Vacancies in Transition
51
52 Metal and Rare Earth Oxides: Current State of Understanding and Remaining
53
54 Challenges. Surface Science Reports 2007, 62, 219–270.
55
56
57
58 31
59 ACS Paragon Plus Environment
60
Page 33 of 41 The Journal of Physical Chemistry

1
2
3 (7) Maeda, Y.; Fujitani, T.; Tsubota, S.; Haruta, M. Size and Density of Au Particles Deposited
4
5
on TiO2 (110)-(1 1) and Cross-Linked (1 2) Surfaces. Surface Science 2004, 562, 1–6.
6
7
8 (8) Wahlstrom, E.; Lopez, N.; Schaub, R.; Thostrup, P.; Ronnau, A.; Africh, C.; Laegsgaard, E.;
9
10
11
Norskov, J. K.; Besenbacher, F. Bonding of Gold NanoClusters to Oxygen
12
13
Vacancies on Rutile TiO2 (110). Phys. Rev. Lett. 2003, 90, 026101.
14
15
(9) Matthey, D.; Wang, J. G.; Wendt, S.; Matthiesen, J.; Schaub, R.; Laegsgaard, E.;
16
17 Hammer, B.; Besenbacher, F. Enhanced Bonding of Gold NanoParticles on
18
19 Oxidized TiO2 (110). Science 2007, 315, 1692–1696.
20
21
22
(10) Shibata, N.; Goto, A.; Matsunaga, K.; Mizoguchi, T.; Findlay, S. D.; Yamamoto, T.;
23
24
25 Ikuhara, Y. Interface Structures of Gold NanoParticles on TiO 2 (110). Phys. Rev.
26
27 Lett. 2009, 102, 136105.
28
29 (11) Kolmakov, A.; Goodman, D. W. Scanning Tunneling Microscopy of Gold Clusters on TiO2
30
31 (110): CO Oxidation at Elevated Pressures. Surface Science 2001, 490, L597–L601.
32
33
(12) Valden, M.; Lai, X.; Goodman, D. W. Onset of Catalytic Activity of Gold Clusters on Titania
34
35
36 with the Appearance of Nonmetallic Properties. Science 1998, 281, 1647–1650.
37
38 (13) Kolmakov, A.; Goodman, D. W. Imaging Gold Clusters on TiO 2 (110) at Elevated
39
40 Pressures and Temperatures. Catalysis Letters 2000, 70, 93–97.
41
42
43 (14) Galhenage, R. P.; Yan, H.; Tenney, S. A.; Park, N.; Henkelman, G.; Albrecht, P.; Mullins, D.
44
45 R.; Chen, D. A. Understanding the Nucleation and Growth of Metals on TiO 2: Co
46
47 Compared to Au, Ni, and Pt. Journal of Physical Chemistry C 2013, 117, 7191–7201.
48
49
(15) Wilson, A.; Bernard, R.; Vlad, A.; Borensztein, Y.; Coati, A.; Croset, B.; Garreau, Y.;
50
51
52 Prévot, G. Epitaxial Growth of Bimetallic Au-Cu Nanoparticles on TiO 2(110)
53
54 Followed in situ by Scanning Tunneling Microscopy and Grazing-Incidence X-Ray
55
56
Diffraction. Phys. Rev. B 2014, 90, 075416.
57
58 32
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 34 of 41

1
2
3
(16) Cosandey, F.; Zhang, L.; Madey, T. E. Effect of Substrate Temperature on the
4
5
Epitaxial Growth of Au on TiO2 (110). Surface Science 2001, 474, 1–13.
6
7
8 (17) Akita, T.; Kohyama, M.; Haruta, M. Electron Microscopy Study of Gold Nanoparticles De-
9
10
posited on Transition Metal Oxides. Accounts of Chemical Research 2013, 46, 1773–1782.
11
12
13 (18) Delannoy, L.; Chantry, R. L.; Casale, S.; Li, Z. Y.; Borensztein, Y.; Louis, C. HRTEM
14
15
and STEM-HAADF Characterisation of Au-TiO 2 and Au-Al2O3 Catalysts for a
16
17 Better Under-standing of the Parameters Influencing their Properties in CO
18
19 Oxidation. Phys. Chem. Chem. Phys. 2013, 15, 3473–3479.
20
21 (19) Saint-Lager, M.-C.; Laoufi, I.; Bailly, A. Operando Atomic Structure and Active Sites
22
23
24 of TiO2 (110)-Supported Gold Nanoparticles during Carbon Monoxide Oxidation.
25
26 Faraday Discus-sions 2013, 162, 179–190.
27
28
(20) Møller, P. J.; Wu, M.-C. Surface Geometrical Structure and Incommensurate
29
30
Growth: Ultra-thin Cu Films on TiO2 (110). Surface Science 1989, 224, 265 – 276.
31
32
33 (21) Charlton, G.; Howes, P. B.; Muryn, C. A.; Raza, H.; Jones, N.; Taylor, J. S. G.;
34
35 Norris, C.; McGrath, R.; Norman, D.; Turner, T. S.; et al:, Copper Interface Induced
36
37 Relaxation of TiO2 (110)-1 1. Phys. Rev. B 2000, 61, 16117–16120.
38
39
40 (22) Chen, D. A.; Bartelt, M. C.; Hwang, R. Q.; McCarty, K. F. Self-limiting Growth of
41
42 Copper Islands on TiO2 (110)-(1 1). Surface Science 2000, 450, 78–97.
43
44
45 (23) Giordano, L.; Pacchioni, G.; Bredow, T.; Sanz, J. F. Cu, Ag, and Au Atoms Adsorbed on
46
47 TiO2 (110): Cluster and Periodic Calculations. Surface Science 2001, 471, 21–31.
48
49
50 (24) Wang, Y.; Hwang, G. S. Adsorption of Au Atoms on Stoichiometric and Reduced TiO 2
51
52 (110) Rutile Surfaces: a First Principles Study. Surface Science 2003, 542, 72–80.
53
54
(25) Vijay, A.; Mills, G.; Metiu, H. Adsorption of Gold on Stoichiometric and Reduced
55
56
Rutile TiO2 (110) Surfaces. Journal of Chemical Physics 2003, 118, 6536–6551.
57
58
59 33
60 ACS Paragon Plus Environment
Page 35 of 41 The Journal of Physical Chemistry

1
2
3
(26) Okazaki, K.; Morikawa, Y.; Tanaka, S.; Tanaka, K.; Kohyama, M. Electronic Structures
4
5
of Au on TiO2 (110) by First-Principles Calculations. Phys. Rev. B 2004, 69, 235404.
6
7
8 (27) Iddir, H.; Ogut, S.; Browning, N. D.; Disko, M. M. Adsorption and Diffusion of Pt and Au on the
9
10
11 Stoichiometric and Reduced TiO2 Rutile (110) Surfaces. Phys. Rev. B 2005, 72, 081407.
12
13 (28) Pillay, D.; Hwang, G. S. Growth and Structure of Small Gold Particles on Rutile
14
15
16
TiO2(110). Phys. Rev. B 2005, 72, 205422.
17
18
19 (29) Pillay, D.; Wang, Y.; Hwang, G. S. Nucleation and Growth of 1B Metal Clusters on
20
21 Ru-tile TiO2 (110): Atomic Level Understanding from First Principles Studies.
22
23 Catalysis Today 2005, 105, 78 – 84.
24
25
(30) Pillay, D.; Hwang, G. S. Structure of small Au n, Agn, and Cun Clusters (n=2-4) on
26
27
28 Rutile TiO2 (110): A density Functional theory Study. Journal of Molecular
29
Structure: THEOCHEM 2006, 771, 129 – 133.
30
31
32 (31) Chrétien, S.; Metiu, H. Density Functional Study of the Interaction between Small
33
34
Au Clus-ters, Aun (n=1-7) and the Rutile TiO2 Surface. I. Adsorption on the
35
36 Stoichiometric Surface. The Journal of Chemical Physics 2007, 127, 084704.
37
38
39 (32) Chrétien, S.; Metiu, H. Density Functional Study of the Interaction between Small
40
41 Au Clus-ters, Aun (n=1-7) and the Rutile TiO 2 Surface. II. Adsorption on a Partially
42
43 Reduced Surface. The Journal of Chemical Physics 2007, 127, 244708.
44
45
46 (33) Chrétien, S.; Metiu, H. Enhanced Adsorption Energy of Au 1 and O2 on the
47
48 Stoichiometric TiO2 (110) Surface by Coadsorption with Other Molecules. The
49
50 Journal of Chemical Physics 2008, 128, 044714.
51
52 (34) Chrétien, S.; Metiu, H. O2 evolution on a clean partially Reduced Rutile TiO 2 (110)
53
54
Surface and on the same Surface precovered with Au 1 and Au2: The importance of
55
56 spin conservation. The Journal of Chemical Physics 2008, 129, 074705.
57
58
59 34
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 36 of 41

1
2
3 (35) Madsen, G. K. H.; Hammer, B. Effect of Subsurface Ti-Interstitials on the Bonding of Small
4
5
Gold Clusters on Rutile TiO2 (110). Journal of Chemical Physics 2009, 130, 044704.
6
7
8
(36) Ammal, S. C.; Heyden, A. Modeling the Noble Metal/TiO 2 (110) Interface with
9
10
11
Hybrid DFT Functionals: A Periodic Electrostatic Embedded Cluster Model Study.
12 Journal of Chemical Physics 2010, 133, 164703.
13
14
15 (37) Pabisiak, T.; Kiejna, A. Stability of Gold Nanostructures on Rutile TiO 2 (110)
16
17 Surface. Sur-face Science 2011, 605, 668–674.
18
19
(38) Camellone, M. F.; Kowalski, P. M.; Marx, D. Ideal, Defective, and Gold-promoted Rutile
20
21
22 TiO2 (110) Surfaces Interacting with CO, H 2, and H2O: Structures, Energies,
23
24 Thermody-namics, and Dynamics from PBE+U. Phys. Rev. B 2011, 84, 035413.
25
26 (39) Goldman, N.; Browning, N. D. Gold Cluster Diffusion Kinetics on Stoichiometric and
27
28
29 Re-duced Surfaces of Rutile TiO 2 (110). Journal of Physical Chemistry C 2011,
30
31
115, 11611– 11617.
32
33 (40) Cai, Y.; Bai, Z.; Chintalapati, S.; Zeng, Q.; Feng, Y. P. Transition Metal Atoms Pathways on
34
35 3+
Rutile TiO2 (110) Surface: Distribution of Ti states and evidence of enhanced peripheral
36
37
charge accumulation. Journal of Chemical Physics 2013, 138, 154711.
38
39
40 (41) Vilhelmsen, L. B.; Hammer, B. A Genetic Algorithm for First Principles Global
41
42 Structure Optimization of Supported Nano Structures. The Journal of Chemical
43
44 Physics 2014, 141, 044711.
45
46
(42) Vilhelmsen, L. B.; Hammer, B. Identification of the Catalytic Site at the Interface
47
48
49
Perimeter of Au Clusters on Rutile TiO2 (110). ACS Catalysis 2014, 4, 1626–1631.
50
51
(43) Parker, S.; Grant, A.; Bondzie, V.; Campbell, C. Island Growth Kinetics during the
52
53
Vapor Deposition of Gold onto TiO2 (110). Surface Science 1999, 441, 10 – 20.
54
55
56
57
58
35
59
ACS Paragon Plus Environment
60
Page 37 of 41 The Journal of Physical Chemistry

1
2
3
(44) Campbell, C. T.; Parker, S. C.; Starr, D. E. The Effect of Size-Dependent
4
5
6
NanoParticle Ener-getics on Catalyst SIntering. Science 2002, 298, 811–814.
7
8 (45) Cheng, J.; Sprik, M. Acidity of the Aqueous Rutile TiO 2 (110) Surface from Density
9
10 Func-tional Theory Based Molecular Dynamics. Journal of Chemical Theory and
11
12 Computation 2010, 6, 880–889.
13
14
(46) Kresse, G.; Hafner, J. Ab initio Molecular Dynamics for Liquid Metals. Phys. Rev. B
15
16
17
1993, 47, 558.
18
19 (47) Kresse, G.; Furthmüller, J. Efficiency of Ab-initio Total Energy Calculations for Metals
20
21 and Semiconductors Using a Plane-wave Basis Set. Comput. Mat. Sci. 1996, 6, 15.
22
23
24 (48) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab-initio Total-energy
25
26 Calculations Using a Plane-wave Basis Set. Phys. Rev. B 1996, 54, 11169.
27
28
29 (49) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made
30
31 Simple. Phys. Rev. Lett. 1996, 77, 3865.
32
33 (50) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector
34
35 Augmented-wave Method. Phys. Rev. B 1999, 59, 1758.
36
37
38 (51) Morgan, B. J.; Watson, G. W. A DFT+U Description of Oxygen Vacancies at the
39
40 TiO2 Rutile (110) Surface. Surface Science 2007, 601, 5034–5041.
41
42
43 (52) Henrich, V. E.; Dresselhaus, G.; Zeiger, H. J. Defect Surface-states On TiO 2 - 2-dimensional
44
45 Surface Phases. Bulletin of the American Physical Society 1976, 21, 940–941.
46
47 (53) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P.
48
49 Electron-energy-loss Spectra and the Structural Stability of Nickel Oxide: An
50
51 LSDA+U Study. Phys. Rev. B 1998, 57, 1505–1509.
52
53
54 (54) Monkhorst, H. J.; Pack, J. D. Special Points For Brillouin-zone Integrations. Phys.
55
56 Rev. B 1976, 13, 5188–5192.
57
58 36
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 38 of 41

1
2
3
(55) Vinet, P.; Ferrante, J.; Smith, J. R.; Rose, J. H. A Universal Equation of State For
4
5
6
Solids. Journal of Physics C-solid State Physics 1986, 19, L467–L473.
7
8 (56) Lee, K.; Murray, E. D.; Kong, L.; Lundqvist, B. I.; Langreth, D. C. Higher-accuracy
9
10
van der Waals density Functional. Phys. Rev. B 2010, 82, 081101.1–081101.4.
11
12
13 (57) Wellendorff, J.; Lundgaard, K. T.; Møgelhoj, A.; Petzold, V.; Landis, D. D.; Nørskov, J. K.;
14
15 Bligaard, T.; Jacobsen, K. W. Density Functionals for Surface Science: Exchange-correlation
16
17 Model Development with Bayesian Error Estimation. Phys. Rev. B 2012, 85, 235149.
18
19
20 (58) Di Valentin, C.; Pacchioni, G.; Selloni, A. Electronic Structure of Defect States in HydrOxy-
21
22 lated and Reduced Rutile TiO2 (110) Surfaces. Phys. Rev. Lett. 2006, 97, 166803.
23
24
25 (59) Janotti, A.; Varley, J. B.; Rinke, P.; Umezawa, N.; Kresse, G.; Van de Walle, C. G. Hybrid
26
27 Functional Studies of the Oxygen Vacancy in TiO2. Phys. Rev. B 2010, 81, 085212.
28
29
30 (60) Malashevich, A.; Jain, M.; Louie, S. G. First-principles DFT + GW Study of Oxygen
31
32 Vacan-cies in Rutile TiO2. Phys. Rev. B 2014, 89, 075205.
33
34
35 (61) Li, H.; Guo, Y.; Robertson, J. Calculation of TiO 2 Surface and SubSurface Oxygen
36
37 Va-cancy by the Screened Exchange Functional. The Journal of Physical
38
39
Chemistry C 2015, 119, 18160–18166.
40
41 (62) Curnan, M. T.; Kitchin, J. R. Investigating the Energetic Ordering of Stable and
42
43
44 Metastable TiO2 Polymorphs Using DFT+U and Hybrid Functionals. The Journal of
45
46 Physical Chemistry C 2015, 119, 21060–21071.
47
48 (63) Gerosa, M.; Bottani, C. E.; Caramella, L.; Onida, G.; Di Valentin, C.; Pacchioni, G.
49
50 Defect Calculations in Semiconductors through a Dielectric-dependent Hybrid DFT
51
52 Functional: The Case of Oxygen Vacancies in Metal Oxides. The Journal of
53
54 Chemical Physics 2015, 143, 134702.
55
56
57
58
59 37
60 ACS Paragon Plus Environment
Page 39 of 41 The Journal of Physical Chemistry

1
2
3
(64) van Santen, R. A.; Niemantsverdiret, J. W. Chemical Kinetics and Catalysis;
4
5
Fundamental and Applied Catalysis, 1995.
6
7
8 (65) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid Functionals Based on a Screened
9
10 Coulomb Potential. The Journal of Chemical Physics 2003, 118, 8207–8215.
11
12
13 (66) Heyd, J.; Peralta, J. E.; Scuseria, G. E.; Martin, R. L. Energy Band Gaps and
14
15 Lattice Param-eters Evaluated with the Heyd-Scuseria-Ernzerhof Screened Hybrid
16
17 Functional. The Journal of Chemical Physics 2005, 123, 174101.
18
19
20 (67) Dovesi, R.; Orlando, R.; Erba, A.; Zicovich-Wilson, C. M.; Civalleri, B.; Casassa, S.;
21
22 Mas-chio, L.; Ferrabone, M.; De La Pierre, M.; D’Arco, P.; et al:, CRYSTAL14: A
23
24 Program for the Ab Initio Investigation of Crystalline Solids. International Journal of
25
26 Quantum Chemistry 2014, 114, 1287–1317.
27
28
29 (68) Piskunov, S.; Heifets, E.; Eglitis, R. I.; Borstel, G. Bulk Properties and Electronic
30
31 Structure of SrTiO3, BaTiO3, PbTiO3 Perovskites: an Ab Initio HF/DFT Study.
32
33 Computational Materials Science 2004, 29, 165 – 178.
34
35 (69) Gatti, C.; Saunders, V. R.; Roetti, C. Crystal Field Effects on the Topological
36
37 Properties of the Electron Density in Molecular Crystals: The Case of Urea. The
38
39
Journal of Chemical Physics 1994, 101, 10686–10696.
40
41
42 (70) Doll, K.; Pyykkö, P.; Stoll, H. Closed-shell Interaction in Silver and Gold Chlorides.
43
44 The Journal of Chemical Physics 1998, 109, 2339–2345.
45
46
47 (71) Chrétien, S.; Metiu, H. Electronic Structure of Partially Reduced Rutile TiO 2 (110)
48
49 Surface: Where Are the Unpaired Electrons Located? The Journal of Physical
50
51 Chemistry C 2011, 115, 4696–4705.
52
53
(72) Krüger, P.; Jupille, J.; Bourgeois, S.; Domenichini, B.; Verdini, A.; Floreano, L.; Mor-
54
55
56
57
58 38
59 ACS Paragon Plus Environment
60
The Journal of Physical Chemistry Page 40 of 41

1
2
3
4
gante, A. Intrinsic Nature of the Excess Electron Distribution at the TiO 2(110) Surface. Phys.
5
Rev. Lett. 2012, 108, 126803.
6
7
8 (73) de Dompablo, M. E. A.; Morales-García, A.; Taravillo, M. DFT + U Calculations of
9
10
11 Crystal Lattice, Electronic Structure, and Phase Stability under Pressure of TiO 2
12
13 polymorphs. The Journal of Chemical Physics 2011, 135, 054503.
14
15 (74) Le Bahers, T.; Rérat, M.; Sautet, P. Semiconductors Used in Photovoltaic and
16
17 PhotoCatalytic Devices: Assessing Fundamental Properties from DFT. The Journal
18
19 of Physical Chemistry C 2014, 118, 5997–6008.
20
21
22 (75) Iachella, M.; Wilson, A.; Naitabdi, A.; Bernard, R.; Prevot, G.; Loffreda, D. Promoter
23
24 effect of hydration on the nucleation of nanoparticles: direct observation for gold
25
26 and copper on rutile TiO2 (110). Nanoscale 2016, 8, 16475–16485.
27
28
29 (76) Takei, T.; Akita, T.; Nakamura, I.; Fujitani, T.; Okumura, M.; Okazaki, K.; Huang, J.;
30
31 Ishida, T.; Haruta, M. In Chapter One - Heterogeneous Catalysis by Gold; Gates,
32
33 B. C., Jentoft, F. C., Eds.; Advances in Catalysis Supplement C; Academic Press,
34
35
2012; Vol. 55; pp 1–126.
36
37
(77) Tada, K.; Koga, H.; Hayashi, A.; Kondo, Y.; Kawakami, T.; Yamanaka, S.; Okumura, M.
38
39
40 Ef-fects of Halogens on Interactions Between a Reduced TiO 2 Surface and Noble
41
42 Metal Atoms: a DFT Study. Applied Surface Science 2017, 411, 149–162.
43
44 (78) Metiu, H.; Chrétien, S.; Hu, Z.; Li, B.; Sun, X. Chemistry of Lewis Acid-Base Pairs
45
46 on Oxide Surfaces. The Journal of Physical Chemistry C 2012, 116, 10439–10450.
47
48
49 (79) Lira, E.; Hansen, J. O.; Merte, L. R.; Sprunger, P. T.; Li, Z.; Besenbacher, F.;
50
51
Wendt, S. Growth of Ag and Au Nanoparticles on Reduced and Oxidized Rutile
52
53 TiO2 (110) Surfaces. Topics In Catalysis 2013, 56, 1460–1476.
54
55
56
57
58
59 39
60 ACS Paragon Plus Environment
Page 41 of 41 The Journal of Physical Chemistry

1
2
3 (80) Zhou, J.; Chen, D. A. Controlling Size Distributions of Copper Islands Grown on
4
5
TiO2 (110)-(1 2). Surface Science 2003, 527, 183–197.
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 40
60 ACS Paragon Plus Environment

Vous aimerez peut-être aussi