Sanchez 2018
Sanchez 2018
Alberto Roldan, *a Ceri Hammond, a Qian He,a Tom Davies,a Alberto Villa*e
and Nikolaos Dimitratos *a
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
48.4 kJ mol1) and dehydration (HCOOH / CO + H2O, DG ¼ the catalytic performance in the case of the liquid phase
28.5 kJ mol1). Acidity or high temperature during the reac- decomposition of formic acid under mild reaction conditions.
tion usually promotes the dehydration pathway. Ultrapure The CNFs used are: (1) pyrolytically stripped (PS-CNF), (2) low-
hydrogen is necessary for the generation of energy with fuel cell heat treated (LHT-CNF) and (3) high-heat treated (HHT-CNF).
devices, and since they are poisoned by CO, inlet CO concen- The choice of the described CNFs allowed us to study the
tration should remain below 20 ppm, otherwise a loss of long- effect of the degree of CNF graphitisation. Furthermore, we
term performance can occur. Mild conditions are also of key studied the effect of these CNFs on the supported Pd particles'
importance since these fuel cells are expected to supply energy morphology prepared by a sol immobilisation method (using
to portable devices with a low heat management prole. polyvinyl alcohol (PVA) as a stabiliser and NaBH4 as a reducing
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018
View Article Online
under vigorous stirring. Aer one hour, the catalyst was ltered a Hitachi TM3030PLUS equipped with a Quantax70 energy-
and washed several times with distilled water (2 L) to remove dispersive X-ray spectroscope (EDX) to study morphology and
species, such as Na+ and Cl, and to reach neutral pH. The determine the palladium content of the samples (fresh and
samples were dried in an oven at 80 C for two hours under used). The BET surface area was determined from the N2
static air. The amount of support added was calculated to adsorption–desorption isotherms in liquid nitrogen at 77 K
obtain a nal nominal metal loading of 1 wt%. The obtained using a Quantachrome NOVA 2200e instrument. The samples
catalysts were labelled PdSI/CNF-HHT, PdSI/CNF-LHT and PdSI/ were outgassed for 3 h under vacuum at 227 C. The total
CNF-PS. surface area was determined using the BET (Brunauer–Emmett–
Impregnation followed by chemical reduction with sodium Teller) equation and the multi-point method.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
tted with a CP-Sil 5CB capillary column (50 m length, 0.32 mm surface properties of the material. The most visible feature in
diameter, and carrier gas: He), a methanator unit and both FID the X-ray diffraction patterns of the fresh series of catalysts is
and TCD detectors with a detection limit of CO below 5 ppm. the diffraction peak at approximately 26 , assigned to the
The gases were quantied using calibration curves created from presence of the (002) plane of graphitic carbon (Fig. 1).35 The
commercial standards (BOC gases). intensity of the corresponding diffraction peak increases during
Computational methods. Periodic plane-wave DFT calcula- the heat treatment process since the utilisation of high-
tions were performed using the Vienna ab-initio simulation temperature post-treatment signicantly enhances the
package (VASP),37,38 the Perdew–Burke–Ernzerhof functional graphitic character of carbon materials and therefore, the
revised for solids39 and a kinetic energy of 500 eV to expand the absence of amorphous carbon. The intensity of this character-
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 28 September 2018. Downloaded on 9/30/2018 11:27:22 AM.
plane-waves of the Kohn–Sham valence states.40 The inner istic diffraction peak is higher for the CNF-HHT samples.
electrons were represented by the projector-augmented wave In Fig. 2, the XRD patterns of the series of catalysts are
(PAW) pseudopotentials also considering non-spherical contri- presented in more detail. Between 42 and 46 , there is a broad
butions from the gradient corrections.41 All the calculations diffraction peak attributed to either (100) or (101) planes of C.
include the long-range dispersion correction approach by However since both hexagonal and rhombohedral graphite
Grimme (D3),42,43 which is an improvement on pure DFT when peaks are present in this region, it is difficult to relate speci-
considering large polarisable atoms.44–49 We included a self- cally both species and diffraction peaks.35
consistent aqueous implicit solvation model.50,51 The optimi- The two catalysts supported on CNF-HHT present intense
sation thresholds were 105 eV and 0.03 eV Å1 for electronic and sharp diffraction peaks at 54 and 78 . They are also
and ionic relaxation, respectively. The Brillouin zone was observed in the other catalysts albeit to a much lesser extent.
sampled with a G-centred k-point mesh generated through The diffraction peaks at 78 correspond to the graphite (110)
a Monkhorst–Pack grid of 5 5 1 k-points, which ensures the plane conrming the presence of rhombohedral graphite.35 The
electronic and ionic convergence.52 In order to improve the assignment of the diffraction peak at 54 is not straightforward
convergence of the Brillouin-zone integrations, the partial since both graphite (004) and PdO (112) planes could be
occupancies were determined using the rst order Methfessel– assigned to the same position.35,56 To clarify this argument, the
Paxton method corrections smearing with a set width for all XRD patterns of the bare supports were analysed, as shown in
calculations of 0.2 eV. Open shell calculations were tested Fig. S1.† It conrms the presence of the graphite (004) plane;
leading to close shell results.
The Pd bulk lattice parameter is 3.893 Å (ref. 53) which is in
very good agreement with the one resulting from our cell opti-
misation (3.836 Å). We have modelled low-Miller index surfaces,
where due to crystal symmetry, we have reduced the number of
surfaces to the {111}, {011} (consisting of the equivalent (011),
(101), and (110) faces), and {001} (which includes the equivalent
(001), (010), and (100) faces). All have a coordination number of
12. We believe that this analysis will provide a deeper under-
standing since most of the literature focus on Pd(111) only.54,55 Fig. 1 XRD patterns of fresh Pd/CNFs. (A) Catalysts synthesised by
These surfaces are simulated by using a slab model containing 5 impregnation. (B) Catalysts synthesised by sol-immobilisation. (a)
atomic layers; the two uppermost layers were relaxed without CNF-HHT, (b) CNF-LHT and (c) CNF-PS.
symmetry restrictions, and the bottom ones were frozen at the
bulk lattice parameter. The slab contains 45 atoms per unit cell
exposing an area of 66.217, 93.645 and 66.217 Å2 for (111), (011)
and (001) respectively. We added a vacuum width of 15 Å
between vertically repeated slabs, to avoid the interaction
between them.
We dened the binding energy (EB) as the difference between
the combined system and the isolated species, and the reaction
energy (ER) of each step is calculated as the total energy differ-
ence between the nal state (product(s)) and the initial state
(reactant(s)). Thus, negative values of EB and ER indicate
favourable adsorption and reaction respectively.
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018
View Article Online
however, this peak shows a lower intensity compared with the Table 2 Atomic content of sp2 and sp3 carbon and sp2/sp3 ratio from
XRD pattern of the Pd/CNFs. Therefore, we can suppose that the XPS and ID/IG ratio from Raman for the bare supports and the catalysts
subjected to different temperature heat treatments
PdO(112) plane is overlaid by the graphite (004) plane and these
results indicate the presence of PdO species. The characteristic ID/IG
planes (111), (200) and (220) of the face-centered cubic structure
of Pd56,57 are assigned to the reections at 2q ¼ 40.4 , 44.9 and Catalyst C sp2 (%) C sp3 (%) sp2/sp3 Fresh Used
68.3 (grey lines in Fig. 2) and are present only in some catalysts
CNF-HHT — — — 0.11 —
as is evident from the XRD patterns. It is well known that one of CNF-LHT — — — 0.71 —
the limitations of the XRD technique is its sensitivity to small CNF-PS — — — 0.75 —
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 28 September 2018. Downloaded on 9/30/2018 11:27:22 AM.
crystallite sizes, with crystallite sizes lower than 5 nm being PdIMP/CNF-HHT 82.10 7.50 10.95 0.26 0.21
below the detection limit of the apparatus.33 Therefore, the PdIMP/CNF-LHT 72.01 17.43 4.13 0.78 0.73
PdIMP/CNF-PS 66.73 22.37 2.98 0.67 0.57
absence of these patterns for the catalysts synthesised by sol-
PdSI/CNF-HHT 81.87 8.11 10.09 0.08 0.23
immobilisation could be attributed to a particle size below 5 PdSI/CNF-LHT 70.44 19.05 3.70 0.80 0.90
mm. The crystallinity will be subsequently studied by means of PdSI/CNF-PS 65.80 23.12 2.85 0.78 0.71
TEM. On the other hand, for the catalysts synthesised by
impregnation, it is easier to observe the presence of metallic Pd
and PdO species, which means that the particles could be
larger. The presence and percentage of metallic Pd and PdII
species have been studied by XPS. Note that the diffraction peak
at 36 only appears for the catalyst whose support has been
treated at the lowest temperature; therefore, the most plausible
reason is the presence of an unidentied contaminant. The
XRD patterns of the used samples are displayed in Fig. S2.† The
appearance and increment of the reections at 2q ¼ 40.4 and
44.9 ascribed to the (111) and (200) planes of metallic Pd
indicate (i) the progressive reduction of PdO to metallic Pd due
to H2 generation and (ii) the increase in particle size due to Fig. 3 XPS spectra of fresh Pd/CNF. (A) Catalysts synthesised by
agglomeration. impregnation. (B) Catalysts synthesised by sol-immobilisation. (a)
XPS analysis of the as-synthesised and selected used samples CNF-HHT, (b) CNF-LHT and (c) CNF-PS.
was carried out to determine the electronic states of the samples
(Tables 1 and 2). The XPS spectra of Pd 3d of the as-synthesised
fresh catalysts are presented in Fig. 3 for impregnation and sol- content of Pd0 species indicating that the presence of the PVA
immobilisation methods, respectively. Pd and O content and ligand may inhibit the oxidation of the metallic Pd surface in
the percentage of Pd0 for both fresh and used series derived ambient air. This feature has a signicant impact on the catalyst
from XPS are shown in Table 1. activity as will be discussed later.
The Pd 3d5/2 component at 335 eV approximately was No apparent trend can be extracted for the variations in Pd
assigned to metallic Pd58 and the component at approximately surface content or the percentage of Pd0 according to modi-
337 eV to PdII mainly present as PdO.59 As presented in Table 1, cations in the morphology through temperature treatment. XPS
the impregnated samples display, in general, a lower Pd atomic analyses of the used catalysts are presented in Fig. S3.† Table 1
percentage (0.40–0.71%) in contrast with the sol- shows a reduction in the atomic Pd content on the CNF surface.
immobilisation prepared samples (0.90–1.44%). From Fig. 3 Since XPS is surface sensitive, the decrease of Pd content for the
and Table 1 it is evident that the majority of the catalysts used catalysts could be explained by (i) Pd leaching during the
prepared by the sol-immobilisation method display a higher reaction, (ii) migration of the nanoparticles to the inner wall of
the nanobers and (iii) increase of the Pd mean particle size.
Table 1 displays the atomic percentage of Pd0 of the used
Table 1 Palladium content, (%) Pd0 on the surface and O content from
catalysts, and it is clear that there is a signicant increase of
XPS data for the different catalysts studied, A: Fresh, and B: Used
metallic Pd aer the reaction, suggesting that the hydrogen
Pd content (%) Pd0 on the O content released during the decomposition of HCOOH can facilitate the
(at%) surface (at%) reduction of PdII to Pd0 species. These results are in agreement
with the appearance or increase of the intensity of Pd0 diffrac-
Catalyst Fresh Used Fresh Used Fresh Used
tion peaks in the XRD patterns of the used catalysts (Fig. S2†).
PdIMP/CNF-HHT 0.71 0.52 30.1 74.1 0.94 1.40 XPS has been used to measure the relative concentration of
PdIMP/CNF-LHT 0.57 0.49 55.0 73.7 2.60 2.94 sp3 and sp2 hybridisation types from the deconvolution of the C
PdIMP/CNF-PS 0.40 0.23 45.1 61.2 2.94 6.34 1s region. The C 1s XPS spectra of the fresh and used series of
PdSI/CNF-HHT 0.93 0.72 51.9 73.1 2.72 4.05
catalysts are presented in Fig. S4.† Table 2 displays the
PdSI/CNF-LHT 1.44 0.98 53.1 69.8 4.49 3.49
PdSI/CNF-PS 0.90 0.04 53.7 84.9 5.07 18.23 concentration of sp3 and sp2 hybridisation and their ratio (sp2/
sp3) since it determines the structural properties of carbon
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018
View Article Online
of sp2 and sp3 from XPS for the catalysts is in agreement with
the results obtained from Raman; the catalysts annealed at
3000 C present the highest graphitisation and a big difference
with LHT samples (annealed at 1500 C) as observable in the
sp2/sp3 ratio. Between LHT and PS samples there is a small
difference conrmed by both the concentrations of sp2 and sp3
and the ID/IG ratio.
The particle size distributions of the catalyst series for both
impregnated and sol-immobilised samples were assessed from
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
Table 3 Statistical mean and standard deviation of particle size distributed in the inner walls of CNFs, besides being deposited
analysis on the external surface.
Table 5 presents the total surface area determined from the
Fresh Used
BET equation. It ranges from approximately 36 m2 g1 to
Mean particle Standard Mean particle Standard 52 m2 g1. This low surface area compared with that of carbon
Catalyst size (nm) deviation size (nm) deviation nanotubes (up to 1200 m2 g1) is caused by the thickness of the
walls (ca. 45 nm). Since BET surface area instruments have
PdIMP/CNF-HHT 5.4 0.9 5.5 0.4
PdIMP/CNF-LHT 5.7 1.3 7.3 0.4 a certain error, in some cases up to 10%, no conclusion can be
PdIMP/CNF-PS 6.9 1.8 6.8 0.3 extracted for the apparent variations in the surface area of these
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 28 September 2018. Downloaded on 9/30/2018 11:27:22 AM.
Table 4 Palladium loading from EDX and AAS data for the different
catalysts studied
Pd loading Pd loading
EDX (wt%) AAS (wt%)
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018
View Article Online
Table 6 Comparative catalytic activity of various Pd-based catalysts for liquid-phase formic acid dehydrogenation under mild conditions
TOF (h1)
Activation energy
Catalyst T ( C) Reagent Initial 2h (kJ mol1) Ref.
PdSI/CNF-PS 30 484.4
Pd/C 21 Formic acid (1.33 M) 18 15a 53.7 25
30 48 28a
Pd/C (citric acid) 25 Formic acid 64b 57
Pd/C 30 Formic acid/sodium formate 1 : 9 228.3 24
Au41Pd59/C 50 Formic acid (1 M) 230 28 2 13
Ag@Pd (1 : 1) 35 Formic acid 156c 30 19
50 252c
Ag/Pd alloy (1 : 1) 20 144c
Ag42Pd58 50 Formic acid (1 M) 382 22 1 65
Pd–MnOx/SiO2–NH2 20 Formic acid (0.265 M) 140 61.9 66
50 1300
Ag0.1Pd0.9/rGO 25 Formic acid 105 67
a b c
TOF calculated aer 50 min. TOF calculated aer 160 min. TOF calculated based on the surface metal sites.
important role regarding the catalytic activity. Analyses by TEM The most common (111), (011) and (001) surfaces have been
conrmed a lower mean particle size for the sol-immobilisation used to simulate each step of formic acid decomposition. The
samples. Therefore, the higher catalytic activity observed could energy prole (Fig. 10) presents the energy requirements of two
be attributed to the smaller Pd mean particle size and as different paths initiated by O–H (1: HCOO formation) and C–H
a consequence the increasing proportion of Pd surface atoms. (2: COOH formation) dissociation.
However, aer 2 hours of reaction, as observed in Fig. 9, the (111) Surface. The HCOO pathway begins with the adsorp-
three pairs of catalysts led to similar nal conversion even tion of trans-HCOOH (EB ¼ 0.77 eV) and the subsequent
though the TOF values differ signicantly during the initial splitting of the O–H bond (ER ¼ 0.32 eV). The separation of the
period. As previously commented, PVA used during the sol- hydrogen atom and breakage of C–H suppose a further stabili-
immobilisation catalyst preparation leads to a higher pres- sation of the system by 0.04 eV. The COOH pathway starts with
ence of C–O–H that might be occupying active sites and leading a reorientation of HCOOH from trans to cis conguration, which
to a slightly faster deactivation than catalysts prepared by the imposes a slight energy increment of 0.02 eV. Aer the
impregnation method. exothermic C–H scission (ER ¼ 1.20 eV), the hydrogen spills
Table S1† presents the H2, CO2 and CO concentrations over the catalyst, which further stabilises the carboxylic species
evolved from formic acid decomposition by the most active by 0.08 eV. From this intermediate, both pathways 1 and 2
catalysts studied in this work. As observed, very low CO either follow route a to produce CO2 or route b to CO. Though
concentration is produced by both preparation methods, with both formate (1a) and carboxylic (2a) intermediates are very
the lowest level of CO production being observed for catalysts close in energy (1.13 and 1.25 eV respectively), HCOO (1a) is
prepared by the colloidal method. the preferable intermediate since C–H scission is largely
unfavourable as previously reported.12 HCOO decomposition
DFT results leads to CO2 and H2 (1a) because the C–O dissociation step (1b)
is highly unfavourable (+0.94 eV from adsorbed HCOO)
The characterisation data of the supported Pd nanoparticles supporting the low ppm level concentration of CO observed
indicate that the majority of the Pd species are metallic and (Table S1†).
there is a variation in the Pd mean particle size in the range of (011) Surface. In this case, the adsorption of trans-HCOOH
4–7 nm. Moreover, the catalytic activity showed a clear depen- supposes a relaxation of EB ¼ 0.94 eV and the splitting of the
dence on the Pd mean particle size. Therefore, in order to O–H bond (ER ¼ 0.36 eV). Both separations of hydrogen and
provide insights into the decomposition of formic acid on breakage of C–H stabilise the system by 0.18 eV. Aer the re-
different sites that are expected to be present on Pd/CNFs with orientation of HCOOH to cis conformation and the C–H scission
different particle sizes, we focused our attention to carry out (ER ¼ 1.52 eV), hydrogen spillover stabilises the carboxylic
DFT calculations. They provided insights into the energetics of species by just 0.01 eV. As within the (111) surface, both path-
the pathways of formic acid decomposition (dehydrogenation ways 1 and 2 split following route a to CO2 and route b to CO
versus dehydration) on the most usual Pd(001), Pd(011) and respectively. The HCOO decomposition also leads to CO2 and
Pd(111) surfaces.
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
Conclusions
A series of monometallic Pd nanoparticles supported on several
carbon nanobers (CNFs) were synthesised. Three CNFs with
different graphitisation grades were used. Sol-immobilisation
and impregnation techniques were selected as model prepara-
tion methods widely used for the deposition of Pd
nanoparticles.
Fig. 10 Potential energy surface for formic acid decomposition on (A) Pd/CNFs prepared by the sol-immobilisation method exhibit
Pd(111), (B) Pd(011) and (C) Pd(001) surfaces. Red and blue lines indi- higher catalytic activity when compared with catalysts prepared
cate HCOO and COOH paths respectively. The solid lines lead to CO2 by the impregnation method due to (i) the higher surface
whereas the dashed line to CO. atomic Pd percentage, (ii) higher percentage of Pd0 and (iii)
smaller Pd particle size. Probably, the sol-immobilisation
method tends to distribute Pd metal nanoparticles on the
H2 (1a) because the C–O dissociation step (1b) is even more surface of the nanobers, whereas the impregnation method
unfavourable (+1.12 eV) for (011) than for the (111) surface and leads to a partial lling of the nanobers and distribution of Pd
thus, CO evolution is largely avoided. nanoparticles in the inner wall, besides the distribution on the
(001) Surface. The adsorption of trans-HCOOH supposes outer CNF surface.
a relaxation of EB ¼ 0.80 eV and the breakage of the O–H bond The heat treatment on CNFs affects catalyst activity. The
(ER ¼ 0.25 eV). The separation of hydrogen and the splitting of catalytic performance of the samples signicantly increases
C–H stabilise the system by 0.30 eV. Aer the re-orientation of with the increase of annealing temperature. A heat treatment at
HCOOH to cis and the C–H scission (ER ¼ 1.55 eV), the 1500 C did not signicantly modify the surface properties of
hydrogen separation destabilises the carboxylic species by CNFs. In contrast, treating CNFs at 3000 C rearranges the
0.04 eV. In this case, the C–O dissociation step (1b) is still structure improving the order and the degree of graphitisation.
unfavourable, but the value is much lower compared with the This increase of activity with annealing temperature has been
other two surfaces (+0.33 eV). Besides this, O–H spillover, the attributed to the presence of smaller Pd nanoparticles formed
last step in 2b is just slightly unfavourable compared as well and improved dispersion. A TOF of 979 h1 was achieved by
with (011) and (111) where it certainly needed more energy. Due PdSI/CNF-HHT, the most active catalyst in this series, with high
to these two facts, (001) is the surface that can lead to a higher selectivity for H2 (>99%) under mild reaction conditions, e.g.
concentration of CO since its evolution is not as unfavourable as 30 C. Finally, DFT studies provide valuable insights into the
in (011) and (111). role of under-coordinated sites in terms of activity and prefer-
Since 1a (via HCOO) is the main pathway followed by formic ence of reaction pathways toward CO2 and H2 evolution against
acid decomposition, according to Fig. 10, it is clear that for both CO formation and provide answers to the important question of
(111) and (011) surfaces, this pathway is exothermic. On the how to avoid/minimise the formation of CO. (001) was found to
other hand, for the (001) surface, pathway 1a nds two steps in be the surface in which CO would evolve at faster rates
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018
View Article Online
compared with (011) and (111). Therefore, a higher degree of 16 P. Sponholz, D. Mellmann, H. Junge and M. Beller,
(111) and (011) surfaces could in our opinion be a possible ChemSusChem, 2013, 6, 1172–1176.
solution in order to minimise CO formation. 17 H. Dai, N. Cao, L. Yang, J. Su, W. Luo and G. Cheng, J. Mater.
Chem. A, 2014, 2, 11060–11064.
Conflicts of interest 18 S.-T. Gao, W. Liu, C. Feng, N.-Z. Shang and C. Wang, Catal.
Sci. Technol., 2016, 6, 869–874.
“There are no conicts to declare”. 19 K. Tedsree, T. Li, S. Jones, C. W. A. Chan, K. M. K. Yu, P. a
J. Bagot, E. a Marquis, G. D. W. Smith and S. C. E. Tsang,
Acknowledgements Nat. Nanotechnol., 2011, 6, 302–307.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 28 September 2018. Downloaded on 9/30/2018 11:27:22 AM.
This journal is © The Royal Society of Chemistry 2018 Sustainable Energy Fuels
View Article Online
40 N. D. Mermin, Phys. Rev., 1965, 137, 1441–1443. 55 J. Cho, S. Lee, J. Han, S. P. Yoon, S. W. Nam, S. H. Choi,
41 G. Kresse, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, K. Y. Lee and H. C. Ham, J. Phys. Chem. C, 2014, 118,
59, 1758–1775. 22553–22560.
42 S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799. 56 C. J. Huang, F. M. Pan, T. C. Tzeng, C. Li and J. T. Sheu,
43 S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., J. Electrochem. Soc., 2009, 156, J28–J31.
2010, 132, 154104. 57 Z.-L. Wang, J.-M. Yan, H.-L. Wang, Y. Ping and Q. Jiang, Sci.
44 S. Irrera, A. Roldan, G. Portalone, N. H. De Leeuw and Rep., 2012, 2, 598–604.
N. H. De Leeuw, J. Phys. Chem. C, 2013, 117, 3949–3957. 58 M. C. Militello and S. J. Simko, Surf. Sci. Spectra, 1994, 3,
45 N. Y. Dzade, A. Roldan and N. H. De Leeuw, J. Chem. Phys., 387–394.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 28 September 2018. Downloaded on 9/30/2018 11:27:22 AM.
2013, 139, 124708. 59 M. C. Militello and S. J. Simko, Surf. Sci. Spectra, 1994, 3,
46 S. S. Tafreshi, A. Roldan, N. Y. Dzade and N. H. De Leeuw, 395–401.
Surf. Sci., 2014, 622, 1–8. 60 F. Y. Xie, W. G. Xie, L. Gong, W. H. Zhang, S. H. Chen,
47 S. Haider, A. Roldan and N. H. De Leeuw, J. Phys. Chem. C, Q. Z. Zhang and J. Chen, Surf. Interface Anal., 2010, 42,
2014, 118, 1958–1967. 1514–1518.
48 N. Dzade, A. Roldan and N. de Leeuw, Minerals, 2014, 4, 89– 61 R. Arrigo, M. Havecker, S. Wrabetz, R. Blume, M. Lerch,
115. J. McGregor, E. P. J. Parrott, J. A. Zeitler, L. F. Gladden,
49 F. Zhang, J. D. Gale, B. P. Uberuaga, C. R. Stanek and A. Knop-Gericke, R. Schlogl and D. S. Su, J. Am. Chem. Soc.,
N. A. Marks, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 132, 9616–9630.
2013, 88, 1–7. 62 O. Baghriche, S. Rtimi, C. Pulgarin, C. Roussel and J. Kiwi,
50 K. Mathew and R. G. Hennig, 2016, eprint arXiv: 1601.03346, Appl. Catal., B, 2013, 130–131, 65–72.
1–6. 63 B. P. Payne, M. C. Biesinger and N. S. McIntyre, J. Electron
51 K. Mathew, R. Sundararaman, K. Letchworth-Weaver, Spectrosc. Relat. Phenom., 2009, 175, 55–65.
T. A. Arias and R. G. Hennig, J. Chem. Phys., 2014, 140, 1–9. 64 F. Tuinstra and J. L. Koenig, J. Chem. Phys., 1970, 53, 1126–
52 J. D. Pack and H. J. Monkhorst, Phys. Rev. B: Condens. Matter 1130.
Mater. Phys., 1976, 13, 5188–5192. 65 S. Zhang, O. Metin, D. Su and S. Sun, Angew. Chem., Int. Ed.,
53 H. W. King and F. D. Manchester, J. Phys. F: Met. Phys., 1978, 2013, 52, 3681–3684.
8, 15–26. 66 A. Bulut, M. Yurderi, Y. Karatas, M. Zahmakiran, H. Kivrak,
54 J. Cho, S. Lee, S. P. Yoon, J. Han, S. W. Nam, K. Y. Lee and M. Gulcan and M. Kaya, Appl. Catal., B, 2015, 164, 324–333.
H. C. Ham, ACS Catal., 2017, 7, 2553–2562. 67 S. F. Ho, A. Mendoza-Garcia, S. Guo, K. He, D. Su, S. Liu,
O. Metin and S. Sun, Nano Lett., 2012, 12, 1102–1106.
Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2018