Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Molecular Catalysis xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

CO2 activation on small Cu-Ni and Cu-Pd bimetallic clusters


Andres Alvarez-Garciaa, Elizabeth Flórezb,*, Andrés Morenoa, Carlos Jimenez-Orozcob,*
a
Química de Recursos Energéticos y Medio Ambiente, Instituto de Química, Facultad de Ciencias Exactas y Naturales, Universidad de Antioquia, Calle 70 No. 52-21,
Medellín, Colombia
b
Grupo de Materiales con Impacto, Mat&mpac. Facultad de Ciencias Básicas, Universidad de Medellín, Carrera 87 No 30-65, Medellín, Colombia

ARTICLE INFO ABSTRACT

Keywords: The use of CO2 to produce methanol is a reaction of growing interest, where bimetallic Cu-M catalysts become
Hydrogenation relevant as an alternative to the known Cu/Zn/Al2O3 catalyst. However, there is a lack in the understanding of
CO2 bimetallic systems at atomic label and its capability towards CO2 activation, a key step in CO2 valorization. In
Bimetallic this work, Cu-Pd and Cu-Ni small clusters are studied using DFT. Among the evaluated bimetallic systems, the
Cluster
binding of CO2 on Cu3Pd has the highest thermodynamics stability (28.82 kcal/mol) and the lowest energy
Catalysis
barrier (40.91 kcal/mol). The activation energy for the dissociation of CO2 (CO2* → CO* + O*) follows the
trend: Cu4 < Cu3Pd < Pd4 < CuPd3 < Cu2Pd2. Therefore, the ideal composition in terms of adsorption energy
and activation barrier is the Cu3Pd bimetallic system. The interaction O-M is weak while C-M is responsible of
the binding, a charge migration from cluster to CO2 was seen, and the band around 1150 cm−1 in the IR was only
found in activated CO2. The results of this work indicate that the Cu3Pd cluster has catalytic potential towards
CO2 activation and dissociation, opening the doors to explore further the Cu3Pd system both theoretically and
experimentally.

1. Introduction producing methanol selectively and limiting the formation of CO. The
composition with higher rate of methanol formation has been assigned
The carbon dioxide (CO2) is the major greenhouse gas responsible of to Pd(0.25)-Cu/SiO2, which is also two times higher as compared to the
global warming and acidification of oceans [1,2]. Therefore, the cap- monometallic Pd and Cu catalysts [18]. For the Cu-Ni bimetallic cluster,
ture, storage and conversion of CO2 to value-added chemical products it has been found that there is a strong interaction between the copper
has received much attention in recent years, as an effective way to and nickel components, where the nickel fulfills the role of helping the
reduce the emission and accumulation of this greenhouse gas in the adsorption of H2 in the cluster [19].
atmosphere [3]. In particular, CO2 catalytic hydrogenation to methanol The atomicity of the Cu-Pd and Cu-Ni systems can be taken as small
is of growing interest [4–10], since methanol can be used as a chemical cluster of atoms [20,21] because they exhibit some advantages to
feedstock, solvent and alternative fuel. bigger systems such as [20,22]: (a) significant changes in reactivity and
The catalytic hydrogenation of CO2 towards methanol is considered selectivity due to small alterations in size and composition; (b) the
an economically feasible process [11]. Nowadays, methanol is in- possibility to reduce the amount of metals used, which can generate an
dustrially produced from a CO−CO2-H2 mixture at 60 bars and 250 °C economic impact, for instance in the case of noble metal particles; (c)
using the Cu/ZnO/Al2O3 catalyst [12,13]. A strategy to overcome the they have a greater surface area, which implies more active sites; (d)
issues related to the industrial catalysts is to study the composition there are methods for synthesis and deposition of clusters that are
effect, particularly the role of a second metal on the performance of capable of creating samples with a precise atomic composition; and (e)
copper-based catalysts [14–16]. Overall, the main purpose of in- high level studies can be carried out with DFT methods to model and
corporating a second metal in catalysis is to avoid sintering of the understand in detail the chemistry at atomic level.
copper particles and increasing the activity of the copper active sites for The cluster with a size of four atoms of copper (Cu4) supported on
CO2 hydrogenation [17]. The bimetallic Cu-Pd and Cu-Ni systems have Al2O3 has been established as the catalyst with the highest catalytic
been reported as potential catalysts for the synthesis of methanol. The performance towards methanol formation [22], which could also op-
composition of Cu-Pd bimetallic system has been established according erate at low pressures (near atmospheric) [21]. Moreover, the use of
to the atomic ratio Pd/(Pd + Cu) within the range 0.25 – 0.34, Cu4 cluster has been reported in several experimental and theoretical


Corresponding authors.
E-mail addresses: elflorez@udem.edu.co (E. Flórez), cjimenez@udem.edu.co (C. Jimenez-Orozco).

https://doi.org/10.1016/j.mcat.2019.110733
Received 3 September 2019; Received in revised form 5 November 2019; Accepted 29 November 2019
2468-8231/ © 2019 Elsevier B.V. All rights reserved.

Please cite this article as: Andres Alvarez-Garcia, et al., Molecular Catalysis, https://doi.org/10.1016/j.mcat.2019.110733
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

studies, not only in methanol formation but in other several systems Table 1
[21–28]. Therefore, a systematic study of the cluster of four metal Structural and electronic parameters of Cu and Pd dimers.
atoms of copper (Cu4) and the bimetallics Cu4-xPdx and Cu4-xNix are Cluster Property PBE0/SDDa Experimental
considered in this work, together with Pd4 and Ni4 monometallic sys-
tems for comparison purposes. The cluster systems are studied in this Cu2 ro (Å) 2.24 2.22b
ω (cm−1) 263 266.00b
work since a systematic approach is necessary before considering sev-
BE (eV) 1.84 2.02c
eral supports, which is beyond of the scope of the current report and EA (eV) 0.78 0.84c
could be a matter of future work. IP (eV) 7.91 7.90d
The understanding of the system at the atomic level is important as Pd2 ω (cm−1) 211 210.00e
it is necessary to know the chemistry in detail, which is the basis to BE (eV) 0.92 1.03e

further improve the performance of catalysts. This understanding is not a


This work.
easy to achieve using experimental techniques, then computational b
From ref [37].
chemistry can be used as an important tool. Using computational c
From ref [38].
methodologies, the geometric issues related to activation of the CO2 d
From [39].
molecule (i.e. CeO bond length elongations, OCO angles and changes in e
From ref [40].
hybridization from sp2 to sp3-like) upon interaction with clusters can be
studied, since CO2 is a stable molecule [29]. This is an important aspect,
since its activation is the first step towards the production of methanol
and, overall, the valorization of CO2 into different products. For in-
stance, Gálvez-González et al. [30] performed a systematic computa-
tional study of CO2 adsorption on gas-phase Cu4-xPtx (x = 0–4) clusters,
where the authors selected the size of cluster M4 according also to ex-
perimental data [21,22]. They found that gas-phase linear CO2 mole-
cule was deformed upon adsorption, with its bend angle varying from
about 134° to 145°. The geometries of the clusters remain unchanged
after CO2 adsorption, except for of Cu3Pt and Pt4 systems. The authors
concluded that Cu3Pt may be a good candidate as a heterogeneous
ultra-small nanocatalyst, because it activates the CO2 molecule, its
adsorption energy is in an optimum range for catalysis and also has a
higher adsorption energy with respect to the monometallic clusters (Cu4 Fig. 1. Initial geometries for homonuclear clusters (Symmetries are included).
and Pt4).
Although there is experimental evidence indicating that Cu4 cluster
is a promising catalyst, there is still a lack in the understanding re- clusters (Cu4, Pd4 and Ni4) considering the first three multiplicities in
garding the inclusion of a second metal. Therefore, the purpose of the every case. For the search of the ground state structures of bimetallic
present work is to study systematically, by means of computational clusters, the potential energy surface (PES) was explored by means of
chemistry, the effect of the inclusion of a second metal to form Cu4-xPdx substitutions of these motifs, characterizing the location of minima in
and Cu4-xNix (x = 0–4) clusters towards the activation (geometry the PES via frequency analysis.
changes and adsorption energies) of CO2, with dissociation of the CO2 Several adsorption modes were studied with an initial distance of
molecule in some representative systems. It is worth to mention that a 1.5 Å between the cluster and the CO2 molecule. The HOMO orbitals of
detailed analysis of the CO2-cluster interaction should be carried out the bare monometallic and bimetallic clusters were used as a guide for
before considering interactions with hydrogen and the concomitant locating CO2 molecule, considering all the binding possibilities. The
reactions towards methanol. natural charge transfer from cluster to CO2 as well as the bond orders
This work is divided as follows: 1) Characterization of bare metallic were carried out using NBO 6 package [41], through the Wiberg bond
and bimetallic clusters, including establishment of the ground state indexes. In the Natural Bond Orbital (NBO) analysis the bond order is
multiplicity, NBO charge analysis, reactivity descriptors, and geometric defined by Wiberg indexes (WBI) [42], which are a measure of electron
analysis. 2) The binding of CO2 on monometallic and bimetallic clusters population overlap between two atoms. Wiberg bond indexes were
is studied, considering both geometric and electronic factors to analyze proposed for calculations in an orthonormal basis set; for A and B atoms
the activation; then, for the most thermodynamically stable systems, the WBI is:
bond orders and atomic charge analyses are carried out; at this point,
WAB = (Pµ ) 2
the activation is also characterized by means of infrared vibrational
µ A B (1)
spectra. 3) The CO2* dissociation into CO*+O* is studied in the most
stable structure of every composition of the Cu4-xPdx (x = 0–4) system In Eq. (1), P is the total density matrix and the summation is over
and compared to the monometallic Cu4 and Pd4 clusters. The data ob- atomic orbitals μ on atom A and atomic orbitals ν on atom B. The
tained in this work is contrasted with those of the Cu4-xPtx (x = 0–4) density matrix depends over orbital coefficients and has different ex-
system, to get some general trends in the bimetallic systems of Cu pressions for closed-shell and open-shell systems [43].
combined with atoms of the group 10 (Ni, Pd, Pt) in the Periodic Table. The binding energy per atom (BE) for bimetallic clusters was cal-
culated as:
2. Computational details BE = [E (Cu x My ) xE (Cu) yE (M )]/4 (2)

The optimization of each structure was carried out using the PBE0 In Eq. (2), x , y = 0 4 , M = Ni , Pd , E (Cu ) is the total energy of the
[31] functional and the SDD basis set [32,33]. This methodology has Cu atoms in ground state, E (M ) is the total energy of the Ni and Pd
been used recently to study the reactivity of metal clusters [34,35] atoms in ground state, and E (Cu x My ) is the total energy of the tetra-
which also provides a rather correct description of Cu2 and Pd2 dimers nuclear system.
(see Table 1). All calculations were performed with the Gaussian 09 The adsorption energy (Eads) was calculated as:
package [36]. Eads = E (CO2) + E (Cu x My ) E (Cu x My CO2) (3)
Several motifs were used (see Fig. 1) to optimize the monometallic

2
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

In Eq. (3), E (Cu x My CO2) is the total energy of the adsorption substitution are the most stable structures for the Cu3Ni cluster, as they
complex, E (CO2 ) is the total energy of an isolated CO2 molecule in the are energy degenerated. There is also a planar triangular structure close
singlet state, and E (Cu x My ) is the total energy of the stable tetranuclear to the ground state (5.65 kcal/mol, see Table S7 in the Supplementary
cluster. For the most stable structure of the Cu4, Pd4, Ni4, Cu4-xPdx (x = Data), where its low stability could be assigned to the fact that the
0–4), and Cu4-xNix (x = 0–4) systems, the infrared spectra were studied atoms in a triangular planar structure have less coordination relative to
to identify signals to characterize the CO2 binding and activation. the rhombic geometry. For the Cu3Pd cluster the most stable structure is
The activation energy for the dissociation of adsorbed CO2* into the rhombic-like with central substitution; however, its change to a
CO*+O* was studied in the most stable structure of every composition lateral substitution could be possible due to a small energy difference
of the Cu4-xPdx (x = 0–4) system, which are compared to Cu4 and Pd4 between them, i.e. 7.53 kcal/mol (see Table S4 in the Supplementary
monometallic systems. The transition state search was carried out by Data). Overall, the stability order in terms of multiplicity for both Cu3Ni
using the Synchronous Transit-Guided Quasi-Newton method [44]. and Cu3Pd clusters is doublet > quadruplet > sextet.
For the bimetallic Cu2Pd2 (Fig. 2c) and CuPd3 (Fig. 2d) systems, the
3. Results and discussion ground state multiplicity belongs to singlet and doublet, respectively,
where the structures have a three-dimensional form. In these structures,
3.1. Geometry and electronic characterization of bare monometallic and there is an absence of isomers to values lower than 10 kcal/mol of
bimetallic clusters energy (see Tables S5 and S6 in the Supplementary Data), which in-
dicates that interconversion to another configuration could not be
The spatial arrangement of nickel, palladium and copper with their possible, as it happens for the Cu3Pd and Cu3Ni discussed above.
self or between them leads to several geometries. Therefore, it is ne- In the case of Cu2Ni2 (Fig. 2c) and CuNi3 (Fig. 2d) clusters, the most
cessary to analyze both the geometry and the energetics involved in the stable structures have a two-dimensional geometry, where the rhombic-
building of monometallic (Cu, Ni, or Pd) and bimetallic (Cu-Ni, Cu-Pd) like structures with lateral and central substitution have the highest
tetrameric clusters. Hence, it is possible to establish the most stable stability (see Tables S8 and S9 in the Supplementary Data). Overall, the
motifs to further analyze the binding of CO2. Cu2Ni2, CuNi3 and Ni4 clusters have a greater stability in excited states,
where the ground state multiplicity for Cu2Ni2 and Ni4 are triplet and
3.1.1. Establishment of the ground state multiplicity quintet, respectively, while CuNi3 has two multiplicities in the ground
Several spin multiplicities (M) were considered for every motif. For state (doublet and quadruplet). These results are expected, since nickel
the monometallic systems (see Fig. 2a), the ground states were found to is an open shell transition metal.
the Cu4, Pd4, and Ni4 clusters in multiplicity 1, 3, and 5, respectively.
The Cu4 and Ni4 clusters have two-dimensional (2D) structures, i.e. a 3.1.2. Properties of the most stable structures
planar geometry, while Pd4 has a 3D structure. For more details please 3.1.2.1. Geometric and electronic characterization. Diverse electronic
refer to the Tables S1–S3 in the Supplementary Data. These results are and structural properties were analyzed for the most stable structures
in agreement with previous reports [45–49]. (see Table 2). The average distance between atoms (denoted
In bimetallic clusters, we analyze the systems as the substitution as < number >) within both monometallic and bimetallic systems
extent increases. The ground state for the Cu3Pd and Cu3Ni systems are was considered accounting for the first neighbors (see Table 2). The
both M = 2. In these systems, the planar structures are preserved Pd-Pd distance in the Pd4 cluster (2.61 Å) is in agreement with previous
(Fig. 2b), where the rhombic-like structure with central and lateral computational studies [50]; a similar behavior was observed for the Cu4

Fig. 2. Spin multiplicity (M) for the optimization of tetrameric clusters, energies relative to the most stable structure in every case. a) Monometallic systems (Cu4, Pd4
and Ni4), b) Cu3M, c) Cu2M2, d) CuM3; M = Ni, Pd. Brown, blue, and green spheres represent Cu, Pd, and Ni atoms, respectively. The dotted vertical lines are used to
separate regions of different clusters. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

3
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

Table 2
Structural and energetic properties of the most stable tetramer Cu-Ni and Cu-Pd
clusters.

Cluster Structure Gap (kcal/mol) 〈d Cu-Cu〉 Å 〈d Cu-M〉 Å 〈d M-M〉 Å

1
Cu4 D2h 53.18 2.38 – –
2
Cu3Ni C2v 53.16 2.40 2.39 –
3
Cu2Ni2 C2v 49.57 – 2.43 2.29
4
CuNi3 C2v 53.83 – 2.42 2.39
5
Ni4 C2h 54.12 – – 2.40
2
Cu3Pd C2v 43.30 2.38 2.48 –
1
Cu2Pd2 C2v 65.64 2.44 2.48 2.72
2
CuPd3 C3v 61.34 – 2.51 2.60
3
Pd4 C3v 52.65 – – 2.61

and Ni4 systems, with distances Cu-Cu = 2.38 Å and Ni-Ni = 2.40 Å,
respectively, which are almost equal due to a similar atomic radii
(1.276 and 1.246 Å respectively) [51,52]. Therefore, similar distances
are expected in Cu-Ni bimetallic systems as compared to the
monometallic Cu4 and Ni4 clusters, which is in fact observed in
Table 2. However, in the copper-palladium bimetallic system the Cu-
Pd distance is higher as compared to the Cu-Ni system, due to the
difference in their atomic radii (Ni = 1.246 Å and Pd = 1.373 Å) [52],
which has a direct impact on Cu-Pd distances in all of the studied
clusters. The largest Cu-Cu distance was observed in the Cu2Pd2 cluster,
which shows that structures in three dimensions usually involve larger Fig. 3. a) Binding energy per atom (BE) for the most stable structure in every
distances. case, b) energy of the highest occupied molecular orbital (HOMO) (isovalue =
In order to get a deeper understanding about the electronic prop- 0.03) for the structures shown in a). The dotted vertical lines are used to se-
parate regions of different kind of clusters.
erties of the evaluated clusters, a NBO charge analysis was performed
for the bare clusters (see Table S13 in the Supplementary Data). The
monometallic clusters (Cu4, Ni4 and Pd4) do not have a significant impact on reactivity by substitution with palladium atoms.
charge polarization, which implies that the electrons are uniformly
distributed throughout the monometallic systems. In the bimetallic Cu- 3.1.2.3. Thermodynamics stability. Up to this point, the geometry and
Pd systems, the inclusion of a second metal leads to a charge polar- electronic issues related to the most stable structures of monometallic
ization, which obeys the difference in electronegativity [53] of the and bimetallic clusters have been analyzed. Before addressing the
studied atoms (Pauling electronegativity: Cu = 1.90 and Pd = 2.20). interaction of these clusters with CO2, their stability should be
For instance, in the Cu3Pd cluster the charge is polarized on the pal- analyzed, as the interaction with an adsorbate could eventually
ladium atom (-0.18e), while in CuPd3 systems the copper atom has a modify the geometry of the cluster. Therefore, the binding energy per
positive charge (0.21e), indicating a migration of electron density from atom (BE, see Fig. 3a) was calculated to analyze the global stability of
copper to palladium atoms. The inclusion of a second Pd atom, i.e. the clusters. The cluster size and its composition have an impact on the
Cu2Pd2, leads to the highest charge polarization, where every copper binding energy. Particularly related to the size effect, the binding
and palladium atoms have a charge of 0.20e and -0.20e, respectively. energy of Pd2 (10.61 kcal) is lower to that of Pd4 (29.74 kcal); these
On the contrary, in bimetallic Cu-Ni clusters charge polarization was results are in agreement to the trend reported by Luo C. et al. [57],
not observed, i.e. the electronic distribution is uniform, due to a neg- therefore, larger the cluster size, higher the binding energy. The BE for
ligible difference in electronegativity of Cu and Ni atoms. The only Cu4 and Ni4 clusters (29.10 and 30.84 kcal/mol respectively), are close
exception was observed in the Cu3Ni cluster, where the partial charge to the values reported by Jung et al. [58]. The BE varies as the
of the lateral copper and nickel atoms is -0.10e and 0.13e respectively. composition in bimetallic systems is modified; overall, the Cu-Pd
clusters are more stable than Cu-Ni clusters. In particular, the most
3.1.2.2. Reactivity descriptor. Overall, the inclusion of a second metal to stable structures are the Cu2Pd2 and CuPd3 clusters, while the Cu3Ni
form a bimetallic system influences a redistribution of the electron and Cu3Pd bimetallic clusters are less stable.
density and, as a consequence, the reactivity of the clusters is different.
A useful descriptor to analyze the global reactivity is the energy 3.1.2.4. Guiding CO2 binding on bare clusters. After the characterization
difference between HOMO and LUMO orbitals, i.e., the gap energy of the clusters discussed above, the next step is to study their interaction
(Egap) [54–56]. It relates to the energy cost for an electron to jump from with CO2. This molecule could interact with different sites of every
HOMO to LUMO orbital, thereby a larger gap usually correlates with a cluster, therefore, several adsorption modes should be considered. The
lower reactivity (see Table 2). The gap of monometallic systems is HOMO orbitals of bare clusters could serve as a guide to locate CO2,
similar, which indicates that the reactivity of Cu4, Pd4, and Ni4 is then, the HOMO schemes for every monometallic and bimetallic cluster
similar. The reactivity of a monometallic cluster is expected to be is shown in Fig. 3b. For the three-dimensional systems (Cu2Pd2, CuPd3
improved with the inclusion of a second atom. This was observed in the and Pd4) the HOMO orbitals are located in the atoms, particularly with
Cu4 system, where the gap decreases from 53.2 kcal/mol to 43.3 kcal/ a larger lobe for the palladium atoms; thus, there is a higher electronic
mol in Cu3Pd. However, this trend is not consistent in Cu2Pd2 and probability over palladium atoms, and the interaction of CO2 over this
CuPd3 systems (indicating lower reactivity), i.e. they are more stable as atom could be preferential to other sites. A different behavior was
compared to the monometallic system. For the Cu-Ni bimetallic system, observed in planar structures (Cu4, Cu3Pd, Cu3Ni, Cu2Ni2, CuNi3 and
the gap is slightly modified. The reactivity of the Cu-Pd clusters is in the Ni4), where the HOMO orbital is distributed in the lateral zones,
range 43.3–65.6 kcal/mol, while in the Cu-Ni clusters the range is including the bridge sites; the only exception is for Ni4 cluster, which
49.6–53.8 kcal/mol. The obtained results suggest that there is a greater has a delocalized distribution. These results indicate that the reactivity

4
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

to their respective monometallic systems (with the exception of Cu3Pt


which has a slightly lower adsorption energy than Pt4), which shows a
synergistic effect in the inclusion of a second metal atom to form a
bimetallic system [29], which is also reflected in the improved binding
energy of CO2.

Fig. 4. Adsorption modes of CO2 on bimetallic clusters. Black, gray, and red 3.2.1. Activation of the CO2 molecule
balls, represent metal, carbon, and oxygen atoms, respectively. (For inter- The activation of the CO2 molecule involves changes in geometry
pretation of the references to colour in this figure legend, the reader is referred and its electronic properties. Several descriptors have been reported to
to the web version of this article.) characterize the activation of CO2 [63]: a) bending of CO2, b) elonga-
tion of a CeO bond, c) redistribution of charges, and d) electron
of monometallic and bimetallic clusters is not always located in atoms, transfer towards CO2. Descriptors a) and b) relate to geometry changes,
then bridge sites should be considered. In the next section, the while c) and d) with electronic properties.
interaction of CO2 on the clusters discussed above is studied.
3.2.1.1. Bending of CO2 and elongation of a CeO bond. The structural
parameters that are mainly involved in the activation and the
3.2. Binding of CO2 on monometallic and bimetallic clusters corresponding adsorption energy for each system are reported in
Table 3, which also includes the values reported by Galvéz-González
Several adsorption modes were evaluated, using Fig. 3b as a guide et al. [30] for the Cu-Pt system with a level of theory (PBE/SDD) similar
to locate the CO2 molecule (see Fig. S2 in the Supplementary Data for to the used in this work, with the purpose to perform a comparison in
an example). The nomenclature used in this work is shown in Fig. 4, the CO2 binding behavior throughout group 10 of the periodic table.
which is consistent with previous reports [59,60]. In η1-C mode only The adsorption energy in monometallic Cu4 (15.69 kcal/mol) and
the C atom interacts directly with a metal atom, while μ2, μ3 and μ4 Pd4 (15.26) systems is improved to energies in the range from 19.36 to
indicate that CO2 is coordinated with 2, 3 and 4 atomic centers of the 28.82 kcal/mol for the bimetallic Cu-Pd systems (see Table 3), i.e. there
cluster respectively, but the C atom does not interact directly with is a synergistic effect between Cu and Pd to bind and activate CO2. A
metal atoms, i.e. it is located in a bridge configuration (μ denotes a similar behavior was observed for the Cu-Ni bimetallic systems as
bridge). compared to monometallic Cu4 and Ni4 clusters [64]. However, the
It is worth to mention that the interaction of CO2 molecule with inclusion of Ni atoms in Cu4 cluster does not imply a remarkable in-
metallic clusters involves the transfer of electron density from HOMO crease in the adsorption energy as observed for the Cu-Pd system.
orbital of cluster to LUMO orbital of CO2, therefore the charge transfer In heterogeneous catalysis, the most common intermediates for the
could be favored at a high spatial orbital overlap. Therefore, it is useful transformation of CO2 to value-added products are the CO2 radical
to highlight some aspects related to the molecular orbital of the ad- anion and the formate anion [65,66], which have a similar geometric
sorbate and those of the metallic clusters. The CO2 molecule has four properties, due to the weak hydrogen bonding. When carrying out the
important interacting orbitals [61]: O-end lone pair, π, nonbonding nπ activation of carbon dioxide, it is expected that adsorbed CO2 has
and the π*orbitals, where the first three are doubly occupied (nπ is the characteristics similar to these intermediates. The adsorbed CO2 pre-
HOMO) and the π*is the LUMO. In the metallic clusters, the HOMO sents elongation in the CeO bond (1.27–1.29 Å) relative to the calcu-
orbital is located in the nd orbitals for nickel and palladium, while in lated value for the isolated CO2 (1.19 Å). The systems with greater
copper it is in the partially occupied 4s orbital (SOMO). Therefore, the elongation of the CeO bond (1.29 Å) are Pt4, Cu3Pt, Cu2Pt2 and Cu2Ni2,
adsorption modes were generated by placing CO2 near the high electron besides all the CeO distances reported for the most stable CO2-cluster
density region in the HOMO orbital of the clusters. systems exceed the experimental CeO distance for the CO2 radical
Despite of the several evaluated modes, only the most stable ones anion (1.24 Å) and the formate anion (1.25 Å) [63].
are discussed; for details about all the evaluated modes, please refer to After adsorption, the angle OeCOe decreases to values between
the Tables S11 and S12 in the Supplementary Data. The most stable 126° to 140°, where the bending is evident. The system with the
structures for the interaction of CO2 with the Cu-Ni and Cu-Pd tetra- greatest CO2 bending is Cu2Ni2, which has a large adsorption energy.
meric clusters are shown in Fig. 5. These structures show structural When comparing these values with the angle for the radical anion of
changes in carbon dioxide due to the interaction with the clusters [62]. CO2 (135.1°) and the formate anion (130.8°), it can be noted that most
It can also be shown that there is a preferential interaction between the systems are between these values. Moreover, the distance of the CO2 to
carbon atom and a cluster metal, particularly with nickel and palladium the cluster is also reported, which is lower than 2.0 Å, reinforcing the
in the case of bimetallic clusters. The details about binding of CO2 in idea of a strong interaction.
every case are discussed below in Section 3.2.1, together with the in- The sites involved in the binding of CO2 are shown in Table 3. For
clusion of data for binding on Cu-Pt systems from a previous report [30] the Cu4 cluster, the carbon dioxide activation was only observed in the
for comparison purposes. adsorption mode η1-C (bond order of 0.55, see Table 4), while in Pd4
The adsorption energies (Eads) are higher than 15 kcal/mol, which and Ni4 the adsorption mode was μ2-O. The adsorption modes reported
indicates a chemical interaction between the clusters and the carbon in Fig. 5 and Table 3 show a preference for binding to metal atoms of
dioxide molecule. In clusters with a single substitution (Cu3M), the clusters, i.e. in the atomic centers. The results also show that there is
order of magnitude of adsorption energy is Cu3Pd > Cu3Ni > Cu3Pt also a bond between the carbon atom of CO2 and metal atoms (η1-C
(see Table 3). It is worth to mention that the most thermodynamically mode, see Table 4). Particularly, there is a higher BO between the C-Pd
stable structure is Cu3Pd, where its composition relates with the ex- relative to the C-Ni and the C-Cu, therefore, a strong interaction be-
perimental atomic ratio Pd/(Pd + Cu) = 0.25 reported by Jiang et al. tween carbon and palladium is observed. In general terms, the BO range
[18], which has the highest catalytic activity among different atomic from 0.37-0.63, which shows an important binding between carbon
ratios for the hydrogenation of CO2 into methanol. For the other sub- (electrophile) and palladium (nucleophile). BOs are not reported for
stitutions (Cu2M2 and CuM3) the adsorption energy shows the opposite OeM, since they present low values (between 0.05 - 0.30), which in-
order (Cu-Pt > Cu-Ni > Cu-Pd). This behavior suggests that in copper- dicates that the interaction between oxygen and metal atoms is weak,
rich clusters it is preferable to substitute with palladium, while in then, the carbon atom is highly responsible of binding to the cluster.
clusters with a low amount of copper it is preferable to use platinum. The bond order (BO) can account for the strength of the binding,
Overall, the bimetallic clusters have higher adsorption energies relative since the larger the bond order, the greater the energy needed to break

5
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

Fig. 5. Most stable structures and adsorption energies for the


interaction of CO2 molecule on monometallic and bimetallic
clusters. The labels of colors are the same as in Fig. 2, ad-
ditionally, gray and red balls represent carbon and oxygen
atoms, respectively. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web
version of this article.)

Table 3 with the elongation of the CeO bond in Table 3, since the higher the
Structural parameters related to the activation of carbon dioxide on the elongation, a lower bond order is expected. In addition, the BO in CO2,
monometallic and bimetallic clusters. i.e. C-O1 vs C-O2 are different once molecule is adsorbed (see Table 4),
Cluster Adsorption Eads Distance C-O Angle O- Distance which indicates that the carbon dioxide molecule is not activated
mode (kcal/ (Å)a C-O(°) Cluster-CO2 symmetrically. The lower BO between CeO is presented with the in-
mol) (Å) teraction with the CuNi3 cluster, which has a correspondence with the
adsorption energy, since this system has the highest adsorption energy
Cu4 η1 - C 15.69 1.27 134.85 1.93
Ni4 μ2 - O 25.04 1.27 136.83 1.96 among the studied systems. A similar behaviour was observed for the
Pd4 μ2 - O 15.26 1.28 138.50 1.95 Cu3Pd and Cu2Ni2 systems, which have a C-O1 bond order of 1.38 and
Pt4b c
26.70 1.29 134.30 c
1.36 respectively, which also corresponds to systems with higher ad-
Cu3Ni η1 - C 25.97 1.28 136.84 1.95 sorption energy. According to the results in Tables 3 and 4, there is a
Cu3Pd η1 - C 28.82 1.28 137.52 1.94
direct correspondence between the bond order and adsorption energy,
Cu3Ptb c
23.36 1.29 132.50 c

Cu2Ni2 μ4 27.08 1.29 126.01 1.94 where lower CeO bond order, higher adsorption energy.
Cu2Pd2 μ2 - O 24.47 1.27 140.18 1.98
Cu2Pt2b c
37.92 1.29 136.60 c
3.2.1.2. Electron transfer towards CO2. The structural properties
CuNi3 μ2 - O 27.32 1.28 136.66 1.96
adopted by the activated CO2 molecule are due to the electronic
CuPd3 μ2 - O 19.36 1.27 140.45 1.98
CuPt3b c
33.30 1.28 137.20 c properties CO2 after adsorption, particularly characterized by the
quantification of the charge migration between cluster and CO2.
a
C-O distance of calculated isolated CO2: 1.19 Å. Therefore, it is necessary to perform an analysis of atomic charges to
b
Value taken from ref [30]. better understand these interactions. Then, cluster-CO2 charge transfer,
c
Data no available from a previous report [30]. and the redistribution of charges after CO2 interaction throughout every
system should be considered.
Table 4 The atomic charges of absorbed CO2 on the Cu-Ni and Cu-Pd clus-
Bond Order (BO) of the absorbed CO2 and of the C-M interaction. ters are shown in Table 5. After adsorption, the CO2 is partially charged,
Cluster BO C-O1 BO C-O2 BO C-M which has values in the range -0.42e to -0.92e, which indicates that

CO2 1.88 1.88 –


Table 5
Cu4 1.39 1.66 0.55
Natural charges of CO2 absorbed on the monometallic and bimetallic clusters.
Ni4 1.45 1.48 0.37
Pd4 1.40 1.66 0.63 Cluster Atomic charges Net charge in CO2 (e)
Cu3Ni 1.40 1.67 0.45
Cu3Pd 1.38 1.68 0.59 CO2 C (1.02), O1 (−0.51), O2 (−0.51) 0.00
Cu2Ni2 1.36 1.43 0.45 Cu4 C (0.67), O1 (−0.73), O2 (−0.58) −0.64
Cu2Pd2 1.45 1.66 0.59 Ni4 C (0.61), O1 (−0.61), O2 (−0.65) −0.65
CuNi3 1.03 1.02 0.43 Pd4 C (0.80), O1 (−0.66), O2 (−0.56) −0.42
CuPd3 1.46 1.66 0.58 Cu3Ni C (0.68), O1 (−0.67), O2 (−0.56) −0.55
Cu3Pd C (0.77), O1 (−0.70), O2 (−0.56) −0.49
Cu2Ni2 C (0.51), O1 (−0.74), O2 (−0.70) −0.93
Cu2Pd2 C (0.80), O1 (−0.69), O2 (−0.56) −0.45
the bond. The bond orders reported in Table 4 indicate that once CO2 is
CuNi3 C (0.61), O1 (−0.65), O2 (−0.61) −0.65
adsorbed, there is a significant decrease in the bond order, which CuPd3 C (0.80), O1 (−0.68), O2 (−0.56) −0.44
ranges from 0.21 to 0.85 relative to isolated CO2. This is in accordance

6
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

there is a charge migration from cluster to CO2. The charge transfer


extent towards CO2 in Cu-Ni clusters is higher than in Cu-Pd systems.
The lowest charge transfer towards CO2 occurs when it interacts with
the Pd4 cluster, while the highest charge transfer is observed in the
Cu2Ni2 cluster. This is due to the higher reactivity of the planar clusters
and the excitation of electrons presented by these systems, which also
relates to the energy gap in Table 2 for the nickel clusters. For the case
of Cu-Pd systems, the difference in electronegativity between palladium
with copper and nickel, does not allow Pd to easily transfer electrons. It
is also important to note that there is a relationship between the charge
transfer extent and the adsorption energy (see Tables 4 and 5).
The activation of CO2 involves charge transfer from the metallic
cluster as mentioned above. Therefore, charge redistribution within the
cluster once CO2 is adsorbed should be analyzed for every metal in all
the studied clusters (see table S13). The atoms that are in a direct in-
teraction with carbon dioxide have a decrease in the negative character
of the charge. It is observed that monometallic clusters are polarized Fig. 6. Infrared CO2 spectra for the most stable adsorbed structures (isolated
after adsorption, presenting positive partial charges in the interacting CO2 is included). The symbols represent: CO2 bending (*), symmetric CeO
atoms; even in palladium which had a negative partial charge before stretching (*), and asymmetric CeO stretching (*).
adsorption, it becomes positive after bonding.
Moreover, it can be observed that after the adsorption of CO2 on the CO2 on the monometallic clusters (Cu4, Ni4, Pd4) and two re-
bimetallic Cu-Ni clusters, there is a charge polarization within metal presentative bimetallic clusters (Cu3Pd and CuNi3) are shown in Fig. 6.
atoms, as compared to the systems before adsorption without a polar- The main bands of the spectra in Fig. 6 are: bending of CO2, symmetric
ization. The Cu3Ni cluster has a positive partial charge on the copper and asymmetric stretching of CeO. Particularly, the isolated CO2 mo-
atoms, which is due to a charge compensation generated by the cluster lecule has two bands centered at 630 cm−1 and 2300 cm−1, which
for the loss of electronic density in the nickel atom. The Cu2Ni2 cluster belongs to the OCO bending and the asymmetric stretching of the C]O
has a high positive partial charge on the nickel atoms (0.56e and 0.43e), bond, respectively.
which is due to the charge transfer from them to CO2, while the copper Once the CO2 is adsorbed on both monometallic and bimetallic
atoms do not remarkable have changes in their charges, which implies clusters, there is a shifting of the signal associated to an asymmetric
that they are not involved in charge compensation within the cluster. CeO stretching. Particularly, for the Cu4, Pd4 and Cu3Pd systems, the
The Cu2Ni2 cluster displays structural conversion, which has been re- signal is around 1730 cm−1 with a shifting of 570 cm−1 relative to the
ported in similar systems [30]. It is expected that the observed high isolated CO2; while for the Ni4 and CuNi3 systems the shifting is larger,
charge transfer in this cluster implies that the cluster is cationic; thus, i.e. 640 cm−1. This behavior is related to the geometric change that the
the structural change could be assigned to a higher stability in this CO2 molecule undergoes when activated on the clusters.
cationic configuration. The CuNi3 and Ni4 clusters also present positive The bending of the CO2 (around 700 cm−1) also has shifting relative
partial charges on the nickel atoms directly bound to the CO2, while the to the isolated molecule, however, there is not a clear shifting trend as it
cluster atoms that do not participate in the interaction do not have happens for the asymmetric CeO stretching. In addition, the bending of
considerable charge changes. adsorbed CO2 has higher intensity relative to the isolated molecule,
The bimetallic Cu-Pd clusters also show an important polarization which indicates that the bending vibrational mode is improved due to
after the adsorption of CO2. In the Cu3Pd cluster, the palladium atom is the interaction of CO2 with the clusters. For the cluster-CO2 systems
directly participating in the interaction with the adsorbate, and its there is an additional band (approximately at 1150 cm−1), which
partial charge goes from negative (-0.18 e) to neutral (0.05 e). The corresponds to a symmetrical CeO stretching. This additional band is
copper atom that is coordinated with the CO2 becomes more positive (it found when CO2 molecule is activated on the clusters, which has also
goes from 0.10 to 0.23 e), while another copper atom being positive been reported experimentally for Pt4 cluster [67]; therefore, this signal
although it is not directly interacting with CO2, which evidences a is a fingerprint, which could also support that the CO2 molecule is ac-
charge retribution within the cluster. In the Cu2Pd2 cluster, all the tivated on all the studied clusters.
atoms are being positive, clearly losing charge; particularly, the palla-
dium atom which is directly bound to the carbon atom of CO2 goes from
a partial negative charge (-0.20 e) to a neutral one (-0.01 e). Finally, the 3.3. Carbon dioxide dissociation in monometallic and bimetallic systems
CuPd3 cluster only shows a positive charge in the copper atom which is
adjacent to oxygen and the Pd atom (responsible of bonding), while the To model CO2 dissociation on clusters, the following reaction was
other atoms remain neutral. In the CuPd3 system there is no significant used: CO2* → CO* + O*. The starting point is the most stable structure
charge redistribution, since its transfer towards CO2 is low, together of the absorbed CO2, named as R. To find the dissociation product (P),
with a low adsorption energy relative to the other clusters studied. the CeO bond of CO2 with the lowest bond order (which is C-O1) was
Overall, the clusters with higher charge transfer have high positive taken, then, the bond length was systematically increased from 1.27 Å
partial charges in their atoms after adsorption, where all atoms with a to 3.0 Å. The most stable composition in the copper-palladium bime-
cluster undergo significant changes in their partial charge, not only tallic system was considered, while Cu4 and Pd4 monometallic clusters
those directly involved in the adsorption, i.e. there is a charge redis- were used for comparison purposes. The structures for initial state (R),
tribution within the metallic cluster after CO2 adsorption. transition states (TS) and products (P) involved in CO2 dissociation on
all the evaluated clusters are shown in Fig. S3.
3.2.2. Infrared vibrational characterization of adsorbed CO2 Overall, all transition states have a single negative imaginary fre-
Infrared spectroscopy is a powerful tool for the identification of quency (I.F) that is between 196 - 315 cm−1, where there is a pre-
chemical compounds, particularly a recent combined experimental and ference for three-dimensional structures. For the Cu4-CO2 system, two
DFT study has been established that the CO2 molecule is activated on dissociation products are shown (P1 and P2 in Fig. S4), where oxygen is
Pt 4 anionic clusters, where DFT was the tool for identifying molecularly inserted into the cluster, causing considerable distortion in the struc-
vibration modes [67]. The infrared spectra for isolated and adsorbed ture; this is in agreement with the smallest cluster binding energy (BE)

7
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

the monometallic Cu4 and Pd4 clusters.


The infrared vibrational spectra were used in the characterization as
it could eventually serve as a bridge between the theoretical results of
this work with further experimental data. Shifting in the infrared vi-
brational spectra were observed for signals associated with the geo-
metric activation of CO2 upon interaction with metal clusters, relative
to the isolated CO2. The asymmetric CeO stretching for adsorbed CO2
has a shifting of 570 cm−1 relative to isolated CO2, where a similar
behavior was also observed for the OCO bending. For the cluster-CO2
systems there is an additional band (approximately at 1150 cm-1),
which corresponds to a symmetrical CeO stretching, which was found
only when CO2 molecule is activated on the clusters, serving as a fin-
gerprint of the activated CO2.
The activation energy for the dissociation of CO2 (CO2* → CO* +
O*) follows the trend: Cu4 < Cu3Pd < Pd4 < CuPd3 < Cu2Pd2.
Fig. 7. Activation energy (TS) for the CO2* (R) dissociation into CO*+O* (P)
Therefore, the ideal composition in terms of adsorption energy and
on bimetallic Cu-Pd and monometallic Cu4 and Pd4 systems.
activation barrier is the Cu3Pd system, which is within the experimental
Pd/(Pd + Cu) atomic ratio found in reported experimental works.
for Cu4 (see Fig. 3a). In P1 the CO* is bound to a copper atom (η1 - C), Among the evaluated clusters the Cu3Pd system has the highest
while the O* is coordinated with three copper atoms (μ3). In P2, the O* potential as catalyst. The formation of a Cu-Pd bimetallic system im-
is coordinated with two copper atoms (μ2), showing favorability for a proves the binding energy as compared to the monometallic clusters,
bridge site. For the other clusters, a single dissociation product was which leads to intermediate activation energies between Cu4 and Pd4.
observed, which corresponds to the dissociation on the closest atom of This behavior indicates that there is a synergistic effect between copper
the cluster to the CeO bond. For comparison to other systems, the and palladium atoms to improve adsorption energies and modulate
lowest value of activation energy was taken into account. activation barriers.
For Cu3Pd system, the dissociation product shows favorability for The results of this work could serve as a basis to further understand
the bridging sites, where the CO* and O* are coordinated in the μ2 the reaction mechanism of the production of methanol on the Cu3Pd
mode. For Pd4 system, the dissociation product displays a μ3 co- cluster, hence, a complete landscape of the CO2 hydrogenation behavior
ordination for CO* and η1 for O*. For the Cu3Pd and Pd4 clusters, there could be achieved in bimetallic vs monometallic systems. Finally, the
is not a geometric distortion regarding the initial structures. results of this work indicate that the Cu3Pd cluster has a catalytic po-
The activation barrier for bimetallic CuPd3 cluster is slightly above tential towards CO2 activation and dissociation, which are key steps in
monometallic Pd4 (see Fig. 7). However, the activation energy for the valorization of CO2 into methanol and other products. The findings
Cu2Pd2 is very high compared to the other bimetallic and monometallic open the doors to further explore the Cu3Pd system experimentally,
clusters. Therefore, the ideal composition to dissociate CO2e is Cu3Pd, alone or supported on several materials, which could also be compared
which also supports and complement the discussion above related with to the commercial Cu/ZnO/Al2O3 catalyst.
the higher CO2 adsorption energy on Cu3Pd cluster. Therefore, we can
conclude that there is a synergistic effect between copper and palla- Declaration of Competing Interest
dium atoms, since the formation of a bimetallic cluster of composition
Cu3Pd leads to an intermediate activation energy in-between mono- The authors declare that they have no known competing financial
metallic Cu4 and Pd4 systems interests or personal relationships that could have appeared to influ-
The activation energy for the dissociation of CO2 on the Cu4, Cu3Pd ence the work reported in this paper.
and Pd4 clusters are between 25.38 kcal/mol and 43.55 kcal/mol,
which is close but above the range reported for the Pt-Ni bimetallic Acknowledgements
system of 45.20–54 kcal/mol [29]. The higher barrier in Pd4 relative to
Cu3Pd indicates that C–O dissociation is kinetically limited in the pal- The authors acknowledge to University of Antioquia and University
ladium cluster, while the barrier on Cu4 is the lowest found. The for- of Medellinfor the financial and technical support. The computational
mation of a Cu-Pd bimetallic systems improves the binding energy as work was performed at University of Medellin. We also appreciate the
compared to the monometallic clusters (see Section 3.2) and leads to support of the QUIREMA group in this research.
intermediate activation energies between Cu4 and Pd4, which suggest
that there is a synergistic effect between copper and palladium atoms to Appendix A. Supplementary data
improve adsorption energies and modulate activation barriers. The
same behavior and trends have been observed in previous reports for Supplementary material related to this article can be found, in the
the Pt4, Pt3Ni, Pt2Ni2, PtNi3, and Ni4 [29]. online version, at doi:https://doi.org/10.1016/j.mcat.2019.110733.
The results regarding dissociation barriers open new windows to
further address a complete hydrogenation route from CO* to methanol References
on the monometallic and bimetallic systems analyzed in this work,
taking all the evaluated compositions of the Cu–Pd system. [1] A.J. Hunt, E.H.K. Sin, R. Marriott, J.H. Clark, Generation, capture, and utilization of
industrial carbon dioxide, ChemSusChem 3 (2010) 306–322.
[2] R. Baciocchi, G. Costa, D. Zingaretti, Transformation and Utilization of Carbon
4. Conclusions Dioxide, (2014).
[3] M. Pérez-fortes, J.C. Schöneberger, A. Boulamanti, E. Tzimas, Methanol synthesis
Several calculations were carried out using the functional PBE0 and using captured CO 2 as raw material : techno-economic and environmental as-
sessment, Appl. Energy 161 (2016) 718–732.
the SDD basis set. The results were divided into three sections: a) [4] H. Ren, C.-H. Xu, H.-Y. Zhao, Y.-X. Wang, J.J.-Y. Liu, J.J.-Y. Liu, Methanol synthesis
Characterization of bare monometallic and bimetallic clusters; b) from CO2 hydrogenation over Cu/γ-Al2O3 catalysts modified by ZnO, ZrO2 and
Binding of CO2 and its characterization over the most stable systems in MgO, J. Ind. Eng. Chem. 28 (2015) 261–267.
[5] K.C. Waugh, Methanol synthesis, Catal. Lett. 142 (2012) 1153–1166.
a); and c) CO2 dissociation in the most stable structure of every com- [6] G.G. Centi, S. Perathoner, Green Carbon Dioxide: Advances in CO2 Utilization,
position of the bimetallic Cu4-xPdx (x = 0–4) system and compared to

8
A. Alvarez-Garcia, et al. Molecular Catalysis xxx (xxxx) xxxx

(2014). metal cluster anions. I. Cu − n, n =1–10, J. Chem. Phys. 86 (2002) 1715–1726.


[7] K. Atsonios, K.D. Panopoulos, E. Kakaras, Investigation of technical and economic [38] W.C.L.J. Ho, K.M. Ervin, Photoelectron spectroscopy of metal cluster anions: Cun,
aspects for methanol production through CO2 hydrogenation, Int. J. Hydrogen Agn, and Aun, J. Chem. Phys. 93 (1990) 6987–7002.
Energy 41 (2016) 2202–2214. [39] A.M. James, G.W. Lemire, P.R. Langridge-Smith, Threshold photoionisation spec-
[8] A. Boretti, Renewable hydrogen to recycle CO2 to methanol, Int. J. Hydrogen troscopy of the CuAg molecule, Chem. Phys. Lett. 227 (1994) 503–510.
Energy 38 (2013) 1806–1812. [40] J. Ho, K.M. Ervin, M.L. Polak, M.K. Gilles, W.C. Lineberger, A study of the electronic
[9] A. Dwivedi, R. Gudi, P. Biswas, An improved tri-reforming based methanol pro- structures of Pd−2 and Pd2 by photoelectron spectroscopy, J. Chem. Phys. 95
duction process for enhanced CO2valorization, Int. J. Hydrogen Energy 42 (2017) (1991) 4845.
23227–23241. [41] A.E. Reed, R.B. Weinstock, F. Weinhold, Natural population analysis, J. Chem. Phys.
[10] M.S. Shaharun, M.A. Alotaibi, A.I. Alharthi, Recent developments on heterogeneous 83 (1985) 735–746.
catalytic CO 2 reduction to methanol, J. CO2 Util. 34 (2019) 20–33. [42] K.B. Wiberg, Application of the pople-santry-segal CNDO method to the cyclopro-
[11] G. Olah, A. Goeppert, S. Prakash, Beyong Oil and Gas: The Methanol Economy, pylcarbinyl and cyclobutyl cation and to bicyclobutane, Tetrahedron. 24 (1968)
Willey-VCH, Germany, 2009. 1083–1096.
[12] M. Behrens, F. Studt, I. Kasatkin, S. Kühl, M. Hävecker, F. Abild-pedersen, [43] O.V. Sizova, L.V. Skripnikov, A.Y. Sokolov, Symmetry decomposition of quantum
S. Zander, F. Girgsdies, P. Kurr, B. Kniep, M. Tovar, R.W. Fischer, J.K. Nørskov, chemical bond orders, J. Mol. Struct. THEOCHEM. 870 (2008) 1–9.
R. Schlögl, The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial [44] C. Peng, H.B. Schlegel, Combining STQN methods to find transition states, Isr. J.
catalysts, Science (80-.) 336 (2012) 893–898. Chem. 33 (1993) 449–454.
[13] X.M. Liu, G.Q. Lu, Z.F. Yan, J. Beltramini, Recent advances in catalysts for methanol [45] E. Florez, W. Tiznado, F. Mondragón, P. Fuentealba, Theoretical study of the in-
synthesis via hydrogenation of CO and CO2, Ind. Eng. Chem. Res. 42 (2003) teraction of molecular oxygen with copper clusters, J. Phys. Chem. A 109 (2005)
6518–6530, https://doi.org/10.1021/ie020979s. 7815–7821.
[14] J. Sinfelt, Bimetallic Catalysts Discoveries, Concepts, and Applications, Wiley, [46] E. Fernandez, M. Boronat, A. Corma, Trends in the reactivity of molecular O2 with
United state, 1983. copper clusters: influence of size and shape, J. Phys. Chem. C 119 (2015)
[15] W. Yu, M.D. Porosoff, J.G. Chen, Review of Pt-based bimetallic catalysis: from 19832–19846.
model surfaces to supported catalysts, Chem. Rev. 112 (2012) 5780–5817. [47] J.C. González-Torres, V. Bertin, E. Poulain, O. Olvera-Neria, The CO oxidation
[16] E. Jeong, Y. Hee, D. Lee, D. Moon, K. Lee, Hydrogenation of CO 2 to methanol over mechanism on small Pd clusters. A theoretical study, J. Mol. Model. 21 (2015).
Pd – Cu/CeO2 catalysts, Mol. Catal. 434 (2017) 146–153. [48] L. Ling, L. Fan, X. Feng, B. Wang, R. Zhang, Effects of the size and Cu modulation of
[17] V. Deerattrakul, P. Dittanet, M. Sawangphruk, P. Kongkachuichay, CO2 hydro- Pdn(n ⩽ 38) clusters on Hg0adsorption, Chem. Eng. J. 308 (2017) 289–298.
genation to methanol using Cu-Zn catalyst supported on reduced graphene oxide [49] S. Goel, A.E. Mansunov, Density functional theory of small nickel clusters, J. Mol.
nanosheets, J. CO2 Util. 16 (2016) 104–113. Model. 18 (2012) 783–790.
[18] X. Jiang, N. Koizumi, X. Guo, C. Song, Bimetallic Pd-Cu catalysts for selective CO2 [50] T.T. Jia, C.H. Lu, K.N. Ding, Y.F. Zhang, W.K. Chen, Oxidation of Pdn(n=1-5)
hydrogenation to methanol, Appl. Catal. B Environ. 170–171 (2015) 173–185. clusters on single vacancy graphene: a first-principles study, Comput. Theor. Chem.
[19] Y. Liu, D. Liu, Study of bimetallic Cu-Ni/-Al2O3 catalysts for carbon dioxide hy- 1020 (2013) 91–99.
drogenation, Int. J. Hydrogen Energy 24 (1999) 351–354. [51] G.V. Gibbs, R.T. Downs, C.T. Prewitt, K.M. Rosso, N.L. Ross, D.F. Cox, Electron
[20] O. Klaja, J. Szczygieł, J. Trawczyński, B.M. Szyja, The CO2 dissociation mechanism density distributions calculated for the nickel sulfides millerite, vaesite, and hea-
on the small copper clusters—the influence of geometry, Theor. Chem. Acc. 136 zlewoodite and nickel metal: a case for the importance of Ni-Ni bond paths for
(2017) 1–9. electron transport, J. Phys. Chem. B 109 (2005) 21788–21795.
[21] C. Liu, B. Yang, E. Tyo, S. Seifert, J. DeBartolo, B. von Issendorff, P. Zapol, S. Vajda, [52] L. Pauling, Atomic radii and interatomic distances in metals, J. Am. Chem. Soc. 69
L.A. Curtiss, Carbon dioxide conversion to methanol over size-selected Cu 4 clusters (1947) 542–553.
at low pressures, J. Am. Chem. Soc. 137 (2015) 8676–8679. [53] R. Jayaprakash, J. Shanker, Correlation between electronegativity and high tem-
[22] B. Yang, C. Liu, A. Halder, E.C. Tyo, A.B.F. Martinson, S. Seifert, P. Zapol, perature superconductivity, J. Phys. Chem. Solids 54 (1993) 365–369.
L.A. Curtiss, S. Vajda, Copper cluster size effect in methanol synthesis from CO2, J. [54] M. Kabir, A. Mookerjee, A.K. Bhattacharya, Structure and stability of copper clus-
Phys. Chem. C 121 (2017) 10406–10412. ters: a tight-binding molecular dynamics study, Phys. Rev. A - At. Mol. Opt. Phys. 69
[23] H. Tao, Y. Li, X. Cai, H. Zhou, Y. Li, W. Lin, S. Huang, K. Ding, W. Chen, Y. Zhang, (2004) 1–10.
What is the best size of subnanometer copper clusters for CO 2 conversion to me- [55] D. Cortés-Arriagada, M.P. Oyarzún, L. Sanhueza, A. Toro-Labbé, Binding of trivalent
thanol at Cu/TiO 2 interfaces? A density functional theory study, J. Phys. Chem. C arsenic onto the tetrahedral Au20 and Au19Pt clusters: implications in adsorption
123 (2019) 24118–24132. and sensing, J. Phys. Chem. A 119 (2015) 6909–6918.
[24] J.A. Rodriguez, J. Evans, L. Feria, A.B. Vidal, P. Liu, K. Nakamura, F. Illas, CO2 [56] G. Li, X. Chen, Z. Zhou, F. Wang, H. Yang, J. Yang, B. Xu, B. Yang, D. Liu,
hydrogenation on Au/TiC, Cu/TiC, and Ni/TiC catalysts: production of CO, me- Theoretical insights into the structural, relative stable, electronic, and gas sensing
thanol, and methane, J. Catal. 307 (2013) 162–169. properties of Pb:NAun (n = 2-12) clusters: a DFT study, RSC Adv. 7 (2017)
[25] F. Maleki, P. Schlexer, G. Pacchioni, Support effects and reaction mechanism of 45432–45441.
acetylene trimerization over silica-supported Cu4 clusters: a DFT study, Surf. Sci. [57] C. Luo, E. Al, First principles study of small palladium cluster growth and iso-
668 (2018) 125–133. merization, Int. J. Quantum Chem. 107 (2007) 1632–1641.
[26] F. Mehmood, J. Greeley, P. Zapol, L.A. Curtiss, Comparative density functional [58] K. Jug, B. Zimmermann, P. Calaminici, A.M. Köster, Structure and stability of small
study of methanol descomposition on Cu4 and Co4, J. Phys. Chem. B 114 (2010) copper clusters, J. Chem. Phys. 116 (2002) 4497–4507.
14458–14466. [59] G. Preda, G. Pacchioni, M. Chiesa, E. Giamello, Formation of CO2− Radical Anions
[27] S. Kansara, S.K. Gupta, Y. Sonvane, Catalytic activity of Cu4-cluster to adsorb H2S from CO2 Adsorption on an Electron-Rich MgO Surface a Combined Ab Initio and
gas: H -BN nanosheet, AIP Conf. Proc. 1961 (2018). Pulse EPR Study, (2008), pp. 19568–19576.
[28] M. Reina, A. Martínez, Silybin interacting with Cu4, Ag4 and Au4 clusters: do these [60] A. Yanagimachi, K. Koyasu, D.Y. Valdivielso, S. Gewinner, W. Schöllkopf,
constitute antioxidant materials? Comput. Theor. Chem. 1112 (2017) 1–9. A. Fielicke, T. Tsukuda, Size-specific, dissociative activation of carbon dioxide by
[29] J. Niu, J. Ran, Z. Ou, X. Du, R. Wang, W. Qi, CO2 dissociation over PtxNi4-x bi- cobalt cluster anions, J. Phys. Chem. C 120 (2016) 14209–14215.
metallic clusters with and without hydrogen sources: a density functional theory [61] S. Gautam, K. Dharamvir, N. Goel, CO2 adsorption and activation over medium
study, J. CO2 Util. 16 (2016) 431–441. sized Cun (n=7, 13 and 19) clusters: a density functional study, Comput. Theor.
[30] L.E. Gálvez-González, J.O. Juárez-Sánchez, R. Pacheco-Contreras, I.L. Garzón, Chem. 1009 (2013) 8–16.
L.O. Paz-Borbón, A. Posada-Amarillas, CO 2 adsorption on gas-phase Cu 4−x Pt x (x [62] A.G. Saputro, M.K. Agusta, T.D.K. Wungu, F.R. Suprijadi, H.K. Dipojono, DFT study
= 0–4) clusters: a DFT study, Phys. Chem. Chem. Phys. 20 (2018) 17071–17080. of adsorption of CO2 on palladium cluster doped by transition metal, J. Phys. Conf.
[31] C. Adamo, V. Barone, Toward reliable density functional methods without ad- Ser. 739 (2016).
justable parameters: the PBE0 model, J. Chem. Phys. 110 (1999) 6158–6170. [63] A. Álvarez, M. Borges, J.J. Corral-Pérez, J.G. Olcina, L. Hu, D. Cornu, R. Huang,
[32] P. Fuentealba, H. Preuss, H. Stoll, L. Von Szentpaly, A proper account of core-po- D. Stoian, A. Urakawa, CO2Activation over catalytic surfaces, ChemPhysChem. 18
larization with pseudopotentials: single valence-electron alkali compounds, Chem. (2017) 3135–3141.
Phys. Lett. 89 (1982) 418–422. [64] S.L. Han, X. Xue, X.C. Nie, H. Zhai, F. Wang, Q. Sun, Y. Jia, S.F. Li, Z.X. Guo, First-
[33] X. Cao, M. Dolg, Segmented contraction scheme for small-core actinide pseudopo- principles calculations on the role of ni-doping in Cun clusters: from geometric and
tential basis sets, J. Mol. Struct. Theochem. 581 (2002) 139–147. electronic structures to chemical activities towards CO2, Phys. Lett. Sect. A Gen. At.
[34] P.L. Rodríguez-Kessler, S. Pan, E. Florez, J.L. Cabellos, G. Merino, Structural evo- Solid State Phys. 374 (2010) 4324–4330.
lution of the rhodium-doped silver clusters AgnRh (n ≤ 15) and their reactivity [65] S. Kattel, P.J. Ramírez, J.G. Chen, J.A. Rodriguez, P. Liu, Active sites for CO2 hy-
toward NO, J. Phys. Chem. C 121 (2017) 19420–19427. drogenation to methanol on Cu/ZnO catalysts, Science (80-.) 355 (2017)
[35] P.L. Rodríguez-Kessler, F. Murillo, A.R. Rodríguez-Domínguez, P. Navarro-Santos, 1296–1299.
G. Merino, Structure of V-doped Pd n (n = 2–12) clusters and their ability for H 2 [66] Y. Li, S.H. Chan, Q. Sun, Heterogeneous catalytic conversion of CO2: a compre-
dissociation, Int. J. Hydrogen Energy 43 (2018) 20636–20644. hensive theoretical review, Nanoscale 7 (2015) 8663–8683.
[36] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, [67] S. Mackenzie, A. Green, A. Fielicke, A.S. Gentleman, J. Justen, W. Schoellkopf, IR
G. Scalmani, V. Barone, G.A. Petersson, H. Nakatsuji, M. Li, et al., Gaussian 09. signature of size-selective CO2 activation on small platinum cluster anions, ptn- (n
Revision A.02, Wallingford CT, (2016). = 4-7), Angew. Chem. Int. Ed. (2018) 4–6.
[37] D.G. Leopold, J. Ho, W.C. Lineberger, Photoelectron spectroscopy of mass‐selected

You might also like